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

Polystability of Stokes representations and differential Galois groups

Philip Boalch and Daisuke Yamakawa
(Date: January 22, 2023)
Abstract.

Polystability of (twisted) Stokes representations (i.e. wild monodromy representations) will be characterised, in terms of the corresponding differential Galois group (generalising the Zariski closure of the monodromy group in the tame case). This extends some results of Richardson. Further, the intrinsic approach to such results will be established, in terms of reductions of Stokes local systems.

1. Introduction

We continue our investigations of the nonabelian moduli spaces in 2d gauge theory (the theory of connections on curves). This article is essentially an appendix to the paper [13] that completed the construction of the wild character varieties of smooth curves as affine algebraic Poisson varieties (completing the sequence [5, 6, 8]). This is a purely algebro-geometric approach, complementary to the earlier analytic approaches [3, 1]. Here we will give an intrinsic characterisation of the points of the wild character varieties, generalising existing results in the tame case, and characterise the stable points (generalising a result of [8] in the untwisted wild case). For more background and applications see the reviews in [7, 9, 11] (the first large class of examples of wild character varieties is due to Birkhoff [2] and the simplest case underlies Uq(𝔤)U_{q}(\mathfrak{g}), [7] §4).

First we will recall the basic statements in the tame case. Let GG be a connected complex reductive group, such as GLn(){\mathop{\rm GL}}_{n}(\mathbb{C}), and let Σ\Sigma^{\circ} be a smooth complex algebraic curve. Thus Σ=Σα\Sigma^{\circ}=\Sigma\setminus\alpha for some smooth compact complex algebraic curve Σ\Sigma (i.e. a compact Riemann surface), and a finite subset αΣ\alpha\subset\Sigma. Given a basepoint bΣb\in\Sigma^{\circ} one can consider the representation variety

=Hom(π1(Σ,b),G).\mathcal{R}=\operatorname{Hom}(\pi_{1}(\Sigma^{\circ},b),G).

which is a complex affine variety equipped with an action of GG, conjugating representations. A point ρ\rho\in\mathcal{R} is a GG-representation, i.e. a group homomorphism ρ:π1(Σ,b)G\rho:\pi_{1}(\Sigma^{\circ},b)\to G. In turn the GG-character variety, or Betti moduli space

B=B(Σ,G)=/G=Hom(π1(Σ,b),G)/G\mathcal{M}_{\text{\rm B}}=\mathcal{M}_{\text{\rm B}}(\Sigma^{\circ},G)=\mathcal{R}/G=\operatorname{Hom}(\pi_{1}(\Sigma^{\circ},b),G)/G

is the (affine) geometric invariant theory quotient of \mathcal{R} by GG. By definition this means that the points of the variety B\mathcal{M}_{\text{\rm B}} are the closed GG orbits in \mathcal{R}. The representations ρ\rho whose GG orbits are closed are called the polystable representations. A basic question is thus to characterise the polystable representations intrinsically. The answer (due to Richardson, building on earlier work in the general linear case), is as follows.

Let Gal(ρ)=A(ρ)G\mathop{\rm Gal}(\rho)=A(\rho)\subset G be the Zariski closure (adhérence) of the image of ρ\rho. A theorem of Schlesinger implies that if ρ\rho is the monodromy representation of an algebraic connection (,E)(\nabla,E) on an algebraic principal GG-bundle EΣE\to\Sigma^{\circ}, with regular singularities at α\alpha, then Gal(ρ)\mathop{\rm Gal}(\rho) is the differential Galois group of (,E)(\nabla,E). Recall that a complex affine algebraic group is a linearly reductive group if its identity component is reductive (i.e. has trivial unipotent radical). The basic characterisation is then:

Theorem 1 ([26]).

A point ρ\rho\in\mathcal{R} is polystable if and only if Gal(ρ)\mathop{\rm Gal}(\rho) is a linearly reductive group. Further, ρ\rho is stable if and only if Gal(ρ)\mathop{\rm Gal}(\rho) is not contained in a proper parabolic subgroup of GG.

A representation ρ\rho\in\mathcal{R} whose Galois group Gal(ρ)\mathop{\rm Gal}(\rho) is linearly reductive is often called a semisimple representation111Indeed if VnV\cong\mathbb{C}^{n} and G=GL(V)GLn()G={\mathop{\rm GL}}(V)\cong{\mathop{\rm GL}}_{n}(\mathbb{C}) then ρ:π1(Σ,b)G\rho:\pi_{1}(\Sigma^{\circ},b)\to G is a semisimple representation if and only if VV is a semisimple π1(Σ,b)\pi_{1}(\Sigma^{\circ},b)-module.. Thus the theorem says that ρ\rho\in\mathcal{R} is polystable if and only if ρ\rho is semisimple, and so the points of B\mathcal{M}_{\text{\rm B}} are the isomorphism classes of semisimple representations. Note that for G=GLn()G={\mathop{\rm GL}}_{n}(\mathbb{C}) this statement is older (Artin/Procesi) and a full account appears in the book of Lubotzky–Magid [23].

This paper is concerned with the extension of this result to the case of Stokes representations, generalising the fundamental group representations, and the characterisation of the points of the wild character varieties, generalising the (tame) character varieties appearing above.

In brief any algebraic connection (,E)(\nabla,E) on an algebraic principal GG-bundle EΣE\to\Sigma^{\circ}, has an invariant, its irregular class, at each point aαa\in\alpha. A connection is regular singular if and only if its irregular class is trivial. Our aim is to give the generalisation of Richardson’s results when the irregular classes are arbitrary. Due to work of many people it is known how to describe algebraic connections completely topologically, so we can work algebraically on the Betti/Stokes side (see the review of this story in [11]). The basic notions from the tame case are generalised as follows:

local system Stokes local system\displaystyle\rightsquigarrow\ \text{Stokes local system}
fundamental group (or groupoid) wild surface group (or groupoid)\displaystyle\rightsquigarrow\ \text{wild surface group (or groupoid)}
fundamental group representation Stokes representation.\displaystyle\rightsquigarrow\ \text{Stokes representation.}

Once these generalisations are understood then the story proceeds similarly to the tame case (defining a wild representation variety \mathcal{R} parameterising framed Stokes local systems, and then acting by a reductive group to forget the framings). A key novelty in the wild setting is that there is a breaking of structure group (“fission”) near the marked points so the group acting involves a (reductive) subgroup of GG. This is intimately related to extra generators in Ramis’ description of the differential Galois group [25], generalising Schlesinger’s density theorem. However, as we will recall, the basic feature of the tame case remains, that \mathcal{R} can be written explicitly in terms of a product of simpler pieces (doubles 𝔻\mathbb{D} or fission spaces 𝒜\mathcal{A}), one for each marked point or handle on Σ\Sigma:

(𝔻g𝒜1𝒜m)//G\mathcal{R}\ \cong\ (\mathbb{D}^{\circledast{}g}\circledast{}\mathcal{A}_{1}\circledast\cdots\circledast\mathcal{A}_{m})/\!\!/G

if Σ\Sigma has genus gg and mm marked points.

1.1. Summary of main results

In this section we will summarise the main results in a short uncluttered form, with links to the references to the full definitions (mainly in [13]). (An overview of the simpler GLn(){\mathop{\rm GL}}_{n}(\mathbb{C}) set-up is in [11] §13.)

Let Σ\Sigma be a smooth compact complex algebraic curve (i.e. a compact Riemann surface) and let αΣ\alpha\subset\Sigma be a finite non-empty subset. Let GG be a connected complex reductive group.

Choose a GG-irregular class Θa\Theta_{a} at each point aαa\in\alpha (possibly twisted) as in [13] §3.5. Let 𝚺=(Σ,α,Θ){\bf\Sigma}=(\Sigma,\alpha,\Theta) be the resulting wild Riemann surface with structure group GG as in [13] §4 (and [8] §8.1, Rmk 10.6).

The topological notion of Stokes local system 𝕃\mathbb{L} on 𝚺{\bf\Sigma} is then well-defined, as in [13] Defn. 13 (and [8] Rmk. 8.4, [9] §4.3, [11] §8).

Although we will not need it here, note that the Stokes local systems on 𝚺{\bf\Sigma} encode the algebraic connections on algebraic principal GG-bundles on the open curve Σ:=Σα\Sigma^{\circ}:=\Sigma\setminus\alpha, with irregular class Θa\Theta_{a} at each aαa\in\alpha.

As in [13] §5, [8] §8, it follows that if we choose suitable basepoints β\beta then the wild surface groupoid Π\Pi is well defined, as is the wild representation variety (the space of Stokes representations of Π\Pi):

(𝚺,β)=Hom𝕊(Π,G)Hom(Π,G).\mathcal{R}({\bf\Sigma},\beta)=\operatorname{Hom}_{\mathbb{S}}(\Pi,G)\subset\operatorname{Hom}(\Pi,G).

It is an affine variety equipped with an action of a complex reductive group 𝐇{\bf H}.

Any framed Stokes GG-local system 𝕃\mathbb{L} determines a Stokes representation ρ=ρ𝕃(𝚺,β)\rho=\rho_{\mathbb{L}}\in\mathcal{R}({\bf\Sigma},\beta). Two Stokes local systems are isomorphic if and only if their Stokes representations are in the same 𝐇{\bf H}-orbit in (𝚺,β)\mathcal{R}({\bf\Sigma},\beta).

The notion of ρ\rho being polystable or stable for the action of 𝐇{\bf H} on \mathcal{R} is well-defined, as for any action of a reductive group on an affine variety (as in[26]). The points of the Poisson wild character variety B(𝚺)\mathcal{M}_{\text{\rm B}}({\bf\Sigma}) are the polystable (i.e. closed) 𝐇{\bf H}-orbits.

On the other hand ρ\rho determines the Galois group Gal(ρ)G\mathop{\rm Gal}(\rho)\subset G, as in Ramis’ density theorem [25] (see §5.2 below). It involves not just the usual monodromy, but also the formal monodromy, Stokes automorphisms and the Ramis tori.

Finally we can define 𝕃\mathbb{L} to be irreducible if it has no proper parabolic reductions, and to be reductive (or “semsimple”) if it has an irreducible Levi reduction.

Theorem 2.

Let 𝕃\mathbb{L} be a Stokes local system on 𝚺{\bf\Sigma}, and let ρ\rho be its Stokes representation. The following are equivalent:

1)\bullet 1) ρ\rho is polystable,

2)\bullet 2) Gal(ρ)\mathop{\rm Gal}(\rho) is linearly reductive,

3)\bullet 3) 𝕃\mathbb{L} is a semisimple Stokes local system.

In the tame case (with each irregular class trivial) the groupoid Π\Pi becomes Poincaré’s fundamental groupoid (with a finite number of basepoints), and the theorem is already known [26].

Further we will consider stability (not just polystability). This requires possibly adding a few extra punctures to control the kernel of the action, but no generality is lost (see the discussion after (5)).

Theorem 3.

Let 𝕃\mathbb{L} be a Stokes local system on 𝚺{\bf\Sigma}, and let ρ\rho be its Stokes representation. The following are equivalent:

4)\bullet 4) ρ\rho is a stable point of (𝚺,β)\mathcal{R}({\bf\Sigma},\beta),

5)\bullet 5) Gal(ρ)\mathop{\rm Gal}(\rho) is not contained in a proper parabolic subgroup of GG,

6)\bullet 6) 𝕃\mathbb{L} is irreducible.

This result was already established in [8] in the case where each irregular class was not twisted.

Recall from [13] that two types of twist are possible: the formal twists (twisted irregular classes, as above), and also interior twists, over the interior of the curve, where we start with a local system of groups 𝒢Σ\mathcal{G}\to\Sigma^{\circ}, with each fibre isomorphic to GG. Similarly we will establish the analogous results in this fully twisted setting. We will suppose that 𝒢\mathcal{G} is “out-finite” in the sense of §5.1 below.

The first main difference (in the presence of interior twists) is that the wild representation variety is replaced by a space of twisted Stokes representations:

=THom𝕊(Π,G)Hom(Π,GAut(G))\mathcal{R}=\operatorname{THom}_{\mathbb{S}}(\Pi,G)\subset\operatorname{Hom}(\Pi,G\ltimes\operatorname{\mathop{\rm Aut}}(G))

which is an affine variety equipped with an action of a complex reductive group 𝐇{\bf H}. Secondly Gal(ρ)\mathop{\rm Gal}(\rho) is not a subgroup of GG, but rather it comes with a homomorphism Gal(ρ)Aut(G)\mathop{\rm Gal}(\rho)\to\operatorname{\mathop{\rm Aut}}(G), so naturally acts on GG by automorphisms. The image of Gal(ρ)\mathop{\rm Gal}(\rho) in Aut(G)\operatorname{\mathop{\rm Aut}}(G) will be denoted Gal¯(ρ)\overline{\mathop{\rm Gal}}(\rho). Then we can define irreducibility (§5.3) and semisimplicity (§5.4) for Stokes 𝒢\mathcal{G}-local systems, and will prove analogues of the above results:

Theorem 4.

Let 𝚺=(Σ,α,Θ){\bf\Sigma}=(\Sigma,\alpha,\Theta) be a wild Riemann surface with group 𝒢Σ\mathcal{G}\to\Sigma^{\circ}. Let 𝕃\mathbb{L} be a Stokes 𝒢\mathcal{G}-local system on 𝚺{\bf\Sigma}, and let ρTHom𝕊(Π,G)\rho\in\operatorname{THom}_{\mathbb{S}}(\Pi,G) be its twisted Stokes representation. The following three conditions are equivalent:

1)\bullet 1^{\prime}) ρ\rho is a polystable point of (𝚺,β)=THom𝕊(Π,G)\mathcal{R}({\bf\Sigma},\beta)=\operatorname{THom}_{\mathbb{S}}(\Pi,G),

2)\bullet 2^{\prime}) Gal¯(ρ)\overline{\mathop{\rm Gal}}(\rho) is linearly reductive,

3)\bullet 3^{\prime}) 𝕃\mathbb{L} is a semisimple Stokes 𝒢\mathcal{G}-local system.

Moreover the following three conditions are also equivalent:

4)\bullet 4^{\prime}) ρ\rho is a stable point of (𝚺,β)=THom𝕊(Π,G)\mathcal{R}({\bf\Sigma},\beta)=\operatorname{THom}_{\mathbb{S}}(\Pi,G),

5)\bullet 5^{\prime}) Gal¯(ρ)Aut(G)\overline{\mathop{\rm Gal}}(\rho)\subset\operatorname{\mathop{\rm Aut}}(G) does not preserve a proper parabolic subgroup of GG,

6)\bullet 6^{\prime}) 𝕃\mathbb{L} is an irreducible Stokes 𝒢\mathcal{G}-local system.

In the set-up of Theorems 2,3 with 𝒢\mathcal{G} constant it is true that a)a)a^{\prime})\Leftrightarrow a) for all a=1,2,6a=1,2,\ldots 6. Thus Thm. 4 implies both Theorems 2,3.

1.2. Layout of the article

Sections 2, 3 generalise some of Richardson’s results, in two steps. These are extrinsic results, and may well have further applications, beyond the wild character varieties of curves.

Section 4 then applies these results to the spaces of Stokes representations leading to the wild character varieties B(𝚺)=/𝐇\mathcal{M}_{\text{\rm B}}({\bf\Sigma})=\mathcal{R}/{\bf H}. The main results are the equivalences 1)2)1^{\prime})\Leftrightarrow 2^{\prime}) in Cor. 25, and 4)5)4^{\prime})\Leftrightarrow 5^{\prime}) in Cor. 28.

Section 5 then discusses the intrinsic objects, Stokes local systems, and how stability/polystability can be read off in terms of (twisted) reductions of structure group. The main results are the equivalences 1)3)1^{\prime})\Leftrightarrow 3^{\prime}) and 4)6)4^{\prime})\Leftrightarrow 6^{\prime}) in parts 2) and 1) of Thm. 29 respectively.

1.3. Some other directions

Note that for constant general linear groups the irreducible Stokes local systems are equivalent to the input data in the construction of wild harmonic metrics in [27]. Further note that in the fully untwisted case (but any GG, as in [8]) the irreducible Stokes local systems fit into the “Betti weight zero” case of the recent extension of the wild nonabelian Hodge correspondence due to Huang–Sun [20] (which looks to be in line with our general conjecture in [10] Rmk. 6, that the “good” meromorphic connections/Higgs fields on parahoric torsors are the right objects to look at).

In another direction one of the key motivations for this work is the fact that any admissible deformation of a wild Riemann surface leads to a local system of wild character varieties, and its monodromy generalises the usual mapping class group actions on character varieties in the tame case. As explained in [3, 4] the motivating examples for this whole line of thought were the Dubrovin–Ugaglia Poisson varieties whose braid group actions come from the braiding of counts of BPS states ([19] Rmk 3.10, related to earlier work of Cecotti–Vafa); these are examples of twisted wild character varieties, involving the non-trivial outer automorphism of GLn(){\mathop{\rm GL}}_{n}(\mathbb{C}) (so we now have an intrinsic general framework encompassing such examples). Indeed it was by forgetting this twist that the Poisson variety GG^{*} underlying Uq(𝔤)U_{q}(\mathfrak{g}) was recognised and identified as a wild character variety [4]. See e.g. [17, 18, 12] for some recent developments concerning the generalised braid groups that act on wild character varieties, from admissible deformations of more general wild Riemann surfaces.

Acknowledgements. These results were completed in 2019 before the first named author moved departments and then learnt of the thesis work leading to [28], that has some overlap with this paper in the tame setting with interior twists, although expressed in a slightly different language. In the intervening years we have not yet managed to incorporate possible simplifications suggested by [28] but thought it reasonable to release our original approach anyway since the scope is larger.

2. Twisted version of Richardson’s results

As in [26] we use the terminology that an affine algebraic group over \mathbb{C} is reductive if it is connected and has trivial unipotent radical. It is linearly reductive if its identity component is reductive.

Let GG be a linearly reductive group over \mathbb{C}. Recall that a point xx of an affine GG-variety is said to be polystable if the orbit GxG\cdot x is closed. It is said to be stable if it is polystable and the kernel of the action has finite index in its stabiliser GxG_{x}.

Let nn be a positive integer. For 𝐱=(xi)i=1nGn{\bf x}=(x_{i})_{i=1}^{n}\in G^{n}, let A(𝐱)GA({\bf x})\subset G be the Zariski closure of the subgroup generated by x1,x2,,xnx_{1},x_{2},\dots,x_{n}. In [26], Richardson examined the simultaneous conjugation action of GG on the product GnG^{n} and obtained the following results:

Theorem 5 ([26, Thm. 3.6]).

If GG is linearly reductive, a point 𝐱Gn{\bf x}\in G^{n} is polystable if and only if A(𝐱)A({\bf x}) is linearly reductive.

Theorem 6 ([26, Thm. 4.1]).

If GG is reductive, a point 𝐱Gn{\bf x}\in G^{n} is stable if and only if A(𝐱)A({\bf x}) is not contained in any proper parabolic subgroup of GG.

In this section we will establish twisted versions of these results, in Thm. 7 and Thm. 12 (2) respectively. Two other characterisations of polystability will also be established, in Thm. 12 (1) and Cor. 18. The subsequent section (§3) will give a further generalisation.

Assume that GG is reductive and choose ϕ1,ϕ2,,ϕnAut(G)\phi_{1},\phi_{2},\dots,\phi_{n}\in\operatorname{\mathop{\rm Aut}}(G). Let Γ\Gamma be the subgroup of Aut(G)\operatorname{\mathop{\rm Aut}}(G) generated by the inner automorphism group Inn(G)\operatorname{Inn}(G) and ϕ1,ϕ2,,ϕn\phi_{1},\phi_{2},\dots,\phi_{n}. We assume that the quotient Γ¯:=Γ/Inn(G)\overline{\Gamma}:=\Gamma/\operatorname{Inn}(G) is finite and regard Γ\Gamma as an algebraic group with identity component Inn(G)G/Z(G)\operatorname{Inn}(G)\cong G/Z(G). For any ϕΓ\phi\in\Gamma let G(ϕ)G(\phi) denote the GG-bitorsor/“twisted group” G×{ϕ}GΓG\times\{\phi\}\subset G\ltimes\Gamma, as in [13] §2 (as explained there, the monodromy of a 𝒢\mathcal{G}-local system lies in such a space, and G(ϕ)G(\phi) embeds in the group of set-theoretic automorphisms of a fibre). Put

X=i=1nG(ϕi)(GΓ)n,X=\prod_{i=1}^{n}G(\phi_{i})\subset(G\ltimes\Gamma)^{n},

on which G=G(Id)G=G(\text{\rm Id}) acts by the simultaneous conjugation. For 𝐱=(xi)i=1nX{\bf x}=(x_{i})_{i=1}^{n}\in X, let A(𝐱)A({\bf x}) be the Zariski closure (in the algebraic group GΓG\ltimes\Gamma) of the subgroup generated by x1,x2,,xnx_{1},x_{2},\dots,x_{n}. Let A¯(𝐱)Γ\overline{A}({\bf x})\subset\Gamma be the image of A(𝐱)A({\bf x}) under the homomorphism

Ad:GΓΓ,x=(g,ϕ)Ad(x)=Ad(g)ϕ,{\mathop{\rm Ad}}\colon G\ltimes\Gamma\to\Gamma,\quad x=(g,\phi)\mapsto{\mathop{\rm Ad}}(x)={\mathop{\rm Ad}}(g)\circ\phi,

where Ad(g)=g()g1{\mathop{\rm Ad}}(g)=g(\,\cdot\,)g^{-1}. Note that A¯(𝐱)\overline{A}({\bf x}) is contained in Aut(G)\operatorname{\mathop{\rm Aut}}(G) and hence naturally acts on GG. Thus in turn, via Ad{\mathop{\rm Ad}}, the group A(𝐱)A({\bf x}) naturally acts on GG.

Theorem 7.

A point 𝐱X{\bf x}\in X is polystable if and only if A¯(𝐱)\overline{A}({\bf x}) is linearly reductive.

Remark 8.

The notation A(𝐱)A({\bf x}) for the Zariski closure presumably stems from [14] §3, where AA stands for adhérence. This may help to avoid possible confusion (since A¯(𝐱)\overline{A}({\bf x}) denotes the image in Aut(G)\operatorname{\mathop{\rm Aut}}(G) here). Note that [14] §3.3 shows that if HGH\subset G is any subgroup then the Zariski closure of the set HH is a Zariski closed subgroup of GG.

To see that Thm. 7 generalises Thm. 5, first note that:

Proposition 9.

Let G~\widetilde{G} be a linearly reductive group with identity component GG, and let AA be a closed subgroup of G~\widetilde{G}. Let A¯\overline{A} be the image of AA under the map Ad|G:G~Aut(G){\mathop{\rm Ad}}\bigl{|}_{G}:\widetilde{G}\to\operatorname{\mathop{\rm Aut}}(G), and regard it as a quotient algebraic group of AA. Then AA is linearly reductive if and only if A¯\overline{A} is linearly reductive.

Proof. Note that we may view Ad|G{\mathop{\rm Ad}}\bigl{|}_{G} as a map of algebraic groups: Aut(G)Inn(G)Out(G)\operatorname{\mathop{\rm Aut}}(G)\cong\operatorname{Inn}(G)\ltimes\mathop{\rm Out}(G) may have an infinite number of components, but Ad|G(G~){\mathop{\rm Ad}}\bigl{|}_{G}(\widetilde{G}) will only encounter a finite number of them. Let KK be the kernel of the homomorphism AA¯A\twoheadrightarrow\overline{A}. Note KGZ(G)K\cap G\subset Z(G), and the identity component K0K^{0} of KK is contained in (G~)0=G(\widetilde{G})^{0}=G. Thus K0Z(G)K^{0}\subset Z(G), which implies K0K^{0} is a torus and so KK is linearly reductive. Now [26] 1.2.2 implies AA is linearly reductive if and only if A¯\overline{A} is linearly reductive. \square

Lemma 10.

Let G~\widetilde{G} be a linearly reductive group with identity component GG and let 𝐱{\bf x} be a point in an affine G~\widetilde{G}-variety YY. Then the following hold.

  1. (1)

    𝐱{\bf x} is polystable for the G~\widetilde{G}-action if and only if 𝐱{\bf x} is polystable for the GG-action.

  2. (2)

    𝐱{\bf x} is stable for the G~\widetilde{G}-action if and only if 𝐱{\bf x} is stable for the GG-action.

Proof. The G~\widetilde{G}-orbit G~𝐱\widetilde{G}\cdot{\bf x} is a disjoint union of a finite number of GG-orbits of the same dimension, one of which is G𝐱G\cdot{\bf x}. Hence G~𝐱\widetilde{G}\cdot{\bf x} is closed if and only if G𝐱G\cdot{\bf x} is closed. The second assertion follows from the equality G𝐱=G~𝐱GG_{\bf x}=\widetilde{G}_{\bf x}\cap G for the stabilisers and a similar one for the kernels. \square

Thus in particular it is now clear that Thm.  7 generalises Thm. 5.

Proof (of Theorem 7). Since Γ¯\overline{\Gamma} is finite, we can find a finite subgroup Γ\Gamma^{\prime} of Γ\Gamma such that Γ=Inn(G)Γ\Gamma=\operatorname{Inn}(G)\cdot\Gamma^{\prime} thanks to a result of Borel–Serre and Brion (see [15, 16]). For each i=1,2,,ni=1,2,\dots,n take giGg_{i}\in G so that ϕi:=Inn(gi)ϕiΓ\phi^{\prime}_{i}:=\operatorname{Inn}(g_{i})\circ\phi_{i}\in\Gamma^{\prime}. Then we have isomorphisms of bitorsors

fi:G(ϕi)G(ϕi),(g,ϕi)(ggi,ϕi)(i=1,2,,n),f_{i}\colon G(\phi^{\prime}_{i})\to G(\phi_{i}),\quad(g,\phi^{\prime}_{i})\mapsto(gg_{i},\phi_{i})\quad(i=1,2,\dots,n),

which induce a GG-equivariant isomorphism

f:X:=i=1nG(ϕi)X,(xi)i=1n(fi(xi))i=1n.f\colon X^{\prime}:=\prod_{i=1}^{n}G(\phi^{\prime}_{i})\to X,\quad(x^{\prime}_{i})_{i=1}^{n}\mapsto(f_{i}(x^{\prime}_{i}))_{i=1}^{n}.

Since ff is equivariant, a point 𝐱X{\bf x}\in X is stable (resp. polystable) if and only if f1(𝐱)Xf^{-1}({\bf x})\in X^{\prime} is stable (resp. polystable). Also, observe that Adfi=Ad{\mathop{\rm Ad}}\circ f_{i}={\mathop{\rm Ad}} (i=1,2,,ni=1,2,\dots,n) and hence A¯(𝐱)=A¯(f1(𝐱))\overline{A}({\bf x})=\overline{A}(f^{-1}({\bf x})) for all 𝐱X{\bf x}\in X. Therefore without loss of generality we may assume that ϕiΓ\phi_{i}\in\Gamma^{\prime} for all i=1,2,,ni=1,2,\dots,n.

Put G~=GΓ\widetilde{G}=G\ltimes\Gamma^{\prime}. It is a linearly reductive group with identity component GG and XX is a closed subvariety of G~n\widetilde{G}^{n}. By Theorem 5, a point 𝐱G~n{\bf x}\in\widetilde{G}^{n} is polystable with respect to the simultaneous G~\widetilde{G}-conjugation if and only if A(𝐱)G~A({\bf x})\subset\widetilde{G} is linearly reductive. By Prop. 9 this happens if and only if A¯(𝐱)\overline{A}({\bf x}) is linearly reductive. Together with Lemma 10 this implies the assertion. \square

Note that in general it really is necessary to work with A¯(𝐱)\overline{A}({\bf x}) rather than A(𝐱){A}({\bf x}):

Lemma 11.

There are examples of 𝐱X{\bf x}\in X which are polystable but A(𝐱){A}({\bf x}) is not linearly reductive.

Proof. Write xi=(gi,ϕi)G(ϕi)x_{i}=(g_{i},\phi_{i})\in G(\phi_{i}). Choose g1,,gnGg_{1},\ldots,g_{n}\in G generating a Zariski dense subgroup of an abelian unipotent subgroup UGU\subset G, such as a root group. Then define ϕi\phi_{i} to be the inner automorphism ϕi(g)=gi1ggi\phi_{i}(g)=g^{-1}_{i}gg_{i}. It follows that A(𝐱)U{A}({\bf x})\cong U is not reductive, but A¯(𝐱)={1}\overline{A}({\bf x})=\{1\} is trivial (and so 𝐱{\bf x} is polystable by Thm. 7). \square

Theorem 12.

For any point 𝐱X{\bf x}\in X the following hold:

  1. (1)

    𝐱{\bf x} is polystable if and only if any A(𝐱){A}({\bf x})-invariant parabolic subgroup PGP\subset G has an A(𝐱){A}({\bf x})-invariant Levi subgroup LPL\subset P.

  2. (2)

    𝐱{\bf x} is stable if and only if there are no proper A(𝐱){A}({\bf x})-invariant parabolic subgroups of GG.

We prepare three lemmas.

Lemma 13.

Let PP be a parabolic subgroup of GG and LL be a Levi subgroup of PP. Then NP(L)=LN_{P}(L)=L.

Proof. Let π:PP/Ru(P)\pi\colon P\to P/R_{u}(P) be the quotient of PP by the unipotent radical. Suppose that gPg\in P normalises LL and decompose it as g=hug=hu, hLh\in L, uRu(P)u\in R_{u}(P). Then for any xLx\in L we have gxg1Lgxg^{-1}\in L and hence uxu1Luxu^{-1}\in L. Since uKerπu\in\mathop{\rm Ker}\pi we have π(uxu1)=π(x)\pi(uxu^{-1})=\pi(x), which implies uxu1=xuxu^{-1}=x since the restriction of π\pi to LL is injective. Thus uu commutes with LL. On the other hand, LL coincides with its centraliser since L=CG(S)L=C_{G}(S) for some torus SLS\subset L. Thus uLu\in L and hence u=1u=1, i.e. gLg\in L. \square

Lemma 14.

The kernel of the GG-action on XX is equal to Z(G)ΓZ(G)^{\Gamma}.

Proof. If kGk\in G lies in the kernel, then for any gGg\in G and i=1,2,,ni=1,2,\dots,n we have k(g,ϕi)=(g,ϕi)kk(g,\phi_{i})=(g,\phi_{i})k, i.e.

kg=gϕi(k).kg=g\phi_{i}(k).

Taking gg to be 11 we obtain ϕi(k)=k\phi_{i}(k)=k (i=1,2,,ni=1,2,\dots,n). Thus kg=gkkg=gk for all gGg\in G, i.e., kZ(G)k\in Z(G). Since Γ\Gamma is generated by Inn(G),ϕ1,ϕ2,,ϕn\operatorname{Inn}(G),\phi_{1},\phi_{2},\dots,\phi_{n} and Inn(G)\operatorname{Inn}(G) trivially acts on Z(G)Z(G) we obtain kZ(G)Γk\in Z(G)^{\Gamma}. The converse is clear. \square

Lemma 15.

Z(G)G𝐱=Z(G)ΓZ(G)\cap G_{\bf x}=Z(G)^{\Gamma} for any 𝐱X{\bf x}\in X.

Proof. Take any 𝐱=(xi)i=1nX{\bf x}=(x_{i})_{i=1}^{n}\in X. For each i=1,2,,ni=1,2,\dots,n we have Ad(xi)ϕi1Inn(G){\mathop{\rm Ad}}(x_{i})\phi_{i}^{-1}\in\operatorname{Inn}(G), which trivially acts on Z(G)Z(G). Thus for any gZ(G)G𝐱g\in Z(G)\cap G_{\bf x} we have

ϕi(g)=Ad(xi)(g)=g(i=1,2,,n),\phi_{i}(g)={\mathop{\rm Ad}}(x_{i})(g)=g\quad(i=1,2,\dots,n),

which implies gZ(G)Γg\in Z(G)^{\Gamma}. The inclusion Z(G)ΓZ(G)G𝐱Z(G)^{\Gamma}\subset Z(G)\cap G_{\bf x} is clear. \square

Proof (of Theorem 12). As in the proof of Theorem 7, we may assume that ϕi\phi_{i}, i=1,2,,ni=1,2,\dots,n are all contained in a common finite subgroup ΓΓ\Gamma^{\prime}\subset\Gamma such that Γ=Inn(G)Γ\Gamma=\operatorname{Inn}(G)\cdot\Gamma^{\prime} and put G~=GΓ\widetilde{G}=G\ltimes\Gamma^{\prime}. Under this assumption A¯(𝐱)\overline{A}({\bf x}) is linearly reductive if and only if A(𝐱)G~A({\bf x})\subset\widetilde{G} is linearly reductive, by Prop. 9.

(1) We should show that a subgroup HG~H\subset\widetilde{G} is linearly reductive if and only if any HH-invariant parabolic in GG has an HH-invariant Levi subgroup. First suppose that HG~H\subset\widetilde{G} is linearly reductive and PGP\subset G is an HH-invariant parabolic subgroup. Since HH is linearly reductive and contained in NG~(P)N_{\widetilde{G}}(P), it is contained in some Levi subgroup L~\widetilde{L} of NG~(P)N_{\widetilde{G}}(P). Then L:=L~GL:=\widetilde{L}\cap G is a Levi subgroup of NG~(P)G=NG(P)=PN_{\widetilde{G}}(P)\cap G=N_{G}(P)=P and normalised by HH (so it is HH-invariant). Conversely, suppose any HH-invariant parabolic in GG has an HH-invariant Levi subgroup. By [26, Prop. 2.6], there exists a one-parameter subgroup λ\lambda of GG such that HPG~(λ)H\subset P_{\widetilde{G}}(\lambda) and Ru(H)UG~(λ)R_{u}(H)\subset U_{\widetilde{G}}(\lambda), where

(1) PG~(λ)\displaystyle P_{\widetilde{G}}(\lambda) ={xG~limt0λ(t)xλ(t)1 exists},\displaystyle=\{\,x\in\widetilde{G}\mid\text{$\lim_{t\to 0}\lambda(t)x\lambda(t)^{-1}$ exists}\,\},
(2) UG~(λ)\displaystyle U_{\widetilde{G}}(\lambda) ={xG~limt0λ(t)xλ(t)1=1},\displaystyle=\{\,x\in\widetilde{G}\mid\lim_{t\to 0}\lambda(t)x\lambda(t)^{-1}=1\,\},

and Ru(H)R_{u}(H) is the unipotent radical of HH. Put P=PG(λ)=PG~(λ)GP=P_{G}(\lambda)=P_{\widetilde{G}}(\lambda)\cap G, which is a parabolic subgroup of GG. Since PP is normalised by PG~(λ)P_{\widetilde{G}}(\lambda), it is normalised by HH. Hence PP has an HH-invariant Levi subgroup LL by assumption. We have HNG~(L)H\subset N_{\widetilde{G}}(L) and hence Ru(H)NG(L)R_{u}(H)\subset N_{G}(L) (recall that the unipotent radical is connected). Note that Ru(H)R_{u}(H) is also contained in UG~(λ)G=Ru(P)U_{\widetilde{G}}(\lambda)\cap G=R_{u}(P), while any non-trivial element of Ru(P)R_{u}(P) does not normalise the Levi subgroup LL by Lemma 13. Thus Ru(H)R_{u}(H) is trivial, i.e. HH is linearly reductive.

(2) Suppose that 𝐱=(xi)i=1nX{\bf x}=(x_{i})_{i=1}^{n}\in X is stable and let PP be a A(𝐱){A}({\bf x})-invariant parabolic subgroup of GG. By Theorem 7, A¯(𝐱)\overline{A}({\bf x}) (and hence A(𝐱)A({\bf x})) is linearly reductive. Since A(𝐱)A({\bf x}) normalises PP, there exists a Levi subgroup L~\widetilde{L} of NG~(P)N_{\widetilde{G}}(P) containing A(𝐱)A({\bf x}). By [26, Prop. 2.4], there exists a one-parameter subgroup λ\lambda of GG such that NG~(P)=PG~(λ)N_{\widetilde{G}}(P)=P_{\widetilde{G}}(\lambda) and L~=CG~(Imλ)\widetilde{L}=C_{\widetilde{G}}(\operatorname{Im}\lambda). Since A(𝐱)CG~(Imλ)A({\bf x})\subset C_{\widetilde{G}}(\operatorname{Im}\lambda) each xix_{i} commutes with Imλ\operatorname{Im}\lambda and hence ImλG𝐱\operatorname{Im}\lambda\subset G_{\bf x}. The stability now implies that Imλ\operatorname{Im}\lambda is contained in the kernel Z(G)ΓZ(G)^{\Gamma} and hence P=GP=G. Conversely, suppose that 𝐱X{\bf x}\in X is not stable. Then if the orbit G𝐱G\cdot{\bf x} is not closed we argue as follows: By the Hilbert–Mumford criterion, there exists a one-parameter subgroup λ\lambda of GG and an element 𝐲=(yi)i=1nX{\bf y}=(y_{i})_{i=1}^{n}\in X such that limt0λ(t)𝐱=𝐲\lim_{t\to 0}\lambda(t)\cdot{\bf x}={\bf y} and G𝐲G\cdot{\bf y} is closed. We show that the parabolic subgroup P:=PG(λ)=PG~(λ)GP:=P_{G}(\lambda)=P_{\widetilde{G}}(\lambda)\cap G is A(𝐱){A}({\bf x})-invariant and proper. For any gPg\in P and i=1,2,,ni=1,2,\dots,n, the limit of

λ(t)Ad(xi)(g)λ(t)1=λ(t)xiλ(t)1λ(t)gλ(t)1λ(t)xi1λ(t)1\lambda(t){\mathop{\rm Ad}}(x_{i})(g)\lambda(t)^{-1}=\lambda(t)x_{i}\lambda(t)^{-1}\cdot\lambda(t)g\lambda(t)^{-1}\cdot\lambda(t)x_{i}^{-1}\lambda(t)^{-1}

as t0t\to 0 exists. Hence PP is A(𝐱){A}({\bf x})-invariant. If PP is not proper, Imλ\operatorname{Im}\lambda is contained in Z(G)G𝐲=Z(G)ΓZ(G)\cap G_{\bf y}=Z(G)^{\Gamma}. Thus λ(t)𝐱=𝐱\lambda(t)\cdot{\bf x}={\bf x} (tt\in\mathbb{C}^{*}) and hence 𝐱=𝐲{\bf x}={\bf y}, which contradicts the assumption that G𝐱G\cdot{\bf x} is not closed. Hence PP is proper. Finally suppose G𝐱G\cdot{\bf x} is closed, but of the wrong dimension. Then the stabiliser G𝐱G_{\bf x} is linearly reductive ([26] 1.3.3) and the quotient G𝐱/Z(G)ΓG_{\bf x}/Z(G)^{\Gamma} has non-trivial identity component. Hence there exists a one-parameter subgroup λ\lambda of G𝐱G_{\bf x} such that ImλZ(G)Γ\operatorname{Im}\lambda\not\subset Z(G)^{\Gamma}. Since each xix_{i} commutes with λ\lambda, the parabolic subgroup P:=PG(λ)P:=P_{G}(\lambda) is A(𝐱){A}({\bf x})-invariant. It is proper since Lemma 15 implies ImλZ(G)\operatorname{Im}\lambda\not\subset Z(G). \square

Let us rephrase/abstract the first part of this in a way that will be useful later. Let GG be a reductive group and Λ\Lambda an algebraic group acting on GG by (algebraic) group automorphisms. Suppose that the action of Λ\Lambda is effective and the identity component Λ0\Lambda^{0} acts by inner automorphisms. Then we may regard Λ\Lambda as a subgroup of Aut(G)\operatorname{\mathop{\rm Aut}}(G) and its image in Out(G)\mathop{\rm Out}(G) is finite.

Proposition 16.

Λ\Lambda is linearly reductive if and only if any Λ\Lambda-invariant parabolic subgroup of GG has a Λ\Lambda-invariant Levi subgroup.

Proof. By the result of Borel–Serre and Brion, there exists a finite subgroup ΛΛ\Lambda^{\prime}\subset\Lambda such that Λ0Λ=Λ\Lambda^{0}\Lambda^{\prime}=\Lambda. Put G~=GΛ\widetilde{G}=G\ltimes\Lambda^{\prime} and let KG~K\subset\widetilde{G} be the preimage of Λ\Lambda under Ad{\mathop{\rm Ad}}, so that Prop. 9 implies Λ\Lambda is linearly reductive if and only if KK is linearly reductive. Thus the equivalence follows from the proof of Thm. 12 (1). \square

In the same setting there is a further characterisation:

Proposition 17.

Λ\Lambda is linearly reductive if and only if there exists a torus SGS\subset G such that CG(S)C_{G}(S) is Λ\Lambda-invariant and has no proper Λ\Lambda-invariant parabolic subgroups.

Proof. Suppose that Λ\Lambda is linearly reductive. We show that there exists a decreasing sequence of Λ\Lambda-invariant closed subgroups

G=L0P1L1P2L2PrLrG=L_{0}\supset P_{1}\supset L_{1}\supset P_{2}\supset L_{2}\supset\cdots\supset P_{r}\supset L_{r}

such that each PiP_{i} is a proper parabolic subgroup of Li1L_{i-1}, each LiL_{i} (i>0i>0) is a Levi subgroup of PiP_{i} and LrL_{r} has no proper Λ\Lambda-invariant parabolic subgroups. If L0=GL_{0}=G has no proper Λ\Lambda-invariant parabolic subgroups we have nothing to do (just put r=0r=0). Otherwise we take any proper Λ\Lambda-invariant parabolic subgroup P1L0P_{1}\subsetneq L_{0}. Then by Prop. 16 there exists a Λ\Lambda-invariant Levi subgroup L1P1L_{1}\subset P_{1}. Let Λ1\Lambda_{1} be the quotient of Λ\Lambda by the kernel of the induced Λ\Lambda-action on L1L_{1}, so that Λ1\Lambda_{1} effectively acts on L1L_{1}. Note that its identity component Λ10\Lambda_{1}^{0} acts by inner automorphisms of L1L_{1}; indeed, the action of any element of Λ10\Lambda_{1}^{0} is induced from some inner automorphism of GG preserving P1,L1P_{1},L_{1} and hence is inner by Lemma 13. Since Λ\Lambda is linearly reductive Λ1\Lambda_{1} is also linearly reductive, and a subgroup of L1L_{1} is Λ1\Lambda_{1}-invariant if and only if it is Λ\Lambda-invariant as a subgroup of L0L_{0}. If L1L_{1} has no proper Λ1\Lambda_{1}-invariant parabolic subgroup, the sequence L0P1L1L_{0}\supset P_{1}\supset L_{1} is as desired. Otherwise we take any proper Λ1\Lambda_{1}-invariant parabolic subgroup P2L1P_{2}\subsetneq L_{1}. Then by Prop. 16 there exists a Λ1\Lambda_{1}-invariant Levi subgroup L2P2L_{2}\subset P_{2}. Iterating this procedure, we obtain a desired decreasing sequence. Since each LiL_{i} (i>0i>0) is a Levi subgroup of PiP_{i} there exists a torus SiLiS_{i}\subset L_{i} such that Li=CLi1(Si)L_{i}=C_{L_{i-1}}(S_{i}). Then LrL_{r} is the common centraliser of S1,S2,,SrS_{1},S_{2},\dots,S_{r} in GG. Hence the torus SGS\subset G generated by S1,S2,,SrS_{1},S_{2},\dots,S_{r} (note that they commute with each other) is as desired.

Conversely, suppose that there exists a torus SGS\subset G such that L:=CG(S)L:=C_{G}(S) is Λ\Lambda-invariant and has no proper Λ\Lambda-invariant parabolic subgroups. Let PGP\subset G be a Λ\Lambda-invariant parabolic subgroup. Then the intersection PLP\cap L is a Λ\Lambda-invariant parabolic subgroup of LL and hence PL=LP\cap L=L, i.e. LPL\subset P. Since LL is reductive, there exists a Levi subgroup MPM\subset P containing LL. For any ψΛ\psi\in\Lambda the image ψ(M)\psi(M) of MM is also a Levi subgroup of PP containing LL. Since LL contains a maximal torus of PP and any maximal torus of PP is contained in a unique Levi subgroup, we have ψ(M)=M\psi(M)=M. Hence MM is Λ\Lambda-invariant, which together with Prop. 16 shows that Λ\Lambda is linearly reductive. \square

Applying this to the set-up of the present section (with Λ=A¯(𝐱)\Lambda=\overline{A}({\bf x})) yields:

Corollary 18.

The following are equivalent:

0) A point 𝐱X{\bf x}\in X is polystable,

1) The group A¯(𝐱)\overline{A}({\bf x}) is linearly reductive,

2) Any A(𝐱){A}({\bf x})-invariant parabolic in GG has an A(𝐱){A}({\bf x})-invariant Levi subgroup,

3) There exists a subtorus SGS\subset G such that CG(S)C_{G}(S) is A(𝐱){A}({\bf x})-invariant and has no proper A(𝐱){A}({\bf x})-invariant parabolic subgroups,

4) There exists an A(𝐱){A}({\bf x})-invariant Levi subgroup LL of a parabolic of GG, such that LL has no proper A(𝐱){A}({\bf x})-invariant parabolic subgroups.

Note that 3) and 4) are trivially equivalent since centralisers of tori in GG are exactly the Levi subgroups of parabolics.

3. More general set-up

The results of the previous section will now be generalised, in a form more directly useful in the context of Stokes local systems. Return to the set-up of Thm. 7 (with n1n\geq 1), but now further choose an integer m1m\geq 1 and tori 𝕋1,,𝕋mG\mathbb{T}_{1},\ldots,\mathbb{T}_{m}\subset G. Write 𝐓=𝕋1××𝕋mGm{\bf T}=\mathbb{T}_{1}\times\cdots\times\mathbb{T}_{m}\subset G^{m}, let Hi=CG(𝕋i)H_{i}=C_{G}(\mathbb{T}_{i}) and 𝐇=H1××Hm=CGm(𝐓)Gm{\bf H}=H_{1}\times\cdots\times H_{m}=C_{G^{m}}({\bf T})\subset G^{m}. We allow some of the 𝕋i\mathbb{T}_{i} to be a point, in which case Hi=GH_{i}=G. In this section we will study the stability and polystability for the action of 𝐇{\bf H} on

(3) X:=Gm1×i=1nG(ϕi)X:=G^{m-1}\times\prod_{i=1}^{n}G(\phi_{i})

given by

𝐡(𝐂,𝐌)=(h2C2h11,,hmCmh11,h1M1h11,,h1Mnh11){\bf h}\cdot({\bf C},{\bf M})=(h_{2}C_{2}h_{1}^{-1},\ldots,h_{m}C_{m}h_{1}^{-1},h_{1}M_{1}h_{1}^{-1},\ldots,h_{1}M_{n}h_{1}^{-1})

where 𝐡=(h1,,hm),𝐂=(C2,,Cm),𝐌=(M1,,Mn),CiG,MiG(ϕi).{\bf h}=(h_{1},\ldots,h_{m}),{\bf C}=(C_{2},\ldots,C_{m}),{\bf M}=(M_{1},\ldots,M_{n}),C_{i}\in G,M_{i}\in G(\phi_{i}).

Thus if m=1m=1 and 𝕋1=1\mathbb{T}_{1}=1 we recover the situation of Thm. 7. The case m=1m=1 and 𝕋1\mathbb{T}_{1} arbitrary but each ϕi=1\phi_{i}=1 was studied by Richardson in [26] Thm. 13.2,14.1 (taking S=𝕋1S=\mathbb{T}_{1} acting on GG by conjugation). More generally, in effect, [8] Cor. 9.6 studied the notion of stability in the case with mm arbitrary and each ϕi=1\phi_{i}=1, making the link to the differential Galois group of irregular connections (whence the 𝕋i\mathbb{T}_{i} are the Ramis/exponential tori).

For 𝐱=(𝐂,𝐌)X{\bf x}=({\bf C},{\bf M})\in X let A(𝐱)GΓA({\bf x})\subset G\ltimes\Gamma be the Zariski closure of the subgroup generated by

M1,M2,,Mn,𝕋1,C21𝕋2C2,,Cm1𝕋mCm.M_{1},M_{2},\ldots,M_{n},\mathbb{T}_{1},C_{2}^{-1}\mathbb{T}_{2}C_{2},\ldots,C_{m}^{-1}\mathbb{T}_{m}C_{m}.

Let A¯(𝐱)Γ\overline{A}({\bf x})\subset\Gamma be the image of A(𝐱)A({\bf x}) in ΓAut(G)\Gamma\subset\operatorname{\mathop{\rm Aut}}(G), as before. Recall that a subset of GG is A(𝐱)A({\bf x})-invariant if it is preserved by A¯(𝐱)Aut(G)\overline{A}({\bf x})\subset\operatorname{\mathop{\rm Aut}}(G).

Theorem 19.

(1) A point 𝐱X{\bf x}\in X is polystable for the 𝐇{\bf H} action if and only if A¯(𝐱)\overline{A}({\bf x}) is linearly reductive.

(2) A point 𝐱X{\bf x}\in X is stable for the 𝐇{\bf H} action if and only if there are no proper A(𝐱)A({\bf x})-invariant parabolic subgroups in GG.

As in the last section, one can rephrase polystability in several different ways:

Corollary 20.

A point 𝐱X{\bf x}\in X is polystable if and only if

1) The group A¯(𝐱)\overline{A}({\bf x}) is linearly reductive, or

2) Any A(𝐱){A}({\bf x})-invariant parabolic in GG has an A(𝐱){A}({\bf x})-invariant Levi subgroup, or

3) There exists a subtorus SGS\subset G such that CG(S)C_{G}(S) is A(𝐱){A}({\bf x})-invariant and has no proper A(𝐱){A}({\bf x})-invariant parabolic subgroups.

Proof (of Thm. 19). Let G×𝐇G\times{\bf H} act on X~:=Gm×1NG(ϕi)\widetilde{X}:=G^{m}\times\prod_{1}^{N}G(\phi_{i}) via

(g,𝐡)(𝐂,𝐌)=(h1C1g1,h2C2g1,,hmCmg1,gM1g1,,gMng1)(g,{\bf h})\cdot({\bf C},{\bf M})=(h_{1}C_{1}g^{-1},h_{2}C_{2}g^{-1},\ldots,h_{m}C_{m}g^{-1},gM_{1}g^{-1},\ldots,gM_{n}g^{-1})

where 𝐡=(h1,,hm),𝐂=(C1,,Cm),𝐌=(M1,,Mn),CiG,MiG(ϕi).{\bf h}=(h_{1},\ldots,h_{m}),{\bf C}=(C_{1},\ldots,C_{m}),{\bf M}=(M_{1},\ldots,M_{n}),C_{i}\in G,M_{i}\in G(\phi_{i}). (Up to relabelling and incrementing mm this is the special case where H1=GH_{1}=G.) Consider the 𝐇{\bf H}-equivariant map X~X\widetilde{X}\to X taking (𝐂;𝐌)({\bf C};{\bf M}) to

(C2C11,,CmC11;C1M1C11,,C1MNC11)X.(C_{2}C_{1}^{-1},\ldots,C_{m}C_{1}^{-1};C_{1}M_{1}C_{1}^{-1},\ldots,C_{1}M_{N}C_{1}^{-1})\in X.

It expresses X~\widetilde{X} as a principal GG-bundle over XX (the fibres are exactly the GG-orbits). Any point 𝐱X{\bf x}\in X has a unique lift 𝐱~X~\widetilde{\bf x}\in\widetilde{X} with C1=1C_{1}=1. It follows that the 𝐇{\bf H} orbit of 𝐱X{\bf x}\in X is closed if and only if the G×𝐇G\times{\bf H} orbit of 𝐱~\widetilde{\bf x} is closed in X~\widetilde{X}.

Now choose 𝐭=(t1,,tm)𝐓{\bf t}=(t_{1},\ldots,t_{m})\in{\bf T} so that tit_{i} generates a Zariski dense subgroup of 𝕋i\mathbb{T}_{i} for each ii. In particular Hi=CG(ti)H_{i}=C_{G}(t_{i}). Consider the simultaneous conjugation action of GG on Y:=Gm×1NG(ϕi)Y:=G^{m}\times\prod_{1}^{N}G(\phi_{i}), and the GG-equivariant embedding

π:X~/𝐇Y;[(𝐂;𝐌)](𝐂1𝐭𝐂,𝐌).\pi:\widetilde{X}/{\bf H}\hookrightarrow Y;\quad[({\bf C};{\bf M})]\mapsto({\bf C}^{-1}{\bf t}{\bf C},{\bf M}).

Thus X~/𝐇\widetilde{X}/{\bf H} is identified with a closed subvariety of YY and X~\widetilde{X} is a GG-equivariant principal 𝐇{\bf H}-bundle over X~/𝐇\widetilde{X}/{\bf H}. It follows that the G×𝐇G\times{\bf H} orbit of 𝐱~\widetilde{\bf x} is closed in X~\widetilde{X} if and only if the GG orbit of π(𝐱~)\pi(\widetilde{\bf x}) is closed in YY.

Hence part (1) of the theorem follows from Thm. 7 (applied to YY). To deal with stability we need to consider the stabilisers and the kernels of the actions.

Lemma 21.

The stabiliser of any point 𝐱X{\bf x}\in X for the 𝐇{\bf H}-action is canonically isomorphic to the stabiliser of the point π(𝐱~)Y\pi(\widetilde{\bf x})\in Y for the GG-action.

Proof. Suppose gGg\in G fixes (𝐂1𝐭𝐂,𝐌)({\bf C}^{-1}{\bf t}{\bf C},{\bf M}) and let hi=CigCi1h_{i}=C_{i}gC_{i}^{-1} (i=1,2,,mi=1,2,\dots,m) and 𝐡=(h1,,hm){\bf h}=(h_{1},\ldots,h_{m}). Note h1=gh_{1}=g because C1=1C_{1}=1. Since gg commutes with each Ci1tiCiC_{i}^{-1}t_{i}C_{i}, each hih_{i} commutes with tit_{i}, which implies 𝐡𝐇{\bf h}\in{\bf H}. We have hiCig1=Cih_{i}C_{i}g^{-1}=C_{i} (i=1,2,,mi=1,2,\ldots,m) by the definition of hih_{i} and gMjg1=MjgM_{j}g^{-1}=M_{j} (j=1,2,,Nj=1,2,\dots,N) as gg centralises 𝐌{\bf M}. Hence the pair (g,𝐡)G×𝐇(g,{\bf h})\in G\times{\bf H} stabilises 𝐱~X~\widetilde{\bf x}\in\widetilde{X}, and hence 𝐡{\bf h} stabilises 𝐱X{\bf x}\in X. Conversely if 𝐡=(h1,,hm)𝐇{\bf h}=(h_{1},\ldots,h_{m})\in{\bf H} fixes 𝐱{\bf x}, then let g=h1g=h_{1}. It follows immediately that (g,𝐡)G×𝐇(g,{\bf h})\in G\times{\bf H} fixes 𝐱~\widetilde{\bf x} and thus that gg fixes π(𝐱~)\pi(\widetilde{\bf x}). Clearly the two correspondences are inverses of each other. \square

As in Lemma 14 one has:

Lemma 22.

(1) The kernel of the GG-action on YY (or X~/𝐇\widetilde{X}/{\bf H}) is the Γ\Gamma-invariant subgroup Z(G)ΓZ(G)^{\Gamma} of the center of GG, and (2) The kernel of the 𝐇{\bf H}-action on XX is the subgroup of elements (h1,,hm)𝐇(h_{1},\ldots,h_{m})\in{\bf H} satisfying h1=h2==hmZ(G)Γh_{1}=h_{2}=\cdots=h_{m}\in Z(G)^{\Gamma}.

Note that these two groups correspond to each other under the correspondence of Lem. 21 (for any 𝐱X{\bf x}\in X). It follows that 𝐱X{\bf x}\in X is stable if and only if π(𝐱~)Y\pi(\widetilde{\bf x})\in Y is stable. By Thm. 12 (2) applied to YY, this happens if and only if there are no proper A(𝐱)A({\bf x}) invariant parabolic subgroups of GG. \square

4. Application to wild character varieties

Recall from [8, 13] that the wild character variety B(𝚺)=THom𝕊(Π,G)/𝐇\mathcal{M}_{\text{\rm B}}({\bf\Sigma})=\operatorname{THom}_{\mathbb{S}}(\Pi,G)/{\bf H} is determined by an irregular curve/wild Riemann surface 𝚺=(Σ,α,Θ){\bf\Sigma}=(\Sigma,\alpha,\Theta) with group 𝒢\mathcal{G}, where Σ\Sigma is a compact smooth complex algebraic curve, αΣ\alpha\subset\Sigma is a non-empty finite subset, 𝒢Σ:=Σα\mathcal{G}\to\Sigma^{\circ}:=\Sigma\setminus\alpha is a local system of groups over the punctured curve (with each fibre isomorphic to some fixed connected complex reductive group GG), and Θ\Theta consists of the data of an irregular class Θa\Theta_{a} at each point aαa\in\alpha (in the sense of [13] §3.5—it is the class of a graded 𝒢\mathcal{G} local system).

As in [13] §4, 𝚺{\bf\Sigma} then determines an auxiliary surface Σ~\widetilde{\Sigma}, equipped with boundary circles \partial, halos Σ~\mathbb{H}\subset\widetilde{\Sigma}, and tangential punctures e(𝔸)e(\mathbb{A}). Further 𝚺{\bf\Sigma} determines a local system 𝕋\mathbb{T}\to\partial of finite dimensional complex tori, the Ramis tori ([13] p.9). Choosing a finite set β={b1,,bm}\beta=\{b_{1},\ldots,b_{m}\}\subset\partial of basepoints (with one point in each component circle, as in [13] §4.1) then determines the wild surface groupoid Π=Π1(Σ~,β)\Pi=\Pi_{1}(\widetilde{\Sigma},\beta), the fundamental groupoid of the auxiliary surface with these basepoints, as in [13] §5. The local system 𝒢\mathcal{G} is determined by a map f:ΠAut(G)f:\Pi\to\operatorname{\mathop{\rm Aut}}(G) and this determines the space of ff-twisted representations

THom(Π,G)={ρHom(Π,GAut(G))|ρ(γ)G(f(γ)) for all γΠ}\operatorname{THom}(\Pi,G)=\{\rho\in\operatorname{Hom}(\Pi,G\ltimes\operatorname{\mathop{\rm Aut}}(G))\ \bigl{|}\ \rho(\gamma)\in G(f(\gamma))\text{ for all }\gamma\in\Pi\}

of Π\Pi, as in [13] §5. Since Σ~\widetilde{\Sigma} is just a punctured real surface with boundary, choosing generating paths in Π\Pi yields an isomorphism THom(Π,G)GN\operatorname{THom}(\Pi,G)\cong G^{N} of spaces, for some integer NN, so it is a smooth affine variety. In turn the wild representation variety =THom𝕊(Π,G)\mathcal{R}=\operatorname{THom}_{\mathbb{S}}(\Pi,G) is the closed subvariety of THom(Π,G)\operatorname{THom}(\Pi,G) cut out by the two Stokes conditions, in [13] Defn. 18. Intrinsically, \mathcal{R} is the moduli space of framed Stokes local systems, as in [13] Prop. 19, framed via a graded isomorphism to a standard fibre i\mathcal{F}_{i} at each basepoint biβb_{i}\in\beta ([13] §4.1). The group Hi=GrAut(i)=CG(𝕋i)G=Aut(i)H_{i}=\operatorname{GrAut}(\mathcal{F}_{i})=C_{G}(\mathbb{T}_{i})\subset G=\operatorname{\mathop{\rm Aut}}(\mathcal{F}_{i}) acts transitively on the set of framings at bib_{i}, where 𝕋i=𝕋bi\mathbb{T}_{i}=\mathbb{T}_{b_{i}}. Thus the group 𝐇=biβHi{\bf H}=\prod_{b_{i}\in\beta}H_{i} acts naturally on \mathcal{R}. The wild character variety B(Σ)\mathcal{M}_{\text{\rm B}}(\Sigma) is the affine geometric invariant theory quotient /𝐇\mathcal{R}/{\bf H}, and so its points are the closed 𝐇{\bf H} orbits in \mathcal{R}. This leads directly to the key statement:

Proposition 23.

The wild representation variety =THom𝕊(Π,G)\mathcal{R}=\operatorname{THom}_{\mathbb{S}}(\Pi,G) may be embedded in an 𝐇{\bf H}-equivariant way, as a closed subvariety of the space X=Gm1×1nG(ϕi)X=G^{m-1}\times\prod_{1}^{n}G(\phi_{i}) of (3), for suitable nn and automorphisms {ϕi}Aut(G)\{\phi_{i}\}\subset\operatorname{\mathop{\rm Aut}}(G), with m=#αm=\#\alpha.

Proof. This comes down to considering the inclusion THom𝕊(Π,G)THom(Π,G)\operatorname{THom}_{\mathbb{S}}(\Pi,G)\subset\operatorname{THom}(\Pi,G) as a closed subvariety (forgetting the Stokes conditions, as in [13] Defn. 18), and then identifying THom(Π,G)X\operatorname{THom}(\Pi,G)\cong X in an 𝐇{\bf H}-equivariant way. Both of these are straightforward. In particular Mj=ρ(γj)M_{j}=\rho(\gamma_{j}) for generators γj\gamma_{j} of π1(Σ~,b1)\pi_{1}(\widetilde{\Sigma},b_{1}), and Ci=ρ(χi)C_{i}=\rho(\chi_{i}) for paths χi\chi_{i} from b1b_{1} to bib_{i}, for i=2,,mi=2,\ldots,m. \square

In order to define the Galois group Gal(ρ)\mathop{\rm Gal}(\rho) of ρTHom𝕊(Π,G)\rho\in\operatorname{THom}_{\mathbb{S}}(\Pi,G) we will assume that 𝒢\mathcal{G} is “Out-finite”, in the sense that the monodromy group f(π1(Σ~,b1))Aut(G)f(\pi_{1}(\widetilde{\Sigma},b_{1}))\subset\operatorname{\mathop{\rm Aut}}(G) of 𝒢\mathcal{G} has finite image in Out(G)=Aut(G)/Inn(G)\mathop{\rm Out}(G)=\operatorname{\mathop{\rm Aut}}(G)/\operatorname{Inn}(G). Then the group Γ\Gamma generated by f(π1(Σ~,b1))f(\pi_{1}(\widetilde{\Sigma},b_{1})) and Inn(G)\operatorname{Inn}(G) is an algebraic group, as in §2 above. Thus any Stokes representation takes values in the algebraic group GΓGAut(G)G\ltimes\Gamma\subset G\ltimes\operatorname{\mathop{\rm Aut}}(G), and so we can consider the Zariski closure of its monodromy. The Galois group is defined by adding the Ramis tori as well: If t𝕋iGt\in\mathbb{T}_{i}\subset G and χ\chi is any path in Σ~\widetilde{\Sigma} from b1b_{1} to bib_{i}, consider the element

(4) C1tCG=G(Id)GΓ,C^{-1}tC\in G=G(\text{\rm Id})\subset G\ltimes\Gamma,

where C=ρ(χ)C=\rho(\chi).

Definition 24.

The differential Galois group Gal(ρ)\mathop{\rm Gal}(\rho) of ρ\rho is the Zariski closure of the subgroup of GΓG\ltimes\Gamma generated by ρ(π1(Σ~,b1))\rho(\pi_{1}(\widetilde{\Sigma},b_{1})) and all of the tori (4) (as i,t,χi,t,\chi vary).

It follows that Gal(ρ)\mathop{\rm Gal}(\rho) acts on GG by group automorphisms, via the adjoint action of GΓG\ltimes\Gamma on G=G(Id)G=G(\text{\rm Id}). Let Gal¯(ρ)ΓAut(G)\overline{\mathop{\rm Gal}}(\rho)\subset\Gamma\subset\operatorname{\mathop{\rm Aut}}(G) be the resulting image of Gal(ρ)\mathop{\rm Gal}(\rho). This definition is (of course) motivated by Ramis’ description ([25], [24] Thm. 21, [22] Thm. III.3.11) of the differential Galois group of an algebraic connection on a vector bundle.

Corollary 25.

A Stokes representation ρTHom𝕊(Π,G)\rho\in\operatorname{THom}_{\mathbb{S}}(\Pi,G) is polystable for the action of 𝐇{\bf H} if and only if Gal¯(ρ)\overline{\mathop{\rm Gal}}(\rho) is a linearly reductive group.

Proof. This now follows from part (1) of Thm. 19, via Prop. 23, since Gal¯(ρ)\overline{\mathop{\rm Gal}}(\rho) matches up with A¯(𝐱)\overline{A}({\bf x}). \square

Special cases include:

\bullet If 𝒢\mathcal{G} has finite monodromy then ρTHom𝕊(Π,G)\rho\in\operatorname{THom}_{\mathbb{S}}(\Pi,G) is polystable if and only if Gal(ρ)\mathop{\rm Gal}(\rho) is a linearly reductive group.

\bullet If 𝒢\mathcal{G} is a constant general linear group then ρHom𝕊(Π,G)\rho\in\operatorname{Hom}_{\mathbb{S}}(\Pi,G) is polystable if and only if ρ\rho is the direct sum of irreducible Stokes representations.

Recall that Z(G)ΓZ(G)^{\Gamma} is the Γ\Gamma invariant subgroup of the centre of GG, and it embeds diagonally in 𝐇{\bf H}. To deal with stability we will avoid degenerate cases by assuming:

(5) The kernel of the action of 𝐇 on THom𝕊(Π,G) is Z(G)Γ\begin{matrix}\text{\em The kernel of the action of ${\bf H}$ on $\operatorname{THom}_{\mathbb{S}}(\Pi,G)$ is $Z(G)^{\Gamma}$. }\end{matrix}

The lemma below shows one can always add one or two punctures to ensure this condition holds. Note that no generality is lost: any symplectic leaf B(𝚺,𝒞)B(𝚺)\mathcal{M}_{\text{\rm B}}({\bf\Sigma},\mathcal{C})\subset\mathcal{M}_{\text{\rm B}}({\bf\Sigma}) will also be a symplectic leaf of the larger Poisson variety obtained by first making such additional punctures (namely that with trivial monodromy around the new punctures). For example the usual character variety of a genus g>0g>0 compact Riemann surface Σ\Sigma is a (very special) symplectic leaf of the character variety of Σa\Sigma\setminus a, for any point aΣa\in\Sigma.

Lemma 26.

Suppose m1m\geq 1 and a1αa_{1}\in\alpha has trivial irregular class, and if g=0g=0 then m2m\geq 2. Then THom𝕊(Π,G)\operatorname{THom}_{\mathbb{S}}(\Pi,G) is a smooth non-empty affine variety and the kernel K of the 𝐇{\bf H} action is Z(G)ΓZ(G)^{\Gamma} embedded diagonally in 𝐇{\bf H}.

Proof. It is nonempty as it is the fusion of some fission spaces and some internally fused doubles: Recall that THom𝕊(Π,G)\operatorname{THom}_{\mathbb{S}}(\Pi,G) can be described as the quasi-Hamiltonian GG-reduction

THom𝕊(Π,G)(𝔻1GG𝔻gG𝒜(Q1)GG𝒜(Qm))//G\operatorname{THom}_{\mathbb{S}}(\Pi,G)\cong\bigl{(}\mathbb{D}_{1}\ \smash{\mathop{\circledast}\limits_{G}}\ \cdots\ \smash{\mathop{\circledast}\limits_{G}}\ \mathbb{D}_{g}\ \smash{\mathop{\circledast}\limits_{G}}\ \mathcal{A}(Q_{1})\ \smash{\mathop{\circledast}\limits_{G}}\ \cdots\ \smash{\mathop{\circledast}\limits_{G}}\ \mathcal{A}(Q_{m})\bigr{)}/\!\!/G

where each 𝔻i\mathbb{D}_{i} is a (twisted) internally fused double ([13] p.23, [8] Thm 8.2). Since 𝒜(Q1)𝐃(G)\mathcal{A}(Q_{1})\cong{\bf D}(G) is the double of GG, it follows that

THom𝕊(Π,G)𝔻1𝔻g𝒜(Q2)𝒜(Qm)\operatorname{THom}_{\mathbb{S}}(\Pi,G)\cong\mathbb{D}_{1}\ \smash{\mathop{\circledast}\limits}\ \cdots\ \smash{\mathop{\circledast}\limits}\ \mathbb{D}_{g}\ \smash{\mathop{\circledast}\limits}\ \mathcal{A}(Q_{2})\ \smash{\mathop{\circledast}\limits}\ \cdots\ \smash{\mathop{\circledast}\limits}\ \mathcal{A}(Q_{m})

so that THom𝕊(Π,G)\operatorname{THom}_{\mathbb{S}}(\Pi,G) is the product of some smooth nonempty affine varieties. Note that 𝐇{\bf H} still acts on THom𝕊(Π,G)\operatorname{THom}_{\mathbb{S}}(\Pi,G) and this includes H1=GH_{1}=G.

Suppose 𝐤K{\bf k}\in{\text{\bf K}} and suppose (as usual) that the framings of 𝒢\mathcal{G} are such that the monodromy in Aut(G)\operatorname{\mathop{\rm Aut}}(G) of 𝒢\mathcal{G} is trivial along the m1m-1 chosen paths χi:b1bi\chi_{i}:b_{1}\to b_{i}. Then 𝐤{\bf k} acts on Ci=MχiC_{i}=M_{\chi_{i}} as kiCik11k_{i}C_{i}k_{1}^{-1}. Taking C2=C3=Cm=1C_{2}=C_{3}=\cdots C_{m}=1 implies k1=k2==kmk_{1}=k_{2}=\cdots=k_{m}. If m2m\geq 2 then the fact that all C2GC_{2}\in G are fixed implies k1Z(G)k_{1}\in Z(G). On the other hand if m=1m=1 then g>0g>0 so looking at 𝔻1\mathbb{D}_{1} there is ϕAut(G)\phi\in\operatorname{\mathop{\rm Aut}}(G) so that k1Aϕ(k11)=Ak_{1}A\phi(k_{1}^{-1})=A for all AGA\in G. This implies k1=ϕ(k1)k_{1}=\phi(k_{1}) and k1Z(G)k_{1}\in Z(G). Thus in all case k1k_{1} is central. Then looking at any loop γ\gamma based at b1b_{1} leads to a relation of the form k1Mγϕγ(k11)=Mγk_{1}M_{\gamma}\phi_{\gamma}(k_{1}^{-1})=M_{\gamma}. Thus since k1k_{1} is central this implies k1=ϕγ(k1)k_{1}=\phi_{\gamma}(k_{1}), so k1Z(G)Γk_{1}\in Z(G)^{\Gamma}. \square

Remark 27.

Note that it follows in general (as in [8] Thm. 8.2) that if THom𝕊(Π,G)\operatorname{THom}_{\mathbb{S}}(\Pi,G) is nonempty (and m>0m>0) then it is a smooth affine variety.

Part (2) of Theorem 12 then implies:

Corollary 28.

A Stokes representation ρTHom𝕊(Π,G)\rho\in\operatorname{THom}_{\mathbb{S}}(\Pi,G) is stable for the action of 𝐇{\bf H} if and only if there is no proper parabolic subgroup PGP\subset G stabilised by the action of Gal(ρ)\mathop{\rm Gal}(\rho).

Proof. This follows from Thm. 12 since \mathcal{R} is closed in XX, Gal(ρ)\mathop{\rm Gal}(\rho) matches up with A(𝐱)A({\bf x}), and the kernel of the 𝐇{\bf H} action on \mathcal{R} and on XX is the same. \square

5. Stability and polystability of Stokes local systems

This section will consider the intrinsic objects (Stokes local systems) underlying Stokes representations, and define the notions of “irreducible” and “reductive” for Stokes local systems, in terms of reductions of structure group. Then we will deduce:

Theorem 29.

Suppose 𝕃\mathbb{L} is a Stokes local system and ρTHom𝕊(Π,G)\rho\in\operatorname{THom}_{\mathbb{S}}(\Pi,G) is the monodromy of 𝕃\mathbb{L}.

1) ρ\rho is stable for the action of 𝐇{\bf H} if and only if 𝕃\mathbb{L} is irreducible,

2) ρ\rho is polystable for the action of 𝐇{\bf H} if and only if 𝕃\mathbb{L} is reductive.

5.1. Graded local systems

A Stokes local system is a special type of 𝕋\mathbb{T}-graded local system (in the sense of Defn. 30 below), so to clarify the ideas we will focus on them here—the results for Stokes local systems follow almost immediately.

Let SS be a connected real oriented surface of finite topological type. Let S\mathbb{H}\subset S be an open subset, and let 𝕋\mathbb{T}\to\mathbb{H} be a local system of complex tori over \mathbb{H}. We allow the fibres of 𝕋\mathbb{T} to have different dimensions in different components of \mathbb{H}. Fix a connected complex reductive group GG, and a local system 𝒢S\mathcal{G}\to S of groups, such that each fibre of 𝒢\mathcal{G} is isomorphic to GG. We will assume throughout that 𝒢\mathcal{G} is “Out-finite”. This means that the monodromy of 𝒢\mathcal{G} has finite image in Out(G)\mathop{\rm Out}(G). In more detail, given a basepoint bSb\in S and a framing G𝒢bG\cong\mathcal{G}_{b} of 𝒢\mathcal{G} at bb, then the monodromy representation f:π1(S,b)Aut(G)f:\pi_{1}(S,b)\to\operatorname{\mathop{\rm Aut}}(G) of 𝒢\mathcal{G} is such that the monodromy group f(π1(S,b))Aut(G)f(\pi_{1}(S,b))\subset\operatorname{\mathop{\rm Aut}}(G) has finite image in Out(G)\mathop{\rm Out}(G). Of course if GG is semisimple then Out(G)\mathop{\rm Out}(G) is finite and so this is no restriction.

Recall that a 𝒢\mathcal{G}-local system over SS is a local system 𝕃S\mathbb{L}\to S which is a 𝒢\mathcal{G}-torsor (cf. e.g. [13] §2.1), and it determines a local system Aut(𝕃)S\operatorname{\mathop{\rm Aut}}(\mathbb{L})\to S of groups (each fibre of which is also isomorphic to GG).

Definition 30.

A 𝕋\mathbb{T}-graded 𝒢\mathcal{G}-local system over SS is a 𝒢\mathcal{G}-local system 𝕃S\mathbb{L}\to S together with an embedding

𝕋Aut(𝕃)|\mathbb{T}\hookrightarrow\operatorname{\mathop{\rm Aut}}(\mathbb{L})\big{|}_{\mathbb{H}}

of local systems of groups over \mathbb{H}.

For brevity we will simply call this a “graded local system on SS”, and write 𝕋Aut(𝕃)\mathbb{T}\hookrightarrow\operatorname{\mathop{\rm Aut}}(\mathbb{L}) for the grading. A Stokes local system 𝕃\mathbb{L} (in the sense of [8, 13]) is a special type of 𝕋\mathbb{T}-graded local system, taking S=Σ~S=\widetilde{\Sigma} to be the auxiliary surface, \mathbb{H} to be the union of the halos, and 𝕋\mathbb{T} to be the image in Aut(𝕃)|\operatorname{\mathop{\rm Aut}}(\mathbb{L})\bigl{|}_{\mathbb{H}} of the exponential torus 𝒯\mathcal{T}. (Note that 𝕋\mathbb{T} is determined just by the irregular class of 𝕃\mathbb{L}, in the sense of [13] §3.5, since as explained there the class determines the finite rank local system II\subset\mathcal{I} of lattices, and 𝕋\mathbb{T} is the local system of tori with character lattice II.)

5.2. Galois group

If 𝕃S\mathbb{L}\to S is a 𝕋\mathbb{T}-graded local system, and bSb\in S is a basepoint, define Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) to be the Zariski closure of the group generated by the monodromy of 𝕃\mathbb{L} and all the tori 𝕋\mathbb{T} (after transporting them to bb). In more detail, first identify 𝒢bG,𝕃b\mathcal{G}_{b}\cong G,\mathbb{L}_{b}\cong\mathcal{F} (the trivial GG-torsor, as in [13] §2) and define ΓAut(G)\Gamma\subset\operatorname{\mathop{\rm Aut}}(G) to be the group generated by the monodromy of 𝒢\mathcal{G} and Inn(G)\operatorname{Inn}(G) (as in §2 above). If γπ1(S,b)\gamma\in\pi_{1}(S,b) let ρ(γ)G(f(γ))GΓPerm()\rho(\gamma)\in G(f(\gamma))\subset G\ltimes\Gamma\hookrightarrow\operatorname{Perm}(\mathcal{F}) be the monodromy of 𝕃\mathbb{L} around γ\gamma (where Perm()\operatorname{Perm}(\mathcal{F}) is the group of all permutations of the fibre \mathcal{F}). Similarly if p,t𝕋pAut(𝕃)pGp\in\mathbb{H},t\in\mathbb{T}_{p}\subset\operatorname{\mathop{\rm Aut}}(\mathbb{L})_{p}\cong G and χ\chi is any path in SS from bb to pp, consider the element

(6) C1tCAut()=G=G(Id)GΓ,C^{-1}tC\in\operatorname{\mathop{\rm Aut}}(\mathcal{F})=G=G(\text{\rm Id})\subset G\ltimes\Gamma,

where CC is the transport of 𝕃\mathbb{L} along χ\chi.

Definition 31.

The differential Galois group Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) of 𝕃\mathbb{L} is the Zariski closure of the subgroup of GΓG\ltimes\Gamma generated by ρ(π1(S,b))\rho(\pi_{1}(S,b)) and all of the tori (6) (as p,t,χp,t,\chi vary).

It follows that Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) acts on GG by group automorphisms, via the adjoint action of GΓG\ltimes\Gamma on G=G(Id)G=G(\text{\rm Id}). Let Gal¯(𝕃)ΓAut(G)\overline{\mathop{\rm Gal}}(\mathbb{L})\subset\Gamma\subset\operatorname{\mathop{\rm Aut}}(G) be the resulting image of Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}). Up to isomorphism the affine algebraic group Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) and its action on GG do not depend on the choice of basepoint bb or framings.

5.3. Irreducible graded local systems

Define a graded local system 𝕃S\mathbb{L}\to S to be reducible if Aut(𝕃)\operatorname{\mathop{\rm Aut}}(\mathbb{L}) has a sublocal system of proper parabolic subgroups, containing 𝕋\mathbb{T}. In other words there is a sublocal system 𝒫Aut(𝕃)\mathcal{P}\subset\operatorname{\mathop{\rm Aut}}(\mathbb{L}) such that 1) each fibre 𝒫b\mathcal{P}_{b} is a proper parabolic subgroup of Aut(𝕃)b\operatorname{\mathop{\rm Aut}}(\mathbb{L})_{b}, and 2) the grading 𝕋Aut(𝕃)\mathbb{T}\hookrightarrow\operatorname{\mathop{\rm Aut}}(\mathbb{L}) factors through 𝒫\mathcal{P}. Such 𝕃\mathbb{L} is irreducible if it is not reducible.

Lemma 32.

𝕃\mathbb{L} is reducible if and only if Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) preserves a proper parabolic subgroup of GG (recalling that Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) naturally acts on GG by group automorphisms).

Proof. Suppose PGAut(𝕃b)P\subset G\cong\operatorname{\mathop{\rm Aut}}(\mathbb{L}_{b}) is preserved by Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}). Then PP is the fibre at bb of a local system of parabolic subgroups 𝒫Aut(𝕃)\mathcal{P}\subset\operatorname{\mathop{\rm Aut}}(\mathbb{L}), since the monodromy of Aut(𝕃)\operatorname{\mathop{\rm Aut}}(\mathbb{L}) is given by the adjoint action of the monodromy of 𝕃\mathbb{L}, which is in Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}). Moreover 𝕋𝒫\mathbb{T}\hookrightarrow\mathcal{P}, since the transport to bb of each fibre of 𝕋\mathbb{T} is in GG and preserves PP (so is in PP, since NG(P)=PN_{G}(P)=P). The converse is similar, taking the fibre at bb of 𝒫Aut(𝕃)\mathcal{P}\subset\operatorname{\mathop{\rm Aut}}(\mathbb{L}). \square

This can be related to (twisted) reductions of structure group as follows (compare [13] Defn. 11).

Definition 33.

Suppose 𝕃\mathbb{L} is a 𝕋\mathbb{T}-graded 𝒢\mathcal{G}-local system.

\bullet A reduction of 𝕃\mathbb{L} is a 𝕋\mathbb{T}-graded 𝒫\mathcal{P}-local system S\mathbb{P}\to S (for some local system of groups 𝒫S\mathcal{P}\to S), such that 𝕃\mathbb{L} is a twisted pushout of \mathbb{P}. This means that there is a 𝒢\mathcal{G} local system 𝕄\mathbb{M} with an embedding 𝒫Aut(𝕄)\mathcal{P}\hookrightarrow\operatorname{\mathop{\rm Aut}}(\mathbb{M}), together with an isomorphism 𝕃×𝒫𝕄\mathbb{L}\cong\mathbb{P}\times_{\mathcal{P}}\mathbb{M} (of graded 𝒢\mathcal{G} local systems),

\bullet A reduction \mathbb{P} of 𝕃\mathbb{L} is a parabolic reduction if the fibres of 𝒫\mathcal{P} embed as parabolic subgroups of the fibres of Aut(𝕄)\operatorname{\mathop{\rm Aut}}(\mathbb{M}) (each of which is isomorphic to GG),

\bullet Similarly it is a Levi reduction if the fibres of 𝒫\mathcal{P} embed as Levi subgroups of parabolic subgroups of the fibres of Aut(𝕄)\operatorname{\mathop{\rm Aut}}(\mathbb{M}),

\bullet The reduction is proper if the fibres of 𝒫\mathcal{P} embed as proper subgroups.

Lemma 34.

𝕃\mathbb{L} is reducible if and only if it has a proper parabolic reduction of structure group.

Proof. Given 𝒫Aut(𝕃)\mathcal{P}\subset\operatorname{\mathop{\rm Aut}}(\mathbb{L}) with 𝕋𝒫|\mathbb{T}\hookrightarrow\mathcal{P}\bigl{|}_{\mathbb{H}}, then taking =𝒫\mathbb{P}=\mathcal{P} and 𝕄=𝕃\mathbb{M}=\mathbb{L} gives the desired reduction. Conversely given 𝕄,𝒫,\mathbb{M},\mathcal{P},\mathbb{P} then Aut()\operatorname{\mathop{\rm Aut}}(\mathbb{P}) gives the desired parabolic sublocal system in Aut(𝕃)\operatorname{\mathop{\rm Aut}}(\mathbb{L}). \square

Note that part 1) of Thm. 29 now follows immediately from Cor. 28, noting that Gal(𝕃)=Gal(ρ)\mathop{\rm Gal}(\mathbb{L})=\mathop{\rm Gal}(\rho).

This irreducibility condition can also be spelt out in terms of Stokes representations and compatible systems of parabolics (as in [8] §9).

If 𝒢\mathcal{G} is constant then we can use the usual (simpler) notion of reduction of structure group (then GalG\mathop{\rm Gal}\subset G and we don’t need twisted reductions, i.e. we can take 𝕄\mathbb{M} to be the trivial GG-torsor).

5.4. Reductive/semisimple graded local systems

If 𝕃S\mathbb{L}\to S is a graded local system and S\mathcal{L}\to S is a Levi reduction of 𝕃\mathbb{L} (in the sense of Defn. 33) then \mathcal{L} is itself a graded local system, and so we can ask if \mathcal{L} is reducible or not. Define 𝕃\mathbb{L} to be reductive (or “semisimple”) if it has an irreducible Levi reduction. Similarly to Lemma 34 one can rephrase this in terms of Aut(𝕃)\operatorname{\mathop{\rm Aut}}(\mathbb{L}):

Lemma 35.

𝕃\mathbb{L} is reductive if and only if there is a sublocal system Aut(𝕃)\mathcal{E}\subset\operatorname{\mathop{\rm Aut}}(\mathbb{L}) such that 1) each fibre p\mathcal{E}_{p} is a Levi subgroup of a parabolic of Aut(𝕃)p\operatorname{\mathop{\rm Aut}}(\mathbb{L})_{p}, 2) 𝕋\mathbb{T}\subset\mathcal{E}, and 3) \mathcal{E} is irreducible in the sense that it has no proper sublocal systems of parabolic subgroups, containing 𝕋\mathbb{T}.

Recall Gal¯(𝕃)ΓAut(G)\overline{\mathop{\rm Gal}}(\mathbb{L})\subset\Gamma\subset\operatorname{\mathop{\rm Aut}}(G) is the image of Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) in Γ\Gamma.

Proposition 36.

𝕃\mathbb{L} is reductive if and only if Gal¯(𝕃)\overline{\mathop{\rm Gal}}(\mathbb{L}) is a linearly reductive group.

Proof. By Prop. 17 Gal¯(𝕃)\overline{\mathop{\rm Gal}}(\mathbb{L}) is a linearly reductive group if and only if there is a subgroup LGL\subset G such that 1) LL is a Levi subgroup of a parabolic subgroup of GG, 2) LL is preserved by the action of Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}), and 3) LL has no proper parabolic subgroups that are Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) invariant. (This uses the fact that the centralisers of tori in GG are exactly the Levi subgroups of parabolics.) As in Lem. 32 the existence of such LL is the same as 𝕃\mathbb{L} having an irreducible Levi reduction. \square

Part 2) of Thm. 29 now follows immediately from Cor. 25.

Note that if 𝒢\mathcal{G} is constant (or has finite monodromy) then this is the same as Gal(𝕃)\mathop{\rm Gal}(\mathbb{L}) being linearly reductive.

Remark 37.

Note that if 𝕃\mathbb{L} is a Stokes 𝒢\mathcal{G}-local system then it makes no difference if we insist on only looking at reductions that are Stokes local systems: i.e. 𝕃\mathbb{L} is semisimple if it has a Levi reduction to an irreducible Stokes \mathcal{L}-local system, and 𝕃\mathbb{L} is reducible if it has a parabolic reduction to a Stokes 𝒫\mathcal{P}-local system. To see this we need to check 1) that the Stokes conditions on the monodromies (of the reductions) around the tangential punctures are automatic, and 2) that there is no loss of generality in assuming that the local systems of parabolic/Levi subgroups are untwisted around the tangential punctures. This is now an easy exercise: 1) follows since the Stokes groups are controlled by 𝕋\mathbb{T}, and 2) follows by considering the proof of Lem. 34, and the analogous proof of Lem. 35.

Remark 38.

In the case where 𝒢\mathcal{G} is a constant general linear group with fibre G=GLn()G={\mathop{\rm GL}}_{n}(\mathbb{C}) for some nn, then a Stokes 𝒢\mathcal{G}-local system 𝕃\mathbb{L} is equivalent to a Stokes local system 𝕍\mathbb{V} of rank nn vector spaces, as in [11]. Then 𝕃\mathbb{L} is irreducible if and only if 𝕍\mathbb{V} has no nontrival proper Stokes sublocal systems, and it is reductive if and only if 𝕍=𝕍i\mathbb{V}=\bigoplus\mathbb{V}_{i} is the direct sum of some irreducible Stokes local systems 𝕍i\mathbb{V}_{i}.

Note that there are thus many simple criteria to ensure points of \mathcal{R} are stable, and they will be studied systematically elsewhere. For example in the constant GLn(){\mathop{\rm GL}}_{n}(\mathbb{C}) setting, all the Stokes local systems on 𝚺{\bf\Sigma} are irreducible if at one of the punctures the irregular class just has one Stokes circle II\subset\mathcal{I} with Ram(I)=n\text{Ram}(I)=n (e.g. if slope(I)=k/n\mathop{\rm slope}(I)=k/n with (k,n)=1(k,n)=1 this is Katz’s irreducibility criterion [21] (2.2.8)).

References

  • [1] O. Biquard and P. P. Boalch, Wild non-abelian Hodge theory on curves, Compositio Math. 140 (2004), no. 1, 179–204.
  • [2] G. D. Birkhoff, The generalized Riemann problem for linear differential equations and allied problems for linear difference and qq-difference equations, Proc. Amer. Acad. Arts and Sci. 49 (1913), 531–568.
  • [3] P. P. Boalch, Symplectic manifolds and isomonodromic deformations, Adv. in Math. 163 (2001), 137–205.
  • [4] by same author, Stokes matrices, Poisson Lie groups and Frobenius manifolds, Invent. Math. 146 (2001), 479–506.
  • [5] by same author, Quasi-Hamiltonian geometry of meromorphic connections, Duke Math. J. 139 (2007), no. 2, 369–405, (Beware section 6 of the published version is not in the 2002 arXiv version).
  • [6] by same author, Through the analytic halo: Fission via irregular singularities, Ann. Inst. Fourier (Grenoble) 59 (2009), no. 7, 2669–2684, Volume in honour of B. Malgrange.
  • [7] by same author, Habilitation memoir, Université Paris-Sud 12/12/12, (arXiv:1305.6593), 2012.
  • [8] by same author, Geometry and braiding of Stokes data; Fission and wild character varieties, Annals of Math. 179 (2014), 301–365.
  • [9] by same author, Poisson varieties from Riemann surfaces, Indag. Math. 25 (2014), no. 5, 872–900.
  • [10] by same author, Wild character varieties, meromorphic Hitchin systems and Dynkin diagrams, (2018), Geometry and Physics: A Festschrift in honour of Nigel Hitchin, pp.433–454, arXiv:1703.10376.
  • [11] by same author, Topology of the Stokes phenomenon, Integrability, Quantization, and Geometry. I, Proc. Sympos. Pure Math., vol. 103, Amer. Math. Soc., 2021, pp. 55–100.
  • [12] P. P. Boalch, J. Douçot, and G. Rembado, Twisted local wild mapping class groups: configuration spaces, fission trees and complex braids, 2022, arXiv:2209.12695.
  • [13] P. P. Boalch and D. Yamakawa, Twisted wild character varieties, arXiv:1512.08091, 2015.
  • [14] A. Borel, Groupes linéaires algébriques, Ann. Math. (2) 64 (1956), 20–82.
  • [15] A. Borel and J.-P. Serre, Théorèmes de finitude en cohomologie galoisienne, Comment. Math. Helv. 39 (1964), 111–164.
  • [16] M. Brion, On extensions of algebraic groups with finite quotient, Pacific J. Math. 279 (2015), no. 1-2, 135–153.
  • [17] J. Douçot, G. Rembado, and M. Tamiozzo, Local wild mapping class groups and cabled braids, 2022, arXiv:2204.08188.
  • [18] J. Douçot and G. Rembado, Topology of irregular isomonodromy times on a fixed pointed curve, 2022, arXiv:2208.02575.
  • [19] B. Dubrovin, Geometry of 2D topological field theories, Integrable Systems and Quantum Groups (M.Francaviglia and S.Greco, eds.), vol. 1620, Springer Lect. Notes Math., 1995, pp. 120–348.
  • [20] Pengfei Huang and Hao Sun, Meromorphic parahoric Higgs torsors and filtered Stokes G-local systems on curves, 2022, arXiv:2212.04939.
  • [21] N. M. Katz, On the calculation of some differential Galois groups, Invent. Math. 87 (1987), no. 1, 13–61.
  • [22] M. Loday-Richaud, Stokes phenomenon, multisummability and differential Galois groups, Ann. Inst. Fourier 44 (1994), no. 3, 849–906.
  • [23] A. Lubotzky and A. R. Magid, Varieties of representations of finitely generated groups, vol. 58, no. 336, Memoirs of the AMS, 1985.
  • [24] J. Martinet and J.P. Ramis, Elementary acceleration and multisummability, Ann. Inst. Henri Poincaré, Physique Théorique 54 (1991), no. 4, 331–401.
  • [25] J.-P. Ramis, Phénomène de Stokes et filtration Gevrey sur le groupe de Picard-Vessiot, C. R. Acad. Sci. Paris Sér. I Math. 301 (1985), no. 5, 165–167.
  • [26] R. W. Richardson, Conjugacy classes of nn-tuples in Lie algebras and algebraic groups, Duke Math. J. 57 (1988), no. 1, 1–35.
  • [27] C. Sabbah, Harmonic metrics and connections with irregular singularities, Ann. Inst. Fourier 49 (1999), no. 4, 1265–1291.
  • [28] Cheng Shu, On character varieties with non-connected structure groups, 2019, arXiv:1912.04360.

IMJ-PRG, Université Paris Cité and CNRS, Bâtiment Sophie Germain, 8 Place Aurélie Nemours, 75205 Paris, France.
[email protected]https://webusers.imj-prg.fr/~philip.boalch

Department of Mathematics, Faculty of Science Division I, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan.
[email protected]