Polystability of Stokes representations and differential Galois groups
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 , [7] §4).
First we will recall the basic statements in the tame case. Let be a connected complex reductive group, such as , and let be a smooth complex algebraic curve. Thus for some smooth compact complex algebraic curve (i.e. a compact Riemann surface), and a finite subset . Given a basepoint one can consider the representation variety
which is a complex affine variety equipped with an action of , conjugating representations. A point is a -representation, i.e. a group homomorphism . In turn the -character variety, or Betti moduli space
is the (affine) geometric invariant theory quotient of by . By definition this means that the points of the variety are the closed orbits in . The representations whose 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 be the Zariski closure (adhérence) of the image of . A theorem of Schlesinger implies that if is the monodromy representation of an algebraic connection on an algebraic principal -bundle , with regular singularities at , then is the differential Galois group of . 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 is polystable if and only if is a linearly reductive group. Further, is stable if and only if is not contained in a proper parabolic subgroup of .
A representation whose Galois group is linearly reductive is often called a semisimple representation111Indeed if and then is a semisimple representation if and only if is a semisimple -module.. Thus the theorem says that is polystable if and only if is semisimple, and so the points of are the isomorphism classes of semisimple representations. Note that for 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 on an algebraic principal -bundle , has an invariant, its irregular class, at each point . 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 | |||
fundamental group (or groupoid) | |||
fundamental group representation |
Once these generalisations are understood then the story proceeds similarly to the tame case (defining a wild representation variety 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 . 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 can be written explicitly in terms of a product of simpler pieces (doubles or fission spaces ), one for each marked point or handle on :
if has genus and 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 set-up is in [11] §13.)
Let be a smooth compact complex algebraic curve (i.e. a compact Riemann surface) and let be a finite non-empty subset. Let be a connected complex reductive group.
Choose a -irregular class at each point (possibly twisted) as in [13] §3.5. Let be the resulting wild Riemann surface with structure group as in [13] §4 (and [8] §8.1, Rmk 10.6).
The topological notion of Stokes local system on 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 encode the algebraic connections on algebraic principal -bundles on the open curve , with irregular class at each .
As in [13] §5, [8] §8, it follows that if we choose suitable basepoints then the wild surface groupoid is well defined, as is the wild representation variety (the space of Stokes representations of ):
It is an affine variety equipped with an action of a complex reductive group .
Any framed Stokes -local system determines a Stokes representation . Two Stokes local systems are isomorphic if and only if their Stokes representations are in the same -orbit in .
The notion of being polystable or stable for the action of on 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 are the polystable (i.e. closed) -orbits.
On the other hand determines the Galois group , 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 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 be a Stokes local system on , and let be its Stokes representation. The following are equivalent:
is polystable,
is linearly reductive,
is a semisimple Stokes local system.
In the tame case (with each irregular class trivial) the groupoid 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 be a Stokes local system on , and let be its Stokes representation. The following are equivalent:
is a stable point of ,
is not contained in a proper parabolic subgroup of ,
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 , with each fibre isomorphic to . Similarly we will establish the analogous results in this fully twisted setting. We will suppose that 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:
which is an affine variety equipped with an action of a complex reductive group . Secondly is not a subgroup of , but rather it comes with a homomorphism , so naturally acts on by automorphisms. The image of in will be denoted . Then we can define irreducibility (§5.3) and semisimplicity (§5.4) for Stokes -local systems, and will prove analogues of the above results:
Theorem 4.
Let be a wild Riemann surface with group . Let be a Stokes -local system on , and let be its twisted Stokes representation. The following three conditions are equivalent:
is a polystable point of ,
is linearly reductive,
is a semisimple Stokes -local system.
Moreover the following three conditions are also equivalent:
is a stable point of ,
does not preserve a proper parabolic subgroup of ,
is an irreducible Stokes -local system.
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.
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 , 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 (so we now have an intrinsic general framework encompassing such examples). Indeed it was by forgetting this twist that the Poisson variety underlying 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 is reductive if it is connected and has trivial unipotent radical. It is linearly reductive if its identity component is reductive.
Let be a linearly reductive group over . Recall that a point of an affine -variety is said to be polystable if the orbit is closed. It is said to be stable if it is polystable and the kernel of the action has finite index in its stabiliser .
Let be a positive integer. For , let be the Zariski closure of the subgroup generated by . In [26], Richardson examined the simultaneous conjugation action of on the product and obtained the following results:
Theorem 5 ([26, Thm. 3.6]).
If is linearly reductive, a point is polystable if and only if is linearly reductive.
Theorem 6 ([26, Thm. 4.1]).
If is reductive, a point is stable if and only if is not contained in any proper parabolic subgroup of .
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 is reductive and choose . Let be the subgroup of generated by the inner automorphism group and . We assume that the quotient is finite and regard as an algebraic group with identity component . For any let denote the -bitorsor/“twisted group” , as in [13] §2 (as explained there, the monodromy of a -local system lies in such a space, and embeds in the group of set-theoretic automorphisms of a fibre). Put
on which acts by the simultaneous conjugation. For , let be the Zariski closure (in the algebraic group ) of the subgroup generated by . Let be the image of under the homomorphism
where . Note that is contained in and hence naturally acts on . Thus in turn, via , the group naturally acts on .
Theorem 7.
A point is polystable if and only if is linearly reductive.
Remark 8.
Proposition 9.
Let be a linearly reductive group with identity component , and let be a closed subgroup of . Let be the image of under the map , and regard it as a quotient algebraic group of . Then is linearly reductive if and only if is linearly reductive.
Proof. Note that we may view as a map of algebraic groups:
may have an infinite number of components, but will only encounter a finite number of them.
Let be the kernel of the homomorphism
.
Note ,
and the identity component
of is contained in .
Thus ,
which implies is a torus and so is linearly reductive.
Now [26] 1.2.2 implies is linearly reductive if and only if is linearly reductive.
Lemma 10.
Let be a linearly reductive group with identity component and let be a point in an affine -variety . Then the following hold.
-
(1)
is polystable for the -action if and only if is polystable for the -action.
-
(2)
is stable for the -action if and only if is stable for the -action.
Proof. The -orbit is a disjoint union
of a finite number of -orbits of the same dimension,
one of which is .
Hence is closed if and only if is closed.
The second assertion follows from
the equality for the stabilisers
and a similar one for the kernels.
Proof (of Theorem 7). Since is finite, we can find a finite subgroup of such that thanks to a result of Borel–Serre and Brion (see [15, 16]). For each take so that . Then we have isomorphisms of bitorsors
which induce a -equivariant isomorphism
Since is equivariant, a point is stable (resp. polystable) if and only if is stable (resp. polystable). Also, observe that () and hence for all . Therefore without loss of generality we may assume that for all .
Put . It is a linearly reductive group
with identity component and is a closed subvariety of .
By Theorem 5, a point is polystable
with respect to the
simultaneous -conjugation if and only if
is linearly reductive.
By Prop. 9 this happens if and only if is linearly reductive.
Together with Lemma 10 this
implies the assertion.
Note that in general it really is necessary to work with rather than :
Lemma 11.
There are examples of which are polystable but is not linearly reductive.
Proof. Write .
Choose generating a Zariski dense subgroup of an abelian
unipotent subgroup , such as a root group.
Then define to be the inner automorphism
.
It follows that
is not reductive, but
is trivial (and so is polystable by Thm. 7).
Theorem 12.
For any point the following hold:
-
(1)
is polystable if and only if any -invariant parabolic subgroup has an -invariant Levi subgroup .
-
(2)
is stable if and only if there are no proper -invariant parabolic subgroups of .
We prepare three lemmas.
Lemma 13.
Let be a parabolic subgroup of and be a Levi subgroup of . Then .
Proof. Let be the quotient of by the unipotent radical.
Suppose that normalises and decompose it
as , , .
Then for any we have and hence .
Since we have ,
which implies since the restriction of to is injective.
Thus commutes with .
On the other hand, coincides with its centraliser
since for some torus .
Thus and hence , i.e. .
Lemma 14.
The kernel of the -action on is equal to .
Proof. If lies in the kernel, then for any and we have , i.e.
Taking to be we obtain ().
Thus for all , i.e., .
Since is generated by
and trivially acts on we obtain .
The converse is clear.
Lemma 15.
for any .
Proof. Take any . For each we have , which trivially acts on . Thus for any we have
which implies .
The inclusion is clear.
Proof (of Theorem 12). As in the proof of Theorem 7, we may assume that , are all contained in a common finite subgroup such that and put . Under this assumption is linearly reductive if and only if is linearly reductive, by Prop. 9.
(1) We should show that a subgroup is linearly reductive if and only if any -invariant parabolic in has an -invariant Levi subgroup. First suppose that is linearly reductive and is an -invariant parabolic subgroup. Since is linearly reductive and contained in , it is contained in some Levi subgroup of . Then is a Levi subgroup of and normalised by (so it is -invariant). Conversely, suppose any -invariant parabolic in has an -invariant Levi subgroup. By [26, Prop. 2.6], there exists a one-parameter subgroup of such that and , where
(1) | ||||
(2) |
and is the unipotent radical of . Put , which is a parabolic subgroup of . Since is normalised by , it is normalised by . Hence has an -invariant Levi subgroup by assumption. We have and hence (recall that the unipotent radical is connected). Note that is also contained in , while any non-trivial element of does not normalise the Levi subgroup by Lemma 13. Thus is trivial, i.e. is linearly reductive.
(2) Suppose that is stable and let be a -invariant parabolic subgroup of . By Theorem 7, (and hence ) is linearly reductive. Since normalises , there exists a Levi subgroup of containing . By [26, Prop. 2.4], there exists a one-parameter subgroup of such that and . Since each commutes with and hence . The stability now implies that is contained in the kernel and hence . Conversely, suppose that is not stable. Then if the orbit is not closed we argue as follows: By the Hilbert–Mumford criterion, there exists a one-parameter subgroup of and an element such that and is closed. We show that the parabolic subgroup is -invariant and proper. For any and , the limit of
as exists.
Hence is -invariant.
If is not proper,
is contained in
.
Thus () and hence
, which contradicts
the assumption that is not closed.
Hence is proper.
Finally suppose is closed, but of the wrong dimension.
Then the stabiliser is linearly reductive
([26] 1.3.3) and
the quotient has non-trivial
identity component.
Hence there exists a one-parameter subgroup of
such that .
Since each commutes with ,
the parabolic subgroup is -invariant.
It is proper since Lemma 15 implies
.
Let us rephrase/abstract the first part of this in a way that will be useful later. Let be a reductive group and an algebraic group acting on by (algebraic) group automorphisms. Suppose that the action of is effective and the identity component acts by inner automorphisms. Then we may regard as a subgroup of and its image in is finite.
Proposition 16.
is linearly reductive if and only if any -invariant parabolic subgroup of has a -invariant Levi subgroup.
Proof. By the result of Borel–Serre and Brion,
there exists a finite subgroup such that
.
Put and let
be the preimage of under ,
so that Prop. 9 implies
is linearly reductive if and only if
is linearly reductive.
Thus the equivalence follows from
the proof of Thm. 12 (1).
In the same setting there is a further characterisation:
Proposition 17.
is linearly reductive if and only if there exists a torus such that is -invariant and has no proper -invariant parabolic subgroups.
Proof. Suppose that is linearly reductive. We show that there exists a decreasing sequence of -invariant closed subgroups
such that each is a proper parabolic subgroup of , each () is a Levi subgroup of and has no proper -invariant parabolic subgroups. If has no proper -invariant parabolic subgroups we have nothing to do (just put ). Otherwise we take any proper -invariant parabolic subgroup . Then by Prop. 16 there exists a -invariant Levi subgroup . Let be the quotient of by the kernel of the induced -action on , so that effectively acts on . Note that its identity component acts by inner automorphisms of ; indeed, the action of any element of is induced from some inner automorphism of preserving and hence is inner by Lemma 13. Since is linearly reductive is also linearly reductive, and a subgroup of is -invariant if and only if it is -invariant as a subgroup of . If has no proper -invariant parabolic subgroup, the sequence is as desired. Otherwise we take any proper -invariant parabolic subgroup . Then by Prop. 16 there exists a -invariant Levi subgroup . Iterating this procedure, we obtain a desired decreasing sequence. Since each () is a Levi subgroup of there exists a torus such that . Then is the common centraliser of in . Hence the torus generated by (note that they commute with each other) is as desired.
Conversely, suppose that there exists a torus
such that is -invariant
and has no proper -invariant parabolic subgroups.
Let be a -invariant parabolic subgroup.
Then the intersection
is a -invariant parabolic subgroup of
and hence , i.e. .
Since is reductive, there exists a Levi subgroup containing .
For any the image of is
also a Levi subgroup of containing . Since
contains a maximal torus of and any maximal torus of is contained
in a unique Levi subgroup, we have .
Hence is -invariant, which together with Prop.
16
shows that is linearly reductive.
Applying this to the set-up of the present section (with ) yields:
Corollary 18.
The following are equivalent:
0) A point is polystable,
1) The group is linearly reductive,
2) Any -invariant parabolic in has an -invariant Levi subgroup,
3) There exists a subtorus such that is -invariant and has no proper -invariant parabolic subgroups,
4) There exists an -invariant Levi subgroup of a parabolic of , such that has no proper -invariant parabolic subgroups.
Note that 3) and 4) are trivially equivalent since centralisers of tori in 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 ), but now further choose an integer and tori . Write , let and . We allow some of the to be a point, in which case . In this section we will study the stability and polystability for the action of on
(3) |
given by
where
Thus if and we recover the situation of Thm. 7. The case and arbitrary but each was studied by Richardson in [26] Thm. 13.2,14.1 (taking acting on by conjugation). More generally, in effect, [8] Cor. 9.6 studied the notion of stability in the case with arbitrary and each , making the link to the differential Galois group of irregular connections (whence the are the Ramis/exponential tori).
For let be the Zariski closure of the subgroup generated by
Let be the image of in , as before. Recall that a subset of is -invariant if it is preserved by .
Theorem 19.
(1) A point is polystable for the action if and only if is linearly reductive.
(2) A point is stable for the action if and only if there are no proper -invariant parabolic subgroups in .
As in the last section, one can rephrase polystability in several different ways:
Corollary 20.
A point is polystable if and only if
1) The group is linearly reductive, or
2) Any -invariant parabolic in has an -invariant Levi subgroup, or
3) There exists a subtorus such that is -invariant and has no proper -invariant parabolic subgroups.
Proof (of Thm. 19). Let act on via
where (Up to relabelling and incrementing this is the special case where .) Consider the -equivariant map taking to
It expresses as a principal -bundle over (the fibres are exactly the -orbits). Any point has a unique lift with . It follows that the orbit of is closed if and only if the orbit of is closed in .
Now choose so that generates a Zariski dense subgroup of for each . In particular . Consider the simultaneous conjugation action of on , and the -equivariant embedding
Thus is identified with a closed subvariety of and is a -equivariant principal -bundle over . It follows that the orbit of is closed in if and only if the orbit of is closed in .
Hence part (1) of the theorem follows from Thm. 7 (applied to ). To deal with stability we need to consider the stabilisers and the kernels of the actions.
Lemma 21.
The stabiliser of any point for the -action is canonically isomorphic to the stabiliser of the point for the -action.
Proof. Suppose fixes
and let
() and
.
Note because .
Since commutes with each ,
each commutes with , which implies .
We have () by the definition of and
() as centralises .
Hence the pair stabilises
,
and hence stabilises .
Conversely if
fixes , then let .
It follows immediately that
fixes and thus that fixes
.
Clearly the two correspondences are inverses of each other.
As in Lemma 14 one has:
Lemma 22.
(1) The kernel of the -action on (or ) is the -invariant subgroup of the center of , and (2) The kernel of the -action on is the subgroup of elements satisfying .
4. Application to wild character varieties
Recall from [8, 13] that the wild character variety is determined by an irregular curve/wild Riemann surface with group , where is a compact smooth complex algebraic curve, is a non-empty finite subset, is a local system of groups over the punctured curve (with each fibre isomorphic to some fixed connected complex reductive group ), and consists of the data of an irregular class at each point (in the sense of [13] §3.5—it is the class of a graded local system).
As in [13] §4, then determines an auxiliary surface , equipped with boundary circles , halos , and tangential punctures . Further determines a local system of finite dimensional complex tori, the Ramis tori ([13] p.9). Choosing a finite set of basepoints (with one point in each component circle, as in [13] §4.1) then determines the wild surface groupoid , the fundamental groupoid of the auxiliary surface with these basepoints, as in [13] §5. The local system is determined by a map and this determines the space of -twisted representations
of , as in [13] §5. Since is just a punctured real surface with boundary, choosing generating paths in yields an isomorphism of spaces, for some integer , so it is a smooth affine variety. In turn the wild representation variety is the closed subvariety of cut out by the two Stokes conditions, in [13] Defn. 18. Intrinsically, is the moduli space of framed Stokes local systems, as in [13] Prop. 19, framed via a graded isomorphism to a standard fibre at each basepoint ([13] §4.1). The group acts transitively on the set of framings at , where . Thus the group acts naturally on . The wild character variety is the affine geometric invariant theory quotient , and so its points are the closed orbits in . This leads directly to the key statement:
Proposition 23.
The wild representation variety may be embedded in an -equivariant way, as a closed subvariety of the space of (3), for suitable and automorphisms , with .
Proof. This comes down to considering the inclusion
as a closed subvariety
(forgetting the Stokes conditions, as in [13] Defn. 18),
and then identifying
in an -equivariant way.
Both of these are straightforward.
In particular
for generators
of , and for paths from to , for .
In order to define the Galois group of we will assume that is “Out-finite”, in the sense that the monodromy group of has finite image in . Then the group generated by and is an algebraic group, as in §2 above. Thus any Stokes representation takes values in the algebraic group , and so we can consider the Zariski closure of its monodromy. The Galois group is defined by adding the Ramis tori as well: If and is any path in from to , consider the element
(4) |
where .
Definition 24.
The differential Galois group of is the Zariski closure of the subgroup of generated by and all of the tori (4) (as vary).
It follows that acts on by group automorphisms, via the adjoint action of on . Let be the resulting image of . 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 is polystable for the action of if and only if is a linearly reductive group.
Special cases include:
If has finite monodromy then is polystable if and only if is a linearly reductive group.
If is a constant general linear group then is polystable if and only if is the direct sum of irreducible Stokes representations.
Recall that is the invariant subgroup of the centre of , and it embeds diagonally in . To deal with stability we will avoid degenerate cases by assuming:
(5) |
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 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 compact Riemann surface is a (very special) symplectic leaf of the character variety of , for any point .
Lemma 26.
Suppose and has trivial irregular class, and if then . Then is a smooth non-empty affine variety and the kernel K of the action is embedded diagonally in .
Proof. It is nonempty as it is the fusion of some fission spaces and some internally fused doubles: Recall that can be described as the quasi-Hamiltonian -reduction
where each is a (twisted) internally fused double ([13] p.23, [8] Thm 8.2). Since is the double of , it follows that
so that is the product of some smooth nonempty affine varieties. Note that still acts on and this includes .
Suppose and suppose (as usual) that the framings of are such that the monodromy in of is trivial along the
chosen paths .
Then acts on as .
Taking implies .
If then the fact that all are fixed implies
.
On the other hand if then so
looking at
there is so that
for all .
This implies and .
Thus in all case is central.
Then looking at any loop based at leads to a relation of the form .
Thus since is central this implies
, so .
Remark 27.
Note that it follows in general (as in [8] Thm. 8.2) that if is nonempty (and ) then it is a smooth affine variety.
Part (2) of Theorem 12 then implies:
Corollary 28.
A Stokes representation is stable for the action of if and only if there is no proper parabolic subgroup stabilised by the action of .
Proof. This follows from Thm. 12
since is closed in ,
matches up with ,
and the kernel of the action on and on is the same.
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 is a Stokes local system and is the monodromy of .
1) is stable for the action of if and only if is irreducible,
2) is polystable for the action of if and only if is reductive.
5.1. Graded local systems
A Stokes local system is a special type of -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 be a connected real oriented surface of finite topological type. Let be an open subset, and let be a local system of complex tori over . We allow the fibres of to have different dimensions in different components of . Fix a connected complex reductive group , and a local system of groups, such that each fibre of is isomorphic to . We will assume throughout that is “Out-finite”. This means that the monodromy of has finite image in . In more detail, given a basepoint and a framing of at , then the monodromy representation of is such that the monodromy group has finite image in . Of course if is semisimple then is finite and so this is no restriction.
Recall that a -local system over is a local system which is a -torsor (cf. e.g. [13] §2.1), and it determines a local system of groups (each fibre of which is also isomorphic to ).
Definition 30.
A -graded -local system over is a -local system together with an embedding
of local systems of groups over .
For brevity we will simply call this a “graded local system on ”, and write for the grading. A Stokes local system (in the sense of [8, 13]) is a special type of -graded local system, taking to be the auxiliary surface, to be the union of the halos, and to be the image in of the exponential torus . (Note that is determined just by the irregular class of , in the sense of [13] §3.5, since as explained there the class determines the finite rank local system of lattices, and is the local system of tori with character lattice .)
5.2. Galois group
If is a -graded local system, and is a basepoint, define to be the Zariski closure of the group generated by the monodromy of and all the tori (after transporting them to ). In more detail, first identify (the trivial -torsor, as in [13] §2) and define to be the group generated by the monodromy of and (as in §2 above). If let be the monodromy of around (where is the group of all permutations of the fibre ). Similarly if and is any path in from to , consider the element
(6) |
where is the transport of along .
Definition 31.
The differential Galois group of is the Zariski closure of the subgroup of generated by and all of the tori (6) (as vary).
It follows that acts on by group automorphisms, via the adjoint action of on . Let be the resulting image of . Up to isomorphism the affine algebraic group and its action on do not depend on the choice of basepoint or framings.
5.3. Irreducible graded local systems
Define a graded local system to be reducible if has a sublocal system of proper parabolic subgroups, containing . In other words there is a sublocal system such that 1) each fibre is a proper parabolic subgroup of , and 2) the grading factors through . Such is irreducible if it is not reducible.
Lemma 32.
is reducible if and only if preserves a proper parabolic subgroup of (recalling that naturally acts on by group automorphisms).
Proof. Suppose is preserved by .
Then is the fibre at of a local system of parabolic subgroups , since the monodromy of is given by the adjoint action of the monodromy of , which is in
.
Moreover , since the transport to
of each fibre of is in and preserves (so is in , since ).
The converse is similar, taking the fibre at of
.
This can be related to (twisted) reductions of structure group as follows (compare [13] Defn. 11).
Definition 33.
Suppose is a -graded -local system.
A reduction of is a -graded -local system (for some local system of groups ), such that is a twisted pushout of . This means that there is a local system with an embedding , together with an isomorphism (of graded local systems),
A reduction of is a parabolic reduction if the fibres of embed as parabolic subgroups of the fibres of (each of which is isomorphic to ),
Similarly it is a Levi reduction if the fibres of embed as Levi subgroups of parabolic subgroups of the fibres of ,
The reduction is proper if the fibres of embed as proper subgroups.
Lemma 34.
is reducible if and only if it has a proper parabolic reduction of structure group.
Proof. Given with ,
then taking and gives the desired reduction.
Conversely given then
gives the desired parabolic
sublocal system in .
This irreducibility condition can also be spelt out in terms of Stokes representations and compatible systems of parabolics (as in [8] §9).
If is constant then we can use the usual (simpler) notion of reduction of structure group (then and we don’t need twisted reductions, i.e. we can take to be the trivial -torsor).
5.4. Reductive/semisimple graded local systems
If is a graded local system and is a Levi reduction of (in the sense of Defn. 33) then is itself a graded local system, and so we can ask if is reducible or not. Define to be reductive (or “semisimple”) if it has an irreducible Levi reduction. Similarly to Lemma 34 one can rephrase this in terms of :
Lemma 35.
is reductive if and only if there is a sublocal system such that 1) each fibre is a Levi subgroup of a parabolic of , 2) , and 3) is irreducible in the sense that it has no proper sublocal systems of parabolic subgroups, containing .
Recall is the image of in .
Proposition 36.
is reductive if and only if is a linearly reductive group.
Proof. By Prop. 17 is a linearly reductive group if and only if there is a subgroup such that 1) is a Levi subgroup of a parabolic subgroup of , 2) is preserved by the action of , and 3) has no proper parabolic subgroups that are invariant.
(This uses the fact that the centralisers of tori in are exactly the Levi subgroups of parabolics.)
As in Lem. 32 the existence of such is the same as having an irreducible Levi reduction.
Note that if is constant (or has finite monodromy) then this is the same as being linearly reductive.
Remark 37.
Note that if is a Stokes -local system then it makes no difference if we insist on only looking at reductions that are Stokes local systems: i.e. is semisimple if it has a Levi reduction to an irreducible Stokes -local system, and is reducible if it has a parabolic reduction to a Stokes -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 , and 2) follows by considering the proof of Lem. 34, and the analogous proof of Lem. 35.
Remark 38.
In the case where is a constant general linear group with fibre for some , then a Stokes -local system is equivalent to a Stokes local system of rank vector spaces, as in [11]. Then is irreducible if and only if has no nontrival proper Stokes sublocal systems, and it is reductive if and only if is the direct sum of some irreducible Stokes local systems .
Note that there are thus many simple criteria to ensure points of are stable, and they will be studied systematically elsewhere. For example in the constant setting, all the Stokes local systems on are irreducible if at one of the punctures the irregular class just has one Stokes circle with (e.g. if with 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 -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 -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]