The Deligne–Simpson problem for connections on with a maximally ramified singularity
Abstract.
The classical additive Deligne–Simpson problem is the existence problem for Fuchsian connections with residues at the singular points in specified adjoint orbits. Crawley-Boevey found the solution in 2003 by reinterpreting the problem in terms of quiver varieties. A more general version of this problem, solved by Hiroe, allows additional unramified irregular singularities. We apply the theory of fundamental and regular strata due to Bremer and Sage to formulate a version of the Deligne–Simpson problem in which certain ramified singularities are allowed. These allowed singular points are called toral singularities; they are singularities whose leading term with respect to a lattice chain filtration is regular semisimple. We solve this problem in the special case of connections on with a maximally ramified singularity at and possibly an additional regular singular point at infinity. Examples of such connections arise from Airy, Bessel, and Kloosterman differential equations. They play an important role in recent work in the geometric Langlands program. We also give a complete characterization of all such connections which are rigid, under the additional hypothesis of unipotent monodromy at infinity.
Key words and phrases:
Deligne-Simpson problem, meromorphic connections, irregular singularities, moduli spaces, parahoric subgroups, fundamental strata, toral connections, rigid connections2020 Mathematics Subject Classification:
34M50, 14D05 (Primary); 22E67, 34M35, 14D24, 20G25 (Secondary)1. Introduction
1.1. The classical Deligne–Simpson problem
A fundamental concern in the study of meromorphic connections is the existence problem for connections with specified singularities. More precisely, this problem poses the question: given points in and formal connections , does there exist a meromorphic connection which is regular away from the ’s and satisfies for all ? The classical Deligne–Simpson problem is a variant of this problem for Fuchsian connections.
From now on, we assume that the underlying vector bundles of all connections on are trivializable. Without loss of generality, we assume that the collection of singular points does not include . A Fuchsian connection with singular points is defined by
where for all . Note that the adjoint orbit of determines the formal isomorphism class at . Since is not a singularity, the residue theorem forces . We say that the collection of matrices is irreducible if they have no common invariant subspaces besides and . We can now state the (additive) Deligne–Simpson problem:
Given adjoint orbits , determine whether there exists an irreducible -tuple with satisfying [Kos03].
In other words, when is there an irreducible Fuchsian connection with residues in the given orbits? Note that the original problem considered by Deligne and Simpson was the multiplicative version, where one looks for Fuchsian connections with monodromies in specified conjugacy classes in [Sim91]. The additive version stated above was originally formulated by Kostov, who solved it in the “generic” case. Crawley-Boevey gave a complete solution by reinterpreting the problem in terms of quiver varieties [CB03]. We remark that while there is an obvious analogue of this problem for arbitrary reductive , little is known about the solution outside of type .
1.2. The unramified Deligne–Simpson problem
In order to generalize the Deligne–Simpson problem to allow for irregular singularities, one considers connections with higher order principal parts at the singularities:
(1) |
Again, we assume that is not a singular point, so . We now require that the singularity at each has a certain specified form called a ”formal type”.
Most previous work on the irregular Deligne–Simpson problem has restricted attention to the ”unramified case” [Kos10, Boa08, HY14, Hir17]. This means that at each singularity, the slope decomposition of the corresponding formal connection only involves integer slopes. More concretely, each such formal connection has Levelt–Turrittin (LT) normal form
(2) |
where the ’s are diagonal, , and the residue term is upper triangular and commutes with each . We view the -form as an unramified formal type. In the regular singular case, one can take to be in Jordan canonical form and view the formal type as .
For Fuchsian connections, the principal part at a singular point is just the residue. Hence, in the classical Deligne–Simpson problem, one requires that the principal part agrees with the formal type after conjugation by a constant matrix (i.e., an element of ). In other words, the principal part lies in the adjoint orbit of the formal type. For unramified formal types of positive slope, one instead requires the principal part to lie in the orbit of the formal type under a certain action of the group . Let denote the space of principal parts of order at most . The group acts on by conjugation followed by truncation at the residue term. Note that this action factors through the finite-dimensional group . If is an unramified formal type of slope , we call the orbit under this action the truncated orbit of . If has slope , may be identified with the usual adjoint orbit of .
We can now state the unramified irregular Deligne–Simpson problem: Given points and unramified formal types of slope , determine when there exists an irreducible connection as in (1) whose principal part at each lies in . This problem can also be restated in the language of moduli spaces. Given the formal types , one can consider the moduli space of “framable” connections on a rank trivial bundle whose singularities have the specified formal types [HY14]. The construction generalizes that of Boalch [Boa01], who assumes that the are all nonresonant, i.e., that the leading term of each is regular semisimple. This moduli space is not necessarily well-behaved, but it is a complex manifold if one restricts to the stable moduli space, i.e., the open subset consisting of irreducible connections. The unramified Deligne–Simpson problem is simply the question of when such a stable moduli space is nonempty.
This problem was solved in 2017 by Hiroe [Hir17], building on earlier work of Boalch [Boa08] and Hiroe and Yamakawa [HY14]. As in the Fuchsian case, the proof involves quiver varieties. Hiroe uses the collection of unramified formal types to define a certain quiver variety and identifies the stable moduli space with a certain open subspace of the quiver variety. He then finds necessary and sufficient conditions for this open subspace to be nonempty. As a corollary, Hiroe shows that the stable moduli space is a connected manifold as long as it is nonempty.
1.3. The ramified Deligne–Simpson problem for toral connections
In this paper, we introduce the study of the ramified Deligne–Simpson problem, where ramified singularities are allowed. A singularity is called ramified if the associated formal connection can only be expressed in LT normal form after passing to a ramified cover. The LT normal form is thus no longer a suitable notion of formal type for ramified singularities. It is possible to formulate the ramified Deligne–Simpson problem by replacing the LT normal form with a “rational canonical form” for connections. Such a form may be obtained from Sabbah’s refined Levelt–Turrittin decomposition [Sab08]; we will discuss this in a future paper.
Here, we only sketch the setup of the ramified Deligne–Simpson problem for a special class of irregular connections called toral connections. Roughly speaking, a formal connection is called toral if its leading term with respect to an appropriate filtration satisfies a graded version of regular semisimplicity. (The precise definition involves the theory of fundamental and regular strata for connections introduced by Bremer and Sage [BS13b, BS18, BS13a].) The terminology reflects the fact that toral connections can be “diagonalized” into a (not necessarily split) Cartan subalgebra of the loop algebra.
First, we describe formal types for toral connections. A rank toral connection has slope , where is a divisor of or and . If , define by ; i.e., is the matrix with ’s in each entry of the superdiagonal, in the lower-left entry, and ’s elsewhere. If , set . Note that . Given such a with (resp. ), we define a block-diagonal Cartan subalgebra (resp. ). There is a natural -filtration induced by assigning degree to . Let denote the corresponding maximal torus in the loop group.
An -formal type of slope (with ) is a -form , where has regular semisimple term in degree and no terms in positive degree. It is a fact that any toral connection of slope is formally isomorphic to a connection with an -formal type; the formal type is unique up to an action of the relative affine Weyl group of [BS13b, BS13a].
In the case of unramified toral connections, is the usual diagonal maximal torus, and the -formal types of slope are those connection matrices in LT normal form (2) with regular (so that is necessarily ). At the opposite extreme, is a ”Coxeter maximal torus”.111Under the bijection between classes of maximal tori in and conjugacy classes in the Weyl group [KL88], corresponds to the Coxeter class consisting of -cycles. The -formal types of slope are the -forms , where is a polynomial of degree .
The unramified Deligne–Simpson problem involves global connections which satisfy a stronger condition than just having specified formal types at the singularities. One also needs the local isomorphisms transforming the matrices of the formal connections into the given formal types to satisfy a global compatibility condition called “framability”. We now explain how this condition can be generalized to toral formal types.
Recall (see, e.g., [Sag00, BS13b]) that the parahoric subgroups of are the local field analogues of the parabolic subgroups of . A parabolic subgroup is the stabilizer of a partial flag of subspaces in , and a parahoric subgroup is the stabilizer of a ”lattice chain” of -lattices in . If is a parahoric subgroup with associated lattice chain , then there is an associated ”lattice chain filtration” on defined by .
To each maximal torus , there is a unique ”standard parahoric subgroup” with the property that the corresponding filtration is compatible with the filtration on , in the sense that for all [BS13b]. In the unramified case, we have . For the Coxeter maximal torus , the corresponding parahoric subgroup is the standard “Iwahori subgroup” ; i.e., is the preimage of the upper-triangular Borel subgroup (consisting of all upper-triangular matrices in ) via the map induced by the “evaluation at zero” map .
The most natural way of describing framability involves coadjoint orbits. One can view the principal part at of a connection as a continuous functional on via . Similarly, an -formal type can be viewed as a functional on . The global connection then is framable at with respect to the -formal type at if for some global trivialization, the restriction to of the principal part at lies in the -coadjoint orbit of . (See Definition 5.1.)
However, one can also give a description more reminiscent of the definition in the unramified case.
Definition 1.1.
Let be a global connection on with a singular point at , and let be a toral formal type of slope . We say that is framable at with respect to if
-
(1)
under some global trivialization , the matrix form satisfies , and ; and
-
(2)
there exists an element such that the nonpositive truncation of equals .
Recall that acts simply transitively on the space of global trivializations. If one starts with a fixed trivialization , then the choice of trivialization in the definition above corresponds to an element ; i.e., there is a unique such that . This matrix is called a compatible framing (or simply, a framing) of at . Framability with respect to a formal type at an arbitrary point is defined similarly, by simply replacing by if is finite, and by if .
We can now state the Deligne–Simpson problem for connections whose irregular singularities are all toral. Note that the statement below can easily be extended to allow for arbitrary unramified singular points.
Toral Deligne–Simpson Problem.
Let be a collection of toral formal types at the points , and let be a collection of adjoint orbits at other points . Does there exist an irreducible rank connection such that
-
(1)
is regular away from the ’s and ’s;
-
(2)
is framable at with respect to the formal type ; and
-
(3)
is regular singular at with residue in ?
If such a connection exists, we call it a “framable connection” with the given formal types.
This problem can be restated in terms of moduli spaces of connections. Suppose that each is nonresonant; i.e., suppose that no pair of the eigenvalues of the orbit differ by a nonzero integer. Further assume that , so that there is at least one irregular singular point. In [BS13b], Bremer and Sage constructed the moduli space of connections satisfying all the above hypotheses except irreducibility. Let be the subset of the moduli space consisting of irreducible connections. In this language, the toral Deligne–Simpson problem poses the question of when is nonempty.
1.4. Coxeter connections
We now restrict attention to a simple special case: connections with a maximally ramified irregular singularity and (possibly) an additional regular singular point. Without loss of generality, we will view such connections as connections on with the irregular singularity at . Following [KS21b], we refer to such connections as Coxeter connections. Well-known classical examples arise from the Airy differential equation and a modified version of the Bessel equation.222In these connections and others described below, the irregular singularity is at . Another important class of examples consists of the generalized Kloosterman connections studied by Katz [Kat88, Kat90]. These hypergeometric connections are the geometric incarnations of certain exponential sums called Kloosterman sums, which are of great importance in number theory.
Coxeter connections and their -connection analogues (for a simple algebraic group) have played a significant role in recent work in the geometric Langlands program. For example, Frenkel and Gross [FG09], building on work of Deligne [Del70] and Katz [Kat88, Kat96], constructed a rigid -connection of this type. This connection, which may be viewed as a -version of a modified Bessel connection, was the first connection with irregular singularities for which the geometric Langlands correspondence was understood explicitly [HNY13, Zhu17]. This connection also arises in Lam and Templier’s proof of mirror symmetry for minuscule flag varieties [LT17]. Other examples include the Airy -connection and more general rigid “Coxeter -connections” constructed in [KS21b]. The Airy -connection and its -adic analogue have also been studied in [JKY21].
Recall that if is a rank formal connection, then every slope of has denominator (when the slope is expressed in lowest form) between and . We say is maximally ramified if all such denominators (or equivalently, at least one) is . In this case, all the slopes are the same — say with — and equal to the slope of the connection. More concretely, the leading term of the LT normal form is of the form with a constant diagonal matrix, and is necessarily regular.
It is shown in [KS21a] that maximally ramified connections are toral connections with respect to a Coxeter maximal torus. Thus any maximally ramified connection of slope has a rational canonical form , where is a polynomial of degree , and the set of formal types is given by . Moreover, any such connection is irreducible.
In this paper, we solve the ramified Deligne–Simpson problem for Coxeter connections. More precisely, let be a maximally ramified formal type, and let be an adjoint orbit (which we will always assume to be nonresonant). We determine necessary and sufficient conditions for the existence of a meromorphic connection on which is framable at with formal type , is regular singular with residue in at , and is otherwise nonsingular. Note that any such connection is automatically irreducible, since its formal connection at is irreducible. Thus, in the language of moduli spaces, we determine when is nonempty.
In order to state our result, we need some facts about adjoint orbits. Fix a monic polynomial of degree with the ’s distinct complex numbers. The set of matrices with characteristic polynomial is a closed subset of which is stable under conjugation. We denote the set of orbits with characteristic polynomial by . This set is partially ordered under the usual Zariski closure ordering: if and only if . The theory of the Jordan canonical form makes it clear that can be identified with the Cartesian product , where denotes the set of partitions of . Moreover, this identification defines a poset isomorphism between the closure ordering and the direct product of the dominance orders.
Given positive integers and , there exists a unique smallest partition with at most parts. Define to be the orbit in corresponding to the element
This tuple of partitions is the (unique) smallest element of such that each component partition has at most parts. Note that is just the regular orbit in . On the other extreme, if for all (as is the case when ), then is the semisimple orbit, the unique minimal orbit in . Let denote the principal filter generated by in , i.e., satisfies if and only if . This filter is proper unless for all . As we will see in Theorem 2.3, the collection of orbits for each fixed satisfies a generalization of one characterization of regular orbits.
We can now give the solution to the Deligne–Simpson problem for Coxeter connections. Given a rank maximally ramified formal type and a monic polynomial of degree that is nonresonant (i.e., no two roots differ by a nonzero integer), let
Theorem 5.4.
Let and be positive integers with , let be a maximally ramified formal type of slope , and let with distinct modulo . Then
In other words, given , then is nonempty if and only if and . Concretely, the condition means that has at most Jordan blocks for each eigenvalue.
Note that the solution depends only on the slope and the residue of the formal type.
Remark 1.2.
If for all , then as long as the trace condition is satisfied. In particular, this is the case if .
There is an obvious analogue of this problem for -connections (as opposed to -connections). Here, maximally ramified formal types are of the form with and . Thus, the trace condition becomes vacuous, and the Deligne–Simpson problem has a positive solution if and only if .
One can define Coxeter -connections for any simple group (or for any reductive group with connected Dynkin diagram) [KS21b]. For such a , Coxeter toral connections have slope , where is the Coxeter number for and . Moreover, there is an analogue of the Deligne–Simpson problem in this more general context. We restrict to the case where the regular singularity at has nilpotent residue (and thus has unipotent monodromy).
Conjecture 1.3.
Let be a simple complex group with Lie algebra . Fix a Coxeter -formal type of slope with . Then there exists a nilpotent orbit such that the Deligne–Simpson problem for Coxeter -connections with initial data and the nilpotent orbit has a positive solution if and only if . Moreover, if , then , so the Deligne–Simpson problem always has a positive solution.
1.5. Rigidity
Our results have applications to the question of when Coxeter connections are rigid. Let be a nonempty open set, and let denote the inclusion. A -connection on is called physically rigid if it is uniquely determined by the formal isomorphism class at each point of . It is called cohomologically rigid if . For irreducible connections, cohomological rigidity implies that has no infinitesimal deformations. For , it is a result of Bloch and Esnault that cohomological and physical rigidity are the same [BE04].
In [KS21b], Kamgarpour and Sage investigated the question of rigidity for “homogeneous” Coxeter -connections with unipotent monodromy. A homogeneous Coxeter -formal type of slope (i.e., for , a formal type of the form with ) gives rise to a Coxeter -connection on with nilpotent residue at infinity. They determined precisely when these connections are (cohomologically) rigid, thus generalizing the work of Frenkel and Gross [FG09]. For , it turns out that such connections are rigid precisely when divides or .
We can now generalize the results of [KS21b] to give a classification of rigid Coxeter connections in type .
Theorem 6.1.
Let be a rank maximally ramified formal type of slope , and let be any nilpotent orbit with . Then there exists a rigid connection with the given formal type and and unipotent monodromy determined by if and only if and .
We expect that the analogous statement is true for Coxeter -connections.
1.6. Organization of the paper
In §2, we discuss some facts about the poset of adjoint orbits in that will be needed in our applications. In particular, we introduce and characterize a sequence of orbits which generalize regular orbits. In §3, we provide a brief review of lattice chain filtrations. In §4, we describe the role of these filtrations in studying formal connections, following earlier work of Bremer and Sage [BS13b, BS12, Sag17]. In particular, we discuss toral connections and characterize maximally ramified formal connections as Coxeter toral connections. In §5, we describe moduli spaces of connections with toral singularities and then state and prove our main result on the Deligne–Simpson problem for Coxeter connections. We conclude the paper in §6 by characterizing rigid Coxeter connections with unipotent monodromy at the regular singular point.
Acknowledgements
The authors are deeply grateful to the American Institute of Mathematics (AIM) for hosting and generously supporting their research during SQuaRE meetings in 2020 and 2021. Many of the key ideas in this paper were conceived during these meetings. The authors would also like to thank an anonymous referee for helpful comments and suggestions that served to improve the manuscript.
2. The poset of adjoint orbits
The solution to the Deligne–Simpson problem for Coxeter connections involves certain distinguished orbits for the adjoint action of the general linear group on (i.e., similarity classes of complex matrices). We will need some facts about these adjoint orbits.
The set of adjoint orbits in is partially ordered via the closure order: if . Let be the map sending a matrix to its characteristic polynomial. Given a monic degree polynomial , is closed and -stable. If we let be the set of adjoint orbits in , it is immediate that as a poset,
The theory of Jordan canonical forms allows us to identify the posets with posets involving partitions. If is a positive integer, a partition of is a nonincreasing sequence of positive integers that sum to . Each integer appearing in this sequence is called a part of . The total number of parts is denoted by . It will sometimes be convenient to use exponential notation for partitions: if the ’s are the distinct parts, each appearing with multiplicity , we will also denote by the multiset . Let be the set of partitions of . We view as a poset via the dominance order
(3) |
Write for distinct and . The set can be identified with , where the partition of is given by the sizes of the Jordan blocks with eigenvalue . It is well-known that the closure order corresponds to the product of the dominance orders under this identification.
The unique maximal orbit in is the orbit with a single Jordan block for each eigenvalue. This is the regular orbit with characteristic polynomial , i.e., the unique orbit in of codimension . We now define a sequence of orbits in which generalize the regular orbit.
Fix , and consider the subset consisting of orbits with at most Jordan blocks for each eigenvalue. It is immediate from (3) that is a filter in the poset . This means that if and , then . It turns out that is a principal filter; i.e., there exists a (unique) element such that . We will also set ; it is a filter in . Note that is the set of all regular (or equivalently, maximal) adjoint orbits.
Proposition 2.1.
Given , write with and . Then the partition
is the (unique) smallest partition of with at most parts.
Proof.
Let be any partition of with at most parts, say with biggest part and smallest part . We will show that there exists a strictly smaller partition with at most parts unless .
If , then , the smallest element of , so the statement is trivial. We thus may assume that . First, suppose that . It follows that , so one obtains a strictly smaller partition with parts by replacing one with and adjoining a new part with value . We may thus assume without loss of generality that .
Next, suppose that . Define a partition with the same parts as except one is replaced by and one is replaced by . It is obvious that and that is strictly bigger than .
It remains to consider the case , so for some with . We then have , so and . Thus, . ∎
Now, define to be the orbit in corresponding to the element
Corollary 2.2.
The filter is principal with generator , i.e., .
Proof.
It suffices to show that the corresponding filter in is principal with generator . This is immediate from the proposition. ∎
Recall that the semisimple orbit in is the unique minimal orbit. If for all (as is the case when ), then is the semisimple orbit so .
We can now give a Lie-theoretic interpretation of which will be important in our applications. Let be the set of matrices with nonzero entries on, and ’s below, the th subdiagonal:
Note that if . It is well-known that every element of is regular and that every regular orbit has a representative in . (It is a famous result of Kostant that the analogous statement holds for any complex simple group [Kos59].)
We now prove a generalization of this result.
Theorem 2.3.
The adjoint orbits which intersect are precisely the orbits in . The minimal such orbits are the ’s.
We begin by proving that intersects . Let be the matrix with ’s on the th subdiagonal and ’s elsewhere. We usually omit from the notation.
Proposition 2.4.
Fix a positive integer . Let , where the eigenvalue appears with multiplicity . Let be any matrix with all entries on the th subdiagonal nonzero and all other entries . Then .
Proof.
It is easy to see that there exists an invertible diagonal matrix such that , so we may assume without loss of generality that . We prove the proposition by induction on the number of distinct eigenvalues (for arbitrary ).
If , the partition for the single eigenvalue is . Indeed, the Jordan strings for are given by for , with the strings of length for and otherwise.
Now assume . Let be the span of . Note that stabilizes ; in fact, , where . It is clear that the Jordan strings for are also Jordan strings for corresponding to the eigenvalues . Also, induces the endomorphism on . It remains to show that each Jordan string for on lifts to a Jordan string for the zero eigenvalues of . (Here, is the smallest integer such that .) Let . This map is invertible, so is the desired lift. ∎
Proof of Theorem 2.3.
We first show that consists of elements whose Jordan forms have at most blocks for each eigenvalue. Let , and let be any eigenvalue of . The bottom rows of the matrix are linearly independent, so . We conclude that . Since this dimension is the number of Jordan blocks for the eigenvalue , the claim follows.
We have shown that the set of adjoint orbits intersecting is contained in . Now take . Let be its characteristic polynomial, so . Take any as in Proposition 2.4. Then . By a theorem of Krupnik [Kru97, Theorem 1], there exists a strictly upper triangular matrix such that . Since , this proves the theorem. ∎
Remark 2.5.
Krupnik’s approach does not lead to explicit constructions of orbit representatives in . However, at least for nilpotent orbits, it is possible to find explicit, simple representatives by means of an algorithm described in [KLM+21].
3. Lattice chain filtrations
Let denote the field of formal Laurent series, and let denote the ring of formal power series. An -lattice in is a finitely generated -module with the property that . A lattice chain in is a collection of lattices satisfying the following properties:
-
(1)
for all ; and
-
(2)
there exists a positive integer , called the period, such that for all .
A parahoric subgroup is the stabilizer of a lattice chain; i.e., . We write to denote the period of the lattice chain stabilized by . A lattice chain in is called complete if its period is as large as possible, i.e., if its period equals . The parahoric subgroups associated to complete lattice chains are called Iwahori subgroups.
Each lattice chain — say with corresponding parahoric — determines a filtration of defined by ; in particular, . One also gets a filtration of defined by and for . In the special case that each lattice in the lattice chain stabilized by admits an -basis of the form , the corresponding filtration is induced by a grading on .
Remark 3.1.
The “lattice chain filtrations” described above may also be obtained through a more general construction. Let be the (reduced) Bruhat–Tits building associated to the loop group . This is a simplicial complex whose simplices are in bijective correspondence with the parahoric subgroups of the loop group. For each point in the building, there is an associated “Moy–Prasad filtration” on the loop algebra [MP94]. Up to rescaling, the lattice chain filtration determined by a parahoric subgroup is the Moy–Prasad filtration associated to the barycenter of the simplex in the Bruhat–Tits building corresponding to [BS18].
The building is the union of -dimensional real affine spaces called apartments, which are in one-to-one correspondence with split maximal tori in . The parahoric subgroups described in the ”special case” above — i.e., the parahoric subgroups where each lattice in the lattice chain stabilized by admits an -basis of the form — are precisely the parahoric subgroups corresponding to the simplices in the ”standard apartment”, i.e., the apartment corresponding to the diagonal torus.
We will focus our attention on two particular parahoric subgroups of : and the ”standard Iwahori subgroup” (defined below). The parahoric subgroup is the stabilizer of the -periodic lattice chain . The associated filtration on is the degree filtration .
The standard Iwahori subgroup is the stabilizer of the standard complete lattice chain in , i.e., the lattice chain
where is the standard basis for . The associated filtration is the standard Iwahori filtration . Note that if is the Borel subgroup of invertible upper triangular matrices, then is the preimage of under the homomorphism induced by the ”evaluation at zero” map .
Define by ; i.e., is the matrix with ’s in each entry of the superdiagonal, in the lower-left entry, and ’s elsewhere. Then [Bus87, Proposition 1.18] for all . Note that for all . Let denote the standard diagonal maximal torus in . Then consists of all diagonal matrices in . The standard Iwahori grading is then given by
(4) |
Example 3.2.
Let . Then
Some steps in the standard Iwahori filtration on are shown below:
If is any -subspace containing some , we denote the space of continuous linear functionals on by . Every continuous functional on extends to a continuous functional on , so , where denotes the annihilator of . The space of -forms can be identified with the space of functionals by associating a -form with the functional . This identification is well-behaved with respect to lattice chain filtrations [BS18, Proposition 3.6]:
When the filtration comes from a grading, one can be even more explicit. In particular, we get
(5) |
For applications to connections, it is important to consider the relationship between filtrations on Cartan subalgebras of and filtrations on parahoric subalgebras [BS13b, BS13a].
Note that, since is not algebraically closed, it is not true that all maximal tori (or equivalently, all Cartan subalgebras) are conjugate. In fact, there is a bijection between the set of conjugacy classes of maximal tori in and the set of conjugacy classes in the Weyl group for (i.e., the symmetric group ) [KL88, Lemma 2]. We also remark that each Cartan subalgebra comes equipped with a natural filtration (see, e.g., [BS13a, Section 3]). In the cases of interest to us in this paper, the filtration will be induced by a grading .
Let be a maximal torus and let be a parahoric subgroup, both in . Let be the Cartan subalgebra associated to . We say that and (or and ) are compatible (resp. graded compatible) if (resp. ) for all . For present purposes, it suffices to consider two examples: the diagonal subalgebra and the “standard Coxeter Cartan subalgebra” . The diagonal Cartan subalgebra (corresponding to the trivial class in the Weyl group ) is endowed with a filtration which comes from the obvious grading . It is immediate that is graded compatible with .
At the opposite extreme, there is a unique class of maximal tori in that are anisotropic modulo the center, meaning that they have no non-central rational cocharacters. Concretely, such tori are as far from being split as possible. This class corresponds to the Coxeter class in , i.e., the class of -cycles. A specific representative of this class is the standard Coxeter torus with Lie algebra . Note that is regular semisimple — its eigenvalues are the distinct th roots of — so its centralizer is indeed a Cartan subalgebra. The natural grading by powers of on induces a filtration on , and it is clear that is graded compatible with .
4. Toral and maximally ramified connections
4.1. Formal connections
A formal connection of rank is a connection on an -vector bundle of rank over the formal punctured disk . Given a trivialization for (which is always trivializable), the connection can be written in matrix form as , where . The loop group acts simply transitively on the set of trivializations via left multiplication. The corresponding action of on the connection matrix is given by the gauge action: if , then . Hence, the set of isomorphism classes of rank formal connections is isomorphic to the orbit space for the gauge action.
A formal connection is called regular singular if the connection matrix with respect to some trivialization has a simple pole. If the matrix has a higher order pole for every trivialization, is said to be irregular singular. Katz defined an invariant of formal connections called the slope which gives one measure of the degree of irregularity of a formal connection [Del70]. The slope is a nonnegative rational number whose denominator in lowest form is at most . The slope is positive if and only if is irregular.
4.2. Fundamental strata
The classical approach to the study of formal connections involves an analysis of the “leading term” of the connection matrix with respect to the degree filtration on [Was76]. To review, suppose that the matrix for with respect to is expanded with respect to the degree filtration on ; i.e., suppose
(6) |
where and for all . When the leading term is well-behaved, it gives useful information about the connection. For example, if is non-nilpotent, then . Moreover, if and is diagonalizable with distinct eigenvalues, then can be diagonalized into a “-formal type of depth ”. This means that there exists such that is an element of
[Was76]. Note that many interesting connections have nilpotent leading terms. For example, the leading term of the formal Frenkel–Gross connection [FG09] is strictly upper triangular (and thus nilpotent). In fact, the leading term is nilpotent no matter what trivialization one chooses for .
More recently, Bremer and Sage — borrowing well-known tools from representation theory developed by Bushnell [Bus87], Moy–Prasad [MP94], and others — have introduced a more general approach to the study of formal connections, where leading terms are replaced by “strata” [BS13b, BS13a, BS18]. A -stratum is a triple with a parahoric subgroup, a nonnegative integer, and a functional on . Consider the special case where corresponds to a simplex in the standard apartment (see Remark 3.1). Here, a functional can be written uniquely as for homogeneous (i.e., for ). The stratum is called fundamental if is non-nilpotent. A formal connection contains the stratum (with respect to a fixed trivialization) if with and induced by . More general definitions of fundamental strata and stratum containment are given in [BS13b, BS18].
Fundamental strata can be viewed as a generalization of the notion of a non-nilpotent leading term. In particular, fundamental strata can be used to compute the slope of any connection, not merely those with integer slopes. Recall that if is a parahoric subgroup, then denotes the period of the lattice chain stabilized by .
Theorem 4.1 ([BS13b, Theorem 4.10], [Sag17, Theorem 1]).
Any formal connection contains a fundamental stratum. If contains the fundamental stratum , then .
We now investigate some examples. The connection in (6) (with for all ) contains the stratum , which is fundamental if and only if is non-nilpotent. The formal Frenkel–Gross connection contains the fundamental stratum . Moreover, any rank formal connection of the form , with , , and , contains the fundamental stratum , and thus has slope .
4.3. Toral connections
The notion of a diagonalizable leading term with distinct eigenvalues is generalized by the notion of a “regular stratum”. For simplicity, we only consider the case where comes from the standard apartment (see Remark 3.1). General definitions can be found in [BS13b] and [BS13a]. For such a , consider the stratum . If is regular semisimple, then its centralizer is regular semisimple, and we say that is a -regular stratum. A connection that contains an -regular stratum is called an -toral connection.
It turns out that toral connections do not exist for every maximal torus . In fact, an -toral connection of slope exists if and only if corresponds to a regular conjugacy class in (in the sense of Springer [Spr74]) and if is a regular eigenvalue of this conjugacy class [BS13a]. The regular classes are parametrized by the partitions (for positive divisors of ) and (for positive divisors of ). Representatives for each of the corresponding conjugacy classes of maximal tori are given by the ’s defined in §1.3. An -toral connection has slope for some with . Note that is the diagonal torus while is the standard Coxeter torus .
Just as for connections whose naive leading term is regular semisimple, there exist “rational canonical forms” for toral connections involving the notion of a formal type. Fix a divisor of either or . We define the set of -formal types of depth (with ) by
Every toral connection of slope is formally isomorphic to a connection of the form with ; we view this as a rational canonical form for .
We will need a more precise variation of this statement. As mentioned in §1.3, for each , there is a standard parahoric subgroup which is compatible (in fact, graded compatible) with .
4.4. Maximally ramified connections
Definition 4.3.
A formal connection of rank is called maximally ramified if it has slope with .
Thus, a formal connection is maximally ramified if the denominator of the slope (in lowest terms) is as big as possible. Another interpretation involves the slope decomposition of . It is a well-known result of Turrittin [Tur55] and Levelt [Lev75] that after extending scalars to for some , there exists a trivialization in which the matrix of is block-diagonal:
here, the ’s are polynomials and the ’s are nilpotent matrices. This is the Levelt–Turrittin normal form of . The slopes of are the rational numbers , each appearing with multiplicity . This collection of invariants gives more detailed information about how irregular is than the single invariant . Indeed, one can define to be the maximum of the slopes of .
One can show that the slopes are nonnegative rational numbers with denominators at most . Moreover, if at least one slope is with , then all slopes are . Thus, is maximally ramified if at least one slope has denominator .
It turns out that maximally ramified connections are the same thing as Coxeter toral connections. If we specialize our results on formal types to the standard Coxeter torus , we see that the -formal types of depth are given by
We thus obtain the following result on rational canonical forms for maximally ramified connections:
Theorem 4.4 ([KS21a]).
Let be a maximally ramified connection of slope with . Then is formally gauge equivalent to a connection of the form with a polynomial of degree .
This theorem may be obtained as a direct corollary of Sabbah’s refined Levelt–Turrittin decomposition [Sab08, Corollary 3.3]. Indeed, since has slope , Sabbah’s theorem shows that it is formally isomorphic to a connection of the form , where is the -fold covering induced by , has degree , and . It is now easy to conclude that has the desired rational canonical form, with the coefficients of determined by and the coefficients of .
However, the theory of toral connections allows one to prove a generalized version of this theorem for -connections, where is a reductive group with connected Dynkin diagram [KS21a]. Here, is replaced by the Coxeter number , is an appropriate fixed Coxeter torus in , and is formally isomorphic to , where is a -formal type of slope . Below, we provide a concise stratum-theoretic proof for the specific case , which is simpler than the general proof.
Proof.
By Theorem 4.1, contains a fundamental stratum with respect to some trivialization. Since , it follows that and is an Iwahori subgroup. By equivariance of stratum containment and the fact that Iwahori subgroups are all conjugate, we may modify the trivialization so that contains the fundamental stratum . The functional is represented by , where is non-nilpotent.
By (4), for some constants . Let . Since , it follows that . The polynomial thus has distinct roots, and is regular semisimple. Let be a fixed th root of . It is easy to see that there exists such that , and since normalizes , contains the stratum . By Theorem 4.2, is formally isomorphic to , where . ∎
5. The Deligne–Simpson problem for Coxeter connections
5.1. Moduli spaces of connections with toral singularities
We now turn our attention to meromorphic connections on a rank trivializable vector bundle over the complex Riemann sphere . To discuss the Deligne–Simpson problem, we need to define what it means for to be framable at a singularity with respect to a given toral formal type. We will assume that the singular point is . The only modification needed if the singularity is at an arbitrary point is to replace the uniformizer by if is finite and by if .
Fix a trivialization of , and write . The principal part of is an element of , and so may be viewed as a continuous functional on by (5). Similarly, the restriction of to is uniquely determined by the truncation of to . Thus, if is an -formal type, then may naturally be viewed as an element of .
Definition 5.1.
Let be a global connection on with a singular point at , and let be an -formal type of depth . We say that is framable at with respect to if
-
(1)
there exists such that and , and
-
(2)
there exists an element such that .
Fix two disjoint subsets and of with and . Let be a collection of toral formal types at the ’s, and let be a collection of adjoint orbits at the ’s. Assume that all of the orbits are nonresonant, meaning that no two eigenvalues of an orbit differ by a nonzero integer. One can now consider the category of meromorphic connections satisfying the following properties:
-
(1)
has irregular singularities at the ’s, regular singularities at the ’s, and no other singular points;
-
(2)
for each , is framable at with respect to the formal type ; and
-
(3)
for each , has residue at in .
In [BS13b], Bremer and Sage constructed the moduli space of this category as a Hamiltonian reduction of a product over the singular points of certain symplectic manifolds, each of which is endowed with a Hamiltonian action of . At a regular singular point with adjoint orbit , the manifold is just , viewed as the coadjoint orbit . To define the symplectic manifold associated to an -toral formal type , we first remark that the parahoric subgroup is the pullback of a certain standard parabolic subgroup under the map induced by . For example, is the pullback of , and is the pullback of . The “extended orbit” is defined by
The group acts on via , with moment map .
Theorem 5.2 ([BS13b, Theorem 5.26]).
The moduli space is given by
Let be the subset of consisting of irreducible connections. One can now restate the toral Deligne–Simpson problem as
Given the toral formal types and the nonresonant adjoint orbits , determine whether is nonempty.
Note that a -toral connection is irreducible, so if any is -toral, then the Deligne–Simpson problem reduces to the question of whether is nonempty.
5.2. Coxeter connections
We now specialize to an important special case: connections on with a maximally ramified singular point at and (possibly) a regular singularity at . Since maximally ramified formal connections are Coxeter toral, we will follow [KS21b] and refer to such connections as Coxeter connections.
It is possible to give a simpler expression for moduli spaces of Coxeter connections.
Proposition 5.3.
Let be a -formal type, and let be a nonresonant adjoint orbit. Then
Proof.
We can now state the solution to the Deligne–Simpson problem for Coxeter connections. For a given -formal type and a monic polynomial of degree , let
Theorem 5.4.
Let and be positive integers with , let be a maximally ramified formal type of slope , and let with distinct modulo . Then
We prove this theorem in the next subsection.
Remark 5.5.
If we write , then is nonempty if and only if and if has at most Jordan blocks for each eigenvalue. This second condition is always satisfied if . Note that the solution only depends on the slope and the residue of the formal type.
This theorem immediately gives the corresponding result for -connections. (In terms of a global trivialization, one may view an -connection on as an operator with .) In this case, maximally ramified formal types are of the form with and , so the trace condition becomes vacuous. Accordingly, we obtain the following solution to the Deligne–Simpson for Coxeter -connections.
Corollary 5.6.
The moduli space of Coxeter -connections with formal type of slope and adjoint orbit is nonempty if and only if .
The notion of a Coxeter -connection makes sense for any reductive group with connected Dynkin diagram, and there is an analogue of the Deligne–Simpson problem in this context. We give a specific conjecture about this problem (under the additional hypothesis of unipotent monodromy) in the introduction.
5.3. Proof of Theorem 5.4
We begin with some preliminaries on -orbits in .
Lemma 5.7.
For any with relatively prime to , the linear map given by has image .
Proof.
Since , we have Note that — indeed, it is a monomial matrix which represents the th power of a Coxeter element in the Weyl group — so the entries in are a reordering of the entries in . It follows that . To show equality, it suffices to check that . We have if and only if commutes with . Since , the centralizers of and coincide and equal . Thus, . ∎
We can now give a convenient representative of certain coadjoint -orbits in .
Proposition 5.8.
Let and be positive integers with . Suppose is given by for some and . Then there exists such that
(7) |
Proof.
All elements of represent the zero functional on , so it suffices to show that if for , then there exists such that .
Corollary 5.9.
Given as in the proposition and with , there exists such that .
Proof.
Given as in (7), we claim that
To see this, write with . Recall that is the matrix with ’s on the th subdiagonal and ’s elsewhere. Similarly, let be the matrix whose only nonzero entries are ’s on the th superdiagonal. (We make the convention that and .) It is easy to verify that . Since , and , we have . In particular, if and equals if . Applying this to (7) gives as desired.
To complete the proof, it suffices to show that if for some , and if satisfies , then there exists such that . Since , it suffices to find such that . This follows from Lemma 5.7, since . ∎
Lemma 5.10.
Let and be positive integers with . Suppose that and is a nonresonant adjoint orbit in . If , then the Jordan form of has at most blocks for each eigenvalue.
Proof.
This is trivial for , so assume that . Choose and such that and . Write for some and for some . Since , we may assume without loss of generality that . This implies that for some . It is easy to see that has ’s in the th subdiagonal and below. In the notation of §2, this means that . By Theorem 2.3, has at most blocks for each eigenvalue. ∎
Lemma 5.11.
Let and be positive integers with , let be a -formal type of depth , and let be a degree polynomial with distinct modulo . If , then .
Proof.
Proof of Theorem 5.4.
If , then there exists and with . Since , we see that .
Remark 5.12.
At least in the case of unipotent monodromy, it is possible to avoid using Krupnik’s Theorem by giving an explicit construction of an element of the moduli space. We discuss this in [KLM+21].
6. Rigidity for Coxeter connections
Let be a meromorphic -connection on which is regular on the Zariski-open set . Let denote the induced formal -connection at . The connection is called physically rigid if, for any meromorphic -connection which is regular on and satisfies for all , we have .
In general, it is very difficult to determine whether a connection is physically rigid. A more accessible notion is given by cohomological rigidity, which means that , where is the inclusion. If is irreducible, then being cohomologically rigid implies that admits no infinitesimal deformations [Yun14]. For , Bloch and Esnault have shown that cohomological rigidity and physical rigidity are equivalent [BE04].
We call a -formal type homogeneous when it is of the form for ; it gives rise to a “homogeneous Coxeter connection” on . This connection has a toral singularity at and (possibly) a regular singularity at with unipotent monodromy. This notion also makes sense for any complex simple group [KS21b]. Again, one can define formal types with respect to a certain maximal torus called the Coxeter torus. Moreover, if is any positive integer relatively prime to the Coxeter number , there exists an element such that may be viewed as a homogeneous formal type. One can again consider the corresponding Coxeter -connection on with a homogeneous -toral irregular singularity of slope at and (possibly) a regular singular point with unipotent monodromy at . The case is the remarkable rigid connection constructed by Frenkel and Gross [FG09].
In [KS21b], Kamgarpour and Sage determined when these homogeneous Coxeter -connections are (cohomologically) rigid for any simple . It turns out and always give rigid connections: the Frenkel–Gross and “Airy -connection” respectively. For the exceptional groups, there are no other such rigid connections except for in . However, for the classical groups, one also has rigidity for with satisfying certain divisibility conditions. For example, in type , these connections are rigid if and only if .
In this paper, we generalize this result in type to give a classification of rigid framable Coxeter connections with unipotent monodromy at . (Framable means that we only consider Coxeter connections contained in the relevant framable moduli space .)
Theorem 6.1.
Let be a rank maximally ramified formal type of slope , and let be any nilpotent orbit with . Then there exists a rigid connection with the given formal type and unipotent monodromy determined by if and only if and .
If is a Coxeter connection, let denote the global differential Galois group, and let and denote the local differential Galois groups at and . These differential Galois groups are all algebraic subgroups of . Also, let denote the irregularity of the formal connection . This is the sum of all the slopes appearing in the slope decomposition of ; it is a nonnegative integer.
Let , and let be the inclusion. It is shown in [FG09, Proposition 11] that . Thus, we get the numerical criterion for rigidity that is rigid if and only if .
We now calculate the numerical criterion as in Section 4 of [KS21b]. The local differential Galois group is given by , where is a certain torus containing a regular semisimple element and is an order element of [KS21a]. The centralizer of is thus a maximal torus , and represents a Coxeter element in the Weyl group. We conclude (as in [KS21b]) that . Since , we also have .
In general, if is a toral -connection with slope , then by Lemma 19 of [KS19], , where is the set of roots with respect to the maximal torus . In our particular case, we obtain .
Finally, if we fix some element , then is regular singular with unipotent monodromy . This means that , so , the centralizer of . Using the fact that , we obtain
Since , we can take with this local behavior. Suppose that . We then have . It follows that in this case, is never rigid.
Finally, take . It was shown in [KS21b] that if and only if . This finishes the proof of the theorem.
References
- [BE04] S. Bloch and H. Esnault, Local Fourier transforms and rigidity for -modules, Asian J. Math. 8 (2004), 587–605.
- [Boa01] P. Boalch, Symplectic manifolds and isomonodromic deformations, Adv. Math. 163 (2001), 137–205.
- [Boa08] by same author, Irregular connections and Kac–Moody root systems, arXiv:0806:1050, 2008.
- [BS12] C. Bremer and D. S. Sage, Isomonodromic deformations of connections with singularities of parahoric formal type, Comm. Math. Phys. 313 (2012), 175–208.
- [BS13a] by same author, Flat -bundles and regular strata for reductive groups, arXiv:1309.6060, 2013.
- [BS13b] by same author, Moduli spaces of irregular singular connections, Int. Math. Res. Not. IMRN (2013), 1800–1872.
- [BS18] by same author, A theory of minimal -types for flat -bundles, Int. Math. Res. Not. IMRN (2018), 3507–3555.
- [Bus87] C. J. Bushnell, Hereditary orders, Gauss sums and supercuspidal representations of , J. Reine Agnew. Math. 375/376 (1987), 184–210.
- [CB03] W. Crawley-Boevey, On matrices in prescribed conjugacy classes with no common invariant subspace and sum zero, Duke Math. J. 118 (2003), 339–352.
- [Del70] P. Deligne, Équations différentielles à points singuliers réguliers, Lect. Notes Math., vol. 163, Springer-Verlag, 1970.
- [FG09] E. Frenkel and B. Gross, A rigid irregular connection on the projective line, Ann. of Math. (2) 170 (2009), 1469–1512.
- [Hir17] K. Hiroe, Linear differential equations on the Riemann sphere and representations of quivers, Duke Math. J. 166 (2017), 855–935.
- [HNY13] J. Heinloth, B.-C. Ngô, and Z. Yun, Kloosterman sheaves for reductive groups, Ann. of Math. (2) 177 (2013), 241–310.
- [HY14] K. Hiroe and D. Yamakawa, Moduli spaces of meromorphic connections and quiver varieties, Adv. Math. 266 (2014), 120–151.
- [JKY21] K. Jakob, M. Kamgarpour, and L. Yi, Airy sheaves for reductive groups, arXiv:2111.02256, 2021.
- [Kat88] N. M. Katz, Gauss sums, Kloosterman sums, and monodromy groups., Ann. Math. Stud., vol. 116, Princeton University Press, 1988.
- [Kat90] by same author, Exponential sums and differential equations., Ann. Math. Stud., vol. 124, Princeton University Press, 1990.
- [Kat96] by same author, Rigid local systems., Ann. Math. Stud., vol. 139, Princeton University Press, 1996.
- [KL88] D. Kazhdan and G. Lusztig, Fixed point varieties on affine flag manifolds, Israel J. Math. 62 (1988), 129–168.
- [KLM+21] M. Kulkarni, N. Livesay, J. Matherne, B. Nguyen, and D. S. Sage, An example of the ramified Deligne–Simpson problem, in preparation, 2021.
- [Kos59] B. Kostant, The principal three-dimensional subgroup and the Betti numbers of a complex simple lie group, Amer. J. Math 81 (1959), 973–1032.
- [Kos03] V. P. Kostov, On some aspects of the Deligne–Simpson Problem, J. Dynam. Control Systems 9 (2003), 393–436.
- [Kos10] by same author, Additive Deligne–Simpson Problem for non-Fuchsian systems, Funkcial. Ekvac. 53 (2010), 395–410.
- [Kru97] M. Krupnik, Jordan structures of upper equivalent matrices, Linear Algebra Appl. 261 (1997), 167–172.
- [KS19] M. Kamgarpour and D. S. Sage, A geometric analogue of a conjecture of Gross and Reeder, Amer. J. Math. 141 (2019), 1457–1476.
- [KS21a] by same author, Differential Galois groups of -connections, in preparation, 2021.
- [KS21b] by same author, Rigid connections on via the Bruhat–Tits building, Proc. Lond. Math. Soc. 122 (2021), 359–376.
- [Lev75] G. Levelt, Jordan decomposition for a class of singular differential operators, Ark. Mat. 13 (1975), 1–27.
- [LT17] T. Lam and N. Templier, Mirror symmetry for minuscule flag varieties, arXiv:1705.00758, 2017.
- [MP94] A. Moy and G. Prasad, Unrefined minimal -types for -adic groups, Invent. Math. 116 (1994), 393–408.
- [Sab08] Claude Sabbah, An explicit stationary phase formula for the local formal Fourier-Laplace transform, Singularities I, Contemp. Math., vol. 474, Amer. Math. Soc., Providence, RI, 2008, pp. 309–330. MR 2454354
- [Sag00] D. S. Sage, The geometry of fixed point varieties on affine flag manifolds, Trans. Amer. Math. Soc. 352 (2000), 2087–2119.
- [Sag17] by same author, Regular strata and moduli spaces of irregular singular connections, New trends in analysis and interdisciplinary applications, Trends Math. Res. Perspect., Birkhäuser/Springer, Cham, 2017, pp. 69–75.
- [Sim91] C. Simpson, Products of matrices, Differential geometry, global analysis, and topology (Halifax, NS, 1990), CMS Conf. Proc., vol. 12, Amer. Math. Soc., 1991, pp. 157–185.
- [Spr74] T. A. Springer, Regular elements of finite reflection groups, Invent. Math. 25 (1974), 159–198.
- [Tur55] H. L. Turrittin, Convergent solutions of ordinary homogeneous differential equations in the neighborhood of an irregular singular point, Acta. Math. 93 (1955), 27–66.
- [Was76] W. Wasow, Asymptotic expansions for ordinary differential equations, Wiley Interscience, New York, 1976.
- [Yun14] Z. Yun, Rigidity in automorphic representations and local systems, Current developments in mathematics 2013, Int. Press, 2014, pp. 73–168.
- [Zhu17] X. Zhu, Frenkel–Gross’ irregular connection and Heinloth–Ngô–Yun’s are the same, Selecta Math. (N.S.) 23 (2017), 245–274.