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

The Deligne–Simpson problem for connections on 𝔾m\mathbb{G}_{m} with a maximally ramified singularity

Maitreyee C. Kulkarni Mathematical Institute, University of Bonn, Bonn, Germany. [email protected] Neal Livesay Institute for Experiential AI, Northeastern University, Boston, MA. [email protected] Jacob P. Matherne Mathematical Institute, University of Bonn, Bonn, Germany and Max Planck Institute for Mathematics, Bonn, Germany. [email protected] Bach Nguyen Department of Mathematics, Xavier University of Louisiana, New Orleans, LA. [email protected]  and  Daniel S. Sage Department of Mathematics, Louisiana State University, Baton Rouge, LA. [email protected]
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 𝔾m\mathbb{G}_{m} with a maximally ramified singularity at 0 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 connections
2020 Mathematics Subject Classification:
34M50, 14D05 (Primary); 22E67, 34M35, 14D24, 20G25 (Secondary)
M.K. received support from the Charles Simonyi Endowment while at the Institute for Advanced Study, and from the Max Planck Institute for Mathematics in Bonn and the Hausdorff Research Institute for Mathematics in Bonn. J.M. received support from NSF Grant DMS-1638352, the Association of Members of the Institute for Advanced Study, and the Hausdorff Research Institute for Mathematics in Bonn. B.N. received support from an AMS-Simons Travel Grant. D.S.S. received support from Simons Collaboration Grant 637367.

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 a1,,ama_{1},\dots,a_{m} in 1\mathbb{P}^{1} and formal connections ^1,,^m\widehat{\nabla}_{1},\dots,\widehat{\nabla}_{m}, does there exist a meromorphic connection \nabla which is regular away from the aia_{i}’s and satisfies ai^i\nabla_{a_{i}}\cong\widehat{\nabla}_{i} for all ii? 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 1\mathbb{P}^{1} are trivializable. Without loss of generality, we assume that the collection of singular points does not include \infty. A Fuchsian connection with singular points a1,,ama_{1},\ldots,a_{m} is defined by

d+(i=1mAizai)dz,d+\Big{(}\sum_{i=1}^{m}\frac{A_{i}}{z-a_{i}}\Big{)}dz,

where Ai𝔤𝔩n()A_{i}\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) for all ii. Note that the adjoint orbit of AiA_{i} determines the formal isomorphism class at aia_{i}. Since \infty is not a singularity, the residue theorem forces Ai=0\sum A_{i}=0. We say that the collection of matrices A1,,AmA_{1},\dots,A_{m} is irreducible if they have no common invariant subspaces besides {0}\{0\} and n\mathbb{C}^{n}. We can now state the (additive) Deligne–Simpson problem:

Given adjoint orbits 𝒪1,,𝒪m\mathscr{O}_{1},\dots,\mathscr{O}_{m}, determine whether there exists an irreducible mm-tuple (A1,,Am)(A_{1},\dots,A_{m}) with Ai𝒪iA_{i}\in\mathscr{O}_{i} satisfying Ai=0\sum A_{i}=0 [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 GLn()\operatorname{GL}_{n}(\mathbb{C}) [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 G{G}, little is known about the solution outside of type AA.

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) d+(i=1mν=0riAν(i)(zai)ν)dzz.d+\Big{(}\sum_{i=1}^{m}\sum_{\nu=0}^{r_{i}}\frac{A^{(i)}_{\nu}}{(z-a_{i})^{\nu}}\Big{)}\frac{dz}{z}.

Again, we assume that \infty is not a singular point, so i=1mA0(i)=0\sum_{i=1}^{m}A^{(i)}_{0}=0. We now require that the singularity at each aia_{i} 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) d+(Drzr++D1z1+R)dzz,d+(D_{r}z^{-r}+\dots+D_{1}z^{-1}+R)\frac{dz}{z},

where the DiD_{i}’s are diagonal, Dr0D_{r}\neq 0, and the residue term RR is upper triangular and commutes with each DiD_{i}. We view the 11-form (Drzr++D1z1+R)dzz(D_{r}z^{-r}+\dots+D_{1}z^{-1}+R)\frac{dz}{z} as an unramified formal type. In the regular singular case, one can take RR to be in Jordan canonical form and view the formal type as RdzzR\frac{dz}{z}.

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 GLn()\operatorname{GL}_{n}(\mathbb{C})). 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 GLn([[z]])\operatorname{GL}_{n}(\mathbb{C}[\![z]\!]). Let r={(Brzr++B0)dzzBi𝔤𝔩n()}\mathcal{B}_{r}=\{(B_{r}z^{-r}+\dots+B_{0})\frac{dz}{z}\mid B_{i}\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})\} denote the space of principal parts of order at most rr. The group GLn([[z]])\operatorname{GL}_{n}(\mathbb{C}[\![z]\!]) acts on r\mathcal{B}_{r} by conjugation followed by truncation at the residue term. Note that this action factors through the finite-dimensional group GLn([[z]]/zr[[z]])\operatorname{GL}_{n}(\mathbb{C}[\![z]\!]/z^{r}\mathbb{C}[\![z]\!]). If 𝒜\mathscr{A} is an unramified formal type of slope rr, we call the orbit 𝒪𝒜\mathscr{O}_{\mathscr{A}} under this action the truncated orbit of 𝒜\mathscr{A}. If 𝒜\mathscr{A} has slope 0, 𝒪𝒜\mathscr{O}_{\mathscr{A}} may be identified with the usual adjoint orbit of 𝒜/dzz\mathscr{A}/\frac{dz}{z}.

We can now state the unramified irregular Deligne–Simpson problem: Given points aia_{i} and unramified formal types 𝒜i\mathscr{A}_{i} of slope rir_{i}, determine when there exists an irreducible connection as in (1) whose principal part at each aia_{i} lies in 𝒪𝒜i\mathscr{O}_{\mathscr{A}_{i}}. This problem can also be restated in the language of moduli spaces. Given the formal types 𝒜i\mathscr{A}_{i}, one can consider the moduli space of “framable” connections on a rank nn trivial bundle whose singularities have the specified formal types [HY14]. The construction generalizes that of Boalch [Boa01], who assumes that the 𝒜i\mathscr{A}_{i} are all nonresonant, i.e., that the leading term of each 𝒜i\mathscr{A}_{i} 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 nn toral connection has slope r/br/b, where bb is a divisor of nn or n1n-1 and gcd(r,b)=1\gcd(r,b)=1. If b>1b>1, define ωb𝔤𝔩b(((z)))\omega_{b}\in\operatorname{\mathfrak{gl}}_{b}(\mathbb{C}(\!(z)\!)) by ωb=i=1b1ei,i+1+zeb,1\omega_{b}=\sum_{i=1}^{b-1}e_{i,i+1}+ze_{b,1}; i.e., ωb\omega_{b} is the matrix with 11’s in each entry of the superdiagonal, zz in the lower-left entry, and 0’s elsewhere. If b=1b=1, set ωb=z\omega_{b}=z. Note that ωbb=zid\omega_{b}^{b}=z\mathrm{id}. Given such a bb with b=nb\ell=n (resp. b=n1b\ell=n-1), we define a block-diagonal Cartan subalgebra 𝔰b=((ωb))\mathfrak{s}^{b}=\mathbb{C}(\!(\omega_{b})\!)^{\ell} (resp. 𝔰b=((ωb))((z))\mathfrak{s}^{b}=\mathbb{C}(\!(\omega_{b})\!)^{\ell}\oplus\mathbb{C}(\!(z)\!)). There is a natural \mathbb{Z}-filtration 𝔰b=i(𝔰b)i\mathfrak{s}^{b}=\bigcup_{i}(\mathfrak{s}^{b})^{i} induced by assigning degree ii to ωbi\omega_{b}^{i}. Let SbS^{{b}} denote the corresponding maximal torus in the loop group.

An SbS^{{b}}-formal type of slope r/br/b (with gcd(r,b)=1\gcd(r,b)=1) is a 11-form AdzzA\frac{dz}{z}, where A(𝔰b)rA\in(\mathfrak{s}^{b})^{-r} has regular semisimple term in degree r-r and no terms in positive degree. It is a fact that any toral connection of slope r/br/b is formally isomorphic to a connection d+Adzzd+A\frac{dz}{z} with AdzzA\frac{dz}{z} an SbS^{{b}}-formal type; the formal type is unique up to an action of the relative affine Weyl group of SbS^{{b}} [BS13b, BS13a].

In the case of unramified toral connections, S1=T(((z)))S^{{1}}=T(\mathbb{C}(\!(z)\!)) is the usual diagonal maximal torus, and the S1S^{{1}}-formal types of slope rr are those connection matrices in LT normal form (2) with DrD_{r} regular (so that RR is necessarily 0). At the opposite extreme, 𝒞Sn=((ωn))\mathcal{C}\coloneqq S^{{n}}=\mathbb{C}(\!(\omega_{n})\!)^{*} is a ”Coxeter maximal torus”.111Under the bijection between classes of maximal tori in GLn(((z)))\operatorname{GL}_{n}(\mathbb{C}(\!(z)\!)) and conjugacy classes in the Weyl group 𝔖n\mathfrak{S}_{{n}} [KL88], 𝒞\mathcal{C} corresponds to the Coxeter class consisting of nn-cycles. The 𝒞\mathcal{C}-formal types of slope r/nr/n are the 11-forms p(ωn1)dzzp(\omega_{n}^{-1})\frac{dz}{z}, where pp is a polynomial of degree rr.

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 GLn(((z)))\operatorname{GL}_{n}(\mathbb{C}(\!(z)\!)) are the local field analogues of the parabolic subgroups of GLn()\operatorname{GL}_{n}(\mathbb{C}). A parabolic subgroup is the stabilizer of a partial flag of subspaces in n\mathbb{C}^{n}, and a parahoric subgroup is the stabilizer of a ”lattice chain” of [[z]]\mathbb{C}[\![z]\!]-lattices in ((z))n\mathbb{C}(\!(z)\!)^{n}. If PP is a parahoric subgroup with associated lattice chain {Lj}j\{L^{j}\}_{j}, then there is an associated ”lattice chain filtration” {𝔭i}i\{\mathfrak{p}^{i}\}_{i\in\mathbb{Z}} on 𝔤𝔩n(((z)))\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}(\!(z)\!)) defined by 𝔭i={XX(Lj)Lj+i j}\mathfrak{p}^{i}=\{X\mid X(L^{j})\subset L^{j+i}\text{ }\forall j\}.

To each maximal torus SbS^{{b}}, there is a unique ”standard parahoric subgroup” PbGLn([[z]])P^{b}\subset\operatorname{GL}_{n}(\mathbb{C}[\![z]\!]) with the property that the corresponding filtration {(𝔭b)i}i\{(\mathfrak{p}^{b})^{i}\}_{i} is compatible with the filtration on 𝔰b\mathfrak{s}^{b}, in the sense that (𝔰b)i=(𝔭b)i𝔰b(\mathfrak{s}^{b})^{i}=(\mathfrak{p}^{b})^{i}\cap\mathfrak{s}^{b} for all ii [BS13b]. In the unramified case, we have P1=GLn([[z]])P^{1}=\operatorname{GL}_{n}(\mathbb{C}[\![z]\!]). For the Coxeter maximal torus SnS^{{n}}, the corresponding parahoric subgroup is the standard “Iwahori subgroup” IPnI\coloneqq P^{n}; i.e., II is the preimage of the upper-triangular Borel subgroup BB (consisting of all upper-triangular matrices in GLn()\operatorname{GL}_{n}(\mathbb{C})) via the map GLn([[z]])GLn()\operatorname{GL}_{n}(\mathbb{C}[\![z]\!])\to\operatorname{GL}_{n}(\mathbb{C}) induced by the “evaluation at zero” map z0z\mapsto 0.

The most natural way of describing framability involves coadjoint orbits. One can view the principal part at 0 of a connection as a continuous functional on 𝔤𝔩n([[z]])\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[\![z]\!]) via YRes(Tr(YXdzz))Y\mapsto\operatorname{\mathrm{Res}}(\operatorname{Tr}(YX\frac{dz}{z})). Similarly, an SbS^{{b}}-formal type can be viewed as a functional on 𝔭b\mathfrak{p}^{b}. The global connection \nabla then is framable at 0 with respect to the SbS^{{b}}-formal type AdzzA\frac{dz}{z} at 0 if for some global trivialization, the restriction to 𝔭b\mathfrak{p}^{b} of the principal part at 0 lies in the PbP^{b}-coadjoint orbit of AdzzA\frac{dz}{z}. (See Definition 5.1.)

However, one can also give a description more reminiscent of the definition in the unramified case.

Definition 1.1.

Let \nabla be a global connection on 1\mathbb{P}^{1} with a singular point at 0, and let 𝒜=Adzz\mathscr{A}=A\frac{dz}{z} be a toral formal type of slope r/br/b. We say that \nabla is framable at 0 with respect to 𝒜\mathscr{A} if

  1. (1)

    under some global trivialization ϕ\phi, the matrix form =d+[]ϕdzz\nabla=d+[\nabla]_{\phi}\frac{dz}{z} satisfies []ϕ(𝔭b)r[\nabla]_{\phi}\in(\mathfrak{p}^{b})^{-r}, and []ϕA(𝔭b)1r[\nabla]_{\phi}-A\in(\mathfrak{p}^{b})^{1-r}; and

  2. (2)

    there exists an element p(Pb)1p\in(P^{b})^{1} such that the nonpositive truncation of Ad(p)[]ϕ\mathrm{Ad}(p)[\nabla]_{\phi} equals AA.

Recall that GLn()\operatorname{GL}_{n}(\mathbb{C}) acts simply transitively on the space of global trivializations. If one starts with a fixed trivialization ϕ\phi^{\prime}, then the choice of trivialization ϕ\phi in the definition above corresponds to an element gGLn()g\in\operatorname{GL}_{n}(\mathbb{C}); i.e., there is a unique gg such that ϕ=gϕ\phi=g\cdot\phi^{\prime}. This matrix gg is called a compatible framing (or simply, a framing) of \nabla at 0. Framability with respect to a formal type at an arbitrary point a1a\in\mathbb{P}^{1} is defined similarly, by simply replacing zz by zaz-a if aa is finite, and by z1z^{-1} if a=a=\infty.

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 𝐀=(𝒜1,,𝒜m)\mathbf{A}=(\mathscr{A}_{1},\dots,\mathscr{A}_{m}) be a collection of toral formal types at the points a1,,am1a_{1},\dots,a_{m}\in\mathbb{P}^{1}, and let 𝐎=(𝒪1,,𝒪)\mathbf{O}=(\mathscr{O}_{1},\dots,\mathscr{O}_{\ell}) be a collection of adjoint orbits at other points b1,,b1b_{1},\dots,b_{\ell}\in\mathbb{P}^{1}. Does there exist an irreducible rank nn connection \nabla such that

  1. (1)

    \nabla is regular away from the aia_{i}’s and bjb_{j}’s;

  2. (2)

    \nabla is framable at aia_{i} with respect to the formal type 𝒜i\mathscr{A}_{i}; and

  3. (3)

    \nabla is regular singular at bjb_{j} with residue in 𝒪j\mathscr{O}_{j}?

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 𝒪j\mathscr{O}_{j} is nonresonant; i.e., suppose that no pair of the eigenvalues of the orbit differ by a nonzero integer. Further assume that m1m\geq 1, so that there is at least one irregular singular point. In [BS13b], Bremer and Sage constructed the moduli space (𝐀,𝐎)\operatorname{\mathcal{M}}(\mathbf{A},\mathbf{O}) of connections satisfying all the above hypotheses except irreducibility. Let irr(𝐀,𝐎)\operatorname{\mathcal{M}}_{\mathrm{irr}}(\mathbf{A},\mathbf{O}) be the subset of the moduli space consisting of irreducible connections. In this language, the toral Deligne–Simpson problem poses the question of when irr(𝐀,𝐎)\operatorname{\mathcal{M}}_{\mathrm{irr}}(\mathbf{A},\mathbf{O}) 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 𝔾m\mathbb{G}_{m} with the irregular singularity at 0. 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 \infty. 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 G{G}-connection analogues (for G{G} 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 G{G}-connection of this type. This connection, which may be viewed as a G{G}-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 G{G}-connection and more general rigid “Coxeter G{G}-connections” constructed in [KS21b]. The Airy G{G}-connection and its \ell-adic analogue have also been studied in [JKY21].

Recall that if ^\widehat{\nabla} is a rank nn formal connection, then every slope of ^\widehat{\nabla} has denominator (when the slope is expressed in lowest form) between 11 and nn. We say ^\widehat{\nabla} is maximally ramified if all such denominators (or equivalently, at least one) is nn. In this case, all the slopes are the same — say r/nr/n with gcd(r,n)=1\gcd(r,n)=1 — and equal to the slope of the connection. More concretely, the leading term of the LT normal form is of the form Drzr/nD_{r}z^{-r/n} with DrD_{r} 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 r/nr/n has a rational canonical form d+p(ωn1)dzzd+p(\omega_{n}^{-1})\frac{dz}{z}, where pp is a polynomial of degree rr, and the set of formal types is given by {p(ωn1)dzzp[x],deg(p)=r}\{p(\omega_{n}^{-1})\frac{dz}{z}\mid p\in\mathbb{C}[x],\deg(p)=r\}. Moreover, any such connection is irreducible.

In this paper, we solve the ramified Deligne–Simpson problem for Coxeter connections. More precisely, let 𝒜\mathscr{A} be a maximally ramified formal type, and let 𝒪\mathscr{O} 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 \nabla on 1\mathbb{P}^{1} which is framable at 0 with formal type 𝒜\mathscr{A}, is regular singular with residue in 𝒪\mathscr{O} at \infty, and is otherwise nonsingular. Note that any such connection is automatically irreducible, since its formal connection at 0 is irreducible. Thus, in the language of moduli spaces, we determine when (𝒜,𝒪)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O}) is nonempty.

In order to state our result, we need some facts about adjoint orbits. Fix a monic polynomial q=i=1s(xai)miq=\prod_{i=1}^{s}(x-a_{i})^{m_{i}} of degree nn with the aia_{i}’s distinct complex numbers. The set {X𝔤𝔩n()char(X)=q}\{X\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})\mid\operatorname{char}(X)=q\} of matrices with characteristic polynomial qq is a closed subset of 𝔤𝔩n()\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) which is stable under conjugation. We denote the set of orbits with characteristic polynomial qq by πq\pi_{q}. This set is partially ordered under the usual Zariski closure ordering: 𝒪𝒪\mathscr{O}\preceq\mathscr{O}^{\prime} if and only if 𝒪𝒪¯\mathscr{O}\subset\overline{\mathscr{O}^{\prime}}. The theory of the Jordan canonical form makes it clear that πq\pi_{q} can be identified with the Cartesian product i=1sPart(mi)\prod_{i=1}^{s}\mathrm{Part}(m_{i}), where Part(mi)\mathrm{Part}(m_{i}) denotes the set of partitions of mim_{i}. Moreover, this identification defines a poset isomorphism between the closure ordering and the direct product of the dominance orders.

Given positive integers rr and mm, there exists a unique smallest partition λm,rPart(m)\lambda^{{m},{r}}\in\mathrm{Part}(m) with at most rr parts. Define 𝒪qr\mathscr{O}_{q}^{r} to be the orbit in πq\pi_{q} corresponding to the element

(λm1,r,λm2,r,,λms,r)i=1sPart(mi).(\lambda^{{m_{1}},{r}},\lambda^{{m_{2}},{r}},\ldots,\lambda^{{m_{s}},{r}})\in\prod_{i=1}^{s}\mathrm{Part}(m_{i}).

This tuple of partitions is the (unique) smallest element of i=1sPart(mi)\prod_{i=1}^{s}\mathrm{Part}(m_{i}) such that each component partition has at most rr parts. Note that 𝒪q1\mathscr{O}_{q}^{1} is just the regular orbit in πq\pi_{q}. On the other extreme, if rmir\geq m_{i} for all ii (as is the case when rnr\geq n), then 𝒪qr\mathscr{O}_{q}^{r} is the semisimple orbit, the unique minimal orbit in πq\pi_{q}. Let 𝒪qr\langle{\mathscr{O}_{q}^{r}}\rangle denote the principal filter generated by 𝒪qr\mathscr{O}_{q}^{r} in πq\pi_{q}, i.e., 𝒪πq\mathscr{O}\in\pi_{q} satisfies 𝒪𝒪qr\mathscr{O}\in\langle{\mathscr{O}_{q}^{r}}\rangle if and only if 𝒪𝒪qr\mathscr{O}\succeq\mathscr{O}_{q}^{r}. This filter is proper unless rmir\geq m_{i} for all ii. As we will see in Theorem 2.3, the collection of orbits 𝒪qr\mathscr{O}_{q}^{r} for each fixed rr 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 nn maximally ramified formal type 𝒜\mathscr{A} and a monic polynomial qq of degree nn that is nonresonant (i.e., no two roots differ by a nonzero integer), let

DS(𝒜,q)={𝒪πq(𝒜,𝒪)}.\mathrm{DS}(\mathscr{A},q)=\{\mathscr{O}\in\pi_{q}\mid\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O})\neq\varnothing\}.
Theorem 5.4.

Let rr and nn be positive integers with gcd(r,n)=1\gcd(r,n)=1, let 𝒜\mathscr{A} be a maximally ramified formal type of slope r/nr/n, and let q=i=1s(xai)mi[x]q=\prod_{i=1}^{s}(x-a_{i})^{m_{i}}\in\mathbb{C}[x] with a1,,asa_{1},\ldots,a_{s}\in\mathbb{C} distinct modulo \mathbb{Z}. Then

DS(𝒜,q)={𝒪qrif Res(Tr(𝒜))=i=1smiai,else.\mathrm{DS}(\mathscr{A},q)=\begin{cases}\langle{\mathscr{O}_{q}^{r}}\rangle&\text{if }\operatorname{\mathrm{Res}}(\operatorname{Tr}(\mathscr{A}))=-\sum_{i=1}^{s}m_{i}a_{i},\\ \varnothing&\text{else}.\end{cases}

In other words, given 𝒜=p(ω1)dzz\mathscr{A}=p(\omega^{-1})\frac{dz}{z}, then (𝒜,𝒪)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O}) is nonempty if and only if np(0)=Tr(𝒪)np(0)=-\operatorname{Tr}(\mathscr{O}) and 𝒪𝒪char(𝒪)r\mathscr{O}\succeq\mathscr{O}_{\operatorname{char}(\mathscr{O})}^{r}. Concretely, the condition 𝒪𝒪char(𝒪)r\mathscr{O}\succeq\mathscr{O}_{\operatorname{char}(\mathscr{O})}^{r} means that 𝒪\mathscr{O} has at most rr 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 rmir\geq m_{i} for all ii, then DS(𝒜,q)=πq\mathrm{DS}(\mathscr{A},q)=\pi_{q} as long as the trace condition is satisfied. In particular, this is the case if r>nr>n.

There is an obvious analogue of this problem for SLn\operatorname{SL}_{n}-connections (as opposed to GLn\operatorname{GL}_{n}-connections). Here, maximally ramified formal types are of the form p(ω1)dzzp(\omega^{-1})\frac{dz}{z} with p(0)=0p(0)=0 and Tr(𝒪)=0\operatorname{Tr}(\mathscr{O})=0. Thus, the trace condition becomes vacuous, and the Deligne–Simpson problem has a positive solution if and only if 𝒪𝒪char(𝒪)r\mathscr{O}\succeq\mathscr{O}_{\operatorname{char}(\mathscr{O})}^{r}.

One can define Coxeter G{G}-connections for any simple group G{G} (or for any reductive group with connected Dynkin diagram) [KS21b]. For such a G{G}, Coxeter toral connections have slope r/hr/h, where hh is the Coxeter number for G{G} and gcd(r,h)=1\gcd(r,h)=1. 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 \infty has nilpotent residue (and thus has unipotent monodromy).

Conjecture 1.3.

Let G{G} be a simple complex group with Lie algebra 𝔤\operatorname{\mathfrak{g}}. Fix a Coxeter G{G}-formal type 𝒜\mathscr{A} of slope r/hr/h with gcd(r,h)=1\gcd(r,h)=1. Then there exists a nilpotent orbit 𝒪r𝔤\mathscr{O}^{r}\subset\operatorname{\mathfrak{g}} such that the Deligne–Simpson problem for Coxeter G{G}-connections with initial data 𝒜\mathscr{A} and the nilpotent orbit 𝒪\mathscr{O} has a positive solution if and only if 𝒪𝒪r\mathscr{O}\succeq\mathscr{O}^{r}. Moreover, if r>hr>h, then 𝒪r=0\mathscr{O}^{r}=0, 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 U1U\subset\mathbb{P}^{1} be a nonempty open set, and let j:U1j:U\hookrightarrow\mathbb{P}^{1} denote the inclusion. A G{G}-connection \nabla on UU is called physically rigid if it is uniquely determined by the formal isomorphism class at each point of 1U\mathbb{P}^{1}\setminus U. It is called cohomologically rigid if H1(1,j!ad)=0H^{1}(\mathbb{P}^{1},j_{!*}\mathrm{ad}_{\nabla})=0. For irreducible connections, cohomological rigidity implies that \nabla has no infinitesimal deformations. For G=GLn(){G}=\operatorname{GL}_{n}(\mathbb{C}), 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 G{G}-connections with unipotent monodromy. A homogeneous Coxeter G{G}-formal type of slope r/hr/h (i.e., for GLn\operatorname{GL}_{n}, a formal type of the form aωnrdzza\omega^{-r}_{n}\frac{dz}{z} with a0a\neq 0) gives rise to a Coxeter G{G}-connection on 𝔾m\mathbb{G}_{m} with nilpotent residue at infinity. They determined precisely when these connections are (cohomologically) rigid, thus generalizing the work of Frenkel and Gross [FG09]. For GLn\operatorname{GL}_{n}, it turns out that such connections are rigid precisely when rr divides n+1n+1 or n1n-1.

We can now generalize the results of [KS21b] to give a classification of rigid Coxeter connections in type AA.

Theorem 6.1.

Let 𝒜\mathscr{A} be a rank nn maximally ramified formal type of slope r/nr/n, and let 𝒪\mathscr{O} be any nilpotent orbit with 𝒪𝒪xnr\mathscr{O}\succeq\mathscr{O}_{x^{n}}^{r}. Then there exists a rigid connection with the given formal type and and unipotent monodromy determined by 𝒪\mathscr{O} if and only if 𝒪=𝒪xnr\mathscr{O}=\mathscr{O}_{x^{n}}^{r} and r|(n±1)r|(n\pm 1).

We expect that the analogous statement is true for Coxeter G{G}-connections.

1.6. Organization of the paper

In §2, we discuss some facts about the poset of adjoint orbits in 𝔤𝔩n()\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) 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 GLn()\operatorname{GL}_{n}(\mathbb{C}) on 𝔤𝔩n()\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) (i.e., similarity classes of n×nn\times n complex matrices). We will need some facts about these adjoint orbits.

The set π\pi of adjoint orbits in 𝔤𝔩n()\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) is partially ordered via the closure order: 𝒪𝒪\mathscr{O}\preceq\mathscr{O}^{\prime} if 𝒪𝒪¯\mathscr{O}\subseteq\overline{\mathscr{O}^{\prime}}. Let char\operatorname{char} be the map sending a matrix to its characteristic polynomial. Given a monic degree nn polynomial qq, char1(q)\operatorname{char}^{-1}(q) is closed and GLn()\operatorname{GL}_{n}(\mathbb{C})-stable. If we let πq\pi_{q} be the set of adjoint orbits in char1(q)\operatorname{char}^{-1}(q), it is immediate that as a poset,

π=q monicdeg(q)=nπq.\pi=\bigsqcup_{\begin{subarray}{c}\text{$q$ monic}\\ \deg(q)=n\end{subarray}}\pi_{q}.

The theory of Jordan canonical forms allows us to identify the posets πq\pi_{q} with posets involving partitions. If nn is a positive integer, a partition of nn is a nonincreasing sequence λ=(λ1,λ2,,λm)\lambda=(\lambda_{1},\lambda_{2},\ldots,\lambda_{m}) of positive integers that sum to nn. Each integer appearing in this sequence is called a part of λ\lambda. The total number of parts is denoted by |λ||\lambda|. It will sometimes be convenient to use exponential notation for partitions: if the bib_{i}’s are the distinct parts, each appearing with multiplicity kik_{i}, we will also denote by λ\lambda the multiset {b1k1,,bsks}\{b_{1}^{k_{1}},\dots,b_{s}^{k_{s}}\}. Let Part(n)\mathrm{Part}(n) be the set of partitions of nn. We view Part(n)\mathrm{Part}(n) as a poset via the dominance order

(3) λμ|λ||μ| and i=1jλii=1jμi for all j[1,|λ|].\text{$\lambda\succeq\mu\quad\iff\quad|\lambda|\leq|\mu|$ and $\sum_{i=1}^{j}\lambda_{i}\geq\sum_{i=1}^{j}\mu_{i}$ for all $j\in[1,|\lambda|]$}.

Write q=i=1s(xai)miq=\prod_{i=1}^{s}(x-a_{i})^{m_{i}} for distinct a1,,asa_{1},\dots,a_{s}\in\mathbb{C} and m1,,ms>0m_{1},\ldots,m_{s}\in\mathbb{Z}_{>0}. The set πq\pi_{q} can be identified with i=1sPart(mi)\prod_{i=1}^{s}\mathrm{Part}(m_{i}), where the partition of mim_{i} is given by the sizes of the Jordan blocks with eigenvalue aia_{i}. It is well-known that the closure order corresponds to the product of the dominance orders under this identification.

The unique maximal orbit in πq\pi_{q} is the orbit with a single Jordan block for each eigenvalue. This is the regular orbit with characteristic polynomial qq, i.e., the unique orbit in πq\pi_{q} of codimension nn. We now define a sequence {𝒪qr}r\{\mathscr{O}_{q}^{r}\}_{r} of orbits in πq\pi_{q} which generalize the regular orbit.

Fix r>0r\in\mathbb{Z}_{>0}, and consider the subset qrπq\mathcal{F}^{r}_{q}\subset\pi_{q} consisting of orbits with at most rr Jordan blocks for each eigenvalue. It is immediate from (3) that qr\mathcal{F}^{r}_{q} is a filter in the poset πq\pi_{q}. This means that if 𝒪qr\mathscr{O}\in\mathcal{F}^{r}_{q} and 𝒪𝒪\mathscr{O}\preceq\mathscr{O}^{\prime}, then 𝒪qr\mathscr{O}^{\prime}\in\mathcal{F}^{r}_{q}. It turns out that qr\mathcal{F}^{r}_{q} is a principal filter; i.e., there exists a (unique) element 𝒪qrπq\mathscr{O}_{q}^{r}\in\pi_{q} such that qr=𝒪qr{𝒪πq𝒪𝒪qr}\mathcal{F}^{r}_{q}=\langle{\mathscr{O}_{q}^{r}}\rangle\coloneqq\{\mathscr{O}\in\pi_{q}\mid\mathscr{O}\succeq\mathscr{O}_{q}^{r}\}. We will also set r=qqr\mathcal{F}^{r}=\bigcup_{q}\mathcal{F}^{r}_{q}; it is a filter in π\pi. Note that 1\mathcal{F}^{1} is the set of all regular (or equivalently, maximal) adjoint orbits.

Proposition 2.1.

Given n,r>0n,r\in\mathbb{Z}_{>0}, write n=kr+rn=kr+r^{\prime} with k,rk,r^{\prime}\in\mathbb{Z} and 0r<r0\leq r^{\prime}<r. Then the partition

λn,r={(k+1)r,krr}\lambda^{{n},{r}}=\{(k+1)^{r^{\prime}},k^{r-r^{\prime}}\}

is the (unique) smallest partition of nn with at most rr parts.

Proof.

Let λ\lambda be any partition of nn with at most rr parts, say with biggest part uu and smallest part vv. We will show that there exists a strictly smaller partition with at most rr parts unless λ=λn,r\lambda=\lambda^{{n},{r}}.

If rnr\geq n, then λn,r={1n}\lambda^{{n},{r}}=\{1^{n}\}, the smallest element of Part(n)\mathrm{Part}(n), so the statement is trivial. We thus may assume that r<nr<n. First, suppose that |λ|<r|\lambda|<r. It follows that u2u\geq 2, so one obtains a strictly smaller partition with |λ|+1r|\lambda|+1\leq r parts by replacing one uu with u1u-1 and adjoining a new part with value 11. We may thus assume without loss of generality that |λ|=r|\lambda|=r.

Next, suppose that uv2u-v\geq 2. Define a partition μ\mu with the same parts as λ\lambda except one uu is replaced by u1u-1 and one vv is replaced by v+1v+1. It is obvious that |λ|=|μ||\lambda|=|\mu| and that λ\lambda is strictly bigger than μ\mu.

It remains to consider the case uv1u-v\leq 1, so λ={(v+1)s,vrs}\lambda=\{(v+1)^{s},v^{r-s}\} for some ss with 0s<r0\leq s<r. We then have n=s(v+1)+(rs)v=vr+sn=s(v+1)+(r-s)v=vr+s, so v=kv=k and s=rs=r^{\prime}. Thus, λ=λn,r\lambda=\lambda^{{n},{r}}. ∎

Now, define 𝒪qr\mathscr{O}_{q}^{r} to be the orbit in πq\pi_{q} corresponding to the element

(λm1,r,λm2,r,,λms,r)i=1sPart(mi).(\lambda^{{m_{1}},{r}},\lambda^{{m_{2}},{r}},\ldots,\lambda^{{m_{s}},{r}})\in\prod_{i=1}^{s}\mathrm{Part}(m_{i}).
Corollary 2.2.

The filter qr\mathcal{F}^{r}_{q} is principal with generator 𝒪qr\mathscr{O}_{q}^{r}, i.e., qr=𝒪qr\mathcal{F}^{r}_{q}=\langle{\mathscr{O}_{q}^{r}}\rangle.

Proof.

It suffices to show that the corresponding filter in i=1sPart(mi)\prod_{i=1}^{s}\mathrm{Part}(m_{i}) is principal with generator (λm1,r,λm2,r,,λms,r)(\lambda^{{m_{1}},{r}},\lambda^{{m_{2}},{r}},\ldots,\lambda^{{m_{s}},{r}}). This is immediate from the proposition. ∎

Recall that the semisimple orbit in πq\pi_{q} is the unique minimal orbit. If rmir\geq m_{i} for all ii (as is the case when rnr\geq n), then 𝒪qr\mathscr{O}_{q}^{r} is the semisimple orbit so 𝒪qr=πq\langle{\mathscr{O}_{q}^{r}}\rangle=\pi_{q}.

We can now give a Lie-theoretic interpretation of r\mathcal{F}^{r} which will be important in our applications. Let VrV^{r} be the set of matrices with nonzero entries on, and 0’s below, the rrth subdiagonal:

Vr={(xij)𝔤𝔩n()xij=0 if ij>r and xij0 if ij=r}.V^{r}=\{(x_{ij})\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})\mid\text{$x_{ij}=0$ if $i-j>r$ and $x_{ij}\neq 0$ if $i-j=r$}\}.

Note that Vr=𝔤𝔩n()V^{r}=\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) if rnr\geq n. It is well-known that every element of V1V^{1} is regular and that every regular orbit has a representative in V1V^{1}. (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 VrV^{r} are precisely the orbits in r\mathcal{F}^{r}. The minimal such orbits are the 𝒪qr\mathscr{O}_{q}^{r}’s.

We begin by proving that 𝒪qr\mathscr{O}_{q}^{r} intersects VrV^{r}. Let Nr,nN_{r,n} be the n×nn\times n matrix with 11’s on the rrth subdiagonal and 0’s elsewhere. We usually omit nn from the notation.

Proposition 2.4.

Fix a positive integer r<nr<n. Let Dq=diag(a1,,a1,a2,,a2,,as,,as)D_{q}=\mathrm{diag}(a_{1},\dots,a_{1},a_{2},\dots,a_{2},\dots,a_{s},\dots,a_{s}), where the eigenvalue aia_{i} appears with multiplicity mim_{i}. Let UrU_{r} be any matrix with all entries on the rrth subdiagonal nonzero and all other entries 0. Then Ur+Dq𝒪qrU_{r}+D_{q}\in\mathscr{O}_{q}^{r}.

Proof.

It is easy to see that there exists an invertible diagonal matrix tt such that Ad(t)(Ur+Dq)=Nr+Dq\mathrm{Ad}(t)(U_{r}+D_{q})=N_{r}+D_{q}, so we may assume without loss of generality that Ur=NrU_{r}=N_{r}. We prove the proposition by induction on the number of distinct eigenvalues ss (for arbitrary nn).

If s=1s=1, the partition for the single eigenvalue is λn,r\lambda^{n,r}. Indeed, the Jordan strings for Nr+Dqa1id=NrN_{r}+D_{q}-a_{1}\mathrm{id}=N_{r} are given by eiei+rei+2re_{i}\mapsto e_{i+r}\mapsto e_{i+2r}\mapsto\cdots for 1ir1\leq i\leq r, with the strings of length k+1k+1 for 1ir1\leq i\leq r^{\prime} and kk otherwise.

Now assume s>1s>1. Let VV be the span of em1+1,,ene_{m_{1}+1},\dots,e_{n}. Note that Nr+DqN_{r}+D_{q} stabilizes VV; in fact, (Nr+Dq)|V=Nr,nm1+Dq^(N_{r}+D_{q})|_{V}=N_{r,n-m_{1}}+D_{\hat{q}}, where q^=i=2s(xai)mi\hat{q}=\prod_{i=2}^{s}(x-a_{i})^{m_{i}}. It is clear that the Jordan strings for (Nr+Dq)|V(N_{r}+D_{q})|_{V} are also Jordan strings for Nr+DqN_{r}+D_{q} corresponding to the eigenvalues a2,,asa_{2},\dots,a_{s}. Also, Nr+DqN_{r}+D_{q} induces the endomorphism Nr,m1+D(xa1)m1N_{r,m_{1}}+D_{(x-a_{1})^{m_{1}}} on n/V\mathbb{C}^{n}/V. It remains to show that each Jordan string e¯ie¯i+re¯i+(1)r0\bar{e}_{i}\mapsto\bar{e}_{i+r}\mapsto\dots\mapsto\bar{e}_{i+(\ell-1)r}\mapsto 0 for Nr,m1N_{r,m_{1}} on n/V\mathbb{C}^{n}/V lifts to a Jordan string for the zero eigenvalues of Nr+Dqa1idN_{r}+D_{q}-a_{1}\mathrm{id}. (Here, \ell is the smallest integer such that i+r>m1i+\ell r>m_{1}.) Let f=(Nr+Dqa1id)|Vf=(N_{r}+D_{q}-a_{1}\mathrm{id})|_{V}. This map is invertible, so eif(ei+r)ei+(1)rf1(ei+r)ker(Nr+Dqa1id)e_{i}-f^{-\ell}(e_{i+\ell r})\mapsto\dots\mapsto e_{i+(\ell-1)r}-f^{-1}(e_{i+\ell r})\in\ker(N_{r}+D_{q}-a_{1}\mathrm{id}) is the desired lift. ∎

Proof of Theorem 2.3.

We first show that VrV^{r} consists of elements whose Jordan forms have at most rr blocks for each eigenvalue. Let XVrX\in V^{r}, and let aa be any eigenvalue of XX. The bottom nrn-r rows of the matrix XaidX-a\mathrm{id} are linearly independent, so rank(Xaid)nr\mathrm{rank}(X-a\mathrm{id})\geq n-r. We conclude that dimker(Xaid)r\dim\ker(X-a\mathrm{id})\leq r. Since this dimension is the number of Jordan blocks for the eigenvalue aa, the claim follows.

We have shown that the set of adjoint orbits intersecting VrV^{r} is contained in r\mathcal{F}^{r}. Now take 𝒪r\mathscr{O}\in\mathcal{F}^{r}. Let qq be its characteristic polynomial, so 𝒪𝒪qr\mathscr{O}\succeq\mathscr{O}_{q}^{r}. Take any X𝒪qrX\in\mathscr{O}_{q}^{r} as in Proposition 2.4. Then XVrX\in V^{r}. By a theorem of Krupnik [Kru97, Theorem 1], there exists a strictly upper triangular matrix Z𝔤𝔩n()Z\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) such that X+Z𝒪X+Z\in\mathscr{O}. Since X+ZVrX+Z\in V^{r}, this proves the theorem. ∎

Remark 2.5.

Krupnik’s approach does not lead to explicit constructions of orbit representatives in VrV^{r}. 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 F=((z))F=\mathbb{C}(\!(z)\!) denote the field of formal Laurent series, and let 𝔬=[[z]]\mathfrak{o}=\mathbb{C}[\![z]\!] denote the ring of formal power series. An 𝔬\mathfrak{o}-lattice LL in FnF^{n} is a finitely generated 𝔬\mathfrak{o}-module with the property that L𝔬FFnL\otimes_{\mathfrak{o}}F\cong F^{n}. A lattice chain in FnF^{n} is a collection {Li}i\{L^{i}\}_{i\in\mathbb{Z}} of lattices satisfying the following properties:

  1. (1)

    LiLi+1L^{i}\supsetneq L^{i+1} for all ii; and

  2. (2)

    there exists a positive integer ee, called the period, such that Li+e=zLiL^{i+e}=zL^{i} for all ii.

A parahoric subgroup PGLn(F)P\subset\operatorname{GL}_{n}(F) is the stabilizer of a lattice chain; i.e., P={gGLn(F)gLi=Li for all i}P=\{g\in\operatorname{GL}_{n}(F)\mid gL^{i}=L^{i}\text{ for all }i\}. We write ePe_{P} to denote the period of the lattice chain stabilized by PP. A lattice chain in FnF^{n} is called complete if its period is as large as possible, i.e., if its period equals nn. The parahoric subgroups associated to complete lattice chains are called Iwahori subgroups.

Each lattice chain {Li}i\{L^{i}\}_{i\in\mathbb{Z}} — say with corresponding parahoric PP — determines a filtration {𝔭i}i\{\mathfrak{p}^{i}\}_{i\in\mathbb{Z}} of 𝔤𝔩n(F)\operatorname{\mathfrak{gl}}_{n}(F) defined by 𝔭i={X𝔤𝔩n(F)XLjLj+i for all j}\mathfrak{p}^{i}=\{X\in\operatorname{\mathfrak{gl}}_{n}(F)\mid XL^{j}\subset L^{j+i}\text{ for all }j\}; in particular, 𝔭0=𝔭Lie(P)\mathfrak{p}^{0}=\mathfrak{p}\coloneqq\operatorname{\mathrm{Lie}}(P). One also gets a filtration {Pi}i0\{P^{i}\}_{i\in\mathbb{Z}_{\geq 0}} of PP defined by P0=PP^{0}=P and Pi=1+𝔭iP^{i}=1+\mathfrak{p}^{i} for i>0i>0. In the special case that each lattice in the lattice chain stabilized by PP admits an 𝔬\mathfrak{o}-basis of the form {zkjej}j=1n\{z^{k_{j}}e_{j}\}_{j=1}^{n}, the corresponding filtration {𝔭i}i\{\mathfrak{p}^{i}\}_{i} is induced by a grading {𝔭(i)}i\{\mathfrak{p}(i)\}_{i} on 𝔤𝔩n([z,z1])\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[z,z^{-1}]).

Remark 3.1.

The “lattice chain filtrations” described above may also be obtained through a more general construction. Let \mathcal{B} be the (reduced) Bruhat–Tits building associated to the loop group GLn(F)\operatorname{GL}_{n}(F). 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 PP is the Moy–Prasad filtration associated to the barycenter of the simplex in the Bruhat–Tits building corresponding to PP [BS18].

The building \mathcal{B} is the union of (n1)(n-1)-dimensional real affine spaces called apartments, which are in one-to-one correspondence with split maximal tori in GLn(F)\operatorname{GL}_{n}(F). The parahoric subgroups described in the ”special case” above — i.e., the parahoric subgroups PP where each lattice in the lattice chain stabilized by PP admits an 𝔬\mathfrak{o}-basis of the form {zkjej}j=1n\{z^{k_{j}}e_{j}\}_{j=1}^{n} — 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 GLn(F)\operatorname{GL}_{n}(F): GLn(𝔬)\operatorname{GL}_{n}(\mathfrak{o}) and the ”standard Iwahori subgroup” II (defined below). The parahoric subgroup GLn(𝔬)\operatorname{GL}_{n}(\mathfrak{o}) is the stabilizer of the 11-periodic lattice chain {zi𝔬n}i\{z^{i}\mathfrak{o}^{n}\}_{i}. The associated filtration on 𝔤𝔩n(F)\operatorname{\mathfrak{gl}}_{n}(F) is the degree filtration {zi𝔤𝔩n(𝔬)}i\{z^{i}\operatorname{\mathfrak{gl}}_{n}(\mathfrak{o})\}_{i\in\mathbb{Z}}.

The standard Iwahori subgroup II is the stabilizer of the standard complete lattice chain in FnF^{n}, i.e., the lattice chain

{𝔬-span{zine1,zi+1ne2,,zi+(n1)nen}}i,\{\mathfrak{o}\text{-span}\{z^{\lfloor{\frac{i}{n}}\rfloor}e_{1},z^{\lfloor{\frac{i+1}{n}}\rfloor}e_{2},\ldots,z^{\lfloor{\frac{i+(n-1)}{n}}\rfloor}e_{n}\}\}_{i\in\mathbb{Z}},

where {ej}j=1n\{e_{j}\}_{j=1}^{n} is the standard basis for FnF^{n}. The associated filtration is the standard Iwahori filtration {𝔦i}i\{\mathfrak{i}^{i}\}_{i\in\mathbb{Z}}. Note that if BGLn()B\subset\operatorname{GL}_{n}(\mathbb{C}) is the Borel subgroup of invertible upper triangular matrices, then II is the preimage of BB under the homomorphism GLn(𝔬)GLn()\operatorname{GL}_{n}(\mathfrak{o})\to\operatorname{GL}_{n}(\mathbb{C}) induced by the ”evaluation at zero” map z0z\mapsto 0.

Define ω𝔤𝔩n(F)\omega\in\operatorname{\mathfrak{gl}}_{n}(F) by ω=i=1n1ei,i+1+zen,1\omega=\sum_{i=1}^{n-1}e_{i,i+1}+ze_{n,1}; i.e., ω\omega is the matrix with 11’s in each entry of the superdiagonal, zz in the lower-left entry, and 0’s elsewhere. Then ωi𝔦j=𝔦jωi=𝔦i+j\omega^{i}\mathfrak{i}^{j}=\mathfrak{i}^{j}\omega^{i}=\mathfrak{i}^{i+j} [Bus87, Proposition 1.18] for all i,ji,j\in\mathbb{Z}. Note that 𝔦i+n=z𝔦i\mathfrak{i}^{i+n}=z\mathfrak{i}^{i} for all ii. Let TT denote the standard diagonal maximal torus in GLn()\operatorname{GL}_{n}(\mathbb{C}). Then 𝔱=Lie(T)\mathfrak{t}=\operatorname{\mathrm{Lie}}(T) consists of all diagonal matrices in 𝔤𝔩n()\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}). The standard Iwahori grading {𝔦(i)}i\{\mathfrak{i}(i)\}_{i\in\mathbb{Z}} is then given by

(4) 𝔦(i)=ωi𝔱=𝔱ωi.\mathfrak{i}(i)=\omega^{i}\mathfrak{t}=\mathfrak{t}\omega^{i}.
Example 3.2.

Let n=3n=3. Then

ω=[010001z00]I=[𝔬𝔬𝔬z𝔬𝔬𝔬z𝔬z𝔬𝔬].\omega=\begin{bmatrix}0&1&0\\ 0&0&1\\ z&0&0\end{bmatrix}I=\begin{bmatrix}\mathfrak{o}^{*}&\mathfrak{o}&\mathfrak{o}\\ z\mathfrak{o}&\mathfrak{o}^{*}&\mathfrak{o}\\ z\mathfrak{o}&z\mathfrak{o}&\mathfrak{o}^{*}\end{bmatrix}.

Some steps in the standard Iwahori filtration on 𝔤𝔩n(F)\operatorname{\mathfrak{gl}}_{n}(F) are shown below:

𝔦2=[𝔬z1𝔬z1𝔬𝔬𝔬z1𝔬𝔬𝔬𝔬]𝔦1=[𝔬𝔬z1𝔬𝔬𝔬𝔬z𝔬𝔬𝔬]𝔦0=[𝔬𝔬𝔬z𝔬𝔬𝔬z𝔬z𝔬𝔬].\mathfrak{i}^{-2}=\begin{bmatrix}\mathfrak{o}&z^{-1}\mathfrak{o}&z^{-1}\mathfrak{o}\\ \mathfrak{o}&\mathfrak{o}&z^{-1}\mathfrak{o}\\ \mathfrak{o}&\mathfrak{o}&\mathfrak{o}\end{bmatrix}\quad\subsetneq\quad\mathfrak{i}^{-1}=\begin{bmatrix}\mathfrak{o}&\mathfrak{o}&z^{-1}\mathfrak{o}\\ \mathfrak{o}&\mathfrak{o}&\mathfrak{o}\\ z\mathfrak{o}&\mathfrak{o}&\mathfrak{o}\end{bmatrix}\quad\subsetneq\quad\mathfrak{i}^{0}=\begin{bmatrix}\mathfrak{o}&\mathfrak{o}&\mathfrak{o}\\ z\mathfrak{o}&\mathfrak{o}&\mathfrak{o}\\ z\mathfrak{o}&z\mathfrak{o}&\mathfrak{o}\end{bmatrix}.

If 𝔥𝔤𝔩n(F)\mathfrak{h}\subset\operatorname{\mathfrak{gl}}_{n}(F) is any \mathbb{C}-subspace containing some 𝔭i\mathfrak{p}^{i}, we denote the space of continuous linear functionals on 𝔥\mathfrak{h} by 𝔥\mathfrak{h}^{\vee}. Every continuous functional on 𝔥\mathfrak{h} extends to a continuous functional on 𝔤𝔩n(F)\operatorname{\mathfrak{gl}}_{n}(F), so 𝔥𝔤𝔩n(F)/𝔥\mathfrak{h}^{\vee}\cong\operatorname{\mathfrak{gl}}_{n}(F)^{\vee}/\mathfrak{h}^{\perp}, where 𝔥\mathfrak{h}^{\perp} denotes the annihilator of 𝔥\mathfrak{h}. The space of 11-forms Ω1(𝔤𝔩n(F))\Omega^{1}(\operatorname{\mathfrak{gl}}_{n}(F)) can be identified with the space of functionals 𝔤𝔩n(F)\operatorname{\mathfrak{gl}}_{n}(F)^{\vee} by associating a 11-form ν\nu with the functional YRes(Tr(Yν))Y\mapsto\operatorname{\mathrm{Res}}(\operatorname{Tr}(Y\nu)). This identification is well-behaved with respect to lattice chain filtrations [BS18, Proposition 3.6]:

(𝔭i)𝔤𝔩n(F)dzz/𝔭i+1dzz.(\mathfrak{p}^{i})^{\vee}\cong\operatorname{\mathfrak{gl}}_{n}(F)\tfrac{dz}{z}/\mathfrak{p}^{-i+1}\tfrac{dz}{z}.

When the filtration comes from a grading, one can be even more explicit. In particular, we get

(5) 𝔤𝔩n(𝔬)𝔤𝔩n([z1])dzzand𝔦𝔱[ω1]dzz.\operatorname{\mathfrak{gl}}_{n}(\mathfrak{o})^{\vee}\cong\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[z^{-1}])\tfrac{dz}{z}\qquad\text{and}\qquad\mathfrak{i}^{\vee}\cong\mathfrak{t}[\omega^{-1}]\tfrac{dz}{z}.

For applications to connections, it is important to consider the relationship between filtrations on Cartan subalgebras of 𝔤𝔩n(F)\operatorname{\mathfrak{gl}}_{n}(F) and filtrations on parahoric subalgebras  [BS13b, BS13a].

Note that, since FF 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 GLn(F)\operatorname{GL}_{n}(F) and the set of conjugacy classes in the Weyl group for GLn()\operatorname{GL}_{n}(\mathbb{C}) (i.e., the symmetric group 𝔖n\mathfrak{S}_{{n}}[KL88, Lemma 2]. We also remark that each Cartan subalgebra 𝔰\mathfrak{s} comes equipped with a natural filtration {𝔰i}i\{\mathfrak{s}^{i}\}_{i\in\mathbb{Z}} (see, e.g., [BS13a, Section 3]). In the cases of interest to us in this paper, the filtration will be induced by a grading {𝔰(i)}i\{\mathfrak{s}(i)\}_{i\in\mathbb{Z}}.

Let SS be a maximal torus and let PP be a parahoric subgroup, both in GLn(F)\operatorname{GL}_{n}(F). Let 𝔰=Lie(S)\mathfrak{s}=\operatorname{\mathrm{Lie}}(S) be the Cartan subalgebra associated to SS. We say that SS and PP (or 𝔰\mathfrak{s} and 𝔭\mathfrak{p}) are compatible (resp. graded compatible) if 𝔰i=𝔭i𝔰\mathfrak{s}^{i}=\mathfrak{p}^{i}\cap\mathfrak{s} (resp. 𝔰(i)=𝔭(i)𝔰\mathfrak{s}(i)=\mathfrak{p}(i)\cap\mathfrak{s}) for all ii. For present purposes, it suffices to consider two examples: the diagonal subalgebra 𝔱(F)\mathfrak{t}(F) and the “standard Coxeter Cartan subalgebra” 𝔠=((ω))\mathfrak{c}=\mathbb{C}(\!(\omega)\!). The diagonal Cartan subalgebra 𝔱(F)\mathfrak{t}(F) (corresponding to the trivial class in the Weyl group 𝔖n\mathfrak{S}_{{n}}) is endowed with a filtration which comes from the obvious grading 𝔱([z,z1])=izi𝔱\mathfrak{t}(\mathbb{C}[z,z^{-1}])=\bigoplus_{i}z^{i}\mathfrak{t}. It is immediate that T(F)T(F) is graded compatible with GLn(𝔬)\operatorname{GL}_{n}(\mathfrak{o}).

At the opposite extreme, there is a unique class of maximal tori in GLn(F)\operatorname{GL}_{n}(F) 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 𝔖n\mathfrak{S}_{{n}}, i.e., the class of nn-cycles. A specific representative of this class is the standard Coxeter torus 𝒞=((ω))\mathcal{C}=\mathbb{C}(\!(\omega)\!)^{*} with Lie algebra 𝔠\mathfrak{c}. Note that ω\omega is regular semisimple — its eigenvalues are the nn distinct nnth roots of zz — so its centralizer ((ω))\mathbb{C}(\!(\omega)\!) is indeed a Cartan subalgebra. The natural grading by powers of ω\omega on [ω,ω1]\mathbb{C}[\omega,\omega^{-1}] induces a filtration on 𝔠\mathfrak{c}, and it is clear that 𝒞\mathcal{C} is graded compatible with II.

4. Toral and maximally ramified connections

4.1. Formal connections

A formal connection of rank nn is a connection ^\widehat{\nabla} on an FF-vector bundle VV of rank nn over the formal punctured disk Spec(F)\mathrm{Spec}(F). Given a trivialization ϕ\phi for VV (which is always trivializable), the connection can be written in matrix form as ^=d+[^]ϕ\widehat{\nabla}=d+[\widehat{\nabla}]_{\phi}, where [^]ϕΩF1(𝔤𝔩n(F))[\widehat{\nabla}]_{\phi}\in\Omega^{1}_{F}(\operatorname{\mathfrak{gl}}_{n}(F)). The loop group GLn(F)\operatorname{GL}_{n}(F) acts simply transitively on the set of trivializations via left multiplication. The corresponding action of GLn(F)\operatorname{GL}_{n}(F) on the connection matrix is given by the gauge action: if gGLn(F)g\in\operatorname{GL}_{n}(F), then g[^]ϕ=[^]gϕ=Ad(g)([^]ϕ)(dg)g1g\cdot[\widehat{\nabla}]_{\phi}=[\widehat{\nabla}]_{g\cdot\phi}=\mathrm{Ad}(g)([\widehat{\nabla}]_{\phi})-(dg)g^{-1}. Hence, the set of isomorphism classes of rank nn formal connections is isomorphic to the orbit space 𝔤𝔩n(F)dzz/GLn(F)\operatorname{\mathfrak{gl}}_{n}(F)\frac{dz}{z}/\operatorname{GL}_{n}(F) for the gauge action.

A formal connection ^\widehat{\nabla} 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, ^\widehat{\nabla} 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 nn. The slope is positive if and only if ^\widehat{\nabla} 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 𝔤𝔩n(F)\operatorname{\mathfrak{gl}}_{n}(F) [Was76]. To review, suppose that the matrix for ^\widehat{\nabla} with respect to ϕ\phi is expanded with respect to the degree filtration on 𝔤𝔩n(F)\operatorname{\mathfrak{gl}}_{n}(F); i.e., suppose

(6) [^]ϕ=(Mrzr+Mr+1zr+1+)dzz,[\widehat{\nabla}]_{\phi}=(M_{-r}z^{-r}+M_{-r+1}z^{-r+1}+\cdots)\tfrac{dz}{z},

where r0r\geq 0 and Mi𝔤𝔩n()M_{i}\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) for all ii. When the leading term MrM_{-r} is well-behaved, it gives useful information about the connection. For example, if MrM_{-r} is non-nilpotent, then slope(^)=r\mathrm{slope}(\widehat{\nabla})=r. Moreover, if r>0r>0 and MrM_{-r} is diagonalizable with distinct eigenvalues, then ^\widehat{\nabla} can be diagonalized into a “T(F)T(F)-formal type of depth rr”. This means that there exists gGL(F)g\in\operatorname{GL}(F) such that [^]gϕ[\widehat{\nabla}]_{g\phi} is an element of

𝒜(T(F),r){(Drzr++D1z+D0)dzzi,Di𝔱 and Dr has distinct eigenvalues}\mathcal{A}(T(F),r)\coloneqq\{(D_{-r}z^{-r}+\cdots+D_{1}z+D_{0})\tfrac{dz}{z}\mid\forall i,D_{i}\in\mathfrak{t}\text{ and }D_{-r}\text{ has distinct eigenvalues}\}

[Was76]. Note that many interesting connections have nilpotent leading terms. For example, the leading term of the formal Frenkel–Gross connection ^FG=d+ω1dzz\widehat{\nabla}_{\text{FG}}=d+\omega^{-1}\frac{dz}{z} [FG09] is strictly upper triangular (and thus nilpotent). In fact, the leading term is nilpotent no matter what trivialization one chooses for VV.

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 GLn\operatorname{GL}_{n}-stratum is a triple (P,r,β)(P,r,\beta) with PGLn(F)P\subset\operatorname{GL}_{n}(F) a parahoric subgroup, rr a nonnegative integer, and β\beta a functional on 𝔭r/𝔭r+1\mathfrak{p}^{r}/\mathfrak{p}^{r+1}. Consider the special case where PP corresponds to a simplex in the standard apartment (see Remark 3.1). Here, a functional β(𝔭r/𝔭r+1)\beta\in(\mathfrak{p}^{r}/\mathfrak{p}^{r+1})^{\vee} can be written uniquely as βdzz\beta^{\flat}\frac{dz}{z} for β\beta^{\flat} homogeneous (i.e., for β𝔭(r)\beta^{\flat}\in\mathfrak{p}(-r)). The stratum is called fundamental if β\beta^{\flat} is non-nilpotent. A formal connection ^\widehat{\nabla} contains the stratum (P,r,β)(P,r,\beta) (with respect to a fixed trivialization) if ^=d+Xdzz\widehat{\nabla}=d+X\frac{dz}{z} with X𝔭rX\in\mathfrak{p}^{-r} and β\beta induced by XdzzX\frac{dz}{z}. 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 PGLn(F)P\subset\operatorname{GL}_{n}(F) is a parahoric subgroup, then ePe_{P} denotes the period of the lattice chain stabilized by PP.

Theorem 4.1 ([BS13b, Theorem 4.10], [Sag17, Theorem 1]).

Any formal connection ^\widehat{\nabla} contains a fundamental stratum. If ^\widehat{\nabla} contains the fundamental stratum (P,r,β)(P,r,\beta), then slope(^)=r/eP\mathrm{slope}(\widehat{\nabla})=r/e_{P}.

We now investigate some examples. The connection in (6) (with Mi𝔤𝔩n()M_{i}\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) for all ii) contains the stratum (GLn(𝔬),r,Mrzrdzz)(\operatorname{GL}_{n}(\mathfrak{o}),r,M_{-r}z^{-r}\frac{dz}{z}), which is fundamental if and only if MrM_{-r} is non-nilpotent. The formal Frenkel–Gross connection ^FG\widehat{\nabla}_{\text{FG}} contains the fundamental stratum (I,1,ω1dzz)(I,1,\omega^{-1}\frac{dz}{z}). Moreover, any rank nn formal connection of the form ^=d+(aωr+X)dzz\widehat{\nabla}=d+(a\omega^{-r}+X)\frac{dz}{z}, with gcd(r,n)=1\gcd(r,n)=1, aa\in\mathbb{C}^{*}, and X𝔦r+1X\in\mathfrak{i}^{-r+1}, contains the fundamental stratum (I,r,aωrdzz)(I,r,a\omega^{-r}\frac{dz}{z}), and thus has slope r/nr/n.

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 PP comes from the standard apartment (see Remark 3.1). General definitions can be found in [BS13b] and [BS13a]. For such a PP, consider the stratum (P,r,βdzz)(P,r,\beta^{\flat}\frac{dz}{z}). If β\beta^{\flat} is regular semisimple, then its centralizer C(β)C(\beta^{\flat}) is regular semisimple, and we say that (P,r,βdzz)(P,r,\beta^{\flat}\frac{dz}{z}) is a C(β)C(\beta^{\flat})-regular stratum. A connection that contains an SS-regular stratum is called an SS-toral connection.

It turns out that toral connections do not exist for every maximal torus SS. In fact, an SS-toral connection of slope ss exists if and only if SS corresponds to a regular conjugacy class in 𝔖n\mathfrak{S}_{{n}} (in the sense of Springer [Spr74]) and if e2πise^{2\pi is} is a regular eigenvalue of this conjugacy class [BS13a]. The regular classes are parametrized by the partitions {bn/b}\{b^{n/b}\} (for positive divisors bb of nn) and {b(n1)/b,1}\{b^{(n-1)/b},1\} (for positive divisors bb of n1n-1). Representatives for each of the corresponding conjugacy classes of maximal tori are given by the SbS^{{b}}’s defined in §1.3. An SbS^{{b}}-toral connection has slope r/br/b for some r>0r>0 with gcd(r,b)=1\gcd(r,b)=1. Note that S1S^{{1}} is the diagonal torus T(F)T(F) while SnS^{{n}} is the standard Coxeter torus 𝒞\mathcal{C}.

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 bb of either nn or n1n-1. We define the set of SbS^{{b}}-formal types of depth rr (with gcd(r,b)=1\gcd(r,b)=1) by

𝒜(Sb,r)={AdzzA=i=0rAii=0r𝔰b(i) with Ar regular semisimple}.\mathcal{A}(S^{{b}},r)=\{A\tfrac{dz}{z}\mid A=\sum_{i=0}^{r}A_{-i}\in\bigoplus_{i=0}^{r}\mathfrak{s}^{b}(-i)\text{ with $A_{-r}$ regular semisimple}\}.

Every toral connection ^\widehat{\nabla} of slope r/br/b is formally isomorphic to a connection of the form d+Adzzd+A\tfrac{dz}{z} with Adzz𝒜(Sb,r)A\tfrac{dz}{z}\in\mathcal{A}(S^{{b}},r); we view this as a rational canonical form for ^\widehat{\nabla}.

We will need a more precise variation of this statement. As mentioned in §1.3, for each bb, there is a standard parahoric subgroup PbP^{b} which is compatible (in fact, graded compatible) with SbS^{{b}}.

Theorem 4.2 ([BS13b, Theorem 4.13], [BS13a, Theorem 5.1]).

Suppose ^\widehat{\nabla} is a formal connection containing an SbS^{{b}}-regular stratum (Pb,r,βdzz)(P^{b},r,\beta^{\flat}\frac{dz}{z}) with β𝔰b(r)\beta^{\flat}\in\mathfrak{s}^{b}(-r). Then there exists pPb,1p\in P^{b,1} such that p[^]p\cdot[\widehat{\nabla}] is a formal type in 𝒜(Sb,r)\mathcal{A}(S^{{b}},r) whose component in 𝔰b(r)dzz\mathfrak{s}^{b}(-r)\frac{dz}{z} is βdzz\beta^{\flat}\frac{dz}{z}.

4.4. Maximally ramified connections

Definition 4.3.

A formal connection ^\widehat{\nabla} of rank nn is called maximally ramified if it has slope r/nr/n with gcd(r,n)=1\gcd(r,n)=1.

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 ^\widehat{\nabla}. It is a well-known result of Turrittin [Tur55] and Levelt [Lev75] that after extending scalars to ((z1/b))\mathbb{C}(\!(z^{1/b})\!) for some b>0b\in\mathbb{Z}_{>0}, there exists a trivialization in which the matrix of ^\widehat{\nabla} is block-diagonal:

^=d+diag(p1(z1/b)idm1+R1,,pk(z1/b)idmk+Rk)dzz;\widehat{\nabla}=d+\mathrm{diag}(p_{1}(z^{-1/b})\mathrm{id}_{m_{1}}+R_{1},\dots,p_{k}(z^{-1/b})\mathrm{id}_{m_{k}}+R_{k})\tfrac{dz}{z};

here, the pip_{i}’s are polynomials and the RiR_{i}’s are nilpotent matrices. This is the Levelt–Turrittin normal form of ^\widehat{\nabla}. The slopes of ^\widehat{\nabla} are the nn rational numbers deg(pi)/b\deg(p_{i})/b, each appearing with multiplicity mim_{i}. This collection of invariants gives more detailed information about how irregular ^\widehat{\nabla} is than the single invariant slope(^)\mathrm{slope}(\widehat{\nabla}). Indeed, one can define slope(^)\mathrm{slope}(\widehat{\nabla}) to be the maximum of the slopes of ^\widehat{\nabla}.

One can show that the slopes are nonnegative rational numbers with denominators at most nn. Moreover, if at least one slope is r/nr/n with gcd(r,n)=1\gcd(r,n)=1, then all slopes are r/nr/n. Thus, ^\widehat{\nabla} is maximally ramified if at least one slope has denominator nn.

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 𝒞\mathcal{C}, we see that the 𝒞\mathcal{C}-formal types of depth rr are given by

𝒜(𝒞,r)={p(ω1)dzzp[x],deg(p)=r}.\mathcal{A}(\mathcal{C},r)=\{p(\omega^{-1})\tfrac{dz}{z}\mid p\in\mathbb{C}[x],\deg(p)=r\}.

We thus obtain the following result on rational canonical forms for maximally ramified connections:

Theorem 4.4 ([KS21a]).

Let ^\widehat{\nabla} be a maximally ramified connection of slope r/nr/n with gcd(r,n)=1\gcd(r,n)=1. Then ^\widehat{\nabla} is formally gauge equivalent to a connection of the form d+p(ω1)dzzd+p(\omega^{-1})\frac{dz}{z} with pp a polynomial of degree rr.

This theorem may be obtained as a direct corollary of Sabbah’s refined Levelt–Turrittin decomposition [Sab08, Corollary 3.3]. Indeed, since ^\widehat{\nabla} has slope r/nr/n, Sabbah’s theorem shows that it is formally isomorphic to a connection of the form [n](d+dϕ+λ)[n]_{*}(d+d\phi+\lambda), where [n]:Spec(F)Spec(F)[n]:\mathrm{Spec}(F)\to\mathrm{Spec}(F) is the nn-fold covering induced by zznz\mapsto z^{n}, ϕz1[z1]\phi\in z^{-1}\mathbb{C}[z^{-1}] has degree rr, and λ\lambda\in\mathbb{C}. It is now easy to conclude that ^\widehat{\nabla} has the desired rational canonical form, with the coefficients of pp determined by λ\lambda and the coefficients of ϕ\phi.

However, the theory of toral connections allows one to prove a generalized version of this theorem for G{G}-connections, where G{G} is a reductive group with connected Dynkin diagram [KS21a]. Here, nn is replaced by the Coxeter number hh, 𝒞\mathcal{C} is an appropriate fixed Coxeter torus in G(F){G}(F), and ^\widehat{\nabla} is formally isomorphic to d+𝒜d+\mathscr{A}, where 𝒜\mathscr{A} is a 𝒞\mathcal{C}-formal type of slope r/hr/h. Below, we provide a concise stratum-theoretic proof for the specific case G=GLn{G}=\operatorname{GL}_{n}, which is simpler than the general proof.

Proof.

By Theorem 4.1, ^\widehat{\nabla} contains a fundamental stratum (P,r,β)(P,r,\beta^{\prime}) with respect to some trivialization. Since slope(^)=r/eP\mathrm{slope}(\widehat{\nabla})=r/e_{P}, it follows that eP=ne_{P}=n and PP 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 ^\widehat{\nabla} contains the fundamental stratum (I,r,β)(I,r,\beta). The functional β\beta is represented by βdzz\beta^{\flat}\frac{dz}{z}, where β𝔦(r)\beta^{\flat}\in\mathfrak{i}(-r) is non-nilpotent.

By (4), β=diag(a1,,an)ωr\beta^{\flat}=\mathrm{diag}(a_{1},\dots,a_{n})\omega^{-r} for some constants aia_{i}. Let a=aia=\prod a_{i}. Since char(β)=xnazr\operatorname{char}(\beta^{\flat})=x^{n}-az^{-r}, it follows that a0a\neq 0. The polynomial thus has distinct roots, and β\beta^{\flat} is regular semisimple. Let a1/na^{1/n} be a fixed nnth root of aa. It is easy to see that there exists tTt\in T such that Ad(t)(β)=a1/nωr\mathrm{Ad}(t)(\beta^{\flat})=a^{1/n}\omega^{-r}, and since tt normalizes II, ^\widehat{\nabla} contains the stratum (I,r,a1/nωrdzz)(I,r,a^{1/n}\omega^{-r}\frac{dz}{z}). By Theorem 4.2, ^\widehat{\nabla} is formally isomorphic to d+(brωr++b0)dzzd+(b_{r}\omega^{-r}+\dots+b_{0})\frac{dz}{z}, where br=a1/nb_{r}=a^{1/n}. ∎

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 \nabla on a rank nn trivializable vector bundle VV over the complex Riemann sphere 1\mathbb{P}^{1}. To discuss the Deligne–Simpson problem, we need to define what it means for \nabla to be framable at a singularity with respect to a given toral formal type. We will assume that the singular point is 0. The only modification needed if the singularity is at an arbitrary point a1a\in\mathbb{P}^{1} is to replace the uniformizer zz by zaz-a if aa is finite and by z1z^{-1} if a=a=\infty.

Fix a trivialization ϕ\phi of VV, and write =d+[]ϕ\nabla=d+[\nabla]_{\phi}. The principal part of []ϕ[\nabla]_{\phi} is an element of 𝔤𝔩n([z1])dzz\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[z^{-1}])\tfrac{dz}{z}, and so may be viewed as a continuous functional on 𝔤𝔩n(𝔬)\operatorname{\mathfrak{gl}}_{n}(\mathfrak{o}) by (5). Similarly, the restriction of []ϕ[\nabla]_{\phi} to 𝔭b\mathfrak{p}^{b} is uniquely determined by the truncation of []ϕ/dzz[\nabla]_{\phi}/\frac{dz}{z} to i=0𝔭b(i)\bigoplus_{i=0}^{\infty}\mathfrak{p}^{b}(-i). Thus, if 𝒜=Adzz\mathscr{A}=A\frac{dz}{z} is an SbS^{{b}}-formal type, then 𝒜\mathscr{A} may naturally be viewed as an element of (𝔭b)(\mathfrak{p}^{b})^{\vee}.

Definition 5.1.

Let =d+[]ϕ\nabla=d+[\nabla]_{\phi} be a global connection on 1\mathbb{P}^{1} with a singular point at 0, and let 𝒜=Adzz\mathscr{A}=A\frac{dz}{z} be an SbS^{{b}}-formal type of depth rr. We say that \nabla is framable at 0 with respect to 𝒜\mathscr{A} if

  1. (1)

    there exists gGLn()g\in\operatorname{GL}_{n}(\mathbb{C}) such that []gϕ=g[]ϕ(𝔭b)r[\nabla]_{g\phi}=g\cdot[\nabla]_{\phi}\in(\mathfrak{p}^{b})^{-r} and []A(𝔭b)1r[\nabla]-A\in(\mathfrak{p}^{b})^{1-r}, and

  2. (2)

    there exists an element p(Pb)1p\in(P^{b})^{1} such that Ad(p)([]gϕdzz)|𝔭b=𝒜\mathrm{Ad}^{*}(p)([\nabla]_{g\phi}\frac{dz}{z})|_{\mathfrak{p}^{b}}=\mathscr{A}.

It is a consequence of Theorem 4.2 that this definition is equivalent to Definition 1.1.

Fix two disjoint subsets {a1,,am}\{a_{1},\dots,a_{m}\} and {b1,,b}\{b_{1},\dots,b_{\ell}\} of 1\mathbb{P}^{1} with m1m\geq 1 and 0\ell\geq 0. Let 𝐀=(𝒜1,,𝒜m)\mathbf{A}=(\mathscr{A}_{1},\dots,\mathscr{A}_{m}) be a collection of toral formal types at the aia_{i}’s, and let 𝐎=(𝒪1,,𝒪)\mathbf{O}=(\mathscr{O}_{1},\dots,\mathscr{O}_{\ell}) be a collection of adjoint orbits at the bjb_{j}’s. Assume that all of the orbits 𝒪j\mathscr{O}_{j} are nonresonant, meaning that no two eigenvalues of an orbit differ by a nonzero integer. One can now consider the category 𝒞(𝐀,𝐎)\mathscr{C}(\mathbf{A},\mathbf{O}) of meromorphic connections \nabla satisfying the following properties:

  1. (1)

    \nabla has irregular singularities at the aia_{i}’s, regular singularities at the bjb_{j}’s, and no other singular points;

  2. (2)

    for each ii, \nabla is framable at aia_{i} with respect to the formal type 𝒜i\mathscr{A}_{i}; and

  3. (3)

    for each jj, \nabla has residue at bjb_{j} in 𝒪j\mathscr{O}_{j}.

In [BS13b], Bremer and Sage constructed the moduli space (𝐀,𝐎)\operatorname{\mathcal{M}}(\mathbf{A},\mathbf{O}) 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 GLn()\operatorname{GL}_{n}(\mathbb{C}). At a regular singular point with adjoint orbit 𝒪\mathscr{O}, the manifold is just 𝒪\mathscr{O}, viewed as the coadjoint orbit 𝒪dzz\mathscr{O}\frac{dz}{z}. To define the symplectic manifold 𝒜\operatorname{\mathcal{M}}_{\mathscr{A}} associated to an SbS^{{b}}-toral formal type 𝒜\mathscr{A}, we first remark that the parahoric subgroup PbP^{b} is the pullback of a certain standard parabolic subgroup QbGLn()Q^{b}\subset\operatorname{GL}_{n}(\mathbb{C}) under the map GLn(𝔬)GLn()\operatorname{GL}_{n}(\mathfrak{o})\to\operatorname{GL}_{n}(\mathbb{C}) induced by z0z\mapsto 0. For example, P1=GLn(𝔬)P^{1}=\operatorname{GL}_{n}(\mathfrak{o}) is the pullback of Q1=GLn()Q^{1}=\operatorname{GL}_{n}(\mathbb{C}), and Pn=IP^{n}=I is the pullback of Qn=BQ^{n}=B. The “extended orbit” 𝒜(Pb\GLn())×𝔤𝔩n(𝔬)\operatorname{\mathcal{M}}_{\mathscr{A}}\subset(P^{b}\backslash\operatorname{GL}_{n}(\mathbb{C}))\times\operatorname{\mathfrak{gl}}_{n}(\mathfrak{o})^{\vee} is defined by

𝒜={(Qbg,α)(Ad(g)α)|𝔭bAd(Pb)(𝒜)}.\operatorname{\mathcal{M}}_{\mathscr{A}}=\{(Q^{b}g,\alpha)\mid(\mathrm{Ad}^{*}(g)\alpha)|_{\mathfrak{p}^{b}}\in\mathrm{Ad}^{*}(P^{b})(\mathscr{A})\}.

The group GLn()\operatorname{GL}_{n}(\mathbb{C}) acts on 𝒜\operatorname{\mathcal{M}}_{\mathscr{A}} via h(Qbg,α)=(Qbgh1,Ad(h)α)h\cdot(Q^{b}g,\alpha)=(Q^{b}gh^{-1},\mathrm{Ad}^{*}(h)\alpha), with moment map (Qbg,α)α|𝔤𝔩n()(Q^{b}g,\alpha)\mapsto\alpha|_{\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})}.

Theorem 5.2 ([BS13b, Theorem 5.26]).

The moduli space (𝐀,𝐎)\operatorname{\mathcal{M}}(\mathbf{A},\mathbf{O}) is given by

(𝐀,𝐎)[(i𝒜i)×(j𝒪j)]0GLn().\operatorname{\mathcal{M}}(\mathbf{A},\mathbf{O})\cong\left[\left(\prod_{i}\operatorname{\mathcal{M}}_{\mathscr{A}_{i}}\right)\times\left(\prod_{j}\mathscr{O}_{j}\right)\right]\sslash_{0}\operatorname{GL}_{n}(\mathbb{C}).

Let irr(𝐀,𝐎)\operatorname{\mathcal{M}}_{\mathrm{irr}}(\mathbf{A},\mathbf{O}) be the subset of (𝐀,𝐎)\operatorname{\mathcal{M}}(\mathbf{A},\mathbf{O}) consisting of irreducible connections. One can now restate the toral Deligne–Simpson problem as

Given the toral formal types 𝐀\mathbf{A} and the nonresonant adjoint orbits 𝐎\mathbf{O}, determine whether irr(𝐀,𝐎)\operatorname{\mathcal{M}}_{\mathrm{irr}}(\mathbf{A},\mathbf{O}) is nonempty.

Note that a 𝒞\mathcal{C}-toral connection is irreducible, so if any 𝒜i\mathscr{A}_{i} is 𝒞\mathcal{C}-toral, then the Deligne–Simpson problem reduces to the question of whether (𝐀,𝐎)\operatorname{\mathcal{M}}(\mathbf{A},\mathbf{O}) is nonempty.

5.2. Coxeter connections

We now specialize to an important special case: connections on 𝔾m\mathbb{G}_{m} with a maximally ramified singular point at 0 and (possibly) a regular singularity at \infty. 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 𝒜\mathscr{A} be a 𝒞\mathcal{C}-formal type, and let 𝒪\mathscr{O} be a nonresonant adjoint orbit. Then

(𝒜,𝒪){(α,Y)α𝔤𝔩n([z1])dzz,Y𝒪,α|𝔦Ad(I)(𝒜), and Res(α)+Y=0}/B.\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O})\cong\{(\alpha,Y)\mid\alpha\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[z^{-1}])\tfrac{dz}{z},Y\in\mathscr{O},\alpha|_{\mathfrak{i}}\in\mathrm{Ad}^{*}(I)(\mathscr{A}),\text{ and }\operatorname{\mathrm{Res}}(\alpha)+Y=0\}/B.
Proof.

Applying Theorem 5.2, we have

(𝒜,𝒪)(𝒜×𝒪)0GLn(){(Bg,α,Y)(Bg,α)𝒜,Y𝒪, and Res(α)+Y=0}/GLn(){(B,α,Y)α𝔤𝔩n([z1])dzz,Y𝒪,α|𝔦Ad(I)(𝒜), and Res(α)+Y=0}/B.\begin{array}[]{rll}\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O})&\cong(\operatorname{\mathcal{M}}_{\mathscr{A}}\times\mathscr{O})\sslash_{0}\operatorname{GL}_{n}(\mathbb{C})\\ &\cong\{(Bg,\alpha,Y)\mid(Bg,\alpha)\in\operatorname{\mathcal{M}}_{\mathscr{A}},Y\in\mathscr{O},\text{ and }\operatorname{\mathrm{Res}}(\alpha)+Y=0\}/\operatorname{GL}_{n}(\mathbb{C})\\ &\cong\{(B,\alpha,Y)\mid\alpha\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[z^{-1}])\frac{dz}{z},Y\in\mathscr{O},\alpha|_{\mathfrak{i}}\in\mathrm{Ad}^{*}(I)(\mathscr{A}),\text{ and }\operatorname{\mathrm{Res}}(\alpha)+Y=0\}/B.\end{array}

We can now state the solution to the Deligne–Simpson problem for Coxeter connections. For a given 𝒞\mathcal{C}-formal type and a monic polynomial qq of degree nn, let

DS(𝒜,q)={𝒪πq(𝒜,𝒪)}.\mathrm{DS}(\mathscr{A},q)=\{\mathscr{O}\in\pi_{q}\mid\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O})\neq\varnothing\}.
Theorem 5.4.

Let rr and nn be positive integers with gcd(r,n)=1\gcd(r,n)=1, let 𝒜\mathscr{A} be a maximally ramified formal type of slope r/nr/n, and let q=i=1s(xai)mi[x]q=\prod_{i=1}^{s}(x-a_{i})^{m_{i}}\in\mathbb{C}[x] with a1,,asa_{1},\ldots,a_{s}\in\mathbb{C} distinct modulo \mathbb{Z}. Then

DS(𝒜,q)={𝒪qrif Res(Tr(𝒜))=i=1smiai,else.\mathrm{DS}(\mathscr{A},q)=\begin{cases}\langle{\mathscr{O}_{q}^{r}}\rangle&\text{if }\operatorname{\mathrm{Res}}(\operatorname{Tr}(\mathscr{A}))=-\sum_{i=1}^{s}m_{i}a_{i},\\ \varnothing&\text{else}.\end{cases}

We prove this theorem in the next subsection.

Remark 5.5.

If we write 𝒜=p(ω1)dzz\mathscr{A}=p(\omega^{-1})\frac{dz}{z}, then (𝒜,𝒪)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O}) is nonempty if and only if np(0)=Tr(𝒪)np(0)=-\operatorname{Tr}(\mathscr{O}) and if 𝒪\mathscr{O} has at most rr Jordan blocks for each eigenvalue. This second condition is always satisfied if r>nr>n. Note that the solution only depends on the slope and the residue of the formal type.

This theorem immediately gives the corresponding result for SLn\operatorname{SL}_{n}-connections. (In terms of a global trivialization, one may view an SLn\operatorname{SL}_{n}-connection on 1\mathbb{P}^{1} as an operator d+Xdzzd+X\frac{dz}{z} with X𝔰𝔩n([z,z1])X\in\operatorname{\mathfrak{sl}}_{n}(\mathbb{C}[z,z^{-1}]).) In this case, maximally ramified formal types are of the form p(ω1)dzzp(\omega^{-1})\frac{dz}{z} with p(0)=0p(0)=0 and Tr(𝒪)=0\operatorname{Tr}(\mathscr{O})=0, so the trace condition becomes vacuous. Accordingly, we obtain the following solution to the Deligne–Simpson for Coxeter SLn\operatorname{SL}_{n}-connections.

Corollary 5.6.

The moduli space of Coxeter SLn\operatorname{SL}_{n}-connections with formal type of slope r/nr/n and adjoint orbit 𝒪\mathscr{O} is nonempty if and only if 𝒪𝒪char(𝒪)r\mathscr{O}\succeq\mathscr{O}_{\operatorname{char}(\mathscr{O})}^{r}.

The notion of a Coxeter G{G}-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 I1I^{1}-orbits in 𝔦\mathfrak{i}^{\vee}.

Lemma 5.7.

For any r,sr,s\in\mathbb{Z} with rr relatively prime to nn, the linear map ϕr,s:𝔱𝔤𝔩n()\phi_{r,s}:\mathfrak{t}\to\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) given by γ[γωrs,ωr]ωs\gamma\mapsto[\gamma\omega^{r-s},\omega^{-r}]\omega^{s} has image 𝔱𝔰𝔩n()\mathfrak{t}\cap\operatorname{\mathfrak{sl}}_{n}(\mathbb{C}).

Proof.

Since [γωrs,ωr]=γωsωrγωrs=(γωrγωr)ωs[\gamma\omega^{r-s},\omega^{-r}]=\gamma\omega^{-s}-\omega^{-r}\gamma\omega^{r-s}=(\gamma-\omega^{-r}\gamma\omega^{r})\omega^{-s}, we have ϕr,s(γ)=γωrγωr.\phi_{r,s}(\gamma)=\gamma-\omega^{-r}\gamma\omega^{r}. Note that ωrN(T)\omega^{-r}\in N(T) — indeed, it is a monomial matrix which represents the r-rth power of a Coxeter element in the Weyl group 𝔖n\mathfrak{S}_{{n}} — so the entries in ωrγωr𝔱-\omega^{-r}\gamma\omega^{r}\in\mathfrak{t} are a reordering of the entries in γ𝔱\gamma\in\mathfrak{t}. It follows that ϕr,s(𝔱)𝔱𝔰𝔩n()\phi_{r,s}(\mathfrak{t})\subset\mathfrak{t}\cap\operatorname{\mathfrak{sl}}_{n}(\mathbb{C}). To show equality, it suffices to check that dimkerϕr,s=1\dim\ker\phi_{r,s}=1. We have γkerϕr,s\gamma\in\ker\phi_{r,s} if and only if γ\gamma commutes with ωr\omega^{-r}. Since gcd(r,n)=1\gcd(r,n)=1, the centralizers of ωr\omega^{-r} and ω\omega coincide and equal ((ω))\mathbb{C}(\!(\omega)\!). Thus, ker(ϕr,s)=((ω))𝔱=id\ker(\phi_{r,s})=\mathbb{C}(\!(\omega)\!)\cap\mathfrak{t}=\mathbb{C}\mathrm{id}. ∎

We can now give a convenient representative of certain coadjoint I1I^{1}-orbits in 𝔦\mathfrak{i}^{\vee}.

Proposition 5.8.

Let rr and nn be positive integers with gcd(r,n)=1\gcd(r,n)=1. Suppose α𝔦\alpha\in\mathfrak{i}^{\vee} is given by α=(aωr+X)dzz\alpha=(a\omega^{-r}+X)\frac{dz}{z} for some aa\in\mathbb{C}^{*} and X𝔦r+1X\in\mathfrak{i}^{-r+1}. Then there exists gI1g\in I^{1} such that

(7) Ad(g)α=(aωr+i=0r1cieiiωi)dzz.\mathrm{Ad}^{*}(g)\alpha=(a\omega^{-r}+\sum_{i=0}^{r-1}c_{i}e_{ii}\omega^{-i})\tfrac{dz}{z}.
Proof.

All elements of 𝔦1dzz\mathfrak{i}^{1}\frac{dz}{z} represent the zero functional on 𝔦\mathfrak{i}, so it suffices to show that if α(aωr+i=s+1r1cieiiωi+βωs+𝔦s+1)dzz\alpha\in(a\omega^{-r}+\sum_{i=s+1}^{r-1}c_{i}e_{ii}\omega^{-i}+\beta\omega^{-s}+\mathfrak{i}^{-s+1})\frac{dz}{z} for 1sr11\leq s\leq r-1, then there exists γ𝔱\gamma\in\mathfrak{t} such that Ad(1+γωrs)(α)(aωr+i=s+1r1cieiiωi+csessωs+𝔦s+1)dzz\mathrm{Ad}^{*}(1+\gamma\omega^{r-s})(\alpha)\in(a\omega^{-r}+\sum_{i=s+1}^{r-1}c_{i}e_{ii}\omega^{-i}+c_{s}e_{ss}\omega^{-s}+\mathfrak{i}^{-s+1})\frac{dz}{z}.

Write β=(b1,,bn)\beta=(b_{1},\dots,b_{n}). By Lemma 5.7, there exists γ𝔱\gamma\in\mathfrak{t} such that

ϕr,s(aγ)=(b1,,bs1,isbi,bs+1,,bn).\phi_{r,s}(a\gamma)=(-b_{1},\ldots,-b_{s-1},\sum_{i\neq s}b_{i},-b_{s+1},\ldots,-b_{n}).

Setting cs=ibic_{s}=\sum_{i}b_{i} gives

Ad(1+γωrs)(α)\displaystyle\mathrm{Ad}^{*}(1+\gamma\omega^{r-s})(\alpha) (aωr+i=s+1r1cieiiωi+βωs+a[γωrs,ωr]+𝔦s+1)dzz\displaystyle\in(a\omega^{-r}+\sum_{i=s+1}^{r-1}c_{i}e_{ii}\omega^{-i}+\beta\omega^{-s}+a[\gamma\omega^{r-s},\omega^{-r}]+\mathfrak{i}^{-s+1})\tfrac{dz}{z}
=(aωr+i=s+1r1cieiiωi+(β+ϕr,s(aγ))ωs+𝔦s+1)dzz\displaystyle=(a\omega^{-r}+\sum_{i=s+1}^{r-1}c_{i}e_{ii}\omega^{-i}+(\beta+\phi_{r,s}(a\gamma))\omega^{-s}+\mathfrak{i}^{-s+1})\tfrac{dz}{z}
=(aωr+i=s+1r1cieiiωi+csessωs+𝔦s+1)dzz.\displaystyle=(a\omega^{-r}+\sum_{i=s+1}^{r-1}c_{i}e_{ii}\omega^{-i}+c_{s}e_{ss}\omega^{-s}+\mathfrak{i}^{-s+1})\tfrac{dz}{z}.

Corollary 5.9.

Given α\alpha as in the proposition and D𝔱D\in\mathfrak{t} with Tr(D)=Res(Tr(α))\operatorname{Tr}(D)=\operatorname{\mathrm{Res}}(\operatorname{Tr}(\alpha)), there exists hI1h\in I^{1} such that Res(Ad(h)α)=Res(aωrdzz)+D\operatorname{\mathrm{Res}}(\mathrm{Ad}^{*}(h)\alpha)=\operatorname{\mathrm{Res}}(a\omega^{-r}\tfrac{dz}{z})+D.

Proof.

Given gg as in (7), we claim that

Res(Ad(g)α)Res(aωrdzz)+𝔱.\operatorname{\mathrm{Res}}(\mathrm{Ad}^{*}(g)\alpha)\in\operatorname{\mathrm{Res}}(a\omega^{-r}\tfrac{dz}{z})+\mathfrak{t}.

To see this, write s=kn+us=kn+u with 0<un0<u\leq n. Recall that NuN_{u} is the matrix with 11’s on the uuth subdiagonal and 0’s elsewhere. Similarly, let EuE_{u} be the matrix whose only nonzero entries are 11’s on the (nu)(n-u)th superdiagonal. (We make the convention that Nn=0N_{n}=0 and En=idE_{n}=\mathrm{id}.) It is easy to verify that ωs=zk(Nu+z1Eu)\omega^{-s}=z^{-k}(N_{u}+z^{-1}E_{u}). Since euuEu=eune_{uu}E_{u}=e_{un}, and euuNu=0e_{uu}N_{u}=0, we have essωs=z(k+1)esne_{ss}\omega^{-s}=z^{-(k+1)}e_{sn}. In particular, Res(essωsdzz)=0\operatorname{\mathrm{Res}}(e_{ss}\omega^{-s}\frac{dz}{z})=0 if s>0s>0 and equals enne_{nn} if s=0s=0. Applying this to (7) gives Res(Ad(g)α)=Res(aωrdzz)+c0enn\operatorname{\mathrm{Res}}(\mathrm{Ad}^{*}(g)\alpha)=\operatorname{\mathrm{Res}}(a\omega^{-r}\frac{dz}{z})+c_{0}e_{nn} as desired.

To complete the proof, it suffices to show that if Res(α)=Res(aωrdzz)+D\operatorname{\mathrm{Res}}(\alpha)=\operatorname{\mathrm{Res}}(a\omega^{-r}\frac{dz}{z})+D^{\prime} for some D𝔱D^{\prime}\in\mathfrak{t}, and if D𝔱D\in\mathfrak{t} satisfies Tr(D)=Tr(D)\operatorname{Tr}(D)=\operatorname{Tr}(D^{\prime}), then there exists γ𝔱\gamma\in\mathfrak{t} such that Res(Ad(1+γωr)α)=Res(aωrdzz)+D\operatorname{\mathrm{Res}}(\mathrm{Ad}^{*}(1+\gamma\omega^{r})\alpha)=\operatorname{\mathrm{Res}}(a\omega^{-r}\frac{dz}{z})+D. Since Res(Ad(1+γωr)α)=Res(aωrdzz)+D+a[γωr,ωr]\operatorname{\mathrm{Res}}(\mathrm{Ad}^{*}(1+\gamma\omega^{r})\alpha)=\operatorname{\mathrm{Res}}(a\omega^{-r}\frac{dz}{z})+D^{\prime}+a[\gamma\omega^{r},\omega^{-r}], it suffices to find γ𝔱\gamma\in\mathfrak{t} such that D+a[γωr,ωr]=DD^{\prime}+a[\gamma\omega^{r},\omega^{-r}]=D. This follows from Lemma 5.7, since a1(DD)𝔱𝔰𝔩n()=Image(ϕr,0)a^{-1}(D-D^{\prime})\in\mathfrak{t}\cap\operatorname{\mathfrak{sl}}_{n}(\mathbb{C})=\operatorname{Image}(\phi_{r,0}). ∎

Lemma 5.10.

Let rr and nn be positive integers with gcd(r,n)=1\gcd(r,n)=1. Suppose that 𝒜𝒜(𝒞,r)\mathscr{A}\in\mathcal{A}(\mathcal{C},r) and 𝒪\mathscr{O} is a nonresonant adjoint orbit in 𝔤𝔩n()\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}). If (𝒜,𝒪)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O})\neq\varnothing, then the Jordan form of 𝒪\mathscr{O} has at most rr blocks for each eigenvalue.

Proof.

This is trivial for r>nr>n, so assume that r<nr<n. Choose α𝔤𝔩n([z1])dzz\alpha\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[z^{-1}])\frac{dz}{z} and Y𝒪Y\in\mathscr{O} such that α|𝔦Ad(I)(𝒜)\alpha|_{\mathfrak{i}}\in\mathrm{Ad}^{*}(I)(\mathscr{A}) and Y=Res(Ad(b)α)Y=-\operatorname{\mathrm{Res}}(\mathrm{Ad}^{*}(b)\alpha). Write 𝒜=(aωr+X)dzz\mathscr{A}=(a\omega^{-r}+X)\frac{dz}{z} for some aa\in\mathbb{C}^{*} and for some X𝔠r+1X\in\mathfrak{c}^{-r+1}. Since I=TI1I=TI^{1}, we may assume without loss of generality that α|𝔦Ad(I1)𝒜\alpha|_{\mathfrak{i}}\in\mathrm{Ad}^{*}(I^{1})\mathscr{A}. This implies that α=(aωr+X)dzz\alpha=(a\omega^{-r}+X^{\prime})\frac{dz}{z} for some X𝔦r+1X^{\prime}\in\mathfrak{i}^{-r+1}. It is easy to see that Res(Xdzz)\operatorname{\mathrm{Res}}(X^{\prime}\frac{dz}{z}) has 0’s in the rrth subdiagonal and below. In the notation of §2, this means that YRes(α)VrY\coloneqq-\operatorname{\mathrm{Res}}(\alpha)\in V^{r}. By Theorem 2.3, YY has at most rr blocks for each eigenvalue. ∎

Lemma 5.11.

Let rr and nn be positive integers with gcd(r,n)=1\gcd(r,n)=1, let 𝒜\mathscr{A} be a 𝒞\mathcal{C}-formal type of depth rr, and let q=i=1s(xai)mi[x]q=\prod_{i=1}^{s}(x-a_{i})^{m_{i}}\in\mathbb{C}[x] be a degree nn polynomial with a1,,asa_{1},\ldots,a_{s}\in\mathbb{C} distinct modulo \mathbb{Z}. If Res(Tr(𝒜))=i=1smiai\operatorname{\mathrm{Res}}(\operatorname{Tr}(\mathscr{A}))=-\sum_{i=1}^{s}m_{i}a_{i}, then 𝒪qrDS(𝒜,q)\mathscr{O}_{q}^{r}\in\mathrm{DS}(\mathscr{A},q).

Proof.

Write 𝒜=(aωr+X)dzz\mathscr{A}=(a\omega^{-r}+X)\frac{dz}{z} for some aa\in\mathbb{C}^{*} and for some X𝔠r+1X\in\mathfrak{c}^{-r+1}. By Proposition 2.4, there exists DD with trace i=1smiai\sum_{i=1}^{s}m_{i}a_{i} such that aNr+D𝒪qr-aN_{r}+D\in\mathscr{O}_{q}^{r}. We now apply Corollary 5.9 to obtain gI1g\in I^{1} such that Res(Ad(g)𝒜)=aNrD\operatorname{\mathrm{Res}}(\mathrm{Ad}^{*}(g)\mathscr{A})=aN_{r}-D. Let α𝔤𝔩n(𝔬)\alpha\in\operatorname{\mathfrak{gl}}_{n}(\mathfrak{o})^{\vee} be any functional extending Ad(g)𝒜𝔦\mathrm{Ad}^{*}(g)\mathscr{A}\in\mathfrak{i}^{\vee}. Then the BB-orbit of (α,aNr+D)(\alpha,-aN_{r}+D) gives a point in (𝒜,𝒪qr)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O}_{q}^{r}), so 𝒪qrDS(𝒜,q)\mathscr{O}_{q}^{r}\in\mathrm{DS}(\mathscr{A},q). ∎

Proof of Theorem 5.4.

If (𝒜,𝒪)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O})\neq\varnothing, then there exists α𝔤𝔩n(𝔬)\alpha\in\operatorname{\mathfrak{gl}}_{n}(\mathfrak{o})^{\vee} and Y𝒪Y\in\mathscr{O} with Res(α)+Y=0\operatorname{\mathrm{Res}}(\alpha)+Y=0. Since Res(Tr(α))=Res(Tr(𝒜))\operatorname{\mathrm{Res}}(\operatorname{Tr}(\alpha))=\operatorname{\mathrm{Res}}(\operatorname{Tr}(\mathscr{A})), we see that Res(Tr(𝒜))=Tr(Y)=i=1smiai\operatorname{\mathrm{Res}}(\operatorname{Tr}(\mathscr{A}))=-\operatorname{Tr}(Y)=-\sum_{i=1}^{s}m_{i}a_{i}.

By Lemma 5.11, 𝒪qrDS(𝒜,q)\mathscr{O}_{q}^{r}\in\mathrm{DS}(\mathscr{A},q), and by Lemma 5.10, DS(𝒜,q)𝒪qr\mathrm{DS}(\mathscr{A},q)\subset\langle{\mathscr{O}_{q}^{r}}\rangle. To show equality, take 𝒪𝒪pr\mathscr{O}\in\langle{\mathscr{O}_{p}^{r}}\rangle. Since 𝒪qrDS(𝒜,q)\mathscr{O}_{q}^{r}\in\mathrm{DS}(\mathscr{A},q), there exists some X𝔤𝔩n([z1])X\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[z^{-1}]) and Y𝒪qrY\in\mathscr{O}_{q}^{r} such that (Xdzz)|𝔦Ad(I)(𝒜)(X\frac{dz}{z})|_{\mathfrak{i}}\in\mathrm{Ad}^{*}(I)(\mathscr{A}) and Res(Xdzz)+Y=0\operatorname{\mathrm{Res}}(X\frac{dz}{z})+Y=0. By a theorem of Krupnik [Kru97, Theorem 1], there exists a strictly upper triangular matrix N𝔤𝔩n()N\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}) such that Y+N𝒪Y+N\in\mathscr{O}. Note that N𝔦1N\subset\mathfrak{i}^{1}, so (Ndzz)|𝔦=0(N\frac{dz}{z})|_{\mathfrak{i}}=0. Thus, (XN)𝔤𝔩n([z1])(X-N)\in\operatorname{\mathfrak{gl}}_{n}(\mathbb{C}[z^{-1}]), ((XN)dzz)|𝔦Ad(I)(𝒜)((X-N)\frac{dz}{z})|_{\mathfrak{i}}\in\mathrm{Ad}^{*}(I)(\mathscr{A}), and Res((XN)dzz)+(Y+N)=0\operatorname{\mathrm{Res}}((X-N)\frac{dz}{z})+(Y+N)=0. Hence (𝒜,𝒪)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O})\neq\varnothing, and the proof is finished. ∎

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 \nabla be a meromorphic G{G}-connection on 1\mathbb{P}^{1} which is regular on the Zariski-open set U=1{x1,,xk}U=\mathbb{P}^{1}\setminus\{x_{1},\dots,x_{k}\}. Let ^xi\widehat{\nabla}_{x_{i}} denote the induced formal G{G}-connection at xix_{i}. The connection is called physically rigid if, for any meromorphic G{G}-connection \nabla^{\prime} which is regular on UU and satisfies ^xi^xi\widehat{\nabla}^{\prime}_{x_{i}}\cong\widehat{\nabla}_{x_{i}} for all ii, we have \nabla^{\prime}\cong\nabla.

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 H1(1,j!ad)=0H^{1}(\mathbb{P}^{1},j_{!*}\mathrm{ad}_{\nabla})=0, where j:U1j:U\hookrightarrow\mathbb{P}^{1} is the inclusion. If \nabla is irreducible, then \nabla being cohomologically rigid implies that \nabla admits no infinitesimal deformations [Yun14]. For G=GLn(){G}=\operatorname{GL}_{n}(\mathbb{C}), Bloch and Esnault have shown that cohomological rigidity and physical rigidity are equivalent [BE04].

We call a 𝒞\mathcal{C}-formal type homogeneous when it is of the form aωrdzza\omega^{-r}\frac{dz}{z} for aa\in\mathbb{C}^{*}; it gives rise to a “homogeneous Coxeter connection” d+aωrdzzd+a\omega^{-r}\frac{dz}{z} on 𝔾m\mathbb{G}_{m}. This connection has a toral singularity at 0 and (possibly) a regular singularity at \infty with unipotent monodromy. This notion also makes sense for any complex simple group G{G} [KS21b]. Again, one can define formal types with respect to a certain maximal torus 𝒞GG(F)\mathcal{C}_{{G}}\subset{G}(F) called the Coxeter torus. Moreover, if rr is any positive integer relatively prime to the Coxeter number hh, there exists an element ωrLie(𝒞G)\omega_{-r}\in\operatorname{\mathrm{Lie}}(\mathcal{C}_{G}) such that aωrdzza\omega_{-r}\frac{dz}{z} may be viewed as a homogeneous formal type. One can again consider the corresponding Coxeter G{G}-connection on 𝔾m\mathbb{G}_{m} with a homogeneous 𝒞G\mathcal{C}_{G}-toral irregular singularity of slope r/hr/h at 0 and (possibly) a regular singular point with unipotent monodromy at \infty. The case r=1r=1 is the remarkable rigid connection constructed by Frenkel and Gross [FG09].

In [KS21b], Kamgarpour and Sage determined when these homogeneous Coxeter G{G}-connections are (cohomologically) rigid for any simple G{G}. It turns out r=1r=1 and r=h+1r=h+1 always give rigid connections: the Frenkel–Gross and “Airy G{G}-connection” respectively. For the exceptional groups, there are no other such rigid connections except for r=7r=7 in E7E_{7}. However, for the classical groups, one also has rigidity for 1<r<h1<r<h with rr satisfying certain divisibility conditions. For example, in type AA, these connections are rigid if and only if r|(n±1)r|(n\pm 1).

In this paper, we generalize this result in type AA to give a classification of rigid framable Coxeter connections with unipotent monodromy at \infty. (Framable means that we only consider Coxeter connections contained in the relevant framable moduli space (𝒜,𝒪)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O}).)

Theorem 6.1.

Let 𝒜\mathscr{A} be a rank nn maximally ramified formal type of slope r/nr/n, and let 𝒪\mathscr{O} be any nilpotent orbit with 𝒪𝒪xnr\mathscr{O}\succeq\mathscr{O}^{r}_{x^{n}}. Then there exists a rigid connection with the given formal type and unipotent monodromy determined by 𝒪\mathscr{O} if and only if 𝒪=𝒪xnr\mathscr{O}=\mathscr{O}^{r}_{x^{n}} and r|(n±1)r|(n\pm 1).

If \nabla is a Coxeter connection, let \mathcal{I} denote the global differential Galois group, and let 0\mathcal{I}_{0} and \mathcal{I}_{\infty} denote the local differential Galois groups at 0 and \infty. These differential Galois groups are all algebraic subgroups of GLn()\operatorname{GL}_{n}(\mathbb{C}). Also, let Irr(ad^0)\operatorname{Irr}(\mathrm{ad}_{\widehat{\nabla}_{0}}) denote the irregularity of the formal connection ad^0\mathrm{ad}_{\widehat{\nabla}_{0}}. This is the sum of all the slopes appearing in the slope decomposition of ad^0\mathrm{ad}_{\widehat{\nabla}_{0}}; it is a nonnegative integer.

Let n()=Irr(ad^0)dim(𝔤𝔩n()0)dim(𝔤𝔩n())+2dim(𝔤𝔩n())n(\nabla)=\operatorname{Irr}(\mathrm{ad}_{\widehat{\nabla}_{0}})-\dim(\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})^{\mathcal{I}_{0}})-\dim(\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})^{\mathcal{I}_{\infty}})+2\dim(\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})^{\mathcal{I}}), and let j:𝔾m1j:\mathbb{G}_{m}\hookrightarrow\mathbb{P}^{1} be the inclusion. It is shown in [FG09, Proposition 11] that dim(H1(1,j!ad))=n()\dim(H^{1}(\mathbb{P}^{1},j_{!*}\mathrm{ad}_{\nabla}))=n(\nabla). Thus, we get the numerical criterion for rigidity that \nabla is rigid if and only if n()=0n(\nabla)=0.

We now calculate the numerical criterion as in Section 4 of [KS21b]. The local differential Galois group 0\mathcal{I}_{0} is given by 0Hθ\mathcal{I}_{0}\cong H\ltimes\langle\theta\rangle, where HH is a certain torus containing a regular semisimple element and θ\theta is an order nn element of N(H)N(H) [KS21a]. The centralizer of HH is thus a maximal torus TT^{\prime}, and θN(T)\theta\in N(T^{\prime}) represents a Coxeter element in the Weyl group. We conclude (as in [KS21b]) that 𝔤𝔩n()0=(𝔤𝔩n()H)θ=Lie(T)θ=id\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})^{\mathcal{I}_{0}}=(\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})^{H})^{\theta}=\operatorname{\mathrm{Lie}}(T^{\prime})^{\theta}=\mathbb{C}\mathrm{id}. Since 0\mathcal{I}_{0}\subset\mathcal{I}, we also have 𝔤𝔩n()=id\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})^{\mathcal{I}}=\mathbb{C}\mathrm{id}.

In general, if ^\widehat{\nabla} is a toral G{G}-connection with slope ss, then by Lemma 19 of [KS19], Irr(ad^)=s|Φ|\operatorname{Irr}(\mathrm{ad}_{\widehat{\nabla}})=s|\Phi|, where Φ\Phi is the set of roots with respect to the maximal torus TT. In our particular case, we obtain Irr(ad^0)=rnn(n1)=r(n1)\operatorname{Irr}(\mathrm{ad}_{\widehat{\nabla}_{0}})=\frac{r}{n}n(n-1)=r(n-1).

Finally, if we fix some element N𝒪𝒪N_{\mathscr{O}}\in\mathscr{O}, then ^\widehat{\nabla}_{\infty} is regular singular with unipotent monodromy exp(2πiN𝒪)\exp(2\pi iN_{\mathscr{O}}). This means that exp(2πiN𝒪)\mathcal{I}_{\infty}\cong\langle\exp(2\pi iN_{\mathscr{O}})\rangle, so 𝔤𝔩n()=C(N𝒪)\operatorname{\mathfrak{gl}}_{n}(\mathbb{C})^{\mathcal{I}_{\infty}}=C(N_{\mathscr{O}}), the centralizer of N𝒪N_{\mathscr{O}}. Using the fact that dim(𝒪)=n2dim(C(N𝒪))\dim(\mathscr{O})=n^{2}-\dim(C(N_{\mathscr{O}})), we obtain

n()=(rn1)(n1)+dim(𝒪).n(\nabla)=(r-n-1)(n-1)+\dim(\mathscr{O}).

Since (𝒜,𝒪xnr)\operatorname{\mathcal{M}}(\mathscr{A},\mathscr{O}_{x^{n}}^{r})\neq\varnothing, we can take \nabla^{\prime} with this local behavior. Suppose that 𝒪𝒪xnr\mathscr{O}\succneqq\mathscr{O}_{x^{n}}^{r}. We then have n()=(rn1)(n1)+dim(𝒪)>(rn1)(n1)+dim(𝒪xnr)=n()=0n(\nabla)=(r-n-1)(n-1)+\dim(\mathscr{O})>(r-n-1)(n-1)+\dim(\mathscr{O}_{x^{n}}^{r})=n(\nabla^{\prime})=0. It follows that in this case, \nabla is never rigid.

Finally, take 𝒪=𝒪xnr\mathscr{O}=\mathscr{O}_{x^{n}}^{r}. It was shown in [KS21b] that dim(𝒪xnr)=(n+1r)(n1)\dim(\mathscr{O}_{x^{n}}^{r})=(n+1-r)(n-1) if and only if r|(n±1)r|(n\pm 1). This finishes the proof of the theorem.

References

  • [BE04] S. Bloch and H. Esnault, Local Fourier transforms and rigidity for 𝒟\mathcal{D}-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 G{{G}}-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 KK-types for flat GG-bundles, Int. Math. Res. Not. IMRN (2018), 3507–3555.
  • [Bus87] C. J. Bushnell, Hereditary orders, Gauss sums and supercuspidal representations of GL(N){GL}({N}), 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 G{G}-connections, in preparation, 2021.
  • [KS21b] by same author, Rigid connections on 1\mathbb{P}^{1} 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 KK-types for pp-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.