1 Introduction

Let G be a connected reductive group over an algebraically closed field \(\Bbbk \) of characteristic p > 0, and let H be a connected reductive subgroup. Recall that (G,H) is said to be a Donkin pair or a good filtration pair if every G-module with a good filtration still has a good filtration when regarded as an H-module.

Now let x be a nilpotent element in the Lie algebra of G, and let GxG be its stabilizer. If p is good for G, then the theory of associated cocharacters is available, and this gives rise to a decomposition

$$ G^{x} = {G}_{\text{red}}^{x} \ltimes {G}_{\text{unip}}^{x} $$

where \({G}_{\text {unip}}^{x}\) is a connected unipotent group, and \({G}_{\text {red}}^{x}\) is a (possibly disconnected) group whose identity component \(({G}_{\text {red}}^{x})^{\circ }\) is reductive (cf. [11, 5.10]). The main result of this paper is the following.

Theorem 1.1

Let G be a connected reductive group over an algebraically closed field \(\Bbbk \) of good characteristic. For any nilpotent element x in its Lie algebra, \((G, ({G}_{\text {red}}^{x})^{\circ })\) is a Donkin pair.

Now suppose that HG is a possibly disconnected reductive subgroup, i.e., a group whose identity component H is reductive. If the characteristic of \(\Bbbk \) does not divide the order of the finite group H/H, then the category of finite-dimensional H-modules is a highest-weight category, as shown in [2]. In particular, it makes sense to speak of good filtrations for H-modules, and so the definition of “Donkin pair” makes sense for (G,H).

In order to apply this notion in the case where \(H = {G}_{\text {red}}^{x}\), we must impose a slightly stronger condition on p: we require it to be pretty good in the sense of [9, Definition 2.11]. (In general, this condition is intermediate between “good” and “very good.” It coincides with “very good” for semisimple simply connected groups, whereas for GLn, all primes are pretty good.) This is equivalent to requiring G to be standard in the sense of [14, §4]. It follows from [8, Theorem 1.8] and [3, Lemma 2.1] that when p is pretty good for G, it does not divide the order of \(G^{x}/(G^{x})^{\circ } \cong {G}_{\text {red}}^{x}/({G}_{\text {red}}^{x})^{\circ }\) for any nilpotent element x. As an immediate consequence of Theorem 1.1 and Lemma 2.2 below, we have the following result.

Corollary 1.2

Let G be a connected reductive group over an algebraically closed field \(\Bbbk \) of pretty good characteristic. For any nilpotent element x in its Lie algebra, \((G, {G}_{\text {red}}^{x})\) is a Donkin pair.

This corollary plays a key role in the proof of the Humphreys conjecture [1].

The paper is organized as follows: Section 2 contains some general lemmas on Donkin pairs, along with a lengthly list of examples (some previously known, and some new). Section 3 gives the proof of Theorem 1.1. The proof consists of a reduction to the quasi-simple case, followed by case-by-case arguments.

Remark 1.3

It would, of course, be desirable to have a uniform proof of Theorem 1.1 that avoids case-by-case arguments, perhaps using the method of Frobenius splittings. Thanks to a fundamental result of Mathieu [13], Theorem 1.1 would come down to showing that the flag variety of G admits a \(({G}_{\text {red}}^{x})^{\circ }\)-canonical splitting. According to a result of van der Kallen [16], this geometric condition is equivalent to a certain linear-algebraic condition (called the “pairing condition”) on the Steinberg modules for G and \(({G}_{\text {red}}^{x})^{\circ }\). Unfortunately, for the moment, the pairing condition for these groups seems to be out of reach.

2 Preliminaries

2.1 General Lemmas on Donkin Pairs

We begin with three easy statements about good filtrations.

Lemma 2.1

Let H be a possibly disconnected reductive group over an algebraically closed field \(\Bbbk \). Assume that the characteristic of \(\Bbbk \) does not divide |H/H|. An H-module M has a good filtration if and only if it has a good filtration as an H-module.

Proof

According to [2, Eq. (3.3)], any costandard H-module regarded as an H-module is a direct sum of costandard H-modules. Hence, any H-module with a good filtration has a good filtration as an H-module.

For the opposite implication, suppose M is an H-module that has a good filtration as an H-module. To show that it has a good filtration as an H-module, we must show that \({\text {Ext}}_{H}^{1}({-},M)\) vanishes on standard H-modules. As explained in the proof of [2, Lemma 2.18], we have

$$ {\text{Ext}}_{H}^{1}({-},M) \cong ({\text{Ext}}_{H^{\circ}}^{1}({-},M))^{H/H^{\circ}}, $$

and the right-hand side clearly vanishes on standard H-modules (using [2, Eq. (3.3)] again). □

Lemma 2.2

Let G be a connected, reductive group, and let HG be a possibly disconnected reductive subgroup. Assume that the characteristic of \(\Bbbk \) does not divide |H/H|. Then, (G,H) is a Donkin pair if and only if (G,H) is a Donkin pair.

Proof

This is an immediate consequence of Lemma 2.1. □

Lemma 2.3

Let G be a connected, reductive group, and let \(G^{\prime }\) be its derived subgroup. Let HG be a connected, reductive subgroup. Then, (G,H) is a Donkin pair if and only if \((G^{\prime },(G^{\prime } \cap H)^{\circ })\) is a Donkin pair.

Proof

Let TBG denote a maximal torus and Borel subgroup, respectively, and suppose that \(G^{\prime \prime }\) is any closed connected subgroup satisfying \(G^{\prime } \subseteq G^{\prime \prime } \subseteq G\). Now let \(T^{\prime \prime } = G^{\prime \prime }\cap T\), \(B^{\prime \prime } = G^{\prime \prime } \cap B\), and observe that by [10, I.6.14(1)], we have

$$ {\text{Res}}_{G^{\prime\prime}}^{G} {\text{Ind}}_{B}^{G} M \cong {\text{Ind}}_{B^{\prime\prime}}^{G^{\prime\prime}} {\text{Res}}_{B^{\prime\prime}}^{B} M $$
(2.1)

for any B-module M. Thus, for any dominant weight λX(T)+ where we set \(\lambda ^{\prime \prime } = {\text {Res}}_{T^{\prime \prime }}^{T} (\lambda ) \in \mathbf {X}(T^{\prime \prime })^{+}\), it follows that \({\text {Res}}_{G^{\prime \prime }}^{G} {\text {Ind}}_{B}^{G}(\lambda ) \cong {\text {Ind}}_{B^{\prime \prime }}^{G^{\prime \prime }} (\lambda ^{\prime \prime })\). Thus, \((G,G^{\prime \prime })\) is always a Donkin pair. Furthermore, if we let \(H^{\prime } \subseteq H\) be the derived subgroup and \(H^{\prime \prime } = (G^{\prime }\cap H)^{\circ }\), then \(H^{\prime } \subseteq H^{\prime \prime } \subseteq H\). We can therefore apply [10, I.6.14(1)] again to show that \((H,H^{\prime \prime })\) is a Donkin pair.

Now suppose (G,H) is a Donkin pair. In this case, it immediately follows from above that \((G, H^{\prime \prime })\) is a Donkin pair. Moreover, if we let \(T^{\prime } = G^{\prime }\cap T\) and \(B^{\prime } = G^{\prime }\cap B\), then Eq. 2.1 actually implies that for any \(\lambda ^{\prime } \in \mathbf {X}(T^{\prime })^{+}\), there exists λX(T)+ with \(\lambda ^{\prime } = \text {Res}^{T}_{T^{\prime }}(\lambda )\) such that

$$ {\text{Res}}_{G^{\prime}}^{G} {\text{Ind}}_{B}^{G} (\lambda) \cong {\text{Ind}}_{B^{\prime}}^{G^{\prime}} (\lambda^{\prime}). $$

In particular,

$$ {\text{Res}}_{H^{\prime\prime}}^{G^{\prime}} {\text{Ind}}_{B^{\prime}}^{G^{\prime}} (\lambda^{\prime}) \cong {\text{Res}}_{H^{\prime\prime}}^{G} {\text{Ind}}_{B}^{G}(\lambda) $$

has a good filtration as an \(H^{\prime \prime }\)-module, and hence, \((G^{\prime }, H^{\prime \prime })\) is also a Donkin pair.

Conversely, suppose that \((G^{\prime }, H^{\prime \prime })\) is a Donkin pair. We can first deduce that \((G, H^{\prime \prime })\) is a Donkin pair from the fact that \((G,G^{\prime })\) is a Donkin pair. Also, by similar arguments as above we can see that for any \(\mu ^{\prime \prime } \in \mathbf {X}(H^{\prime \prime }\cap T)^{+}\), there exists some μX(HT)+ with \(\mu ^{\prime \prime } = {\text {Res}}_{H^{\prime \prime }\cap T}^{H\cap T}(\mu )\), such that

$$ {\text{Res}}_{H^{\prime\prime}}^{H} {\text{Ind}}_{H\cap B}^{H} (\mu) \cong {\text{Ind}}_{H^{\prime\prime}\cap B}^{H^{\prime\prime}}(\mu^{\prime\prime}). $$

This implies that an H-module M has a good filtration if and only if the \(H^{\prime \prime }\)-module \({\text {Res}}_{H^{\prime \prime }}^{H} M\) has a good filtration. Therefore, (G,H) is also a Donkin pair. □

2.2 Examples of Donkin Pairs

The following proposition collects a number of known examples of Donkin pairs. The last five parts of the proposition deal with various examples where G is quasi-simple and simply connected. For pairs of the form (Spinn,H), it is usually more convenient to describe the image \(H^{\prime }\) of H under the map π : Spinn →SOn. Of course, H can be recovered from \(H^{\prime }\), as the identity component of π− 1(H). We use the notation that

$$ (\text{SO}_{n},H^{\prime})^{\sim} = (\text{Spin}_{n}, H). $$

It should be noted that the following proposition does not exhaust the known examples in the literature: for instance, according to [5], there is a Donkin pair of type (B3,G2), but this example is not needed in the present paper.

Proposition 2.4

Let G be a connected, reductive group, and let HG be a closed, connected, reductive subgroup. If the pair (G,H) satisfies one of the following conditions, then it is a Donkin pair.

  1. (1)

    G = H ×⋯ × H, and HG is the diagonal embedding.

  2. (2)

    H is a Levi subgroup of G.

For the remaining parts, assume that G is quasi-simple and simply connected.

  1. (3)

    G is of simply laced type, and H is the fixed-point set of a diagram automorphism of G:

    $$ \begin{array}{lc} (\mathrm{A}_{2n-1}, \mathrm{C}_{n}) = (\text{SL}_{2n},\text{Sp}_{2n}) & \qquad (\mathrm{D}_{4},\mathrm{G}_{2}) = (\text{Spin}_{8}, \mathrm{G}_{2}) \\ ~~(\mathrm{D}_{n}, \mathrm{B}_{n-1}) = (\text{SO}_{2n}, \text{SO}_{2n-1})^{\sim} & \qquad (\mathrm{E}_{6}, \mathrm{F}_{4}) \end{array} $$
  2. (4)

    Certain embeddings of classical groups:

    $$ \begin{array}{@{}rcl@{}} \left.\begin{array}{c} (\mathrm{A}_{2n}, \mathrm{B}_{n}) \\ (\mathrm{A}_{2n-1}, \mathrm{D}_{n}) \end{array}\right\} &=& (\text{SL}_{r}, \text{SO}_{r}) \qquad (p > 2) \\ (\mathrm{A}_{2n-1}, \mathrm{C}_{n}) &=& (\text{SL}_{2n}, \text{Sp}_{2n}) \\ \left.\begin{array}{c} (\mathrm{B}_{n+m}, \mathrm{B}_{n}\mathrm{D}_{m}) \\ (\mathrm{D}_{n+m}, \mathrm{D}_{n}\mathrm{D}_{m}) \\ (\mathrm{D}_{n+m+1}, \mathrm{B}_{n}\mathrm{B}_{m}) \end{array}\right\} &=& (\text{SO}_{r+s}, \text{SO}_{r} \times \text{SO}_{s})^{\sim} \quad(p > 2) \\ (\mathrm{C}_{n+m}, \mathrm{C}_{n}\mathrm{C}_{m}) &=& (\text{Sp}_{2n+2m}, \text{Sp}_{2n} \times \text{Sp}_{2m}) \end{array} $$
  3. (5)

    Certain maximal-rank subgroups of exceptional groups:

    $$ \begin{array}{ll} & \qquad\qquad (\mathrm{E}_{8} , \mathrm{A}_{2}\mathrm{E}_{6})\quad (p > 5) \\ (\mathrm{E}_{8} , \mathrm{D}_{8}) \quad (p > 2) & \qquad\qquad (\mathrm{E}_{8} , \mathrm{A}_{1}\mathrm{A}_{2}\mathrm{A}_{5})\quad (p > 5) \\ (\mathrm{E}_{8} , \mathrm{A}_{1}\mathrm{E}_{7}) \quad (p > 2 ) & \qquad\qquad (\mathrm{E}_{8} , \mathrm{A}_{3}\mathrm{D}_{5})\quad (p > 5) \\ (\mathrm{E}_{7} , \mathrm{A}_{1}\mathrm{D}_{6}) \quad (p > 2) & \qquad\qquad (\mathrm{E}_{8} , \mathrm{A}_{4}\mathrm{A}_{4})\quad (p > 5) \\ (\mathrm{F}_{4} , \mathrm{B}_{4}) \quad (p > 2) & \qquad\qquad (\mathrm{F}_{4} , \mathrm{A}_{3}\mathrm{A}_{1})\quad (p > 3) \\ & \qquad\qquad (\mathrm{G}_{2} , \mathrm{A}_{1}\mathrm{A}_{1}) \end{array} $$
  4. (6)

    Certain restricted irreducible representations:

    $$ \begin{array}{@{}rcl@{}} (\mathrm{A}_{n} , \mathrm{A}_{1}) \quad (p > n) \\ (\mathrm{A}_{7} , \mathrm{A}_{2}) \quad (p > 3) \\ (\mathrm{A}_{6} , \mathrm{G}_{2}) \quad (p > 3) \end{array} $$
  5. (7)

    Tensor product embeddings of classical groups (p > 2):

    $$ \begin{array}{@{}rcl@{}} \left.\begin{array}{c} (\mathrm{C}_{(2n+1)m}, \mathrm{B}_{n})\\ (\mathrm{C}_{2nm} , \mathrm{D}_{n}) \end{array}\right\} &=& (\text{Sp}_{2rm}, \text{SO}_{r}) \qquad \left.\begin{array}{c} (\mathrm{B}_{n+m+2nm}, \mathrm{B}_{n}) \\ (\mathrm{D}_{(2n+1)m}, \mathrm{B}_{n})\\ (\mathrm{D}_{nm} , \mathrm{D}_{n}) \end{array}\right\} = (\text{SO}_{rs}, \text{SO}_{r})^{\sim} \\ (\mathrm{D}_{2nm}, \mathrm{C}_{n}) &=& (\text{SO}_{4nm}, \text{Sp}_{2n})^{\sim} \qquad\qquad\quad (\mathrm{C}_{nm}, \mathrm{C}_{n}) = (\text{Sp}_{2nm}, \text{Sp}_{2n}) \end{array} $$

The details of the embeddings in parts (6) and (7) will be described below.

Proof Proofs for parts (2)–(5)

Parts (1) and (2) are due to Mathieu [13] (following earlier work of Donkin [6] that covered most cases). Parts (3) and (4), with the exception of the pair (E6,F4), are due to Brundan [5]. The pair (G2,A1A1) in part (5) is also due to Brundan [5]. The pair (E6,F4) and the pairs in the first column of part (5) are due to van der Kallen [16]. The pairs in the second column of part (5) are due to Hague–McNinch [7]. □

Proof Proof of part (6)

Each pair (An,H) = (SLn+ 1,H) in this statement arises from some (n + 1)-dimensional representation of H. Call that representation V. The representations V are as follows:

  • (An,A1): the dual Weyl module for SL2 of highest weight n

  • (A7,A2): the adjoint representation of PGL3

  • (A6,G2): the 7-dimensional dual Weyl module whose highest weight is the short dominant root

According to [5, Lemma 3.2(iv)] or [7, §3.2.6], to prove the claim, we must show that each exterior algebra \(\bigwedge ^{\bullet } V\) has a good filtration as an H-module. For (An,A1), this is shown in [7, §3.4.3]. For (A7,A2) and (A6,G2), explicit calculations using the LiE software package [17] show that the character of \(\bigwedge ^{\bullet } V\) is the sum of characters of dual Weyl modules whose highest weights are restricted weights when p > 3. □

Proof Proof of part (7)

To define the group embeddings in this statement, we will assume that G is either Sp2n or SOn. However, in the latter case, the proof that (G,H) is a Donkin pair will also imply the corresponding statement for G = Spinn.

Let V1 be a vector space equipped with a nondegenerate bilinear form B1 satisfying B1(v,w) = ε1B1(w,v), where ε1 = ± 1, and let Aut(V1,B1) be the connected group of linear automorphisms of V1 that preserve B1. This group is either \(\text {SO}_{\dim V}\) or \(\text {Sp}_{\dim V}\), depending on ε1. Let V2, B2, and ε2 be another collection of similar data. Then, B1B2 is a nondegenerate pairing on V1V2, with sign ε1ε2. We obtain an embedding

$$ \text{Aut}(V_{1},B_{1})^{\circ} \times \text{Aut}(V_{2},B_{2})^{\circ} \hookrightarrow \text{Aut}(V_{1} \otimes V_{2}, B_{1} \otimes B_{2})^{\circ}. $$

Now restrict to just one factor:

$$ \text{Aut}(V_{1},B_{1})^{\circ} \hookrightarrow \text{Aut}(V_{1} \otimes V_{2}, B_{1} \otimes B_{2})^{\circ}. $$
(2.2)

The four kinds of pairs listed in the statement are all instances of this embedding, depending on the signs ε1 and ε2. We will now prove that

$$ (G,H) = (\text{Aut}(V_{1} \otimes V_{2}, B_{1} \otimes B_{2})^{\circ}, \text{Aut}(V_{1},B_{1})^{\circ}) $$

is a Donkin pair. Let \(r = \dim V_{1}\) and \(s = \dim V_{2}\).

Suppose first that ε2 = 1. Then, the embedding (2.2) corresponds to either (SOrs,SOr) or (Sprs,Spr). In this case, V2 admits an orthonormal basis x1,…,xs, where

$$ B_{2}(x_{i},x_{j}) = \delta_{ij}. $$

Then, the group Aut(V1,B1) preserves each V1xiV1V2. In this case, the embedding (2.2) can be factored as

$$ \text{Aut}(V_{1},B_{1})^{\circ} \overset{1}{\hookrightarrow} \underbrace{\text{Aut}(V_{1},B_{1})^{\circ} \times {\cdots} \times \text{Aut}(V_{1},B_{1})^{\circ}}_{s\text{ copies}} \overset{4}{\hookrightarrow} \text{Aut}(V_{1} \otimes V_{2}, B_{1} \otimes B_{2})^{\circ}. $$

The first map is a diagonal embedding; it results in a Donkin pair by part (1) of the proposition. The second embedding gives a Donkin pair by part (4).

Next, suppose that ε2 = − 1, and assume for now that \(\dim V_{2} = 2\). Choose a basis {x,y} for V2 such that B2(x,y) = 1. Then, V1x and V1y are both maximal isotropic subspaces of V1V2. Define an action of GL(V1) on V1V2 as follows:

$$ \begin{array}{l} g \cdot (v \otimes x) = (gv) \otimes x, \\ g \cdot (v \otimes y) = ((g^{\mathrm{t}})^{-1}v) \otimes y \end{array} \qquad \text{for } g \in \text{GL}(V_{1}), $$

where gt denotes the adjoint operator to g with respect to the nondegenerate form on V1. This action defines an embedding of GL(V1) in Aut(V1V2,B1B2). In fact, it identifies GL(V1) with a Levi subgroup of Aut(V1V2,B1B2). (This is the usual embedding of GLr as a Levi subgroup in either SO2r or Sp2r.) The embedding (2.2) then factors as

$$ \text{Aut}(V_{1},B_{1})^{\circ} \overset{4}{\hookrightarrow} \text{GL}(V_{1}) \overset{2}{\hookrightarrow} \text{Aut}(V_{1} \otimes V_{2}, B_{1} \otimes B_{2})^{\circ}. $$

The first embedding gives a Donkin pair by part (4) of the proposition, and the second by part (2).

Finally, suppose ε2 = − 1 and \(s = \dim V_{2} > 2\). This dimension must still be even, say s = 2m. Choose a basis x1,…,xm,y1,…,ym for V2 such that

$$ B_{2}(x_{i},x_{j}) = B_{2}(y_{i},y_{j}) = 0, \qquad B_{2}(x_{i},y_{j}) = \delta_{ij}. $$

Let \({V}_{2}^{(i)}\) be the 2-dimensional subspace spanned by xi and yi. Then, B2 restricts to a nondegenerate symplectic form \(B_{2}^{(i)}\) on \({V}_{2}^{(i)}\). We factor the map (2.2) as follows:

$$ \begin{array}{@{}rcl@{}} \text{Aut}(V_{1},B_{1})^{\circ} &\overset{1}{\hookrightarrow}& \underbrace{\text{Aut}(V_{1},B_{1})^{\circ} \times {\cdots} \times \text{Aut}(V_{1},B_{1})^{\circ}}_{m\text{ copies}} \\ &\hookrightarrow& \text{Aut}(V_{1} \otimes {V}_{2}^{(1)}, B_{1} \otimes {B}_{2}^{(1)})^{\circ} \!\times\! {\cdots} \!\times\! \text{Aut}(V_{1} \otimes {V}_{2}^{(m)}, B_{1} \otimes {B}_{2}^{(m)})^{\circ} \\ &&\!\!\!\!\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad \overset{4}{\hookrightarrow} \text{Aut}(V_{1} \otimes V_{2}, B_{1} \otimes B_{2})^{\circ}. \end{array} $$

Here, the first arrow is a diagonal embedding (part (1) of the proposition); the second arrow is several instances of the embedding from the previous paragraph (since \(\dim {V}_{2}^{(i)} = 2\)); and the last arrow comes from part (4) of the proposition. We thus again obtain a Donkin pair. □

3 Proof of Theorem 1.1

3.1 Reduction to the Quasi-Simple Case

Let G be an arbitrary connected reductive group in good characteristic. For any nilpotent element x ∈Lie(G), there exists a cocharacter \(\tau : \mathbb {G}_{m} \rightarrow G\) and a Levi subgroup LτG such that Lτ is the centralizer of the subgroup \(\tau (\mathbb {G}_{m})\), where \( {G}_{\text {red}}^{x} = L_{\tau } \cap G^{x}\) (cf. [11, 5.10]).

If we let \(G^{\prime }\) be the derived subgroup of G, then any nilpotent element x for G also satisfies \(x \in \text {Lie}(G^{\prime })\), and by [11, 5.9], \(\tau (\mathbb {G}_{m}) \subset G^{\prime } \subseteq G\). In particular, \((G^{\prime })^{x} = G^{\prime }\cap G^{x}\) and \({L}_{\tau }^{\prime } = G^{\prime }\cap L_{\tau }\) is the centralizer of \(\tau (\mathbb {G}_{m})\) in \(G^{\prime }\). Thus,

$$ (G^{\prime})^{x}_{\text{red}} = G^{\prime} \cap {G}_{\text{red}}^{x}. $$

It now follows from Lemma 2.3 that \((G,({G}_{\text {red}}^{x})^{\circ })\) is a Donkin pair if and only if \((G^{\prime }, ((G^{\prime })^{x}_{\text {red}})^{\circ })\) is a Donkin pair. So we can reduce to the case where G is semisimple.

Suppose now that \(\pi : G \twoheadrightarrow \bar {G}\) is an isogeny (i.e., surjective with finite central kernel), where G is an arbitrary connected reductive group in good characteristic. Then, by [11, Proposition 2.7(a)], π induces a bijection between the nilpotent elements in Lie(G) and those in \(\text {Lie}(\bar {G})\), and for any nilpotent element x ∈Lie(G), we have \(\pi (G^{x}) = \bar {G}^{\pi (x)}\). Moreover, by similar arguments as above, we can also deduce that \(\pi ({G}_{\text {red}}^{x}) = \bar {G}^{\pi (x)}_{\text {red}}\) (cf. [11, 5.9]). In particular,

$$ \pi(({G}_{\text{red}}^{x})^{\circ}) = \pi({G}_{\text{red}}^{x})^{\circ} = ({\bar{G}}_{\text{red}}^{\pi(x)})^{\circ}, $$

since any surjective morphism of algebraic groups takes the identity component to the identity component.

Let \(H = ({G}_{\text {red}}^{x})^{\circ }\) and \(\bar {H} = (\bar {G}^{\pi (x)}_{\text {red}})^{\circ }\), and note that for any \(\bar {G}\)-module M, there is a natural isomorphism

$$ {\text{Res}}_{H}^{G} {\text{Res}}_{G}^{\bar{G}} M \cong {\text{Res}}_{H}^{\bar{H}} {\text{Res}}_{\bar{H}}^{\bar{G}} M. $$

From this, we can see that if (G,H) is a Donkin pair, then \((\bar {G},\bar {H})\) must also be a Donkin pair, since it is straightforward to check that a \(\bar {G}\)-module M (resp. an \(\bar {H}\)-module N) has a good filtration if and only if \({\text {Res}}_{G}^{\bar {G}} M\) (resp. \({\text {Res}}_{H}^{\bar {H}} N\)) has a good filtration. This allows us to reduce to the case where G is semisimple and simply connected.

Finally, suppose that that G = G1 × G2 where G1, G2 are connected reductive groups in good characteristic. Let x = (x1,x2) ∈Lie(G1) ⊕Lie(G2) be an arbitrary nilpotent element. We can immediately see that

$$ ({G}_{\text{red}}^{x})^{\circ} = ((G_{1})^{x_{1}}_{\text{red}})^{\circ} \times ((G_{2})^{x_{2}}_{\text{red}})^{\circ}. $$

It now follows from the general properties of induction for direct products (see [10, I.3.8]) that \((G,({G}_{\text {red}}^{x})^{\circ })\) is a Donkin pair if and only if \((G_{1},((G_{1})^{x_{1}}_{\text {red}})^{\circ })\) and \((G_{2},((G_{2})^{x_{2}}_{\text {red}})^{\circ })\) are Donkin pairs. Therefore, by the well-known fact that any simply connected semisimple group is a direct product of quasi-simple simply connected groups, we can reduce the proof of Theorem 1.1 to the case where G is quasi-simple.

3.2 Proof for Classical Groups

We now prove the theorem for the groups GLn, Spn, and Spinn. For the last case, we will actually describe the group \(({G}_{\text {red}}^{x})^{\circ }\) and its embedding in G for SOn instead, but the proof of the Donkin pair property will also apply to Spinn.

Let x be a nilpotent element in the Lie algebra of one of GLn, Spn, or SOn. Let \(\mathbf {s} = [{s}_{1}^{r_{1}}, {s}_{2}^{r_{2}}, \ldots , {s}_{k}^{r_{k}}]\) be the partition of n that records the sizes of the Jordan blocks of x. (This means that x has r1 Jordan blocks of size s1, and r2 Jordan blocks of size s2, etc.) The vector space \(V = \Bbbk ^{n}\) can be decomposed as

$$ V = V^{(1)} \oplus V^{(2)} \oplus {\cdots} \oplus V^{(k)} $$

where each V(i) is preserved by x, and x acts on V(i) by Jordan blocks of size si. (Thus, \(\dim V^{(i)} = r_{i}s_{i}\).) When G is Spn or SOn, the nondegenerate bilinear form on V restricts to a nondegenerate form of the same type on each V(i).

The description of \(({G}_{\text {red}}^{x})^{\circ }\) in [12, Chapter 3] shows that it factors through the appropriate embedding below:

$$ \begin{array}{@{}rcl@{}} \text{GL}(V^{(1)}) \times {\cdots} \times \text{GL}(V^{(k)}) &\hookrightarrow& \text{GL}(V) \\ \text{Sp}(V^{(1)}) \times {\cdots} \times \text{Sp}(V^{(k)}) &\hookrightarrow& \text{Sp}(V) \\ \text{SO}(V^{(1)}) \times {\cdots} \times \text{SO}(V^{(k)}) &\hookrightarrow& \text{SO}(V) \end{array} $$

All three of these embeddings give Donkin pairs: in the case of GLn, it is an inclusion of a Levi subgroup (Proposition 2.4(2)); and in the case of Spn or SOn, it falls under Proposition 2.4(4).

We can therefore reduce to the case where x has Jordan blocks of a single size. Suppose from now on that s = [sr]. Then, there exists a vector space isomorphism

$$ V \cong V_{1} \otimes V_{2} $$

where \(\dim V_{1} = r\) and \(\dim V_{2} = s\), and such that x corresponds to \(\text {id}_{V_{1}} \otimes N\), where N : V2V2 is a nilpotent operator with a single Jordan block (of size s).

Suppose now that G = GL(V ). Then, according to [12, Proposition 3.8], we have \(({G}_{\text {red}}^{x})^{\circ } \cong \text {GL}(V_{1})\). Choose a basis {v1,…,vs} for V2. The embedding of \(({G}_{\text {red}}^{x})^{\circ }\) in G factors as

$$ \text{GL}(V_{1}) \hookrightarrow \text{GL}(V_{1} \otimes v_{1}) \times {\cdots} \times \text{GL}(V_{1} \otimes v_{s}) \hookrightarrow \text{GL}(V). $$

The first map above is a diagonal embedding (Proposition 2.4(1)), and the second is the inclusion of a Levi subgroup (Proposition 2.4(2)), so \((G, ({G}_{\text {red}}^{x})^{\circ })\) is a Donkin pair in this case.

Next, suppose G = Sp(V ) or SO(V ). According to [12, Proposition 3.10], both V1 and V2 can be equipped with nondegenerate bilinear forms B1 and B2 such that B1B2 agrees with the given bilinear form on V. Moreover, \(({G}_{\text {red}}^{x})^{\circ } = \text {Aut}(V_{1},B_{1})^{\circ }\). We are thus in the setting of Proposition 2.4(7).

3.3 Proof for E8

When x is distinguished, \(({G}_{\text {red}}^{x})^{\circ }\) is the trivial group; and when x = 0, \(({G}_{\text {red}}^{x})^{\circ } = G\). For all remaining nilpotent orbits, we rely on the very detailed case-by-case descriptions of \(({G}_{\text {red}}^{x})^{\circ }\) given in [12, Chapter 15]. In each case, that description shows that the embedding \(({G}_{\text {red}}^{x})^{\circ } \hookrightarrow G\) factors as a composition of various cases from Proposition 2.4.

These factorizations are shown in Tables 1 and 2. Here is a brief explanation of the notation used in these tables. Nearly all groups mentioned are semisimple, and they are recorded in the tables by their root systems. However, the notation “T1” indicates a 1-dimensional torus; this is used to indicate a reductive group with a 1-dimensional center. In a few cases, nonstandard names for root systems—such as B1 or C1, in place of A1—are used when it is convenient to emphasize the role of a certain classical group. The notation D1 (meant to evoke SO2) is occasionally used as a synonym for T1.

Table 1 Nilpotent centralizers in E8
Table 2 Nilpotent centralizers in E8, continued

Finally, we remark that there are two orbits—labeled by A6 and by A6A1—where the information given in [12] is insufficient to finish the argument. In each of these cases, \(({G}_{\text {red}}^{x})^{\circ }\) contains a copy of A1 that is the centralizer of a certain copy of G2 inside E7, and [12] does not give further details on this embedding A1↪E7. However, according to [15, §3.12], this A1 is in fact (the derived subgroup of) a Levi subgroup of E7.

3.4 Proof for E7 and E6

Recall that if \(H \subseteq G\) is a closed subgroup, then there is a natural embedding \(\mathcal {N}_{H} \hookrightarrow \mathcal {N}_{G}\) of nilpotent cones. We also recall that a subgroup \(L \subseteq G\) is a Levi subgroup if and only if it is the centralizer of a torus SG.

Lemma 3.1

Let G be a reductive group, \(S \subseteq G\) a torus with L = CG(S) a Levi subgroup, and suppose \(x \in \mathcal {N}_{L} \subseteq \mathcal {N}_{G}\) is such that \(S \subseteq ({G}_{\text {red}}^{x})^{\circ }\). Then, \(({L}_{\text {red}}^{x})^{\circ }\) is a Levi subgroup of \(({G}_{\text {red}}^{x})^{\circ }\).

Proof

We clearly have \(C_{G^{x}}(S) = G^{x} \cap L = L^{x}\). Moreover, the identity component (Lx) must be contained in \((G^{x})^{\circ } \cap L = C_{(G^{x})^{\circ }}(S)\). But since the centralizer of a torus in a connected group is connected, we actually have

$$ (L^{x})^{\circ} = C_{(G^{x})^{\circ}}(S). $$

Next, consider the semidirect product decomposition \((G^{x})^{\circ } = ({G}_{\text {red}}^{x})^{\circ } \ltimes {G}_{\text {unip}}^{x}\). Let gGx, and write it as g = (gr,gu), with \(g_{r} \in ({G}_{\text {red}}^{x})^{\circ }\), \(g_{u} \in G^{x}_{\text {unip}}\). If g centralizes S, then gr and gu must individually centralize S as well. In other words,

$$ C_{(G^{x})^{\circ}}(S) = C_{({G}_{\text{red}}^{x})^{\circ}}(S) \ltimes C_{{G}_{\text{unip}}^{x}}(S). $$
(3.1)

Here, \(C_{({G}_{\text {red}}^{x})^{\circ }}(S)\) is a connected reductive group, and \(C_{{G}_{\text {unip}}^{x}}(S)\) is a normal unipotent group (which must be connected, because \(C_{(G^{x})^{\circ }}(S)\) is connected). We conclude that Eq. 3.1 is a Levi decomposition of (Lx). In particular, we see that \(({L}_{\text {red}}^{x})^{\circ } = C_{({G}_{\text {red}}^{x})^{\circ }}(S)\). □

We now let G denote the simple, simply connected group of type E8. If G0 is the simple, simply connected group of type E7 or E6, then as explained in [12, Lemma 11.14], there is a simple subgroup H of type A1 or A2, respectively, such that G0 = CG(H). Moreover, by [12, 16.1.2], there exists a torus SHG such that G0 is the derived subgroup of the Levi subgroup L = CG(S). Explicitly, let α1,…,α8 be the simple roots for G, labelled as in [4], and let α0 be the highest root. The groups H and G0 can be described as follows.

(3.2)

Now it is explained in [12, 16.1.1], that if \(x \in \mathcal {N}_{G_{0}} = \mathcal {N}_{L} \subset \mathcal {N}_{G}\), then the subgroup \(({G}_{\text {red}}^{x})^{\circ }\) must contain a conjugate of H. Without loss of generality we can assume that x is chosen so that \(H \subseteq ({G}_{\text {red}}^{x})^{\circ }\). Hence, we can also assume that \(S \subseteq ({G}_{\text {red}}^{x})^{\circ }\). Thus, by Lemma 3.1, \(({L}_{\text {red}}^{x})^{\circ }\) is a Levi subgroup of \(({G}_{\text {red}}^{x})^{\circ }\) and we also have

$$ ({L}_{\text{red}}^{x})^{\circ \prime} \subseteq (({G}_{0}^{x})_{\text{red}})^{\circ} \subseteq ({L}_{\text{red}}^{x})^{\circ}. $$

By Lemma 2.3, Proposition 2.4(2) and Section 3.3 we deduce that \((G,({G}_{0}^{x})_{\text {red}}))^{\circ })\) is a Donkin pair.

Finally, to show that \((G_{0}, (({G}_{0}^{x})_{\text {red}})^{\circ })\) is a Donkin pair, it will be sufficient to show that every fundamental tilting module for G0 is a summand of the restriction of a tilting module for G. In more detail, let \(\pi : \mathbf {X}_{G} \twoheadrightarrow \mathbf {X}_{G_{0}}\) be the map on weight lattices. It is well known that if λ is a dominant weight for G, then the G0-tilting module \(T_{G_{0}}(\pi (\lambda ))\) occurs as a direct summand of \(\text {Res}^{G}_{G_{0}}(T_{G}(\lambda ))\). So it is enough to show that every fundamental weight for G0 occurs as π(λ) for some dominant G-weight λ. Let ϖ1,…,ϖ8 be the fundamental weights for G. A short calculation with the table (3.2) shows that π(ϖ1),…,π(ϖ7) are precisely the fundamental weights for E7, and that π(ϖ1),…,π(ϖ6) are the fundamental weights for E6.

Remark 3.2

One can also prove the theorem for E7 and E6 directly by writing down the embedding of each centralizer, as we did for E8. Here is a brief summary of how to carry out this approach. Let G be of type E8, and let G0 and H be as in the discussion above. As explained in [12, §16.1], we have

$$ (({G}_{0}^{x})_{\text{red}})^{\circ} = C_{({G}_{\text{red}}^{x})^{\circ}}(H). $$

The computation of \(C_{({G}_{\text {red}}^{x})^{\circ }}(H)\) is explained in [12, §16.1.4], and the results are recorded in Tables 34, and 5, following the same notational conventions as in the E8 case.

Table 3 Nilpotent centralizers in E7
Table 4 Nilpotent centralizers in E7, continued
Table 5 Nilpotent centralizers in E6

3.5 Proof for F4

Let G be the simple, simply connected group of type E8. Then, [12, Lemma 11.7] implies that G contains a simple subgroup H of type G2, and that its centralizer G0 = CG(H) is a simple group of type F4. The embeddings of centralizers of nilpotent elements for G0 = F4 can then be computed using the method explained in Remark 3.2. One caveat is that the name (i.e., the Bala–Carter label) of a nilpotent orbit usually changes when passing from F4 to E8. The correspondence between these names is given in [12, Proposition 16.10].

We remark that in some cases, the book [12] does not quite give enough details about embeddings of subgroups to establish our result, but in these cases, the relevant details can be found in [15, §3.16]. Here is an example illustrating this. The F4-orbit labelled \(\tilde {\mathrm {A}}_{1}\) corresponds (by [12, Proposition 16.10]) to the E8-orbit labelled \({\mathrm {A}}_{1}^{2}\). Let x be an element of this orbit. We have see that in E8, \(({G}_{\text {red}}^{x})^{\circ } = \mathrm {B}_{6}\), which embeds in the Levi subgroup D7 ⊂E8. The group \(({G}_{\text {red}}^{x})^{\circ }\) has a subgroup of type D3B3, which embeds in D3D4 ⊂D7 ⊂D8. The explicit construction of H = G2 in [15, §3.16] shows that it is contained in the second factor in each of D3B3 ⊂D3D4. It follows that D3 = A3 is contained in \((({G}_{0}^{x})_{\text {red}})^{\circ } = C_{({G}_{\text {red}}^{x})^{\circ }}(H)\), and then a dimension calculation shows that in fact \((({G}_{0}^{x})_{\text {red}})^{\circ } = \mathrm {A}_{3}\).

The results of these calculations are recorded in Table 6.

Table 6 Nilpotent centralizers in F4

3.6 Proof for G2

In this case, there are only two nilpotent orbits that are neither distinguished nor trivial. From the classification, both of these orbits meet the maximal reductive subgroup \(\mathrm {A}_{1}\tilde {\mathrm {A}}_{1} \subset \mathrm {G}_{2}\), and an argument explained in [12, §16.1.4] shows that if x belongs to either of these orbits, then the reductive part of its centralizer in \(\mathrm {A}_{1}\tilde {\mathrm {A}}_{1}\) is equal to the reductive part of its centralizer in G2. See Table 7.

Table 7 Nilpotent centralizers in G2