789
Views
2
CrossRef citations to date
0
Altmetric
Articles

Representatives for unipotent classes and nilpotent orbits

, & ORCID Icon
Pages 1641-1661 | Received 23 May 2021, Accepted 19 Sep 2021, Published online: 12 Nov 2021

Abstract

Let G be a simple algebraic group over an algebraically closed field k of characteristic p. The classification of the conjugacy classes of unipotent elements of G(k) and nilpotent orbits of G on Lie(G) is well-established. One knows there are representatives of every unipotent class as a product of root group elements and every nilpotent orbit as a sum of root elements. We give explicit representatives in terms of a Chevalley basis for the eminent classes. A unipotent (resp. nilpotent) element is said to be eminent if it is not contained in any subsystem subgroup (resp. subalgebra), or a natural generalization if G is of type Dn. From these representatives, it is straightforward to generate representatives for any given class. Along the way we also prove recognition theorems for identifying both the unipotent classes and nilpotent orbits of exceptional algebraic groups.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

Let G be a connected reductive algebraic group over an algebraically closed field k of characteristic p. Fix a maximal torus T of G, and positive system of roots; then it is well known that every unipotent class has a representative which is a product of root group elements and that every nilpotent orbit has a representative which is a sum of root elements. However, it is not easy to extract such representatives from the existing literature and there are a few errors (some perpetrated by the second author). This note gives a definitive list when G is simple. For compactness, we provide the representatives of certain distinguished classes we call eminent classes. An eminent unipotent (resp. nilpotent) element is one which is not contained in any proper subsystem subgroup (resp. subalgebra) of G, or a natural generalization if G has a factor of type Dn—the precise definition is given in the next section. Given representatives of the eminent classes in the simple case, it is then a routine task to find representatives of all classes for G. (See Section 2.4.)

Theorem 1.1.

Let G be a simple k-group over a field k=k¯ of characteristic p with Lie algebra Lie(G).

  1. A nilpotent element eLie(G) is eminent if and only if it is G-conjugate to one of the representatives in (G of classical type) or (G of exceptional type). Moreover, no two representatives in and are G-conjugate.

    Table 1. Eminent nilpotent representatives in classical types.

    Table 2. Eminent nilpotent representatives in exceptional types.

  2. A unipotent element uG is eminent if and only if it is G-conjugate to one of the representatives in (G of classical type) or (G of exceptional type). Moreover, no two representatives in and are G-conjugate.

    Table 3. Eminent unipotent representatives in classical types.

    Table 4. Eminent unipotent representatives in exceptional types.

We note that the only time a regular element is not eminent is when G is of type Dn. Indeed, such elements are contained in generalized subsystem subgroups or subalgebras of type Bn1.

We make a few remarks on the tables. An explanation of the notation used is given in Section 2.1; we observe Bourbaki doctrine when exhibiting representatives. When G=SL(V),SO(V) or Sp(V) is a classical group, we also supply the action of the elements on the natural G-module V, which characterizes them up to conjugacy in G. Lastly, since we use the same representatives as [Citation17, Tables 13.3, 14.1], one can invoke [Citation17, Theorems 19.1, 19.2] to generate representatives for the classes in from those in as follows: a representative of the unipotent class with label X is jJxαj(1), where jJeαj is the representative of the nilpotent element with label X.

It is often useful to have a means of recognizing unipotent and nilpotent elements in exceptional types from easily computable invariants. For unipotent classes, this was accomplished by Lawther [Citation15, Section 3]. Let C=CLie(G)(x) for x a unipotent or nilpotent element. We use the following abbreviations.

JBS: List of Jordan block sizes

DS: Dimensions of the terms in the derived series of C

LS: Dimensions of the terms in the lower central series of C

ALG: Dimension of the Lie algebra of derivations of C

ALG′: Dimension of the derived subalgebra of the Lie algebra of derivations of C

NIL: Dimension of the Nilradical of C

NDS: Dimensions of the Lie normalizer of each term in the derived series of C

Theorem 1.2.

Let G be an exceptional k-group. Then gives a list of data which is sufficient to identify the class of any nilpotent or unipotent element of G.

Table 5. Recognition data for nilpotent orbits and unipotent classes.

Remark 1.3.

(i) One important application of Theorem 1.2 is its use in the proof of Theorem 1.1(i) for G of exceptional type. To find the eminent nilpotent elements in Lie(G) the strategy is to list all maximal subsystem subalgebras up to conjugacy; for each m in that list, construct representatives of all distinguished orbits; and then use Theorem 1.2 to identify their class in G.

(ii) We supply auxiliary tables in Section 5; whenever at least two nilpotent orbits have the same Jordan block structure on Vmin (when it exists) and Lie(G) we list them in a table along with the data required to distinguish them, using the same abbreviations given before Proposition 1.2.

2. Preliminaries

2.1. Notation

Throughout G is a reductive k-group, and most often G is simple. Fix a Borel subgroup BG containing a maximal torus T. This defines a root system Φ with base Δ={α1,,αl} of simple roots generating the positive roots Φ+ as positive integral sums. We use the Bourbaki ordering for a Dynkin diagram a corresponding roots and follow [Citation2] in denoting roots by a string of numbers in the form of a Dynkin diagram corresponding to coefficients of a root expressed in terms of the simple roots. For example, the highest short root of an F4 system is denoted 1232.

The adjoint action of G on Lie(G) defines a Cartan decomposition Lie(G)Lie(T)αΦkeα. Corresponding to the root spaces keα one has the one-parameter subgroups Uα<G, where each UαGa is a root groups. In case G is simple and simply connected, we may insist that the eα form a Chevalley basis; usually, this is only recognized as canonical up to a choice of the sign of a coefficient λ for certain triples of roots (α,β,γ) such that [eα,eβ]=λeγ. These choices determine the structure constants for Lie(G) [Citation3, Proposition 4.2.2]. Through exponentiation one can fix an isomorphism keαUα in which ceαxα(c) for all ck.

The computations done in Section 3 were performed with GAP4 [Citation5] in the simply connected case, using the representatives for the nilpotent orbits given in [Citation21] (see the latest arXiv version for a correction in type F4). In the adjoint case we used Magma [Citation1]. Although Magma uses a different choice of signs in the structure constants compared to GAP4, the same representatives can still be used and this is justified as follows.

All representatives listed in [Citation21] are of the form e=i=1teβi, where β1,,βtΦ+ are Z-independent roots. Thus for any choice of signs c1,,ct{1,1} there exists a semisimple element in G that conjugates e to the nilpotent element i=1tcieβi; see for example [Citation8, Lemma 16.2 C]. In particular, the representatives in [Citation21] do not depend on the signs of structure constants in the Chevalley basis of Lie(G).

2.2. Eminent elements and generalized subsystem subgroups and subalgebras

Recall that a subgroup M of G is a subsystem subgroup if it is semisimple (the definition of which we assume to include connected) and normalized by a maximal torus.Footnote1 Note that the derived subgroup of any Levi subgroup is a subsystem subgroup. Define also a subsystem subalgebra to be a subalgebra mLie(G) with m=Lie(M) for some subsystem subgroup M.

Define a generalized subsystem subgroup as follows. If M is a subsystem subgroup of G containing some simple factors M1,,Mt of type Dr1,,Drt where t0, then there is a semisimple subgroup Mi,l of Mi of type BlBril1 with 0lri1 (see the proof of Lemma 2.2). Then the generalized subsystem subgroups are obtained by replacing some subset of the Mi with the subgroups Mi,l. Accordingly, we define a generalized subsystem subalgebra m to be Lie(M) for a generalized subsystem subgroup M.

Recall that a unipotent or nilpotent element x is distinguished if it is contained in no proper Levi of G. The following definition is more restrictive.

Definition 2.1.

A unipotent element uG is called eminent if it is not contained in any proper generalized subsystem subgroup of G. A nilpotent element eLie(G) is called eminent if it is not contained in any proper generalized subsystem subalgebra of Lie(G).

For use in the proof of Theorem 1.1 we record the conjugacy classes of maximal generalized subsystem subgroups which are not Levi subgroups. The maximal generalized subsystem subalgebras follow immediately.

Lemma 2.2.

Suppose that G is a classical group. Let M be a generalized subsystem subgroup which is maximal amongst proper ones, such that M is not a Levi subgroup. Then M is conjugate to precisely one subgroup with root system as given in . Moreover, every subgroup in is a maximal generalized subsystem subgroup.

Table 6. Maximal generalized subsystem subgroups.

Proof.

For the subsystem subgroups this is a routine use of the Borel–de Siebenthal algorithm for closed subsystems [Citation18, Proposition 13.12] and [Citation18, Proposition 13.15(i),(ii)] for the 2-closed non-closed subsystems. When G has type Dn further consideration is required. The definition of generalized subsystem subgroups immediately implies that they will not be maximal unless we replace Dn itself with BmBnm1 with 0mn1. When p = 2, the subgroups of type BmBnm1 with 1mn1 are contained in Bn1 (indeed they are subsystem subgroups of Bn1 since p = 2) and so they are not maximal. We need to check that these subgroups are unique up to conjugacy and maximal amongst proper generalized subsystem subgroups. When p2, a subgroup of type BmBnm1 is the connected stabilizer of a decomposition of V into non-degenerate orthogonal subspaces of dimensions 2m+1 and 2n2m1 and thus is unique up to conjugacy since G acts transitively on non-degenerate orthogonal subspaces of dimension k. Moreover, this implies they are maximal amongst all connected proper subgroups of G and thus amongst all proper generalized subsystem subgroups. When p = 2, a subgroup of type Bn1 is the stabilizer of a nonsingular vector and is again unique up to conjugacy and maximal amongst all connected proper subgroups of G. □

We observe that eminence is invariant under central isogenies.

Lemma 2.3.

Let ϕ:GH be a central isogeny of reductive k-groups.

  1. The map ϕ induces bijections between the unipotent varieties of G and H, and the set of generalized subsystem subgroups of G and H.

  2. The map dϕ induces a bijection between the nilpotent cones of G and H, and the set of generalized subsystem subalgebras of G and H.

Proof.

The bijections of unipotent varieties and nilpotent cones follow from [Citation4, Proposition 5.1.1] and [Citation11, 2.7]. Since the kernel of ϕ is central, it is contained in any maximal torus; as ϕ is surjective with finite kernel it follows that ϕ is a bijection on maximal tori. Thus M is a subsystem subgroup if and only if ϕ(M) is. If M is a generalized subsystem subgroup, the result follows from the definition. □

2.3. Unipotent classes and nilpotent orbits in classical groups

Let G be simple of classical type. In light of Lemma 2.3, for the purposes of this paper it does no harm for us to assume that G=SL(V),Sp(V), or SO(V).

We will choose forms preserved by G on V consistent with [Citation9]. Let V have k-basis {v1,v2,,vm}. For G=Sp(V) and m=2n, we choose a non-degenerate alternating bilinear form on V defined by (vi,vn+i)=1=(vn+i,vi) and (vi,vj)=0 for all other i, j. For G=SO(V) with V of odd dimension m=2n+1, choose a non-degenerate symmetric bilinear form on V defined by (vi,vn+i)=1=(vn+i,vi),(v2n+1,v2n+1)=2 and (vi,vj)=0 for all other i, j. When p = 2, we also define a quadratic form on V by Q(vi)=0 for i2n,Q(v2n+1)=1, such that Q has polarization (,) on V. Finally, for SO(V) with V of even dimension m=2n choose a non-degenerate symmetric bilinear form on V defined by (vi,vn+i)=1=(vn+i,vi) and (vi,vj)=0 for all other i, j. And in this case for p = 2, define a quadratic form on V by Q(vi)=0 for all i, with polarization (,) on V.

If G=SL(V), the class of a unipotent or nilpotent element xG is determined by its Jordan block sizes on V. In what follows, we will describe the unipotent classes and nilpotent orbits for G=Sp(V) and G=SO(V). In this case for x a unipotent element or a nilpotent element, we have a decomposition Vk[x]=V1Vt, where the Vi are orthogonally indecomposable k[x]-modules. Here orthogonally indecomposable means that if Vi=WW as k[x]-modules, then W = 0 or W=0. Hesselink [Citation6, 3.5] determined the orthogonally indecomposable k[x]-modules that can occur, and thus classified the unipotent classes and nilpotent orbits for Sp(V) and O(V). The possibilities for orthogonally indecomposable k[x]-modules are as follows:

  • The k[x]-module V(m) is non-degenerate of dimension m and x acts on V(m) with a single Jordan block of size m.

  • The k[x]-module W(m) has dimension 2m, with W(m)=W1W2 where Wi are totally singular of dimension m and x acts on each Wi with a single Jordan block of size m.

  • The k[x]-module Wl(m) only occurs when p = 2 and x is a nilpotent element. It is non-degenerate of dimension 2m and x acts with two Jordan blocks of size m. If G=Sp(V) then 0<l<m/2 and l=min{n:(xn+1(v),xn(v))=0 for all vWl(m)}. If G=SO(V) then (m+1)/2<lm and l=min{n:Q(xn(v))=0 for all vWl(m)}.

  • The k[x]-module D(m) only occurs when G=SO(V) with dim(V) odd, p = 2 and x is a nilpotent element. It is degenerate of dimension 2m1 and x acts with two Jordan blocks of sizes m and m – 1.

  • The k[x]-module R only occurs when G=SO(V) with dim(V) odd, p = 2 and x is a unipotent element. It is the 1-dimensional radical of V and so x acts trivially on it.

It is clear that the decomposition of Vk[x] into orthogonally indecomposable summands determines the class of x. There are various ways of writing down the decomposition Vk[x], we will use the distinguished normal form introduced in [Citation17]. Since we are interested in the simple group SO(V), we state when a class in O(V) meets the subgroup SO(V) and if so, whether it splits into two SO(V)-classes; see [Citation17, Lemma 3.11, Propositions 5.25, 6.22]. The following lemma is a combination of results from [Citation17, Chapters 5.2, 5.3, 5.6, 6.2, 6.8].

Lemma 2.4.

Let G=Sp(V) or O(V) and let x be unipotent or nilpotent element of G or Lie(G), respectively. The following list states which orthogonal decompositions Vk[x] can occur. Conversely, the multiset of orthogonal factors determines x uniquely up to conjugacy.

Furthermore, x is distinguished if and only if r = 0.

  1. If p2, then Vk[x]=i=1rW(mi)aij=1sV(nj), where the nj are distinct, and all are even (resp. odd) when G=Sp(V) (resp. O(V)). For G=O(V) and x unipotent, the class meets SO(V). For G=O(V), a class splits into two SO(V)-classes if and only if s = 0 and ai = 0 for all odd mi.

  2. If x is unipotent, p = 2 and V is of even dimension, then Vk[x]=i=1rW(mi)aij=1tV(2nj)bj where the nj are distinct and bj2 for all j. Such a class meets SO(V) if jbj is even and it splits if and only if t = 0.

  3. If x is unipotent, p = 2 and V is of odd dimension, then Vk[x]=i=1rW(mi)aij=1tV(2nj)bjR where the nj are distinct and bj2.

  4. If x is nilpotent, p = 2 and G=Sp(V), then Vk[x]=i=1rW(mi)aij=1sWlj(nj)k=1tV(2qk)bk where the sequences (nj),(lj) and (njlj) are strictly decreasing, bk2 and for all j, k either qk>njlj or qk < lj.

  5. If x is nilpotent, p = 2 and G=O(V) with V of even dimension, then Vk[x]=i=1rW(mi)aij=1sWlj(nj) where the sequences (nj),(lj) and (njlj) are strictly decreasing. Each class splits into two SO(V)-classes if and only if s = 0.

  6. If x is nilpotent, p = 2 and V is of odd dimension, then Vk[x]=i=1rW(mi)aij=1sWlj(nj)D(m)

    where the sequences (nj),(lj) and (njlj) are strictly decreasing and m<ls.

Remark 2.5.

Each simple k-group of classical type has a unique class of nilpotent and unipotent elements of highest dimension called regular. A representative of the regular class is well-known to be ixαi(1) for unipotent elements [Citation20, 3.2] and ieαi for nilpotent elements by [Citation19, 5.9] and [Citation12], where the product and sum are supported on all simple roots. From [Citation17, 3.3.6 and pp. 60–61] we see that the regular elements are usually the only ones which have a Jordan block of largest possible size. In type A, this means that x acts with a single Jordan block. For the other classical cases from the lemma, with m=dim(V):

  1. Vk[x]=V(m1)+V(1) if G=SO(V) and m is even, otherwise Vk[x]=V(m);

  2. Vk[x]=V(m) if G=Sp(V), and Vk[x]=V(m2)+V(2) if G=SO(V);

  3. Vk[x]=V(m1)+R;

  4. Vk[x]=V(m);

  5. Vk[x]=Wm/2(m/2);

  6. Vk[x]=D((m+1)/2).

2.4. Recovering all classes from the eminent ones

We present a brief discussion of how one can find a full set of representatives of the unipotent classes and nilpotent orbits iteratively from a set of representatives of the eminent classes and orbits. (If G is of exceptional type, one can simply look at the full list in [Citation21].) Since an element x is eminent in some generalized subsystem it suffices to give a set of Chevalley generators for maximal generalized subsystem subgroups. The Borel-de Siebenthal algorithm solves the problem in the case of subsystems. Hence the following lemma completes the picture.

Lemma 2.6.

Let G be a simple k-group of type Dn with p2. Let β=αm+αm+1++αn2 and define: H1=U±α1,,U±αm1,xβ+αn1(t)xβ+αn(t),xβαn1(t)xβαn(t)|tk,H2=U±αm+1,,U±αn2,xαn1(t)xαn(t),xαn1(t)xαn(t)|tk.

Then the subgroups H1 and H2 commute, H1 has type Bm and H2 has type Bnm1.

Proof.

This is proved in [Citation22, pp. 67–68], with the generators amended to reflect our choice of structure constants. □

If p2, then Theorem 1.1 implies that the unique eminent class of x in type Bm is regular. Then in view of the lemma, an eminent unipotent element in the subsystem subgroup H:=H1H2 can be represented by u=u1u2, where u1=xα1(1)xαm1(1)xβ+αn1(1)xβ+αn(1)andu2=xαm+1(1)xαn2(1)xαn1(1)xαn(1).

Similarly, if e is a regular nilpotent element of Lie(H), then e is conjugate to eβ+αn1eβ+αn+i=1,imn(eαi). For even more representatives that work regardless of the choice of signs of the structure constants, see Remark 4.6.

2.5. Representatives for classical groups in small dimensions

To illustrate our main theorem, provide representatives for the unipotent classes and nilpotent orbits in classical algebraic groups of small dimension, as considered in [Citation17, Section 8].

Table 7. Representatives for nilpotent orbits and unipotent classes of Sp4(k).

Table 8. Representatives for nilpotent orbits and unipotent classes of Sp6(k).

Table 9. Representatives for nilpotent orbits and unipotent classes of Sp8(k).

Table 10. Representatives for unipotent classes of SO7(k).

Table 11. Representatives for nilpotent orbits of SO7(k).

Table 12. Representatives for nilpotent orbits and unipotent classes of SO8(k).

Table 13. Representatives for nilpotent orbits and unipotent classes of SO10(k).

Type An is easily described. The class of x corresponds to a partition d1++dt=n+1. Viewed as an element of SLn+1 and put in Jordan normal form, the elements of the superdiagonal that x is supported on correspond to the simple roots that it is supported on. More specifically, a unipotent representative is k=1nxαk(εk), and a nilpotent representative is k=1nεkeαk, where εk=1 if j=1i1dj<k<j=1idj for some 1it, and εk=0 otherwise.

In the remaining explicit examples we give, there is an injective map from the set of unipotent classes to the set of nilpotent orbits and we have chosen representatives in such a way that a representative for a unipotent class is given by i=1txβi(1), where i=1teβi is the corresponding representative for the corresponding nilpotent orbit. Therefore we only list representatives for nilpotent orbits, apart from SO7(k) where the injection is not as natural. When p2, the characteristic is good for G, and so there is a Springer map, hence a fortiori a bijection from the set of unipotent classes to the set of nilpotent orbits. When p = 2 and G=Sp2n, a unipotent class is mapped to a nilpotent orbit such that the decompositions in Lemma 2.4(ii) and (iv) coincide. For a unipotent element of SO2n(k), the decomposition of the corresponding nilpotent element is described in [Citation17, 6.3]. In the tables for the orthogonal groups, some representatives have two rows of decompositions. The first row corresponds to the case where p2, and the second row corresponds to the case where p = 2.

3. Proof of Theorem 1.2

For unipotent elements, the required result is given in [Citation15, Section 3]. We describe the calculations in the nilpotent case. First we need the following lemma, which shows that for the action of nilpotent elements on Lie(G), the coincidences of Jordan block sizes for nilpotent elements on different orbits are the same in the adjoint case and the simply connected case.

Lemma 3.1.

Let Gsc be a simply connected simple algebraic group over k of exceptional type, and let Gad be the corresponding simple algebraic group of adjoint type. Furthermore, let x be either a unipotent element of Gsc or a nilpotent element of Lie(Gsc). Then the Jordan block sizes of x on Lie(Gsc) and Lie(Gad) are the same.

Proof.

For types G2, F4, and E8, we have Gsc = Gad, so it suffices to consider the case where Gsc is of type E6 or E7. In this case, we show that Lie(Gad)Lie(Gsc)* as Gsc-modules, from which the lemma follows. If Lie(Gsc) is simple, then Lie(Gsc) is an irreducible Weyl module and thus Lie(Gsc)*Lie(Gsc) and Lie(Gad)Lie(Gsc) by [Citation10, Lemma II.2.13(b)].

If Lie(Gsc) is not simple (p = 3 for E6 and p = 2 for E7), then by [Citation7] there is a nonsplit short exact sequence 0kLie(Gsc)W0 of Gsc-modules, where W is irreducible. Since W is self-dual, it follows that we have a nonsplit short exact sequence 0WLie(Gsc)*k0.

Because Lie(Gsc) is a Weyl module with highest weight the highest root, by [Citation10, II.2.12(4), Proposition II.2.14] we have ExtGsc1(k,W)ExtGsc1(W,k)k. Furthermore, the group ExtGsc1(k,W) is a k-vector space, and equivalence classes of extensions which are scalar multiples of each other correspond to isomorphic Gsc-modules. It follows then from ExtGsc1(k,W)k that a nonsplit extension of k by W is unique up to isomorphism of Gsc-modules, and so must be isomorphic to Lie(Gsc)*.

By [Citation7], there is a nonsplit short exact sequence 0WLie(Gad)k0 of Gsc-modules, so we conclude that Lie(Gad)Lie(Gsc)* as Gsc-modules. □

The representatives and tables of Jordan blocks on the minimal and adjoint modules is given in [Citation21]. The remaining data is supplied via in-built functions in GAP4 [Citation5] and Magma such as LieDerivedSeries. It is then straightforward to confirm there are no coincidences.

4. Proof of Theorem 1.1

We first treat the case where G is classical, making heavy use of Lemma 2.4. Throughout the section x and x are unipotent elements of G or nilpotent elements of Lie(G). In light of Lemma 2.3, we take G to be SL(V),Sp(V), or SO(V) accordingly with natural module V. The proof is completed in the following two lemmas.

Lemma 4.1.

Suppose that x is not conjugate to an element in or . Then x is not eminent.

Proof.

If x is not distinguished, then x is contained in a Levi subgroup or a Levi subalgebra and thus is not eminent. Therefore in what follows, we will assume that x is distinguished. This already completes the proof for G=SL(V), since the only distinguished elements are the regular ones. Suppose then that G=Sp(V) or G=SO(V). Comparing and with Lemma 2.4, we can assume that Vk[x] has at least two summands, and r = 0 in the notation of Lemma 2.4.

If p2, this means that Vk[x]=j=1sV(nj) with s2. It follows that x is contained in the stabilizer of a non-trivial decomposition of V into non-degenerate subspaces, and thus x is contained in a proper generalized subsystem subgroup or subalgebra.

For p = 2, we must consider the unipotent and nilpotent cases separately. Suppose first that x is unipotent. If G is of type Cn, then the argument from the previous paragraph applies.

If G has type Dn, then by Lemma 2.4(ii) we have Vk[x]=j=1sV(2nj), where s is even. When there are at least four summands, it follows that x is contained in a maximal subsystem subgroup of type DmDnm (one may choose m=n1+n2). For s = 2 there is one case remaining which is not in , namely Vk[x]=V(2n2)+V(2), corresponding to the regular unipotent class. It follows from the description of unipotent classes of Bn1 in [Citation17, Section 6.8] that x is contained in the generalized subsystem subgroup of type Bn1 (conjugate to the element x in SO(W) with Wk[x]=V(2n2)R).

If G has type Bn, then in the notation of Lemma 2.4(iii) we have Vk[x]=Vk[x]=j=1sV(2nj)+R, where s > 1. Similarly to the previous cases, it follows that x is contained in a subsystem subgroup of type DmBnm1 (again one may choose m=n1+n2).

It remains to consider the case where x is nilpotent. If G has type Cn, then Vk[x] has at least two summands, each of them isomorphic to V(2qk) or Wlj(nj). Therefore x is contained in the stabilizer of the non-trivial decomposition of V into two non-degenerate subspaces, which is a subsystem subalgebra of type CmCnm. If G has type Dn, then Vk[x] either has at least two summands isomorphic to Wlj(nj), or is isomorphic to Wn(n). In the first case, x is contained in a subsystem subalgebra of type DnjDnnj. In the second case, it follows from [Citation17, Section 5.6] that x is contained in the generalized subsystem subalgebra of type Bn1 (conjugate to the element x in so(W) with Wk[x]=D(n)). Finally if G has type Bn, then Vk[x] has at least one summand isomorphic to Wlj(nj), and so x is contained in a subsystem subalgebra of type DnjBnnj.

Lemma 4.2.

The elements in and are eminent.

Proof.

Let x be such an element. It follows from Lemma 2.4 that x is distinguished. Therefore if x is not eminent, then it must be contained one of the maximal generalized subsystem subgroups or subalgebras described in Lemma 2.2.

Suppose first that x acts on V with a single Jordan block. Elements in the subgroups and subalgebras of Lemma 2.2 act on V with at least two Jordan blocks, so it follows that x is eminent. The remaining cases occur when p = 2 and in all of them x acts with two Jordan blocks on V. We treat them in turn.

Consider first the case where G=SO(V) is of type Bn, and let R be the 1-dimensional radical of V. We will show that elements from the subgroups and subalgebras of Lemma 2.2 act on V with at least three Jordan blocks, which proves that x must be eminent. For M < G of type Dn, we have VM=WR, where W is the natural module for M=SO(W). Unipotent elements of M and nilpotent elements of Lie(M) act on W with an even number of Jordan blocks by Lemma 2.4 (ii) and (v), so in particular they have at least three Jordan blocks on V. The other possibility in Lemma 2.2 is M < G of type BmBnm for 1mn/2. In this case (V/R)M=V1V2, where V1 is 2 m-dimensional and V2 is 2(nm)-dimensional. Thus unipotent elements of M and nilpotent elements of Lie(M) act on V/R with at least three Jordan blocks by Lemma 2.4(iii) and (vi), and therefore they also act on V with at least three Jordan blocks.

If G=Sp(V) of type Cn, then x is nilpotent and Vk[x]=Wl(n) for some 0<l<n2. Such an element x acts with two Jordan blocks of size n on V. By considering each of the maximal subsystem subalgebras, we see that only those of type Cn2Cn2 (if n is even) and Dn contain elements that act with two Jordan blocks of size n. In type Cn2Cn2 the only element which does that acts as V(n)+V(n) on V, which is already in canonical form and thus not conjugate to x. In the second case, the elements x which act with two Jordan blocks in SO(V) have Vk[x]=Wl(n) for some n+12<ln. The definition of Wl(n) for orthogonal groups in even dimension, given in [Citation17, p. 66], shows that all of these elements x are conjugate to a nilpotent element acting via W(n) in Lie(G), and thus not conjugate to x.

If G=SO(V) of type Dn, then there are both unipotent classes and nilpotent orbits to consider. In all of the cases the elements x being considered act with two Jordan blocks on V. Using Lemma 2.4 we see that every element contained in subsystem subgroups and subalgebras of type DmDnm (1mm2) will act with at least four Jordan blocks on V. Thus they are not conjugate to x. By Lemma 2.2, what remains is to show that x is not conjugate to an element of the maximal generalized subsystem subgroup M (or subalgebra m) of type Bn1. When x is unipotent, this follows from [Citation13, Lemma 3.8], which shows that the only unipotent elements xM that act with two Jordan blocks on V are those with Vk[x]=V(2)+V(2n2). Such elements x are regular and not conjugate to x. Now let x be nilpotent. We consider the distinguished nilpotent elements x in m=so(W). By Lemma 2.4(vi), we have Wk[x]=j=1sWlj(nj)D(m) for some integers lj, nj and m. If s = 0, then Wk[x]=D(n) and x is a regular nilpotent element of Lie(G), and therefore is not conjugate to x. When s1, the elements x are contained in generalized subsystem subalgebras of type DnmBm1, which are contained in maximal subsystem subalgebras of type DnmDm. We have already seen that any nilpotent element contained in such a subalgebra will act with at least four Jordan blocks on V, and thus cannot be conjugate to x. This completes the proof of the lemma. □

The remaining task for G of classical type is to prove that the representatives of the eminent elements do indeed act with the claimed decompositions on V. When x is regular, a representative is provided by Remark 2.5. The non-regular eminent classes occur in characteristic p = 2 for types Cn and Dn, the representatives for these classes are constructed in the following sections.

4.1. Nilpotent representatives in type Cn

Let G=Sp(V) with V a k-vector space of dimension 2n. We need to prove that the nilpotent elements el=(i=1n1eαi)+e2αl++2αn1+αn in act as Wl(n) on the natural module V. We do this in the next lemma, but first present our choice of Chevalley basis for sp(V).

Recall the form described at the beginning of Section 2.3. For this choice of form, one checks that there is a Cartan subalgebra h of diagonal matrices in sp(V)gl(V) of the form diag(h1,,hn,hn,,h1). For 1in, define maps εi:hk by εi(h)=hi.

For all i, j let Ei,j be the linear endomorphism on V such that Ei,j(ej)=ei and Ei,j(ek)=0 for kj. Then one checks that the endomorphisms eλ in are elements of sp(V), and are simultaneous eigenvectors for adh.

Table 14. Chevalley basis for sp2n.

From a dimension count, we deduce that concatenating the elements in together with a basis of h gives a basis of sp2n. Moreover, Φ={±(εi±εj):1i<jn}{±2εi:1in}, and Φ+={εi±εj:1i<jn}{2εi:1in} is a system of positive roots. We let Δ={α1,,αn} be our base of Φ corresponding to Φ+, where αi=εiεi+1 for 1i<n and αn=2εn. We give the expressions of roots αΦ+ in terms of Δ in . Lastly, checking commutator relations reveals the basis in is in fact a Chevalley basis, with positive structure constants for extraspecial pairs.

Table 15. Type Cn, expressions for positive roots in terms of base Δ.

In particular, eαi=Ei,i+1En+i+1,n+i for 1in1,eαn=En,2n and e2αl++2αn1+αn=El,n+l for l < n.

Lemma 4.3.

For 0<l<n/2, the element el=i=1n1eαi+e2αl++2αn1+αnsp2n(k) acts on V as Vk[el]=Wl(n).

Proof.

For ease of notation we set e = el. From the description of the k[e]-module Wl(n) in Section 2.3, it suffices to show that e acts on the natural 2n-dimensional module V with two Jordan blocks of size n and that l=min{m:(em+1(v),em(v))=0 for all vV}.

Using and , we write e=El,n+l+i=1n1(Ei,i+1+En+i+1,n+i), and therefore the action of e on V is described as follows. We have e(v1)=0e(vi)=vi1(2in)e(vn+l)=vl+vn+l+1e(vj)=vj+1(n+1j<2n,jn+l)e(v2n)=0.

We immediately see that en(vi)=0 for all 1i2n. Furthermore, the relations above readily imply that the kernel of e is the 2-dimensional subspace of V generated by v1 and v2n. Thus the Jordan normal form of e has two Jordan blocks of size n.

To show that l=min{m:(em+1(v),em(v))=0 for all vV} we start by observing (ej+1(vn+lj),ej(vn+lj))=(vl+vn+l+1,vn+l)=1 for all j<l.

It remains to show that (el+1(v),el(v))=0 for all vV. Since (el+1(vi),el(vj))+(el+1(vj),el(vi))=0 for all i, j, by definition of sp(V) preserving the alternating form, it suffices to show that (el+1(vi),el(vi))=0 for all basis vectors vi.

If 1in or n+l<i2n we immediately have that (el+1(vi),el(vi))=0 since e stabilizes the totally isotropic subspaces v1,,vn and vn+l+1,,v2n. Now suppose that 1jl. Then el(vn+j)=vlj+1+vn+l+j and so el+1(vn+j)=vlj+vn+l+j+1 and so (el+1(vn+j),el(vn+j))=0, as required. □

4.2. Nilpotent and unipotent representatives in type Dn

Let G=SO(V) with V a k-vector space of dimension 2n. We need to establish the correctness of the representatives in and . Recall we define the root βl=αn+i=2ln1n2αi for (n+1)/2<ln.

We mimic the process from the last section. Using the form from the beginning of Section 2.3 we check we may choose a Cartan subalgebra h of the form diag(h1,,hn,hn,,h1) with hik. For 1in, define maps εi:hk by εi(h)=hi. A Chevalley basis for so(V) with positive structure constants for extraspecial pairs is given in .

Table 16. Chevalley basis of so(2n).

Now Φ={±(εi±εj):1i<jn} is the root system of so(V), and Φ+={εi±εj:1i<jn} is a system of positive roots. Here the base Δ of Φ corresponding to Φ+ is Δ={α1,,αn}, where αi=εiεi+1 for 1i<n and αn=εn1+εn. We give the expressions of roots αΦ+ in terms of Δ in .

Table 17. Type Dn, expressions for positive roots in terms of base Δ.

Lemma 4.4.

The element ul=i=1n1xαi(1)·xβl(1) acts on V as Vk[ul]=V(2l2)V(2n2l+2).

Proof.

From and , we have xαi(1)=I+Ei,i+1+En+i+1,n+i for all 1in1, where I is the identity and Ei,j is defined as before. Furthermore, we see that xβl(1)=I+E2ln1,2n+En,2l1. Now u:=ul acts on the basis elements of V as follows: u(vi)=vi+vi1++v1(1in)u(vj)=vj+vj+1(n+1j<2n,j2l1)u(v2l1)=v2l1+v2l+vn+vn1++v1u(v2n)=v2n+v2ln1+v2ln2++v1.

A calculation shows that the fixed point space of u has dimension 2, and that it is spanned by v1 and v2ln+v2n. Therefore the Jordan normal form of u has two Jordan blocks. To see that the Jordan block sizes are 2l2 and 2n2l+2, it suffices to show that (u1)2l2=0 and (u1)2l30, as then the largest Jordan block size in u is 2l2. To this end, a calculation shows that (u1)2l2(vi)=0 for all i and (u1)2l3(vn+1)={v1, if 2l2>n,v2n+v2+v1, if 2l2=n.

It is now clear that ((u1)2l3(vn+1),vn+1)=1 and (u1)2l2(vn+1)=0. Therefore V(2l2) occurs as summand of Vk[u] by [Citation14, Lemma 6.9], and we must have Vk[u]=V(2l2)V(2n2l+2), as required. □

Lemma 4.5.

The element el=eβl+i=1n1eαi acts on V via Vk[el]=Wl(n).

Proof.

The tables above yield eαi=Ei,i+1+En+i+1,n+i for i < n and eβl=E2ln1,n+En,2l1, so that e:=el=E2ln1,2n+En,2l1+i=1n1(Ei,i+1+En+i+1,n+i).

We need to show that e has two Jordan blocks of size n and that l=min{m|Q(em(v))=0vV}. We start by calculating the action of e on the basis vectors of V: e(v1)=0e(vi)=vi1(2in)e(vj)=vj+1(n+1j<2n,j2l1)e(v2l1)=v2l+vne(v2n)=v2ln1.

One sees that the kernel of e has dimension 2, and it is spanned by v1 and v2ln+v2n. Therefore the Jordan normal form of e has two Jordan blocks and a routine calculation with the basis {vi} shows that en(v)=0 for all vV.

Finally, we must show that l=min{m|Q(em(v))=0vV}. Since el1(vn+1)=vl+vn+l, it follows that Q(el1(vm+1))=1 and so Q(ej(el1j(vm+1)))=1 for all jl1. It remains to prove that Q(el(v))=0 for all vV. Since 2ln, it follows that (el(v),el(w))=(v,e2l(w))=0. Therefore, it suffices to show that Q(el(vi))=0 for all basis vectors vi. If 1in or 2li2n then el(vi)=vj for some j and so Q(el(vi))=0. When n+1i2l1, we have el(vi)=vl+i+vni+l and so again, Q(el(vi))=0.

Remark 4.6.

Interpreting the coefficients of the unipotent element ul of Lemma 4.4 as integers, reduction mod p for p > 2 gives an element acting on the natural module with Jordan block sizes (2l1,2n2l+1). Indeed, a calculation shows that the fixed point space is spanned by v1 and v2ln+v2n, and furthermore (u1)2l2(vj)={(1)n2v1, if j=n+10, if jn+1 so the largest Jordan block size of ul is 2l1.

Therefore ul is regular in the subsystem BmBnm+1 of Dn and is therefore another representative alongside that provided in Lemma 2.6. Similarly one can calculate that the nilpotent element el of Lemma 4.5 has Jordan block sizes (2l1,2n2l+1) in characteristic p2. Since the roots involved in ul and el are Z-independent, by [Citation8, Lemma 16.2C] these representatives have the advantage that they work regardless of the choice of signs of the structure constants.

4.3. Exceptional types

In this section we prove Theorem 1.1 for G of exceptional type. If x is unipotent, this follows from work of Lawther in [Citation15]. It is routine to refine his lists to remove those distinguished unipotent classes of G which are not eminent. For each non-eminent unipotent class we list its maximal subsystem overgroups in , for completeness. We note that a representative of a unipotent class with label X is simply jJxαj(1), where jJeαj is the representative of the nilpotent element with label X given in [Citation21]. This follows from the proofs in [Citation17] (see the introduction of Sections 17 and 18 in ibid.) since the representatives in [Citation21] are deduced from [Citation17].

Table 18. Maximal subsystem overgroups of non-eminent distinguished unipotent elements in exceptional types.

In good characteristic, [Citation16, p. 24] provides enough information to find the non-eminent distinguished nilpotent orbits. As might be expected, these classes are in bijection with the non-eminent distinguished unipotent classes (sending a unipotent class to the nilpotent orbit of the same label). It remains to consider the nilpotent orbits of Lie(G) in bad characteristic. We follow Lawther’s approach, which entails constructing representatives of each distinguished nilpotent orbit of Lie(M) for M a maximal subsystem subgroup of G and determining its G-class. To do this we make heavy use of use Theorem 1.2.

Carrying out this calculation provides a proof of Theorem 1.1. Lastly, we mention . It provides a list of distinguished, non-eminent nilpotent orbits, giving their maximal subsystem overalgebras, for all characteristics. Specializing to good characteristic, we recover the analogous result in [Citation15].

Table 19. Maximal subsystem overalgebras of non-eminent distinguished nilpotent elements in exceptional types.

Table A1. G = F4, p = 2.

Table A2. G = E6, p = 2.

Table A3. G = E6, p = 3.

Table A4. G = E7, p = 3.

Table A5. G = E7 (simply connected), p = 2.

Table A6. G = E7 (adjoint), p = 2.

Table A7. G = E8, p = 5.

Table A8. G = E8, p = 3.

Table A9. G = E8, p = 2.

Additional information

Funding

Mikko Korhonen was partially supported by SNSF grant 200021_146223 and NSFC grants 11771200 and 11931005. All authors thank the Isaac Newton Institute for Mathematical Sciences for support and hospitality during the programme Groups, Representations and Applications: New perspectives, when some of the work on this paper was undertaken (EPSRC grant number EP/R014604/1).

Notes

1 Unless G has a factor with root system Φ such that the pair (Φ,p) is (Bn,2),(Cn,2),(F4,2), or (G2,3), this implies it has a symmetric root system which is closed under sums. In any case, the relevant root systems can be obtained from the Dynkin diagram using a version of the Borel–de Siebenthal algorithm; see [18, Proposition 13.15].

References

Appendix:

Auxiliary tables for Theorem 1.2

Below follows some auxiliary data for Theorem 1.2. Let G be exceptional. In each table we use a horizontal line to separate the sets of nilpotent orbits (represented by their labels) for which the Jordan block structure on Vmin (when it exists) and Lie(G) coincide. For each class we then provide the additional data required to distinguish them. See the introduction for the terminology (DS, ALG, etc.).