Nilpotent coadjoint orbits in small characteristic
Abstract.
We show that the numbers of nilpotent coadjoint orbits in the dual of exceptional Lie algebra in characteristic and in the dual of exceptional Lie algebra in characteristic are finite. We determine the closure relation among nilpotent coadjoint orbits in the dual of Lie algebras of type in characteristic and in the dual of Lie algebra of type in characteristic . In each case we give an explicit description of the nilpotent pieces in the dual defined in [CP], which are in general unions of nilpotent coadjoint orbits, coincide with the earlier case-by-case definition in [L5, X4] in the case of classical groups and have nice properties independent of the characteristic of the base field. This completes the classification of nilpotent coadjoint orbits in the dual of Lie algebras of reductive algebraic groups and the determination of closure relation among such orbits in all characteristic.
Keywords: Nilpotent coadjoint orbits; closure relation; nilpotent pieces; Springer correspondence.
1. introduction
Let be a connected reductive algebraic group defined over an algebraically closed field k of characteristic . Let be the Lie algebra of and the dual vector space of . Denote by the set of nilpotent elements in (recall that an element is called nilpotent if it annihilates some Borel subalgebra of , see [KW]). Note that acts on by coadjoint action. The -orbits in under this action are called nilpotent coadjoint orbits. Such orbits play an important role in representation theory. When is not special for (i.e. does not equal the ratio of the squared lengths of long and short roots in any irreducible component of the root system of ), there exists a -equivariant bijection between the nilpotent variety of and (see [PS, section 5.6]). Hence in these cases nilpotent coadjoint orbits in can be and are often identified with nilpotent orbits in (-orbits in under adjoint action). The latter has been extensively studied in all characteristic. For example, the number of nilpotent orbits in is well-known to be finite and closure relation among nilpotent orbits in has been determined.
It remains to study the nilpotent coadjoint orbits in when is special, more specifically, when is of type , or and , and when is of type and . The nilpotent coadjoint orbits in type when have been classified in [X1]. We give the classification in the remaining cases here. In particular, it follows from the classification that the number of nilpotent coadjoint orbits in is finite for any as in the beginning of the introduction. This result has appeared in the PhD thesis of the author and we include it here for completeness. We determine the closure relation among nilpotent coadjoint orbits in when is special. For this we can and will assume that is adjoint in type and simply connected in type .
In [L5, X4] Lusztig and the author give a case-by-case definition of a partition of into nilpotent pieces for classical groups, indexed by the set of unipotent orbits in the group over complex numbers of the same type as . In [CP, Section 7] Clarke and Premet define nilpotent pieces in uniformly across all types, in the same way as Lusztigβs original definition of unipotent pieces (a partition of unipotent variety in ). Moreover the nilpotent pieces in [CP] equal the Hesselink strata [H2] on (considered as the null-cone of under coadjoint -action), and coincide with the nilpotent pieces defined in [L5, X4] for classical types. These pieces have nice properties independent of . On the other hand, there is a natural injective map from the set to the set of nilpotent coadjoint orbits in given by Springer correspondence. Using this map and the closure relation on one can define a partition of into locally closed pieces. We show that for classical groups these pieces are the same as nilpotent pieces defined by Lusztig and the author, and thus also the same as nilpotent pieces defined by Clarke and Premet. In particular we determine which nilpotent coadjoint orbits lie in the same piece (Proposition 6.1). The analogous result for nilpotent pieces in is given in [X2]. We also describe nilpotent pieces in for type when and for type when .
This paper is organized as follows. Sections 2-6 study the cases when is of type , and . In Section 2 we recall a natural partial order on the set of irreducible characters of the Weyl group of (used by Spaltenstein [Sp2] to describe closure relation among nilpotent orbits in ), the classification of nilpotent coadjoint orbits in , the combinatorial description of the Springer correspondence maps, and the definition of nilpotent pieces in given in [L5, X4]. Section 3 and Section 4 are preparation for Section 5, where we describe the Springer fibers at elements in and induction for nilpotent coadjoint orbits in (by an easy adaptation of [LS, Sp2]) respectively. In Section 5 we determine the closure relation on which turns out to correspond to the natural partial order on recalled in Subsection 2.1 via Springer correspondence map. In Section 6 we describe the nilpotent pieces in explicitly. In Section 7 we classify the nilpotent coadjoint orbits in when is of type and , and when is of type and . We determine the closure relation on and describe the nilpotent pieces in explicitly.
Acknowledgement The author wish to thank George Lusztig and Kari Vilonen for helpful discussions and for encouragement. Thanks are also due to the referee for carefully reading the manuscript and for many suggestions that helped improve the exposition of the paper.
2. Notations and recollections
Although nilpotent coadjoint orbits in type are not under consideration, we include the information for type in Subsections 2.1-2.3 for future use in the inductive proof of Theorem 5.1 in type case.
2.1. A partial order on the set of irreducible characters of Weyl groups of type
For a finite group we denote by the set of irreducible characters of (over ).
Let denote the set of all partitions of an integer . For a partition and each , we set
Let denote the set of pairs of partitions. If is a Weyl group of type or (resp. ), we can identify with the set (resp. the set with each pair counted twice) ([L1]). There is a natural partial order on the set as follows. We say that
if
and that if and . This gives rise to a partial order on (in the case of type , the two degenerate characters corresponding to a pair are incomparable).
2.2. Classification of nilpotent coadjoint orbits in (type )
Let be a finite dimensional vector space over k equipped with a fixed non-degenerate symplectic from (resp. a fixed non-degenerate quadratic form with the associated bilinear form denoted by ). We can assume that (resp. ), the subgroup of (resp. identity component of the subgroup of ) that preserves (resp. ). Then (resp. ).
Let (resp. ) denote the vector space of all quadratic forms (resp. all symplectic forms ). We have a vector space isomorphism (see [L5, X4])
(a) ββ
β(resp. ),
where is such that for all (resp. ).
Suppose that with and . Let always denote the quadratic form corresponding to under the isomorphism in (a) and let always denote the bilinear form associated to (given by )). Let be defined by for all . Assume that . Then the -orbit of is characterized by a pair as follows ([X1]):
(i) the partition given by the sizes of Jordan blocks of (we have even for all );
(ii) the map given by (we have , and for all ).
Suppose that and . Let always denote the symplectic form corresponding to under the isomorphism in (a).
If , and , let be defined by for all . The -orbit of is characterized by a pair as follows ([H1]):
(i) the partition given by the sizes of Jordan blocks of (we have even for all );
(ii) the map given by (we have , and for all ).
Assume that , and . Let be the unique integer such that there exists a (unique) set of vectors with
(b)
If , let be a complementary subspace of in ; if , let be a set of vectors such that
(c)
and let . Then and is non-degenerate. Define by
for all and let be given by .
Then the orbit of is characterized by as follows ([X1]):
(i) the integer ;
(ii) the partition given by the sizes of Jordan blocks of (we have even for all );
(iii) the map given by (we have , and for all ).
Note that does not depend on the choice of and .
Let (resp. , ) be the set of all (resp. ) corresponding to nilpotent coadjoint orbits () in (resp. , ). Note that in the case of , there are two orbits corresponding to each pair with for all . Let
2.3. Combinatorial description of Springer correspondence maps (type )
Let denote the Weyl group of . Recall that we have an injective Springer correspondence map [X1]:
which maps an orbit to the Weyl group character corresponding to the pair under Springer correspondence. When , the Springer correspondence maps are given as follows ([X3])
We denote the image of (resp. , ) (when ) by (resp. and ). Let , and . We have
(here we use the identification of with ).
2.4. Nilpotent pieces in and
Suppose that (resp. ) as in Subsection 2.2. Let and
. Let be the canonical filtration of associated to as in [L5, X4], where . If , let be defined as in Subsection 2.2 (resp. by for all ), then
ββ(a) .
The definitions of when are recalled in 2.4.1 (resp. 2.4.3). We
define
Then for finitely many and . The sequence of numbers () depends only on and not on the choice of ; we denote this sequence by . Two sequences and are equal iff for all .
Note that if , the orbit of is characterized by the partition given by the sizes of Jordan blocks of . It follows from (a) that
ββ(aβ²) if , and , then .
Moreover each nilpotent piece consists of one orbit when .
2.4.1. Canonical filtrations for ()
Assume that and . Let
and let be defined for as in Subsection 2.2. The canonical filtration associated to is defined by induction on as follows ([L5]), where . If , we set for all and for all . Hence is defined when . Assume now that and . Let be the smallest integer such that , the smallest integer such that and
ββββββββββββ
Then . We set
Let . Then induces a nondegenerate symplectic form on and induces a quadratic form corresponding to , where . By induction hypothesis, a canonical filtration of is defined for . For we set to be the inverse image of under the natural map (note that and ). This completes the definition of .
2.4.2.
Suppose that and the -orbit of corresponds to . Recall that ([X1]) we have a decomposition of into mutually orthogonal -stable subspaces such that , , where . Moreover, , ,
where
(a) means that there exist such that ; we have .
We state some facts which will be used later (see [X1]).
(i) Let be a -stable subspace of such that with . Let , , and let be the quadratic form on induced by . Using the basis for chosen as in (a), one can easily check that
ββββββββββββ
(ii) Let be a -stable subspace of such that (an orthogonal decomposition into copies of ), where . Let and be the quadratic form on induced by . Using the basis for chosen as in (a), one can easily check that
ββββββββββββ
2.4.3. Canonical filtrations for ()
Assume that and . Let
and let , be defined for as in Subsection 2.2. The canonical filtration associated to is defined by induction on as follows (see
[X4]), where and for all . If we set for all and for all . Hence is defined when . Assume now that and . Let be the smallest integer such that , and
ββββββββββββ
Then . We set
where is the map and is such that for all .
Let . Then induces a non-degenerate quadratic form on and induces a symplectic form corresponding to , where . By induction hypothesis, a canonical filtration of is defined for . For we set to be the inverse image of under the natural map (note that and ). This completes the definition of .
3. Springer fibers
Suppose that and is of type or in this section.
3.1.
For a Borel subgroup of , we denote by the Lie algebra of and define . For , denote by the -orbit of and by the centralizer of in . Let be the variety of all Borel subgroups of and let be the Springer fiber at . One can easily adapt the proofs in [St, p. 132] and [Sp1] to show that is connected and all its irreducible components have the same dimension.
Proposition 3.1.
Suppose that . We have
3.2.
Suppose that with in this subsection. We can identify the variety with the set of all complete flags such that for all .
Lemma 3.2.
Suppose that . We can identify with the set
Proof.
Let , and . There exists a basis , , of such that and for all . Assume that . We have and
Let be the Borel subgroup corresponding to . For , we have for , and for . Hence if and only if for all , for all , and for all . The lemma follows. β
Let and suppose that . Let . Then we have and . Let . The non-degenerate symplectic form on induces a non-degenerate symplectic form on ; we write . Moreover the quadratic form induces a quadratic form on which corresponds to an orbit in . Let and for each that arises in this way, let
Then the fibers of the morphism
are isomorphic to , where and the -orbit of is . Now we have that and by induction hypothesis that . For corresponding to , we have (see [X1]). It is then easy to check that using the following lemma ([X3, 5.3]).
Lemma 3.3.
We have if and only if and are related as follows: and for , , and satisfies , (this can happen if ). Moreover if and if .
3.3.
Suppose that with in this subsection. We can identify the variety with the set of all complete flags such that and for all .
Lemma 3.4.
Suppose that . We can identify with the set
Proof.
Let , and . There exists a basis , , of such that , , and for all . Assume that . We have for all , and for all , and
Let be the Borel subgroup corresponding to . For , we have for , for , and for . Hence if and only if for all , for all , and for all . The lemma follows.β
Suppose that and that . Let . Then we have and . Let . The non-degenerate quadratic form on induces a non-degenerate quadratic form on ; we write . The symplectic form induces a symplectic form on , which corresponds to an orbit in . Let and for each that arises in this way, let
Then the fibers of the morphism
are isomorphic to , where and the -orbit of is . For corresponding to we have (see [X1]). Using the same argument as in the case of and the following lemma ([X3, 6.3]), one can easily check that .
Lemma 3.5.
We have if and only if and are related as follows.
(a) ,
and (this can happen if ). We have ;
(b) ,
and for ,
,
and
satisfies ,
(this can happen if ).
We have if
and if
.
4. Induction for nilpotent coadjoint orbits
Suppose that and is of type or in this section.
4.1.
In this subsection we recall the notion of -induction in (see [L4] and the references therein). For , let be the smallest integer such that appears in the th symmetric power of the reflection representation of . For a parabolic subgroup of and , there is a unique such that and appears in ; we write .
Let (where we use the notation that ). Let and for we denote (resp. ) a Weyl group of type or (resp. ). Let sgn denote the sign character of . For each , we have ([L4])
ββ(a)
where is defined by
It is easy to see that is injective and .
4.2.
Let be a Levi subgroup of a parabolic subgroup of . Denote by and the Lie algebra of and respectively. Choose a maximal torus and a Borel subgroup of such that . Let and be the root system of and respectively. We define
where denotes the root space corresponding to . Note that .
Let be an -orbit in . Since consists of finitely many -orbits, there exists a unique -orbit in such that is dense in (note that ). Following [LS] we say that is obtained by inducing from to and denote .
Proposition 4.1.
Suppose that and . We have , where is the Weyl group of .
The proposition is an analog of [LS, 3.5]. To prove it one can adapt the proof in [Sp4, 4.1] (another proof in unipotent case is given in [LS]). We outline the proof here. Suppose that and . Let and . One can show that and by direct computation (using Proposition 3.1 and the known information on and on ) or by adapting the proof in [Sp4]. It is easy to adapt the proof in [LS] to show that and . It then follows that . Now let and the set of irreducible components of of dimension . Let denote the component group of . Then acts on ; we denote the corresponding representation by . We have and (see [L2, X3]). It now follows from the definition of -induction that .
Note that it follows from Proposition 4.1 and the injectivity of that does not depend on the choice of . Hence we can write .
4.3.
Let (resp. ) be the Lie algebra of (resp. ). Let be a parabolic subgroup of such that (resp. ) for a Levi subgroup of . Let be the nilpotent orbit in corresponding to the nilpotent orbit (resp. ) and to the orbit in . Let (resp. ) correspond to the orbit . We write (resp. ) in view of the following commutative diagrams (see 4.1 (a) and Proposition 4.1)
where is the Springer correspondence map. Using this one easily sees that (see also [Sp2])
(a) For every (resp. ), there exists a sequence of integers such that (resp. ) is of the form for some sequence .
Here in the expression , denotes the empty partition, not the zero orbit.
5. Closure relation among nilpotent coadjoint orbits in type , in characteristic 2
5.1.
Assume that is of type or . By identifying with the set of we get a partial order on (see Subsection 2.1). For , we say that if is contained in the Zariski closure of , and that if and . We have
Theorem 5.1.
Suppose that and . We have if and only if .
Note that theorem is true when is of type by [Sp2] or when (see [X2]) (in these cases we can identify with ). We prove the theorem in the remainder of this section (where we assume that ) using similar arguments as in [Sp2] most of the time. In the reduction process of 5.3 we use a different argument without using the theory of sheets and packets.
5.2.
Assume that (resp. ). Let and be orthogonal subspaces of such that . Then the restrictions of (resp. ) on and are non-degenerate. Let (resp. ) and (resp. ). Using the isomorphism (resp. ) (see 2.2 (a)) we have a natural inclusion .
Lemma 5.2.
If and are such that and , then .
5.3.
For , we denote the corresponding orbit, where stands for or (type of ). We show by induction on that if then (see Subsection 5.4 for and Subsection 5.5 for ). We may assume that this is true for classical groups of strictly smaller dimension and that , . We have the following reduction process.
Reduction 1. Suppose in the case of (resp. ), we have with and , (resp. , ). Assume that we can find , with corresponding to in and corresponding to in , such that and . Then by induction hypothesis and Lemma 5.2 we have .
Reduction 2. Assume that we can find in such that and for some positive integers . Then by induction hypothesis and the fact that induction for orbits preserves order, we have .
5.4.
Assume that . Recall that . Let be such that and .
Lemma 5.3.
Suppose that reduction 1 does not apply. One of the following is true:
(a) ,Β ;
(b) ,Β .
Proof.
Note that we can apply Reduction 1 if or , and for some , with or for some . We denote by the partition with all parts and multiplicity . Since Reduction 1 does not apply, if , then we have and thus are as in case (a). From now on we assume that . Let and . Then (otherwise is minimal). We have the following cases.
1) . Let and . Then and .
i) If , then there exists such that . One easily verifies that (note that for ). Hence .
ii) If and , then there exists such that . One easily verifies that (note that and for ). Hence .
iii) If , and , then there exists such that . One easily verifies that (note that and for ). Hence .
iv) If , and , then let with and except that and . Then one easily verifies that (note that and for ). Hence .
Since , in case (i) we have , cases (ii) and (iii) do not happen, in case (iv) we have . As Reduction 1 does not apply, in case (i) we have , and ; in case (iv) we have , and , but notice that ( since ) so case (iv) does not happen. It follows that and are as in (a).
2) . Then . Let and .
i) If , then there exists such that . One easily verifies that (note that for ). Hence .
ii) If and , then there exists such that . One easily verifies that (note that , and for ). Hence .
iii) If , , then there exists such that .
Let with and except that and . Then one easily verifies that (note that , and for ). Hence .
Since , we have that in case (i) , in cases (ii) and (iii) . As Reduction 1 does not apply, in case (i) we have , and ; case (ii) does not happen (); in case (iii) we have , and (but notice that , so this case does not happen). It follows that and are as in (b). β
Proposition 5.4.
Suppose that Reductions 1 and 2 do not apply. One of the following is true:
-
(i)
, ;
-
(ii)
, ;
-
(iii)
, ;
-
(iv)
, ;
-
(v)
, Β Β (for even).
Proof.
It remains to show that in each case of Proposition 5.4 we have . In case (i) it is obvious that as corresponds to the orbit. In each case (ii)-(v), we choose an element and define a family such that . Then it follows that . The elements (or equivalently the quadratic form associated to ) and the family in each case are given as follows. We have .
Let be a basis of such that .
(ii) . Then
(iii) ;
. Then .
(iv) ;
. Then .
(v) , where ;
Then .
5.5.
Assume that . Recall that . Let be such that and .
Proposition 5.5.
Suppose that reduction 1 and 2 do not apply. Then , .
Proof.
Note that we can apply Reduction 1 if or if for some , with or for some .
Since Reduction 1 does not apply, we have . Let and . Let with for and for . Then one easily verifies that and . Hence . Since Reduction 1 does not apply, we have , and . We have and if ; and if . Since Reduction 2 does not apply, and are as in the proposition. β
Let and . We show that . If , this is obvious. Assume that . Let , be a basis of such that and . Let be the symplectic form corresponding to such that . Then . Let , be defined by:
We have . Thus and .
5.6.
We show that if then . Since induction for orbits preserves order and iff , we may use the operation as often as needed. We show that
As pointed out by the referee, for the condition follows by considering the partitions corresponding to and , which are and .
We write for . For each , let , , , and
It is enough to show that and .
It follows from the definitions that we have
Lemma 5.6.
(a) if (resp. ) then (resp. ).
(b) if , and (or ), then or .
Assume that (resp. ). Let or . We can assume that for some . Let be an isotropic subspace of with and let be such that . Let be the parabolic subgroup that stabilizes and the Levi subgroup of that stabilizes . Then (resp. ). We have .
Since , there exists such that , (resp. ) and such that (resp. ) induces a quadratic form (resp. symplectic form) on which corresponds to an element . Let correspond to the -orbit (resp. -orbit) of . We have and thus
ββeither a)
ββor b) .
In case a) we have and or . Applying when needed we can assume that for some sequence (see 4.3 (a)) and moreover:
(i) is even;
(ii) for and if ;
(iii) .
We apply the previous construction with , and repeat the argument with replaced by (conditions (i)-(iii) above are still satisfied) until we reach a point that . Hence we are reduced to the case that (or ), , ( satisfy (i)-(iii) above), where and are related as in the above construction. We will study more closely the relation between and and derive a contradiction.
Suppose that . Let , . We show that (see 5.7 and 5.8)
c) for all ;
d) for all ;
e) if , then and for (resp. if , then and for ).
If , it follows from c) and d) that and from e) that ; if it follows from c) and d) that and from e) that . This is the required contradiction.
Suppose that . We have if is odd, then , , ; if is even, then , , ; if is even, then , , , ; if is odd, then , , , ; where in each case and .
5.7.
We prove 5.6 c)-e) for by studying the quadratic form on induced by . Let .
If , then is indecomposable (as -module), and (where , ; see 2.4.2 (a)). Then one easily verifies that (if ), or (if ). Thus and ; in this case (c) and (d) hold.
If , then is decomposable, and there exists an orthogonal decomposition with indecomposable, , such that . We apply the previous result for each factor in the decomposition . Assume that (where ). Then and for some . Suppose that . We have and ; thus c) and d) hold.
Now we prove e). Assume that . Then and . Note that for ; hence if is a permutation such that , then for . Moreover since and for , for all . It follows that for . Now we have and for , for some ; hence for . Then e) follows.
5.8.
We prove 5.6 c)-e) for by studying the symplectic form on induced by . Let .
If , then , (where are as in 2.2 (b) (c)), and thus , ; in this case (c) and (d) hold.
If , then there exists an orthogonal decomposition with indecomposable (as -module), , such that . Assume that (where ; notation as in [X2, 5.6]). One easily verifies that (if ), or (if ). Write . Then and . We have , and ; thus if and if (note that ). Hence c) and d) follow.
We prove e). Assume that . Then and . Note that for ; hence if is a permutation such that , then for . Moreover for all , since and . It follows that for all . Now we have and for , for some or (if ) ; hence and for . This completes the proof of Theorem 5.1.
6. Nilpotent pieces in in type , in characteristic 2
Assume that and is of type or throughout this section unless otherwise stated.
6.1.
Suppose that (resp. ). Let denote the image of the Springer correspondence map , where as before stands for or (the type of or ). We have (see [L1, X2])
ββ, .
Let and let be such that . Define
to be the set of all orbits such that
and
for any
with
. We show that
ββ(a) form a partition of .
Following [X2], we define maps
where if , then
if , then and
It is easy to verify that in each case we get a well-defined element
. We have
(b) if , , then , ; if moreover , then .
In fact, if , (b) follows from [X2, 4.2]; if , one can prove (b) by the same argument. Now in view of Theorem 5.1, it follows from the definition of and (b), (c) that for each ,
ββ(d) .
Then (a) follows from (d).
6.2.
Assume that and . If , then , if and , then
if and , then
otherwise .
Assume that and . Let
For , if and , then
if and , then
otherwise , .
We show that
ββ(a)
Let (resp. ) and . Assume that , and . Using the definition of , one easily shows that iff and is even, and iff (resp. iff , and iff and is odd). Using this and the description of the map (see [X2, 2.4]), one easily verifies that
It is then easy to verify that using the description of (see 2.3).
Note that the proposition computes the nilpotent pieces in explicitly. Now in view of (a) and 6.1 (d), it follows from Proposition 6.1 that each set is a nilpotent piece defined in [L5, X4]. One can also define a partition of into special pieces as in [L3, X2] and show that each special piece is a union of nilpotent pieces. The proof of the proposition is given in the remainder of this section following the argument used in [X2].
6.3.
Suppose that (resp. ). Let and
. Suppose that and (see Subsection 2.4).
We show that
ββ(a) for all .
Then Proposition 6.1 follows from (a) and Lemma
2.1.
We prove (a) by induction
on . Let .
If , (a) is obvious. Assume from now on that .
Let , , be associated to and , , defined for as in
2.4.1 (resp. 2.4.3). Let be the orbit of in and let
.
Suppose that and Since , by induction hypothesis
for all . By the definition of we have that for all , and thus .
We show that
ββ(b)
It then follows from (b) and 2.4 (aβ²) that for all , , and that for all . Hence (a) follows (note that for all ).
6.4.
Assume that throughout subsection 6.6. We keep the notations in 6.3. Suppose that . We show that
ββ(a)
Recall that and (see 2.4.1). We describe in various cases in the following and then (a) follows.
Suppose that . Then the map is linear and thus is a line.
Suppose that , or and . Then .
Suppose that and . Let be subspace of such that and let . Then is non-degenerate (in fact, if , where , then and thus ; now use induction and similar argument one shows that ). Thus , is -stable and ( implies that ). We have (note that ). Thus , where is a line (we apply the result in the first case for ).
6.5.
We describe in various cases as follows. Let be the unique integer such that
(i) , or and . We have
(if ),, for , if and if .
(ii) and . We have
(if ), , (if ), , for .
(iii) . We have
, (if ),, for .
Let and be defined for as before. Let .
Assume that we are in case (i). We have a decomposition of into mutually orthogonal -stable subspaces such that
βββββββββ and .
Then and . Hence we have a natural decomposition of into mutually orthogonal -stable subspaces, where , and (see 2.4.2 (ii))
βββββββββ .
We have and . For , and thus , which implies that and thus . Now and , thus the assertion on holds; since .
Assume that we are in case (ii). Then . We have a decomposition of into mutually orthogonal -stable subspaces such that (we use [X1, Lemma 2.9])
ββ, , and .
Then and where and . Hence we have a natural decomposition of into mutually orthogonal -stable subspaces, where , and (see 2.4.2 (i) (ii))
ββ.
We have and . Thus for , (note that ); for , (since ) and thus .
Assume that we are in case (iii). We have a decomposition of into mutually orthogonal -stable subspaces such that (we use [X1, Lemma 2.9])
βββββββββ and .
Then and where and . Hence we have a natural decomposition of into mutually orthogonal -stable subspaces, where . Moreover (see 2.4.2 (i))
βββββββββ and .
We have and . Thus for , (as ); for , and thus .
6.6.
Using the definition of and the description of in 6.5, we compute and in each case (i)-(iii) as follows. It is then easy to check that 6.3 (b) holds in each case.
Let in case (i) and in cases (ii) and (iii). We have for all , since and for all . In case (i), if , then since and . In cases (ii) and (iii), for , since , and . Let and . We have
(i)
(ii)
(iii)
6.7.
Assume that and in the remainder of this section. Suppose that . We keep the notations in 6.3.
We show in this subsection that
ββ(a)
Recall that and (see 2.4.3). We describe in various cases in the following and then (a) follows.
Suppose that . Then since . We have , where is a chosen square root on k.
Suppose that . Then .
Suppose that . Then , where is a line (see [X2, 5.5]).
Suppose that . Then , where , and if , if (see [X2, 5.5]).
Suppose that and . Then , where . Same argument as in [X2, 5.5] shows that we have a decomposition of into -stable orthogonal subspaces such that and . Hence and .
Suppose that and . Then .
6.8.
We describe in various cases as follows. Let be the unique integer such that
(i) , or and . We have
, , (if ), , for .
(ii) . We have
, , for , if , and if .
(iii) . We have
(if ),, for , if and if .
(iv) and . We have
, (if ), (if ), , for .
(v) and . We have
, for .
Recall that we can choose (or if ) such that in the decomposition . In the following or is chosen as such. Let .
Assume that we are in case (i). Suppose first that . Then . There exists such that for all . Let . Then and (note that ). The description for follows.
Suppose now that and . Let and let be defined accordingly. Then for all , . Then , where . Thus . Let be defined for . We have . Suppose that .
We have a decomposition of into mutually orthogonal -stable subspaces such that (see [X3, 5.6])
βββββββββ and .
Then , where and . We have and . Thus and (note that and ) for all .
Assume that we are in case (ii). Then . Thus and . We have . If , then ; if , then and thus , (note that for ) and .
In cases (iii)-(v), we have , where for some subspace . Thus . Let be defined for . We write . We can apply the results in [X3, 5.6] for and and then in each case the assertions on follow. The assertions on also hold since (see below for the description of ).
(iii) In this case and . Note that for .
(iv) In this case , and . Since , and thus for all .
(v) In this case and Note that for , .
6.9.
Using the definition of and the description of in 6.8, we compute and in each case (i)-(v) as follows. It is then easy to check that 6.3 (b) holds in each case.
Let in cases (i) and (v); in case (ii); and in cases (iii) and (iv). Then for all since and for all . In case (i), if , then and thus ; if moreover , then since , . In case (v) or in case (ii) with (then ), since and ; if , then for since and . In case (iii) and (iv), for since and , .
Let and . We have
(i)
(ii)
(iii)
(iv)
(v)
7. Nilpotent coadjoint orbits in type in characteristic 3 and in type in characteristic 2
Assume that is of type and , or of type and in this section unless otherwise stated. We classify the nilpotent coadjoint orbits in and determine the closure relation among them. We describe explicitly the nilpotent pieces in defined in [CP]. In particular, it follows from the classification (see Subsections 7.2-7.5) that
Proposition 7.1.
The number of nilpotent coadjoint orbits in is finite.
As mentioned in the Introduction, Proposition is true now for any connected reductive algebraic group in any characteristic.
7.1.
Following a suggestion of the referee, we include here references for unipotent orbits in and nilpotent orbits in . When , we can identify nilpotent orbits in with unipotent orbits in , which are thus both classified by Bala-Carter theory. If is of type and , both unipotent and nilpotent orbits are classified by Stuhler in [S]. If is of type , the unipotent orbits are classified by Shoji in [Sh] when and by Shinoda in [Shi] when , and the nilpotent orbits in are classified by Spaltenstein in [Sp5] when and by Holt and Spaltenstein in [HS] when . The closure relation among unipotent orbits in is determined by Spaltenstein in [Sp3] and that among nilpotent orbits in when is of type and is given in [Sp5].
7.2.
Let be a finite field of characteristic (resp. ). Let be of type (resp. ) defined over . We prove Proposition 7.1 by studying -orbits in . The strategy is as follows. We specify various elements which lie in different -orbits and compute (the number of rational points in the centralizer ). Then by a direct computation one verifies that the numbers of rational points in all nilpotent coadjoint orbits add up to , where is the number of positive roots. As (private communication by G. Lusztig, see also [CP]), we get all the -orbits in .
7.3.
We describe how to compute for in this subsection.
Let be a maximal torus of , the root system of , a set of simple roots, and the corresponding set of positive roots. We have a Chevalley basis of satisfying
where and are constant integers (for determination of structural constants see [C]). For each , there is a unique 1-dimensional connected closed unipotent subgroup and an isomorphism
such that for all and . We assume that and normalizes . Define . Then is generated by . Let be the Borel subgroup of , where . By Bruhat decomposition, each can be written uniquely in the form for some , some and some , where is a representative of in . We can choose to be the representative of the simple reflection , . Let , be the Lie algebra of , respectively. We define by
Then form a basis of . The coadjoint actions of , , and , on , are given as follows
(a) | ||||
where (here the equality is in and we then reduce mod to regard as a constant in k).
Since , we can find representatives of nilpotent coadjoint orbits in , namely, we can choose elements of the form , . Now we can compute using the Bruhat decomposition
and (a). In particular, we need knowledge on the set . Let
be the set of minimal elements in the set
under the order relation on , where if can be written as a sum of positive roots. If , then by (a), for any , there exists such that , more precisely,
where we write for .
7.4.
Suppose that is of type and in this subsection. We denote by (resp. ) the short (resp. long) simple root. The structural constants can be chosen as follows
Fix and . The representatives for nilpotent coadjoint orbits over and are listed in Table 1.
Orbit | Representative | |
---|---|---|
One can easily verify that , , and are in the same -orbit. Thus form a set of representatives for -obits in . This proves Proposition 7.1 for type .
It is easy to verify that the closure relation among nilpotent coadjoint orbits in is as in Figure 1 and the nilpotent pieces in coincide with nilpotent coadjoint orbits.

7.5.
Suppose that is of type and in this subsection. We denote by (resp. ) the long (resp. short) simple roots with . We denote by the root . The structural constants can be chosen as follows:
Fix and . The representatives for nilpotent coadjoint orbits over and are listed in Table 2.
Orbit | Representative | |
---|---|---|
The computations of are long and follow the strategy described in 7.3. We give one example here and omit the details. We denote by the simple reflections in . Consider . We have and (see 7.3). Assume . Then and . In fact a further look at the formulas in 7.3 (a) shows that if either , or both and , then . This forces . Now it is easy to verify that
Thus
We need to show that are not in the same -orbit for those in Table 2 with . We verify this for and . The verification of others is entirely similar. Assume that there exists such that and . By a similar argument as in the last paragraph of Subsection 7.3, we have and . A closer look at the formulas in 7.3 (a) shows that and can not both be negative if . This forces or . It is clear that . Thus for some and . When using (a) to solve , the following equation appears
which has no solution in by our choice of . Moreover the computation shows that and lie in the same -orbit.
Similarly one can verify that , , , are in the same -orbit respectively. Hence , form a set of representatives for -obits in . This completes the classification of nilpotent coadjoint orbits in for type , and the proof of Proposition 7.1.
7.6.
Suppose again that is of type and in this subsection. For , we define
Now let , and , where
For , we define . Using [CP] and well-known results on nilpotent coadjoint orbits in in characteristic (identified with nilpotent orbits in ), we get that the nilpotent pieces in defined in [CP] are as follows
(5) |
where
Moreover
(6) |
and
It is easy to see that
(7) | |||
One can also verify that
(8) | |||
For example, since there exists no such that , , and if either , or both and .
It follows from (5)-(8) and dimension consideration that the nilpotent pieces in are as follows
where is the -orbit of , .
By [CP] and [H1], we have . It follows that , , and moreover , in view of (6) and (8). We show that
(9) | |||
(10) |
Let be two nilpotent coadjoint orbits and . We have if and only if since is complete. Using in 7.3, one can show that
(11) | |||
(12) |
Now (10) follows from (12) and follows from (11). Suppose that . Then there exists such that . Suppose that . By (11), . It follows that
(13) | |||
Suppose . Write . Then . Since , by (11), there exist greater than or such that . But this is impossible. Similarly one shows that can not equal any set of pairs in the right hand side of (13). This gives us a contradiction. Thus (9) is proved.
It follows from the above discussion that the closure relation on nilpotent coadjoint orbits in is as in Figure 2.

References
- [C] C. Chevalley, Sur certains groupes simples. (French) Thoku Math. J. (2) 7 (1955), 14-66.
- [CP] M. C. Clarke and A. Premet, The Hesselink stratification of nullcones and base change. Invent. Math. 191 (2013), 631-669.
- [H1] W. H. Hesselink, Nilpotency in classical groups over a field of characteristic 2. Math. Z. 166 (1979), no. 2, 165-181.
- [H2] W. H. Hesselink, Desingularizations of varieties of nullforms. Invent. Math. 55 (1979), no. 2, 141-163.
- [HS] D. F. Holt and N. Spaltenstein, Nilpotent orbits of exceptional Lie algebras over algebraically closed fields of bad characteristic. J. Austral. Math. Soc. Ser. A 38 (1985), no. 3, 330-350.
- [K] V. Kac, Infinite-dimensional Lie algebras. Third edition. Cambridge University Press, Cambridge, 1990.
- [KW] V. Kac and B. Weisfeiler, Coadjoint action of a semi-simple algebraic group and the center of the enveloping algebra in characteristic . Nederl. Akad. Wetensch. Proc. Ser. A 79=Indag. Math. 38 (1976), no. 2, 136-151.
- [L1] G. Lusztig, A class of irreducible representations of a Weyl group. Nederl. Akad. Wetensch. Indag. Math. 41 (1979), no. 3, 323-335.
- [L2] G. Lusztig, Intersection cohomology complexes on a reductive group. Invent. Math. 75 (1984), no.2, 205-272.
- [L3] G. Lusztig, Notes on unipotent classes. Asian J. Math. 1 (1997), no. 1, 194-207.
- [L4] G. Lusztig, Unipotent classes and special Weyl group representations. J. Algebra 321 (2009), no. 11, 3418-3449.
- [L5] G. Lusztig, Unipotent elements in small characteristic, IV. Transform. Groups 15 (2010), no. 4, 921-936.
- [LS] G. Lusztig and N. Spaltenstein, Induced unipotent classes. J. London Math. Soc. (2) 19 (1979), no. 1, 41-52.
- [PS] A. Premet and S. Skryabin, Representations of restricted Lie algebras and families of associative L-algebras. J. Reine Angew. Math. 507 (1999), 189-218.
- [Sh] T. Shoji, The conjugacy classes of Chevalley groups of type over finite fields of characteristic . J. Fac. Sci. Univ. Tokyo Sect. IA Math. 21 (1974), 1-17.
- [Shi] Ken-ichi Shinoda, The conjugacy classes of Chevalley groups of type over finite fields of characteristic 2. J. Fac. Sci. Univ. Tokyo Sect. I A Math. 21 (1974), 133-159.
- [Sp1] N. Spaltenstein, On the fixed point set of a unipotent element on the variety of Borel subgroups. Topology 16 (1977), no. 2, 203-204.
- [Sp2] N. Spaltenstein, Nilpotent Classes and Sheets of Lie Algebras in Bad Characteristic. Math. Z. 181 (1982), 31-48.
- [Sp3] N. Spaltenstein, Classes unipotentes et sous-groupes de Borel. (French) [Unipotent classes and Borel subgroups] Lecture Notes in Mathematics, 946. Springer-Verlag, Berlin-New York, 1982.
- [Sp4] N. Spaltenstein, On the generalized Springer correspondence for exceptional groups. Algebraic groups and related topics (Kyoto/Nagoya, 1983), 317-338, Adv. Stud. Pure Math., 6, North-Holland, Amsterdam, 1985.
- [Sp5] N. Spaltenstein, Nilpotent classes in Lie algebras of type over fields of characteristic 2. J. Fac. Sci. Univ. Tokyo Sect. IA Math. 30 (1984), no. 3, 517-524.
- [St] R. Steinberg, Conjugacy classes in algebraic groups. Lecture Notes in Mathematics, Vol. 366. Springer-Verlag, Berlin-New York, 1974.
- [S] U. Stuhler, Unipotente und nilpotente Klassen in einfachen Gruppen und Liealgebren vom Typ G2. (German) Nederl. Akad. Wetensch. Proc. Ser. A 74=Indag. Math. 33 (1971), 365-378.
- [X1] T. Xue, Nilpotent orbits in the dual of classical Lie algebras in characteristic 2 and the Springer correspondence. Represent. Theory 13 (2009), 609-635 (electronic).
- [X2] T. Xue, On unipotent and nilpotent pieces for classical groups. J. Algebra 349 (2012), no. 1, 165-184.
- [X3] T. Xue, Combinatorics of the Springer correspondence for classical Lie algebras and their duals in characteristic 2. Adv. Math. 230 (2012) 229-262.
- [X4] T. Xue, Nilpotent elements in the dual of odd orthogonal algebras. Transform. Groups. 17 (2012), no. 2 (2012), 571-592.