Adjoint Jordan blocks for simple algebraic groups of type in characteristic two
Abstract.
Let be a simple algebraic group over an algebraically closed field with Lie algebra . For unipotent elements and nilpotent elements , the Jordan block sizes of and are known in most cases. In the cases that remain, the group is of classical type in bad characteristic, so and is of type , , or .
In this paper, we consider the case where is of type and . As our main result, we determine the Jordan block sizes of and for all unipotent and nilpotent . In the case where is of adjoint type, we will also describe the Jordan block sizes on .
1. Introduction
Let be a simple algebraic group over an algebraically closed field , and denote the Lie algebra of by . Recall that an element is unipotent, if is a unipotent linear map for every rational representation . Similarly is said to be nilpotent, if is a nilpotent linear map for every rational representation .
We denote the adjoint representations of and by and , respectively. In this paper, we will consider the following two problems.
Problem 1.1.
Let be a unipotent element. What is the Jordan normal form of ?
Problem 1.2.
Let be a nilpotent element. What is the Jordan normal form of ?
In most cases, the answer to both questions is known. When is simply connected of exceptional type, the Jordan block sizes were computed in the unipotent case by Lawther [21, 22] and in the nilpotent case by Stewart [33]. In the case where is exceptional of adjoint type, both questions are easily settled using the results of Lawther and Stewart, see [20, Lemma 3.1].
When is of type , solutions to both questions follow from the main results of [16] and [19], see [19, Remark 1.3]. For of type , , or in good characteristic, the Jordan block sizes are described by well-known results on decompositions of tensor products, symmetric squares, and exterior squares — see for example [19, Remark 3.5].
Thus it remains to solve Problem 1.1 and Problem 1.2 in the case where is of type , , or in characteristic . In this paper, we consider the case where is of type . We make the following assumption for the rest of this paper.
Assume that .
Let be a simple algebraic group of type over . As our main result, we determine the Jordan block sizes of and for all unipotent and nilpotent . In type either is simply connected or is adjoint; we deal with these two possibilities as follows.
Let be simply connected and simple of type with Lie algebra , and let be simple of adjoint type with Lie algebra . There exists an isogeny , which induces a bijection between the unipotent variety of and . Similarly the differential induces a bijection between the nilpotent cone of and .
Thus for the solution of Problem 1.1 and Problem 1.2 in type , it will suffice to consider the Jordan block sizes of unipotent on and , and the Jordan block sizes of nilpotent on and .
We can assume that with Lie algebra , where . It is well known (Lemma 8.1) that as -modules. The Jordan block sizes of the action of unipotent on and nilpotent on are described in [18]. Taking the dual does not change the Jordan block sizes, so the Jordan block sizes on are known by previous results.
In our main results, we will describe the Jordan block sizes on in terms of the Jordan block sizes on . To state the result in the unipotent case, we first need describe the classification of unipotent conjugacy classes in , which in characteristic two is due to Hesselink [8]. We discuss this in some more detail in Section 5.
Let be a cyclic group of order , where for some . Then there are a total of indecomposable -modules , , up to isomorphism, where and a generator of acts on with a single unipotent Jordan block. For convenience we will denote . If is a vector space over , we denote , and ( summands) for an integer .
Suppose then that is unipotent, so has order for some . We denote the group algebra of by . Then for some integers , where for all . Equivalently, the Jordan normal form of has Jordan blocks of sizes , , , and a block of size has multiplicity .
For unipotent , we have a decomposition , where are orthogonally indecomposable -modules. Here orthogonally indecomposable means that if as -modules, then or . The orthogonally indecomposable -modules fall into two types: one denoted by (for even) and another denoted by (for ). We define these in Section 5, for now we only mention that and as -modules.
Our first main result is the following. (Below we denote by the -adic valuation on the integers, so is the largest integer such that divides .)
Theorem 1.3.
Let be unipotent, with orthogonal decomposition
Denote by the largest integer such that for all and . Then the following hold:
-
(i)
Suppose that . Then:
-
(a)
If , then as -modules.
-
(b)
If , then
for some -module .
-
(a)
-
(ii)
Suppose that , and let . Then:
-
(a)
If for all and for all , then as -modules.
-
(b)
If for some , or for some , then
for some -module .
-
(a)
To describe our results in the nilpotent case, we recall the classification of nilpotent orbits in , due to Hesselink [8]. For more details we refer to Section 6.
Suppose that for some . Let be the abelian -Lie algebra over generated by a single nilpotent element such that and . There are a total of indecomposable -modules , , up to isomorphism, where and acts on with a single nilpotent Jordan block. Throughout we will denote .
Consider then a nilpotent linear map , and denote the -Lie subalgebra generated by with . Then , where is the smallest power of two such that . We have for some integers , where for all . As in the unipotent case, this amounts to the statement that the Jordan normal form of has Jordan blocks of sizes , , , and a block of size has multiplicity .
For nilpotent , we have an orthogonal decomposition , where the are orthogonally indecomposable -modules. Here orthogonally indecomposable is defined as in the unipotent case. For the orthogonally indecomposable -modules, there are several different types: (for even), (for and ), and (for ). We give the definitions and more details in Section 6, for now we just note that , , and as -modules.
Our main result in the nilpotent case is the following.
Theorem 1.4.
Let be nilpotent, with orthogonal decomposition
Denote by the largest integer such that for all , , and . Then the following statements hold:
-
(i)
Suppose that . Then:
-
(a)
If , then as -modules.
-
(b)
If , then
for some -module .
-
(a)
-
(ii)
Suppose that is not of the form . Then:
-
(a)
If , then as -modules.
-
(b)
If , then
for some -module .
-
(a)
Example 1.5.
In Table 1 and Table 2, we illustrate Theorem 1.3 and Theorem 1.4 in the case where is of type for . In the tables we have also included the Jordan block sizes on , which we prove as an intermediate result in Proposition 9.5 and Proposition 10.5. Furthermore, the values of and that appear in Theorem 1.3 and Theorem 1.4 are included.
In the tables, we use the notation
for and to denote that the Jordan block sizes are , , , and a block of size has multiplicity . For the orthogonal decompositions in the first column, notation such as denotes an orthogonal direct sum ( summands).
In particular, from Theorem 1.3 and Theorem 1.4 we get the number of Jordan blocks of and , which is equal to the dimension of the Lie algebra centralizer of and , respectively.
Corollary 1.6.
Let be unipotent and denote , as in Theorem 1.3. Then the following hold:
-
(i)
Suppose that . Then
-
(ii)
Suppose that . Then
Corollary 1.7.
Let be nilpotent and denote as in Theorem 1.3. Then the following hold:
-
(i)
Suppose that . Then
-
(ii)
Suppose that is not of the form . Then
In the case of regular elements, this amounts to the following. (An element or is regular, if . In the case of , a regular unipotent element is characterized by , and a regular nilpotent element is characterized by .)
Corollary 1.8.
Suppose that is of type in characteristic two (adjoint or simply connected). Let be a regular unipotent element and a regular nilpotent element. Then and .
For the proofs of our main results, our basic approach is as follows. We will first describe the Jordan block sizes of unipotent and nilpotent on , in terms of the Jordan block sizes on . This is attained in Section 9 and Section 10.
It is known that as -modules (Lemma 8.2), so we then have the Jordan block sizes on as well. In Section 11 and Section 12 we will describe the Jordan block sizes of unipotent and nilpotent on , in terms of the Jordan block sizes on . Combining these results, we get the Jordan block sizes of and on , in terms of the Jordan block sizes on (Theorem 1.3, Theorem 1.4).
The other sections of this paper are organized as follows. The notation, terminology, and some preliminary results are stated in Section 2 and Section 3. In Section 4, we state results on Jordan block sizes of unipotent and nilpotent elements on tensor products, symmetric squares, and exterior squares. The classification of unipotent classes in and nilpotent orbits in is described in Section 5 and Section 6, respectively.
In Section 7, we discuss the Chevalley construction of simple algebraic groups, and in particular the action of on . We then make some observations about the structure of and as a -module in Section 8.
As mentioned earlier, in Section 9 and Section 10 we describe the Jordan block sizes of unipotent and nilpotent elements on , in terms of the Jordan block sizes on . In Section 11 and Section 12 we similarly describe the Jordan block sizes on , in terms of Jordan block sizes on . These results allow us to prove our main results, and the proofs of the results stated in this introduction are given in Section 13.
2. Notation and terminology
We will use the following notation and terminology throughout the paper.
2.1. Generalities
We will always assume that is an algebraically closed field of characteristic two.
For a -vector space and integer , we denote , where occurs times. Furthermore, we denote .
We will always use to denote a group, and all the -modules that we consider will be finite-dimensional. Suppose that is the socle filtration of a -module , so for all . Then we will denote this by , and is said to be uniserial if is irreducible for all .
We denote by the -adic valuation on the integers, so is the largest integer such that divides .
2.2. Algebraic groups
Suppose that is a simple linear algebraic group over . In the context of algebraic groups, the notation that we use will be as in [14]. We will denote the Lie algebra of by .
When is an algebraic group, by a -module we will always mean a finite-dimensional rational -module. Fix a maximal torus of with character group , and a base for the root system of , where . We will always use the standard Bourbaki labeling of the simple roots , as given in [11, 11.4, p. 58]. We denote the dominant weights with respect to by , and the fundamental dominant weight corresponding to is denoted by . For a dominant weight , we denote the rational irreducible -module with highest weight by .
An element is unipotent, if is a unipotent linear map for every rational representation . Since , equivalently is unipotent if and only if has order for some .
Similarly is said to be nilpotent, if is a nilpotent linear map for every rational representation . Equivalently is nilpotent if and only if for some , where is the canonical -mapping on .
We will mostly be concerned with the case where , in which case is of type , where . In this case is unipotent if and only if is a unipotent linear map on . Furthermore , and is nilpotent if and only if is a nilpotent linear map on .
2.3. Actions of unipotent elements
Let for some . Then a cyclic group has indecomposable -modules, up to isomorphism. We denote them by , , , where and acts on as a single unipotent Jordan block. For convenience we denote .
Let be an element of a group and let be a finite-dimensional -module on which acts as a unipotent linear map. We denote , which is the fixed point space of on . Note that is the number of Jordan blocks of on .
If is unipotent, we denote by the group algebra of . Then , where are the Jordan block sizes of on , and is the multiplicity of a Jordan block of size .
2.4. Actions of nilpotent elements
Let for some . We denote by the abelian -Lie algebra over generated by a nilpotent element with and . Then as a -vector space
There are a total of indecomposable -modules, up to isomorphism. We denote them by , , , where and acts on as a single nilpotent Jordan block. For convenience we denote .
If is a finite-dimensional module for a Lie algebra and acts on as a nilpotent element, we denote . Then is the number of Jordan blocks of on .
If is nilpotent, we denote by the -Lie algebra generated by . Then , where is the smallest power of two such that . Furthermore , where are the Jordan block sizes of on , and is the multiplicity of a Jordan block of size .
3. Jordan normal forms on subspaces and quotients
In the proofs of our main results, the basic tool that we use to describe Jordan block sizes are the following two lemmas.
Lemma 3.1 ([16, Lemma 3.3]).
Let be a nilpotent linear map. Suppose that is a subspace invariant under such that . Let be such that and . Then
for some -module .
Lemma 3.2 ([17, Lemma 3.4]).
Let be a nilpotent linear map. Suppose that is a subspace invariant under such that . Let be such that and . Then
for some -module .
4. Decomposition of tensor products and symmetric squares
Let for some . In this section, we state some basic results about the decomposition of tensor products, exterior squares, and symmetric squares of -modules and -modules.
4.1. Unipotent case
Let be a generator for the cyclic group . To describe the indecomposable summands of tensor products of -modules, it is clear that it suffices to do so in the case where and are indecomposable.
The decomposition of into indecomposable summands has been studied extensively in all characteristics, see for example (in chronological order) [31], [28], [24], [30], [26], [10], [27], [12], and [1]. Although there is no closed formula in general, there are various recursive formulae which can be used to determine the indecomposable summands efficiently. We are assuming , in which case we have the following result.
Theorem 4.1 ([7, (2.5a)], [6, Lemma 1, Corollary 3]).
Let and suppose that . Then the following statements hold:
-
(i)
If , then as -modules.
-
(ii)
If , then as -modules.
-
(iii)
If , then as -modules, where .
With Theorem 4.1, every tensor product is either described explicitly (case (i)), or in terms of another tensor product with . Thus by repeated applications of Theorem 4.1, we can rapidly find the indecomposable summands of for any given and .
For exterior squares and symmetric squares, similarly it suffices to consider the indecomposable case. This follows from the fact that for all -modules and , we have isomorphisms
as -modules. For the decomposition of and , we have the following results.
Theorem 4.2 ([6, Theorem 2]).
Suppose that . Then we have
as -modules.
Theorem 4.3 ([18, Theorem 1.3]).
Suppose that . Then we have
as -modules.
Similarly to Theorem 4.1, with Theorem 4.2 and Theorem 4.3 we can quickly decompose and for any given .
Lemma 4.4.
Let be an integer. Then the following hold:
-
(i)
.
-
(ii)
.
Proof.
We first consider (i). If , then has dimension and thus (i) holds. Suppose then that and proceed by induction on . Let be a power of two such that . If , it follows from Theorem 4.2 that , so . If , we have
by Theorem 4.2. Then by induction
as claimed by (i).
Next we will prove (ii). If , then so (ii) holds. Suppose then that , and let be a power of two such that . If , then , and thus . Suppose then that . It follows from Theorem 4.3 and (i) that
as claimed by (ii).∎
Lemma 4.5.
Let and define . Then the smallest Jordan block size in is , occurring with multiplicity one.
4.2. Nilpotent case
As in the previous section, to determine the indecomposable summands of tensor products, exterior squares, and symmetric squares of -modules, it suffices to do so in the indecomposable case.
For the indecomposable summands of , it turns out that we get the same decomposition as for in the unipotent case.
Proposition 4.6 ([4, Section III], [25, Corollary 5 (a)]).
Let and suppose that we have as -modules for some . Then as -modules.
Thus we can apply Theorem 4.1 to find the decomposition of into indecomposable summands.
Following [5, p. 231], we call the consecutive-ones binary expansion of an integer the alternating sum such that and is minimal. (Here if .)
The decomposition of into indecomposable summands can be given explicitly in terms of the consecutive-ones binary expansion of [5, Theorem 15]. Such descriptions can also be given for and , by using Theorem 4.2 and Theorem 4.3.
For the decomposition of , , and we have the following result.
Theorem 4.7 ([18, Theorem 1.6, Theorem 1.7, Theorem 3.7]).
Let be an integer, with consecutive-ones binary expansion , where . For , define . Then
as -modules.
5. Unipotent elements in
Consider with . In this section, we will recall the description of unipotent conjugacy classes in , due to Hesselink [8]. For more details, see for example [8], [23, Chapter 4, Chapter 6], or [17, Section 6].
For a group , a bilinear -module is a finite-dimensional -module equipped with a -invariant bilinear form . Two bilinear -modules and are said to be isomorphic, if there exists an isomorphism of -modules which is also an isometry.
Let be unipotent. Let be a power of two such that , so that and are -modules. Then and are conjugate in if and only if as bilinear -modules.
For unipotent, it is clear that we can write , where are orthogonally indecomposable -modules. Here orthogonally indecomposable means that if as -modules, then or . There are two basic types of orthogonally indecomposable -modules, which we can define as follows. (Similar definitions are given in [23, Section 6.1].)
Definition 5.1.
For , we define the module as follows. Let , and suppose that has basis , , with if and otherwise. Define by
Then , and we define as the bilinear -module .
Definition 5.2.
For , we define the module as follows. Let , and suppose that has basis , , with if and otherwise. Define by
Then , and we define as the bilinear -module .
The fact that the modules are isomorphic to those described by Hesselink [8, Proposition 3.5] is seen as follows.
- •
- •
The classification of unipotent conjugacy classes in is based on the following result.
Theorem 5.3 ([8, Proposition 3.5]).
Let be unipotent such that is orthogonally indecomposable. Then is isomorphic to or , where .
By Theorem 5.3, for every unipotent we have an orthogonal decomposition
where for all we have or for some integer .
In general there can be several different ways to decompose into orthogonally indecomposable summands, and even the number of summands is not uniquely determined. This is due to the fact that for even , we have an isomorphism
of bilinear -modules. However, there are normal forms which are uniquely determined, such as the Hesselink normal form [8, 3.7] [17, Theorem 6.4] or the distinguished normal form defined by Liebeck and Seitz in [23, p. 61].
6. Nilpotent elements in
We consider with Lie algebra , where . We recall the classification of nilpotent orbits in due to Hesselink [8]. For more details, we refer to [8] and [23, Chapter 4, Chapter 5].
For a nilpotent element , define the index function corresponding to by
Let be the Jordan block sizes of , and let be the multiplicity of Jordan block size for . By a result of Hesselink [8, Theorem 3.8], the nilpotent orbit of is determined by the integers , , and the function .
Hesselink also proved that it suffices to only consider the values of on the Jordan block sizes , , .
Theorem 6.1 ([8, 3.9]).
Let be nilpotent with index function on . Let be the Jordan block sizes of , and let be the multiplicity of a Jordan block of size for . Then the following statements hold:
-
(i)
The nilpotent orbit of is determined by the symbol .
-
(ii)
and .
-
(iii)
for all .
-
(iv)
if is odd.
Remark 6.2.
Similarly to the unipotent case, we can phrase the classification in terms of bilinear modules. For a Lie algebra , a bilinear -module is a finite-dimensional -module equipped with a -invariant bilinear form , so for all and . Two bilinear -modules and are said to be isomorphic, if there exists an isomorphism of -modules which is also an isometry.
Let be nilpotent. Choose a power of two such that , so that and are -modules. Then and are conjugate under the action of if and only if as bilinear -modules.
For nilpotent , it is clear that we can write , where are orthogonally indecomposable -modules. (Here orthogonally indecomposable is defined similarly to the group case.) By the next lemma, the index function is determined by its restriction to the orthogonally indecomposable summands of .
Lemma 6.3 ([23, Lemma 5.2]).
Let be nilpotent and assume as -modules. Then for all .
The orthogonally indecomposable modules were classified by Hesselink. In the case of , there are three types of orthogonally indecomposable modules, defined as follows. (Similar definitions are given in [23, Section 5.1].)
Definition 6.4.
For , we define the module as follows. Let , and suppose that has basis , , with if and otherwise. Define by
Then , and we define as the bilinear -module .
Definition 6.5.
For , we define the module as follows. Let , and suppose that has basis , , with if and otherwise. Define by
Then , and we define as the bilinear -module .
Definition 6.6.
For and we define the module as follows. Let , and suppose that has basis , , with if and otherwise. Define by
Then , and we define as the bilinear -module .
The fact that the modules in Definition 6.4 – 6.6 agree with those described by Hesselink in [8, Proposition 3.5] is seen as follows.
-
•
In Definition 6.4, this is clear from the fact that , so as defined by Hesselink.
- •
- •
Theorem 6.7 ([8, Proposition 3.5]).
Let be nilpotent such that is orthogonally indecomposable. Then is isomorphic to , , or for some . These modules are characterized by the following properties:
Jordan normal form on | ||
---|---|---|
() |
By Theorem 6.7, for every nilpotent we have an orthogonal decomposition
where for all we have , , or () for some integer .
As in the unipotent case, the orthogonally indecomposable summands and their number is not uniquely determined. For example, by Lemma 6.3 and Theorem 6.1 (i), we have isomorphisms
of bilinear -modules.
We end this section with two observations about orthogonally indecomposable modules of the form .
Lemma 6.8.
Let be nilpotent. Then the following statements are equivalent:
-
(i)
There is an orthogonal decomposition for some integers , , .
-
(ii)
There is a totally singular decomposition , where and are -submodules of .
-
(iii)
for all .
Proof.
(cf. [17, Lemma 6.12]) We prove that (i) (ii) (iii) (i).
(i) (ii): It is clear from Definition 6.5 that each has a decomposition , where and are totally singular -submodules with . Therefore (i) implies (ii).
(ii) (iii): We have for all and since and are -invariant and totally singular. Thus for all and , since .
Lemma 6.9.
Let be nilpotent. Then for some integers , , .
Proof.
We have for all since , so for all . Now the claim follows from Lemma 6.8.∎
7. Chevalley construction
In this section, we will recall basics of the Chevalley construction of Lie algebras and simple algebraic groups, and in particular how it applies for groups of type in characteristic two. For more details, see for example [32] or [11, Chapter VII]. We will also make some preliminary observations about the actions of unipotent and nilpotent elements on the adjoint Lie algebra of type .
7.1. Chevalley construction
Let be a finite-dimensional simple Lie algebra over . Fix a Cartan subalgebra of , and let be the corresponding root system, so
where The Killing form is non-degenerate on , so for all there exists such that for all . We define for all .
It was shown by Chevalley [2, Théorème 1] that one can choose such that the following properties hold:
-
(a)
for all ,
-
(b)
If and , then , where is the largest integer such that is a root.
-
(c)
If and , then .
For a choice of root vectors satisfying (a) – (c) above, we define to be the -span of all and for . Let be a base for and let be the corresponding system of positive roots. Then is a -basis of , called a Chevalley basis for .
Fix a Chevalley basis for . Let be the corresponding Kostant -form, which is the subring of the universal enveloping algebra generated by and for all and .
Let be a faithful finite-dimensional -module over . We denote the set of weights of in by . Then , where is the weight lattice.
A lattice in is the -span of a basis of . We say that a lattice is admissible, if is -invariant.
Let be an algebraically closed field of characteristic , and an admissible lattice in . Set . Then for all , the reduction modulo of is the -linear map defined by , where is the action of on .
In particular, for all and we have a linear map which is the reduction modulo of on . Since acts nilpotently on , we can define for all the root element as the exponential . We have , and the Chevalley group (over ) corresponding to and is defined as
Then is a simple algebraic group over with root system [32, Theorem 6]. Furthermore, we have a maximal torus
where is defined as in [32, Lemma 19, p. 22]. Then the weights of on can be identified with , and the character group can be identified with [32, p. 39].
We say that is simply connected if , and adjoint if .
Lemma 7.1 ([32, Corollary 1, p. 41]).
Suppose that is another faithful finite-dimensional -module with admissible lattice , and denote the corresponding root elements in by . Then:
-
(i)
If is simply connected, there exists a morphism of algebraic groups with for all and .
-
(ii)
If is adjoint, there exists a morphism of algebraic groups with for all and .
The Lie algebra of is identified as follows. The stabilizer of in and is given by [32, Corollary 2, p. 16]
In particular and only depend on the weights in , not the choice of the admissible lattice . Furthermore, under the adjoint action is an admissible lattice in [11, Proposition 27.2].
Then is a -Lie algebra which acts on . The Chevalley algebra
is a Lie algebra over which acts faithfully on , and is precisely the Lie algebra of . The adjoint action of can be realized by applying the Chevalley construction with and , and then taking the morphism of Lemma 7.1 (ii).
Note that then the Lie algebra is a -module for a simply connected Chevalley group with root system (Lemma 7.1 (i)). In all cases, the structure of as a Lie algebra and as a -module is described by Hogeweij in [9]. Here the composition factors of are completely determined by , but the submodule structure depends on the choice of the admissible lattice .
In the simply connected case we have . In the adjoint case we denote , so is the -span of and .
7.2. Chevalley construction for type
We will now setup the Chevalley construction for groups of type over . (Recall that by we always denote an algebraically closed field of characteristic two.)
Let be a -vector space of dimension , with basis , , . We define a non-degenerate alternating bilinear form on by for all , and if .
Let , so is a simple Lie algebra of type . Let be the Cartan subalgebra formed by the diagonal matrices in . Then .
For , define linear maps by where is a diagonal matrix with diagonal entries . Then
is the root system for , and
is a system of positive roots. The set of simple roots corresponding to is , where for and .
For all let be the linear endomorphism on such that and for . Throughout we will use the following Chevalley basis of , which is taken from [13, Section 11, p. 38].
Let be the weight lattice. The maps , , form -basis for , and for , the th fundamental highest weight is equal to .
As in the previous subsection, we denote by be the -span of the Chevalley basis above. Then is the -span of and , where is the diagonal matrix . In terms of the Chevalley basis, we have
7.3. Simply connected groups of type
Let be the -span of the basis , , of . It is clear that is an admissible lattice. Denote . By abuse of notation we denote for all , so , , is a basis of . The alternating bilinear form induces an alternating bilinear form on , with if and otherwise.
Then the Chevalley group corresponding to and is simply connected of type , and it is precisely the symplectic group corresponding to [29, 5]. We denote . Then the Lie algebra of is
We denote by the adjoint Lie algebra of type . For any -module with admissible lattice , we will consider as a -module via the action provided by Lemma 7.1 (i).
7.4. Lie algebra of adjoint type
To compute with the action of on , it will be convenient to use the following well-known identification of with the symmetric square .
Lemma 7.2.
For , define a linear map by . Then we have an isomorphism of -modules, defined by for all .
Proof.
We have an isomorphism of -modules, where is the linear map for all and . Moreover, we have an isomorphism of -modules defined by , where for all .
Now identifying via , the restriction of to is a map defined by . Then is an injective morphism of -modules, and it is clear that for all . Since and have the same dimension, we conclude that defines an isomorphism of -modules.∎
We denote by the isomorphism as in Lemma 7.2. Let with . Then under the map , the root vectors in the Chevalley basis correspond to elements of as follows.
Furthermore, define
Then .
We define
so that and .
Then are admissible lattices in , since and are admissible lattices in .
From the Chevalley construction and as -modules, so induces isomorphisms
(7.1) | ||||
of -modules.
By [9, Lemma 2.2] we have and is spanned by . Similarly by [9, Table 1], we have and is spanned by with . Thus under the isomorphisms in (7.1), we get
(7.2) | ||||
Note that
(7.3) |
so for the linear map induced by the inclusion , we have and .
7.5. Action of unipotent elements on , orthogonally indecomposable case
Let as in the previous sections, and let be a unipotent element.
We will describe how to construct a representative for the conjugacy class of in terms of root elements, and how to compute the action of on . We first do this in the case where is orthogonally indecomposable, so or (Theorem 5.3). In this case it is well known that by replacing with a conjugate, we can take
For , let
Then is the reduction modulo of the action of on . Furthermore, since , the action of on is the reduction modulo of the action of on .
Therefore we define
If , we get
Similarly for , we get
Thus the action of on is exactly as described in Definition 5.1 or Definition 5.2.
Since , for the action on we have
for all . Thus
for all .
7.6. Action of unipotent elements on , general case
For all unipotent elements , in general we can proceed as follows. We have an orthogonal decomposition , where and is an orthogonally indecomposable -module for all . Write for all .
Then , where is the action of on . For all , we have or (Theorem 5.3).
Relabel the basis , , of as
where if and , and otherwise. Then the Cartan subalgebra consists of diagonal matrices of the form
We define then for and the linear map by .
Denote by (resp. ) the -span (resp. -span) of , so we have
Since , we can assume that for all .
We have , and has a root system of type with simple roots , where
Note that is a Chevalley basis of , and the corresponding Kostant -form is a subring of . Then is an admissible lattice for , and applying the Chevalley construction we get .
As in the orthogonally indecomposable case, we define
Then , and the reduction modulo of the action of on is precisely .
Thus we can define
so that is the reduction modulo of the action of on . Furthermore, the action of on is the reduction modulo of the action of on .
7.7. Action of nilpotent elements on
Let be nilpotent. Similarly to the unipotent case, we can find such that is the reduction modulo of the action of on .
We first consider the orthogonally indecomposable case, so , , or for some (Theorem 6.7). In this case we define
Then for the reduction modulo of , the action of is exactly as in Definition 6.4 – 6.6. (Here the expression for is taken from [20, Lemma 3.4].)
In the general case, we have an orthogonal decomposition , where is an orthogonally indecomposable -module, with . We have , where is the action of on . In the notation of the previous section, we take . We can then define
where is defined as in the orthogonally indecomposable case, such that is the reduction modulo of .
Then is the reduction modulo of . Moreover, given any irreducible -module and admissible lattice , the action of on is precisely the reduction modulo of the action of on .
7.8. Elements of acting nilpotently
In this section, we consider which act nilpotently on . Then by reduction modulo , we get an action on , and on . We will describe when and when . For our purposes this is mostly relevant for in the unipotent case, and in the nilpotent case, for integer .
Lemma 7.3.
Let be such that acts nilpotently on , and denote the action of on by . Then if and only if there exists such that .
Proof.
Conversely, suppose that . Then there exists with . Since where is the action of on , we can assume that with . Then
implies that .∎
For actions of unipotent elements, Lemma 7.3 gives the following.
Lemma 7.4.
Let be unipotent, and suppose that is the reduction modulo of . Denote the action of on by . Let be an integer. Then if and only if there exists such that .
Proof.
Follows from Lemma 7.3, with . ∎
Lemma 7.5.
Let be such that acts nilpotently on , and denote the action of on by . Then the following statements are equivalent:
-
(i)
.
-
(ii)
There exists such that .
-
(iii)
There exists such that one of the following holds:
-
(a)
.
-
(b)
.
-
(a)
Proof.
We first prove that (i) and (ii) are equivalent. We identify via (7.1), in which case as -vector spaces. Here is the subspace spanned by with , as seen in (7.2).
If there exists such that , we have
so .
Conversely, suppose that . Then there exists such that . We have where is the action of on , so we can assume that for some . Then
implies that .
Next we show that (ii) and (iii) are equivalent. It follows from (7.3) that contains only two elements, and . Therefore if and only if (iii)(a) or (iii)(b) holds. ∎
Next we make some observations that will allow us to reduce the proofs of our main results to the orthogonally indecomposable case. Continue with the notation as in Section 7.6, so with , and is the Kostant -form for . Define
for all . (Note that .) Furthermore, for we define
Then
(7.4) |
as -modules.
Moreover is an admissible lattice for the action of , and
is an adjoint Lie algebra of type . Here acts on via the Chevalley construction.
Lemma 7.6.
Let , where acts nilpotently on for all . Denote by the linear map acting on , given by reducing the action of on modulo . For , let be the linear map acting on , given by reducing the action of on modulo .
Then the following statements hold.
-
(i)
if and only if there exist , , with such that one of the following holds:
-
(a)
for all .
-
(b)
for all .
-
(a)
-
(ii)
If , then for all .
Proof.
For (i), suppose first that there exist , , with such that (i)(a) or (i)(b) holds. For , we have
Thus if (a) holds, and if (b) holds. Therefore by Lemma 7.5.
Conversely, suppose that . Then by Lemma 7.5, there exists such that or . By (7.4) we can write , where for all and .
Note that and are -invariant. Thus
with for all , and . Since , it follows that . Moreover we have assumed that or , so either (i)(a) or (i)(b) holds.
Claim (ii) follows from (i) and Lemma 7.5. ∎
For the action of unipotent elements, we have the following result.
Lemma 7.7.
Let be unipotent with as in Section 7.6. Denote the action of on by , and denote the action of on by .
Let be an integer. Then the following statements hold.
-
(i)
if and only if there exists , , with such that one of the following holds:
-
(a)
for all .
-
(b)
for all .
-
(a)
-
(ii)
If , then for all .
8. Lie algebras of type as -modules
In the notation of Section 7.3, let be a simply connected simple algebraic group of type , with Lie algebra . In this section, we will make some initial observations about the structure of and as -modules.
Lemma 8.1.
Let with Lie algebra . Then as -modules.
Proof.
Lemma 8.2.
We have as -modules.
Proof.
Let the linear map defined by for all . Since is -invariant, it follows that is a surjective morphism of -modules.
Lemma 8.3.
Let with Lie algebra . Then as -modules.
Proof.
Lemma 8.4.
Let be simply connected and simple of type . Then:
-
(i)
There exists a uniserial -module with .
-
(ii)
A -module as in (i) is unique up to isomorphism.
-
(iii)
Let be a unipotent element. Then
Proof.
Let with , so is a simple algebraic group of adjoint type . Let be an exceptional isogeny as in [32, Theorem 28]. We can embed into a simple algebraic group of type as the stabilizer of a nonsingular vector, see for example [23, Section 6.8]. Here , and as in [23, Section 6.8], we identify , where is a nonsingular vector.
It is straightforward to see that is a uniserial -module with
where is the first fundamental highest weight for . Then the twist of by is a uniserial -module as in (i). For a unipotent element , the fixed point space of on the Frobenius twist is the same as on , because the Frobenius endomorphism preserves unipotent conjugacy classes. Therefore (iii) holds for by [15, Lemma 3.8].
It remains to check that is unique. To this end, note first that
[14, II.2.14] and [3, 5.4]. Here , since has a unique -invariant alternating bilinear form up to a scalar. Thus there exists a unique nonsplit extension
up to isomorphism of -modules.
Since [14, II.4.11] and , we have . Hence there exists a unique nonsplit extension
up to isomorphism of -modules. Every as in (i) is such an extension, so we conclude that is unique up to isomorphism.∎
Lemma 8.5.
Let , so is simply connected and simple of type . Assume that is even. Then there is a short exact sequence
of -modules, where is as in Lemma 8.4 (i).
Proof.
Lemma 8.6.
Let be simply connected and simple of type . Assume that and let be unipotent with . Then .
9. Jordan block sizes of unipotent elements on
Let be simply connected of type , and let be the -invariant alternating bilinear form defining . We have by Lemma 8.1, so for every unipotent element the Jordan block sizes of on are known by the results described in Section 4. In this section, we will describe the Jordan block sizes of unipotent elements on , in terms of Jordan block sizes of on .
Throughout this section, we will denote by the linear map defined by for all . Since is -invariant, it follows that is a surjective morphism of -modules. By Lemma 8.3, the Jordan normal form of a unipotent element on is the same as on .
We will then describe the Jordan block sizes of unipotent elements on . By Lemma 3.1, this amounts to finding the largest integer such that , where is the action of on .
Lemma 9.1.
Let be unipotent. Suppose that we have a orthogonal decomposition as -modules. Denote the action of on by , and the action of on by . Let be an integer.
Then if and only if for all .
Proof.
The orthogonal decomposition into -modules induces a decomposition
into -modules, where . We have for all , from which the lemma follows.∎
With Lemma 9.1, we reduce to the case where is orthogonally indecomposable. We first consider the case where , so . Let , , be a basis as in the definition of (Definition 5.1). Then if , and otherwise. Furthermore, the action of is defined by
We will denote for all and .
Define . We have a short exact sequence
where is the subspace generated by for all . As noted in [17, Section 9], the image of in is fixed by the action of . Thus for some , and more precisely we have the following.
Lemma 9.2.
Let and be as above. Then , and is fixed by the action of .
Proof.
Denote . First we note that
(9.1) |
By another calculation, we get
(9.2) |
Since , it follows that is fixed by the action of . ∎
Lemma 9.3.
Let be unipotent such that . Let be the action of on , and denote . Then the following hold:
-
(i)
.
-
(ii)
.
Proof.
For (ii), we first consider the case where , so is odd. In the notation used before the lemma, by Lemma 9.2 the vector is fixed by the action of . Furthermore , so , as claimed.
Lemma 9.4.
Let be unipotent such that . Let be the action of on , and denote . Then the following hold:
-
(i)
.
-
(ii)
.
Proof.
We have as -modules, where and are totally isotropic subspaces. This induces a decomposition
of -modules, where is the subspace . We have and the smallest Jordan block size in is [16, Lemma 4.2], so (i) follows from Lemma 3.1.
For (ii), we first consider the case where , so is odd. Choose a basis , , of , and let , , be the corresponding dual basis in , so for all . Then is fixed by the action of , see for example [16, Lemma 3.7]. We have , so , as claimed.
Proposition 9.5.
Let be unipotent, with orthogonal decomposition
Denote by the largest integer such that for all and . Then
for some -module .
10. Jordan block sizes of nilpotent elements on
We continue with the setup of the previous section. Let be simply connected of type with Lie algebra . In this section, we will describe the Jordan block sizes of nilpotent elements on (Lemma 8.3). As in the previous section, we describe the Jordan block sizes of on in terms of Jordan block sizes of on . Note that the Jordan normal form of on is known by the results described in Section 4.
We begin by reducing the calculation to the orthogonally indecomposable case. After this we consider the different orthogonally indecomposable -modules in turn, and by combining the results we obtain Proposition 10.5, which is analogous to Proposition 9.5.
Lemma 10.1.
Let be unipotent. Suppose that we have a orthogonal decomposition as -modules. Denote the action of on by , and the action of on by . Let be an integer.
Then if and only if for all .
Proof.
Follows with the same proof as Lemma 9.1.∎
Lemma 10.2.
Let be nilpotent such that . Let be the action of on , and denote . Then the following hold:
-
(i)
.
-
(ii)
.
Proof.
We proceed similarly to the proof of Lemma 9.4. By definition of , we have a totally singular decomposition , where and are -modules on which acts with a single Jordan block of size . This induces a decomposition
of -modules, where . Thus
We have since and are totally singular. Furthermore, the smallest Jordan block size in is by Proposition 4.6 and [16, Lemma 4.2], so (i) follows from Lemma 3.1.
Lemma 10.3.
Let be nilpotent such that , where . Let be the action of on , and denote . Then the following hold:
-
(i)
.
-
(ii)
.
Proof.
Let , , be the basis used in the definition of . Then if and otherwise. Furthermore, the action of is defined by
We will denote for all and .
Let and . Then , where and are -modules on which acts as a single Jordan block of size . We take , , as a basis of , where for all .
For the claims, we will first consider the case where , so is odd. In this case (i) is trivial. For (ii), note that the vector is annihilated by the action of . For all , we have or for some . Therefore , so , as claimed.
Suppose then that . For (i), we will first prove that . To this end, note that the decomposition induces a decomposition
of -modules, where is isomorphic to . It follows from [18, Proof of Theorem 1.6] that and , so
It is clear that , and since the smallest Jordan block size in is (Proposition 4.6 and [16, Lemma 4.2]).
Thus . This also proves (i) in the case where , so suppose next that . We have by Lemma 6.9. Since is even, it follows from Lemma 10.2 and Lemma 10.1 that
(10.1) |
Lemma 10.4.
Let be nilpotent such that . Let be the action of on , and denote . Then the following hold:
-
(i)
.
-
(ii)
.
Proof.
Proposition 10.5.
Let be nilpotent, with orthogonal decomposition
Denote by the largest integer such that for all , , and .
Then
for some -module .
11. Jordan block sizes of unipotent elements on
Let be simply connected and simple of type . In this section, we will describe the Jordan block sizes of unipotent elements acting on . Our approach is to describe the Jordan block sizes of on in terms of the Jordan block sizes on , using Lemma 3.1. Since the Jordan block sizes of on (Lemma 8.2) are known by the results of Section 9, we then get an explicit description of the Jordan block sizes of on .
Throughout this section we will consider as constructed in Section 7.2 – 7.4, and we use the notation established there.
Lemma 11.1.
Let be the reduction modulo of as in Section 7.5, so that . Then the following hold:
-
(i)
.
-
(ii)
There exists such that if and only if is even.
Proof.
For claim (i), consider for from the Chevalley basis of . It is clear that , so for all , from which (i) follows.
Lemma 11.2.
Assume that is odd. Let be the reduction modulo of as in Section 7.5, so that . Then the following hold:
-
(i)
There exists such that .
-
(ii)
There does not exist such that .
Proof.
For (i), first we calculate
Furthermore
so
(11.1) |
Next define
For , we denote . Then modulo , we have
Combining this with (11.1), we get
We conclude then that (i) holds, with .
Lemma 11.3.
Let be unipotent such that , where for all . Then .
Lemma 11.4.
Let be unipotent such that , where is odd for all . Assume that .
Denote the action of on by . Then if and only if is even for all .
Proof.
Lemma 11.5.
Let be unipotent such that . Let be the action of on , and denote . Then the following hold:
-
(i)
.
-
(ii)
.
Proof.
We begin with the proof of (i), which is trivially true when , so suppose that .
We first consider , in which case . Then
by Lemma 8.6 | ||||
by Lemma 8.2 and Proposition 9.5 | ||||
Thus we conclude that It follows then from Lemma 3.1 that , as claimed by (i).
Suppose then that . We have by [17, Lemma 6.13]. It follows then from Lemma 7.7 (ii) and the case that
In particular, we have
(11.2) |
Next we note that it follows from Lemma 8.2, Proposition 9.5, and Lemma 4.5 that the smallest Jordan block size of on is equal to . In particular, there are no Jordan blocks of size for on , which combined with Lemma 3.1 and (11.2) implies .
Proposition 11.6.
Let be unipotent, with orthogonal decomposition . Assume that , and let . Let be the action of on . Then the following statements hold:
-
(i)
.
-
(ii)
if and only if for all , and for all .
-
(iii)
.
Proof.
Next we will prove (ii). In the case where we have odd for all . Then the claim of (ii) is that if and only if is even for all , which is precisely Lemma 11.4.
Suppose then that . Write with and with . Then for the restrictions of the summands in , we have
for all and , by [17, Lemma 3.1, Lemma 6.12, Lemma 6.13]. In the above we denote .
We have for some , so appears as a summand of . Thus by Lemma 11.4, we have if and only if all the summands of the form in have even.
Claim 1: Assume that for some . Then .
It suffices to show that has a summand with odd. If , then has and as a summand. If , then with odd, since .
Claim 2: Assume that for some . Then .
It suffices to show that has a summand with odd. If , then has and as a summand. Suppose then that . Then , with is odd since .
The “only if” part of (ii) follows from Claim 1 and Claim 2. Conversely, if for all and for all , we have
Thus by Lemma 11.4.
For (iii), let be as in Section 7.5, so that is the reduction modulo of the action of on . Denote , , as in Section 7.5. It follows from Lemma 7.7 (i), Lemma 11.1, and Lemma 11.5 that there exist such that
Then
On the other hand, we have for some , as seen in the beginning of the proof of Lemma 11.2. Thus for all , and consequently
It follows then from Lemma 7.7 that , as claimed by (iii).∎
Proposition 11.7.
Let be unipotent, with orthogonal decomposition . Then the following statements hold:
-
(i)
Suppose that . Then
as -modules.
-
(ii)
Suppose that , and let . Then:
-
(a)
If for all and for all , then
for some -module .
-
(b)
If for some , or for some , then
for some -module .
-
(a)
12. Jordan block sizes of nilpotent elements on
Continuing with the setup of the previous section, in this section we describe the Jordan block sizes of nilpotent on , in terms of the Jordan block sizes on . The basic approach is similar to the previous section, but the proofs will be more simple due to the fact that is always has an orthogonal decomposition of the form (Lemma 6.9).
Lemma 12.1.
Let be nilpotent such that , and let be the action of on . Then .
Proof.
Lemma 12.2.
Let be nilpotent with or for some . Let be the action of on . Then .
Proof.
Suppose that . Then it follows from Lemma 7.5 that there exists such that . We have so
(12.1) |
for some .
Note that
for all . Therefore . Since also , it follows from (12.1) that , which is impossible. We have a contradiction, so we conclude that .∎
Lemma 12.3.
Let be nilpotent, and let be the action of on . Then , except possibly when for some , , .
Proof.
Write , where is orthogonally indecomposable for all , and with . Let be as in Section 7.7, so that is the reduction modulo of .
Lemma 12.4.
Let be nilpotent, and let be the action of on . Then .
Proposition 12.5.
Let be nilpotent. Then the following hold:
-
(i)
If , then
as -modules.
-
(ii)
If is not of the form , then
for some -module .
13. Proofs of main results
We can now prove the main results stated in the introduction; all of them are straightforward consequences of the results from previous sections.
Proof of Theorem 1.3.
Proof of Theorem 1.4.
References
- [1] M. J. J. Barry. Decomposing tensor products and exterior and symmetric squares. J. Group Theory, 14(1):59–82, 2011.
- [2] C. Chevalley. Sur certains groupes simples. Tohoku Math. J. (2), 7:14–66, 1955.
- [3] M. F. Dowd and P. Sin. On representations of algebraic groups in characteristic two. Comm. Algebra, 24(8):2597–2686, 1996.
- [4] R. M. Fossum. One-dimensional formal group actions. In Invariant theory (Denton, TX, 1986), volume 88 of Contemp. Math., pages 227–397. Amer. Math. Soc., Providence, RI, 1989.
- [5] S. P. Glasby, C. E. Praeger, and B. Xia. Decomposing modular tensor products: ‘Jordan partitions’, their parts and -parts. Israel J. Math., 209(1):215–233, 2015.
- [6] R. Gow and T. J. Laffey. On the decomposition of the exterior square of an indecomposable module of a cyclic -group. J. Group Theory, 9(5):659–672, 2006.
- [7] J. A. Green. The modular representation algebra of a finite group. Illinois J. Math., 6:607–619, 1962.
- [8] W. H. Hesselink. Nilpotency in classical groups over a field of characteristic . Math. Z., 166(2):165–181, 1979.
- [9] G. M. D. Hogeweij. Almost-classical Lie algebras. I, II. Nederl. Akad. Wetensch. Indag. Math., 44(4):441–452, 453–460, 1982.
- [10] X. D. Hou. Elementary divisors of tensor products and -ranks of binomial matrices. Linear Algebra Appl., 374:255–274, 2003.
- [11] J. E. Humphreys. Introduction to Lie algebras and representation theory. Springer-Verlag, New York-Berlin, 1972. Graduate Texts in Mathematics, Vol. 9.
- [12] K.-i. Iima and R. Iwamatsu. On the Jordan decomposition of tensored matrices of Jordan canonical forms. Math. J. Okayama Univ., 51:133–148, 2009.
- [13] J. C. Jantzen. Darstellungen halbeinfacher algebraischer Gruppen und zugeordnete kontravariante Formen. Bonn. Math. Schr., (67):v+124, 1973.
- [14] J. C. Jantzen. Representations of algebraic groups, volume 107 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, second edition, 2003.
- [15] M. Korhonen. Unipotent elements forcing irreducibility in linear algebraic groups. J. Group Theory, 21(3):365–396, 2018.
- [16] M. Korhonen. Jordan blocks of unipotent elements in some irreducible representations of classical groups in good characteristic. Proc. Amer. Math. Soc., 147(10):4205–4219, 2019.
- [17] M. Korhonen. Hesselink normal forms of unipotent elements in some representations of classical groups in characteristic two. J. Algebra, 559:268–319, 2020.
- [18] M. Korhonen. Decomposition of exterior and symmetric squares in characteristic two. Linear Algebra Appl., 624:349–363, 2021.
- [19] M. Korhonen. Jordan blocks of nilpotent elements in some irreducible representations of classical groups in good characteristic. J. Pure Appl. Algebra, 225(8):Paper No. 106694, 10, 2021.
- [20] M. Korhonen, D. I. Stewart, and A. R. Thomas. Representatives for unipotent classes and nilpotent orbits. Comm. Algebra, 50(4):1641–1661, 2022.
- [21] R. Lawther. Jordan block sizes of unipotent elements in exceptional algebraic groups. Comm. Algebra, 23(11):4125–4156, 1995.
- [22] R. Lawther. Correction to: “Jordan block sizes of unipotent elements in exceptional algebraic groups”. Comm. Algebra, 26(8):2709, 1998.
- [23] M. W. Liebeck and G. M. Seitz. Unipotent and nilpotent classes in simple algebraic groups and Lie algebras, volume 180 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2012.
- [24] J. D. McFall. How to compute the elementary divisors of the tensor product of two matrices. Linear and Multilinear Algebra, 7(3):193–201, 1979.
- [25] C. W. Norman. On Jordan bases for two related linear mappings. J. Math. Anal. Appl., 175(1):96–104, 1993.
- [26] C. W. Norman. On the Jordan form of the tensor product over fields of prime characteristic. Linear and Multilinear Algebra, 38(4):351–371, 1995.
- [27] C. W. Norman. On Jordan bases for the tensor product and Kronecker sum and their elementary divisors over fields of prime characteristic. Linear Multilinear Algebra, 56(4):415–451, 2008.
- [28] T. Ralley. Decomposition of products of modular representations. Bull. Amer. Math. Soc., 72:1012–1013, 1966.
- [29] R. Ree. On some simple groups defined by C. Chevalley. Trans. Amer. Math. Soc., 84:392–400, 1957.
- [30] J.-C. Renaud. The decomposition of products in the modular representation ring of a cyclic group of prime power order. J. Algebra, 58(1):1–11, 1979.
- [31] B. Srinivasan. The modular representation ring of a cyclic -group. Proc. London Math. Soc. (3), 14:677–688, 1964.
- [32] R. Steinberg. Lectures on Chevalley groups, volume 66 of University Lecture Series. American Mathematical Society, Providence, RI, 2016.
- [33] D. I. Stewart. On the minimal modules for exceptional Lie algebras: Jordan blocks and stabilizers. LMS J. Comput. Math., 19(1):235–258, 2016. Corrected version in arXiv:1508.02918.