Overgroups of regular unipotent elements in simple algebraic groups

By Gunter Malle and Donna M. Testerman

Abstract

We investigate positive-dimensional closed reductive subgroups of almost simple algebraic groups containing a regular unipotent element. Our main result states that such subgroups do not lie inside proper parabolic subgroups unless possibly when their connected component is a torus. This extends the earlier result of Testerman and Zalesski treating connected reductive subgroups.

1. Introduction

Let be a simple linear algebraic group defined over an algebraically closed field. The regular unipotent elements of are those whose centraliser has minimal possible dimension (the rank of ) and these form a single conjugacy class which is dense in the variety of unipotent elements of . The main result of our paper is a contribution to the study of positive-dimensional subgroups of which meet the class of regular unipotent elements. Since any parabolic subgroup must contain representatives from every unipotent conjugacy class, the question arises only for reductive, not necessarily connected subgroups, where we establish the following:

Theorem 1.

Let be a simple linear algebraic group over an algebraically closed field, a closed reductive subgroup containing a regular unipotent element of . If , then lies in no proper parabolic subgroup of .

In addition, we show that for many simple groups , there exists a closed reductive subgroup with a torus and such that meets the class of regular unipotent elements of . (See Proposition 7.2 and Examples 7.7, 7.11.) Finally, we go on to consider subgroups of non-simple almost simple algebraic groups where there is a well-defined notion of regular unipotent elements in unipotent cosets of . We establish the corresponding result in this setting; see Corollary 6.2.

The investigation of the possible overgroups of regular unipotent elements in simple linear algebraic groups has a long history. The maximal closed positive-dimensional reductive subgroups of which meet the class of regular unipotent elements were classified by Saxl and Seitz Reference 17 in 1997. In earlier work, see Reference 21, Thm 1.9, Suprunenko obtained a particular case of their result. In order to derive from the Saxl–Seitz classification an inductive description of all closed positive-dimensional reductive subgroups containing regular unipotent elements, one needs to exclude that any of these can lie in proper parabolic subgroups. For connected this was shown by Testerman and Zalesski in Reference 22, Thm 1.2 in 2013. They then went on to determine all connected reductive subgroups of simple algebraic groups which meet the class of regular unipotent elements. Our result generalises Reference 22, Thm 1.2 to the disconnected case and thus makes the inductive approach possible. It is worth pointing out that the analogous result is no longer true even for simple subgroups once one relaxes the condition of positive-dimensionality. For example, there exist reducible indecomposable representations of the group whose image in the corresponding contains a matrix with a single Jordan block, i.e., the image meets the class of regular unipotent elements in . In Reference 3, Burness and Testerman consider -subgroups of exceptional type simple algebraic groups which meet the class of regular unipotent elements and show that with the exception of two precise configurations, such a subgroup does not lie in a proper parabolic subgroup of (see Reference 3, Thms 1 and 2).

Our proof of Theorem 1 relies on the result of Testerman–Zalesski Reference 22 in the connected case, which actually implies our theorem in characteristic 0 (see Remark 2.1) as well as on results of Saxl–Seitz Reference 17 classifying almost simple irreducible and tensor indecomposable subgroups of classical groups containing regular unipotent elements and maximal reductive subgroups in exceptional groups with this property. For the exceptional groups we also use information on centralisers of unipotent elements and detailed knowledge of Jordan block sizes of unipotent elements acting on small modules, as found in Lawther Reference 6. For establishing the existence of positive-dimensional reductive subgroups , with a torus, and meeting the class of regular unipotent elements, we produce subgroups which centralise a non-trivial unipotent element and hence necessarily lie in a proper parabolic subgroup of . (See Reference 15, Thm 17.10, Cor. 17.15.)

After collecting some useful preliminary results we deal with the case of in Section 3, with the orthogonal case in Section 4, and with the simple groups of exceptional type in Section 5. The case of almost simple groups is deduced from the connected case in Corollary 6.2. Finally, in Section 7 we discuss the case when is a torus.

2. Preliminary results

In this paper we consider almost simple algebraic groups defined over an algebraically closed field of characteristic and investigate closed positive-dimensional subgroups that contain a regular unipotent element. For us, throughout “algebraic group” will mean “linear algebraic group”, and all vector spaces will be finite-dimensional vector spaces over . An algebraic group is called an almost simple algebraic group if is simple and embeds into . Thus, is an extension of by a subgroup of its group of graph automorphisms (see, e.g., Reference 15, Thm 11.11). As a matter of convention, a “reductive subgroup” of an algebraic group will always mean a closed subgroup whose unipotent radical is trivial. In particular, a reductive group may be disconnected. For an algebraic group , we write to denote the unipotent radical of . Throughout, all -modules are rational, as are all extensions, and cohomology groups are those associated to rational cocycles.

Let us point out that for the question treated here, the precise isogeny type of the ambient simple algebraic group will not matter, as isogenies preserve parabolic subgroups as well as regular unipotent elements. (If is almost simple and does not divide the order of the fundamental group of , the natural map induces an isogeny of onto its adjoint quotient, preserving regular unipotent elements in ; in the general case, a reduction to of adjoint type is given in Reference 18, I.1.7.) In particular, for a classical type simple algebraic group we will argue for the groups , and , and for the groups of type and defined over of characteristic , we may choose to work with whichever group is more convenient under the given circumstances.

We start by making two useful observations which will simplify the later analysis.

Remark 2.1.

In the situation of Theorem 1, assume that . As is finite, some power of a regular unipotent element will lie in . In characteristic 0 any power of a regular unipotent element is again regular unipotent, so here we are thus reduced to studying the connected reductive subgroup satisfying the same assumptions. In that case, the conclusion of Theorem 1 was established in Reference 22, Thm 1.2. Hence, in proving Theorem 1 we may assume whenever convenient. Furthermore, we will assume without loss of generality that .

Remark 2.2.

Let be a reductive subgroup of a connected reductive group such that is regular unipotent in and . Let be one of the simple components of and set . Then , so acts transitively on the set of simple components of , and if lies in a proper parabolic subgroup of , then so does . Thus, when proving Theorem 1 we may as well assume that the simple components of are permuted transitively by .

2.1. Jordan forms and tensor products

The following elementary fact will be used throughout (see also Reference 17, Lemma 1.3(i)):

Lemma 2.3.

Assume that and let be unipotent with a single Jordan block. Write with . Then has Jordan blocks, of size  and the other of size .

Lemma 2.4.

Let be a unipotent element with a single Jordan block of size , or with two Jordan blocks of sizes or . If preserves the factors in a non-trivial tensor product decomposition of then and has two Jordan blocks on . If these are of sizes .

Proof.

Using the description of Jordan block sizes of unipotent elements in tensor products given in Reference 17, Lemma 1.5 we see that necessarily and either has Jordan block sizes , or and has Jordan block sizes , as claimed.

Before establishing a useful consequence of Lemma 2.4, we recall the following well-known Clifford-theoretic result, see, e.g., Reference 2, Prop. 2.6.2:

Lemma 2.5.

Let be groups with finite cyclic and be a finite-dimensional irreducible -module. Then , where is any irreducible -submodule of and runs over a system of coset representatives of the stabiliser of in . Moreover, the are pairwise non-isomorphic -modules.

We now show the desired corollary of Lemma 2.4:

Lemma 2.6.

Let be connected reductive with non-trivial derived subgroup and assume that is completely reducible and homogeneous. If is normalised by a unipotent element with a single Jordan block of size , or with two Jordan blocks of sizes or then either is an irreducible -module, or , preserves a non-trivial tensor product decomposition of , and has two Jordan blocks on , of sizes if .

Proof.

Let be the unipotent element normalising as in the assumption. Let be an irreducible -submodule of . As is homogeneous as an -module, Lemma 2.5 shows that is irreducible. Since , we have . Then with we have as an -module, and this decomposition is stabilised by (see e.g. Reference 15, Prop. 18.1). Applying Lemma 2.4, this implies that either , whence is irreducible for , or we are in the exceptional case of that result, as in the conclusion.

The proof of the next result is modelled after the proof of Reference 17, Prop. 2.1 which treats a more special situation:

Lemma 2.7.

Assume and let be a reductive subgroup and a unipotent element with a single Jordan block of size , or with two Jordan blocks of sizes or . If acts irreducibly on , then either is simple, or one of the following holds:

(1)

, preserves a non-trivial tensor decomposition of  and has two Jordan blocks on . If these are of size ;

(2)

, as an -module with , , permutes both sets of factors transitively and has a single Jordan block on . Moreover, has a single Jordan block on each ; or

(3)

, as an -module with , has Jordan blocks of sizes on and has a single Jordan block on each .

Here, in and , does not preserve the stated tensor product decomposition of .

Proof.

Note that as and acts irreducibly. Write with simple algebraic groups , so with non-trivial irreducible -modules . Now permutes the factors and their corresponding tensor factors . Assume that . If has at least two orbits on the set of , this yields a corresponding -invariant tensor decomposition of . By Lemma 2.4 we reach case (1).

Henceforth, we may assume that permutes the , and thus the , transitively. In particular all have the same dimension , that is, , and is a power of . Let be minimal with . Now, stabilises all , so is a tensor product of matrices of size and thus of order at most . Hence divides . On the other hand,

The above conditions imply that either , and , or , and .

In the first case, , , and as all simple factors of must have type , as in (2). The statement about the Jordan form of follows from Lemma 2.3.

In the second case we have , and our inequalities force that and hence has Jordan blocks of sizes or and by Lemma 2.3, has Jordan blocks of sizes , respectively . The latter cannot arise as the block sizes of a tensor product of two unipotent matrices by Reference 17, Lemma 1.5, so we are in the former case and has a single Jordan block on each , as in (3).

2.2. On subgroups containing regular unipotent elements

For connected groups, the following result from Reference 22, Lemma 2.6 will be useful:

Lemma 2.8.

Let be connected reductive, a parabolic subgroup with Levi complement and assume that is regular unipotent in . Then the image of is regular unipotent in and hence in each simple factor of .

Lemma 2.9.

Let be simple and be a connected reductive subgroup normalised by a regular unipotent element of . Assume that is a central product with such that acts by an inner automorphism on . Then contains a regular unipotent element of .

Proof.

By assumption, acts as an inner automorphism on , say by . Thus, and are contained in and so . But then Reference 22, Prop. 2.3 implies that . Now is regular unipotent. Replacing and by their unipotent parts respectively, we may assume both to be unipotent and lying in a common Borel subgroup of . As centralises and thus isn’t regular, must be regular by Reference 22, Lemma 2.4.

Lemma 2.10.

Let be simple and a connected reductive subgroup containing a regular unipotent element of . Then is regular in .

Proof.

Let be a Borel subgroup of containing . Assume is not regular unipotent in . By Reference 19, Ch. III, 1.13 it may be written as a product of root elements

where the first product runs over a proper subset of the simple roots of the root system of with respect to the pair where is some maximal torus, and the second one over the roots in of height at least 2. Thus, lies in the unipotent radical of the parabolic subgroup of whose Levi factor is generated by the root subgroups for the simple roots not occurring in the representation of and their negatives, which thus is not a torus.

By Borel–Tits, then also lies in the unipotent radical of a proper parabolic subgroup of with non-toral Levi factor. But then, when writing as a product of root elements for with respect to a Borel subgroup contained in , not all simple roots can occur, whence is not regular in . This contradiction achieves the proof.

We will make frequent use of the following result, the second part of which was essentially shown by Saxl and Seitz Reference 17, Prop. 2.2:

Proposition 2.11.

Let be a reductive subgroup of the simple classical group , or with simple and irreducible on , where when is of orthogonal type. If contains a regular unipotent element of , then acts tensor indecomposably on .

Furthermore, either , or and the highest weight of on are as in Table 1 (up to Frobenius twists and taking duals) and has a single Jordan block.

Proof.

Assume that for non-trivial irreducible -modules . If , then Lemma 2.4 gives that and , but this is not simple. If some power acts as an inner element on , then centralises , hence, as acts irreducibly, we must have that equals the unipotent part of and so lies in . So now we may assume that and hence (), () or , with , or and . Unipotent elements in have order at most 32, but there is no faithful representation of of dimension less than 54 (the 27-dimensional modules for are not invariant under the graph automorphism). Thus is of classical type. Let denote the dimension of its natural module. Then, e.g., by Reference 12, Tab. 2 we have , so , but unipotent elements of have order less than . Thus . Note that with cannot occur, as here unipotent elements have order at most  (see Lemma 2.12). This only leaves the possibility , , and . But there is no 9-dimensional irreducible orthogonal module in characteristic 2, so in fact must be tensor indecomposable for . The remaining assertions are now shown in Reference 17, Prop. 2.2.

We will obtain a similar classification for unipotent elements in with a Jordan block of size  when in Proposition 4.1.

2.3. Jordan forms and orders of regular unipotent elements

While the notion of regular unipotent element is well known for connected reductive groups, this is much less so for non-connected reductive groups. Still, similar results hold.

Let be a not necessarily connected reductive algebraic group and let be unipotent. Spaltenstein Reference 18, p. 41 and II.10.1 has shown (generalising a result of Steinberg in the connected case) that the coset of the connected component contains a unipotent -conjugacy class that is dense in the variety of unipotent elements of , called the class of regular unipotent elements of . Since this variety is irreducible Reference 18, Cor. I.1.6, is also the unique class of unipotent elements in of maximal dimension.

We now describe the Jordan block structure of regular unipotent elements in classical type almost simple groups on their natural representation. For us, the natural representation for the extension of , , by its graph automorphism of order 2 is defined by its embedding into the stabiliser in of a pair of complementary totally singular subspaces, with acting in its natural representation, respectively its dual, on these subspaces. We do not consider or as being of classical type.

Lemma 2.12.

Let be almost simple of classical type with regular unipotent in . Then in the natural representation of :

(a)

has a single Jordan block for , (when ), , and for when ;

(b)

has two Jordan blocks of sizes when for ;

(c)

has two Jordan blocks of sizes when , respectively of sizes when for ;

(d)

has a single Jordan block of size when for with odd; and

(e)

has two Jordan blocks of sizes when for with even.

Proof.

Only (d) and (e) are not shown in Reference 17, Lemma 1.2. Spaltenstein Reference 18, I.2.7, I.2.8(c) gives a description of the unipotent classes in in terms of the Jordan normal form of the square of the elements on the natural -module, and a formula for the centraliser dimension. From this it can be seen that elements with minimal centraliser dimension are those for which has one Jordan block of size if is odd, and two blocks of sizes if is even. Thus, in the natural -dimensional orthogonal representation of , the element has two Jordan blocks of size , respectively four of sizes . Given the possible Jordan block shapes of unipotent elements in in its natural representation Reference 18, I.2.6, the claim for follows with Lemma 2.3.

Since the natural representations are faithful, the above result also allows one to read off the orders of regular unipotent elements of almost simple classical groups.

2.4. Some results on extensions

We conclude this preparatory section by collecting some basic properties on Ext-groups. We state the following well-known result for future reference (see Reference 24, Prop. 3.3.4).

Lemma 2.13.

Let be a group, and with , , be -modules. Then

We thank Jacques Thévenaz for pointing out the following result:

Lemma 2.14.

Let be a field, be groups and two finite-dimensional -modules on which acts trivially. Assume that . Then

Proof.

We use the (exact) inflation-restriction sequence for cohomology (see Reference 24, 6.8.3) for a -module :

which by Reference 24, 6.1.2 can be interpreted as the -sequence

Applying this with and using (see Reference 1, Cor. 1) the previous sequence becomes

As acts trivially on and , the first term equals , while the third is by our hypothesis on and Lemma 2.13, whence exactness of the sequence implies our claim.

Lemma 2.15.
(a)

Let be a semisimple algebraic group. Then there are no non-trivial self-extensions between irreducible -modules.

(b)

Let be a semisimple group acting on , with acting trivially on . Then

Proof.

Part (a) is Reference 4, II.2.12(1). In (b) by Lemma 2.13 we have

As acts trivially on , the first summand is isomorphic to

by two applications of Lemma 2.14, and similarly for the second summand.

3. The case of

In this section we prove Theorem 1 for those classical type simple algebraic groups for which regular unipotent elements have a single Jordan block on their natural module (see Lemma 2.12). We are in the following situation: is a (not necessarily connected) reductive subgroup of the form for a regular unipotent element of . We also assume that is not a torus; this case will be considered in Section 7.1. We will show that cannot be contained in a proper parabolic subgroup of , that is, acts irreducibly on . For this, we may whenever convenient, assume that is semisimple, since if is not contained in a proper parabolic subgroup of then neither is . As has a single Jordan block on , if are -invariant subspaces, then has a single Jordan block on .

3.1. The completely reducible case

We first deal with the case when is completely reducible for .

Lemma 3.1.

Let be a reductive subgroup of the form for a regular unipotent element of , such that is not a torus. Assume that is completely reducible. Then is an irreducible -module.

Proof.

Decompose into its homogeneous components . Then acts on the set of components, transitively since otherwise we obtain a -invariant decomposition of , contradicting regularity of . Thus stabilises each component , acting as a single Jordan block on it by Lemma 2.3, and is homogeneous as an -module. Since is not a torus, Lemma 2.6 implies that is an irreducible -module. As the are permuted transitively by , the module is irreducible for .

In view of Lemma 3.1, it seems interesting to determine the structure of irreducible subgroups containing regular unipotent elements.

Proposition 3.2.

Assume . Let be a reductive subgroup of the form for a regular unipotent element of , such that is semisimple. Assume that acts irreducibly on . Then

for submodules , transitively permuted by , on which acts with a single Jordan block; and

for distinct semisimple normal subgroups transitively permuted by , with , where acts trivially on for , and in addition one of the following holds:

(1)

the are isomorphic simple algebraic groups and is an irreducible tensor indecomposable -module, for ; if then , for some , and ; or

(2)

, , for all , with permuting the factors transitively, and each , as a -module, is a tensor product of irreducible -dimensional -modules.

Proof.

As is irreducible for , by Lemma 2.5, is a direct sum of non-isomorphic irreducible -modules permuted transitively by , with for all . Thus stabilises each and acts with a single Jordan block on it by Lemma 2.3. The semisimple group is a product of simple normal subgroups. Let denote the product of those simple factors of acting non-trivially on . By Lemma 2.7 applied to acting on , either is simple, and then it acts tensor indecomposably on by Proposition 2.11; or , , permutes the factors transitively, and is a tensor product of irreducible 2-dimensional -modules.

Now acts non-trivially on for . Thus is generated by, and hence equal to the product of the , which are permuted transitively by . Since each is an orbit of on the set of simple factors of , these sets are mutually disjoint. Let be the smallest power of stabilising . Then and .

Assume is simple and . As normalises but does not stabilise , it must act as an outer automorphism of the simple group . So and moreover, acts as an inner automorphism on and stabilises the -module . So , and as this implies that . By Proposition 2.11 the group acting on occurs in Table 1. If , so , there is no such entry. So we have . The case on the natural module does not occur here, as is not -stable. So with , and while stabilises all , as in (1).

Finally, assume that and . Then is normalised by , and permutes its factors transitively, where . This forces , as in (2).

Remark 3.3.

In the situation of Proposition 3.2, for set if , and if . Then is an irreducible -module and acts trivially on for . So is the trivial -homogeneous component of and the decomposition is stabilised by . Moreover, acts by a single Jordan block on each .

3.2. The not completely reducible case

Lemma 3.4.

Let be a reductive subgroup of the form for a regular unipotent element of , with . Then is an irreducible -module. That is, the socle series of as -module is a composition series as -module.

Proof.

By definition, is a completely reducible -module, on which acts as a single Jordan block since that property passes to quotients and submodules. So by Lemma 3.1, it is an irreducible -module. The claim now follows by induction on the length of the socle series of .

Lemma 3.5.

Let be groups with . Assume that has order and acts by an inner automorphism on . If does not contain elements of order then for some element of infinite order or of order a multiple of .

Proof.

By assumption there is such that centralises . As centralises , either both and are of infinite order, or both have finite order and then the order of divides the least common multiple of and . In the latter case, is not divisible by by assumption, which implies that divides . Clearly, , whence our claim.

Lemma 3.6.

Let be a group, and assume that all composition factors of are mutually non-isomorphic. Then contains no non-trivial unipotent elements.

Proof.

Let be unipotent. Set , so acts as 0 on any irreducible -subquotient of . Let and set . If then , so . Let be an irreducible -submodule. Then where is the full preimage of in . But then

is a simple submodule of isomorphic to , contradicting our assumption. Thus and so .

We now use results of McNinch on semisimplicity of low-dimensional modules in order to study extensions of irreducible modules for simple algebraic groups on which a full Jordan block acts.

Proposition 3.7.

Let with a simple algebraic group and regular unipotent in . If has at most two -composition factors then acts irreducibly on .

Proof.

By Reference 22, Prop. 2.11 we may assume that is not connected. We argue by contradiction and so assume has two -composition factors , . Then by Lemma 3.4 these are the socle and the head of as an -module. Applying Proposition 3.2 to and invoking the hypothesis that is simple, we find that 3.2(1) holds with . In particular, for each we have with irreducible -modules where , or and . Note that for any there must be a non-trivial extension between and some .

First assume that , so is an irreducible -module for . Since is simple, must act as an inner automorphism on . We claim that the order of in its action on is bounded above by the order of a unipotent element of . Indeed, if the order of is larger than that, then by Lemma 3.5 there is a non-trivial central unipotent element in . On the other hand by Lemma 2.15(a) the two composition factors are not isomorphic as -modules. This contradicts Lemma 3.6.

Now acts by a single Jordan block on both -composition factors, and both are tensor indecomposable for by Proposition 2.11. Thus each of these two -modules is either trivial or occurs in Table 1.

By a result of McNinch Reference 16, Thm 1 the possibilities for non-split extensions between such -modules are very restricted. Namely, he shows that all -modules of dimension below a certain explicit bound are either completely reducible or contained on a short list of exceptions in Reference 16, Tab. 5.1.1. Based on his dimension bounds one sees that there is no non-trivial extension between a module in Table 1 and the trivial -module with disconnected.

Hence and are non-trivial. Then again combining Table 1 with Reference 16, Thm 1 we only arrive at the case that acts as on , with , . Here, has dimension 16, so has order 16 so this case cannot occur by our above considerations.

Next consider the case that . Then we have , with and . Then cannot be trivial as argued in the preceding case. So if , then is an irreducible - and -module on which acts with a single Jordan block. By Table 1 this implies that and . Recall that here . A regular unipotent element of has order 16. But unipotent elements of have order at most 8. Arguing as in paragraph 2 using Lemmas 3.5 and 3.6 we again reach a contradiction.

If , then as well. Now unipotent elements of have order less than . So has bigger order than any unipotent element of . Arguing as before this shows with Lemma 3.6 that (after possibly renumbering), and so also . By untwisting we may assume that is the natural module. But there is no non-trivial extension between the natural -module and itself or its dual. This final contradiction completes the proof.

The extension question for the exceptional modules showing up in Proposition 3.2(2) can be discussed in a similar manner:

Proposition 3.8.

Let with and regular unipotent in . If has at most two -composition factors, one of which is the -fold tensor product of isomorphic -dimensional -modules, then is an irreducible -module.

Proof.

Assume that has two -composition factors and , on one of which acts non-trivially. By passing to the dual, we may assume that this is . First consider . Then and , so the regular unipotent element has order at least . But it normalises an , so by Lemma 3.5 that is centralised by an element of order 4. Then Lemma 3.6 shows that the two -composition factors must be isomorphic. As there are no non-trivial self-extensions for by Lemma 2.15(a), is in fact completely reducible as an -module, and hence an irreducible -module by Lemma 3.4, contrary to our assumption.

When then note that there cannot be a non-trivial extension between the -module and the trivial module, since the zero weight is not subdominant to the highest weight of . Thus we may assume that also acts by a non-trivial semisimple group on . Since the smallest dimension of a faithful -module is 6, , and so has order at least 27. We can now argue exactly as in the case , with an element of order 9 centralising .

We can now prove the main result of this section, establishing Theorem 1 for the classical algebraic groups of type , and . Recall that it suffices to consider any group isogenous to .

Theorem 3.9.

Let be simple of type or and be a reductive subgroup of the form for a regular unipotent element of with . Then does not lie in any proper parabolic subgroup of .

Proof.

As noted in Remark 2.1 we may assume . If then we need not consider as it is isogenous to . Now embed via its natural representation. Then by Lemma 2.12, has a single Jordan block on . If lies in some proper parabolic subgroup of , with unipotent radical , then , which by Borel–Tits lies in a proper parabolic subgroup of . Thus it is sufficient to establish the result in the case that . Here the claim is equivalent to showing that is irreducible on . For this we may replace by and therefore assume that is a non-trivial semisimple group. Furthermore, by Remark 2.2 we may also assume that is transitive on the set of simple components of by replacing the latter by one of the -orbits. In particular, any component of acts non-trivially on any -composition factor of on which acts non-trivially.

For a contradiction, assume is reducible on . By passing to a suitable quotient of , where still acts by a single Jordan block, we may assume that has exactly two composition factors on . By Lemma 3.4, any composition series for is a socle series for , which thus also has two layers.

As there is at least one term in the socle series on which does not act trivially (which we may assume to be the top layer , by going to the dual if necessary). By Proposition 3.2, acts on through with isomorphic semisimple groups and . Let be the corresponding decomposition into modules as in Remark 3.3, with an irreducible -module and acting trivially on for .

We claim that also acts non-trivially on . Assume otherwise, so . Now permutes the semisimple factors transitively, as well as the . For let denote the full preimage of in . As stabilises , it acts on . If is simple, then by Proposition 3.7, cannot act with a single Jordan block on any of the . If is not simple, so we are in case (2) of Proposition 3.2, we reach the same conclusion by Proposition 3.8. With this contradicts the fact that by Lemma 2.3, has at least one Jordan block of size on .

Thus, acts non-trivially on and on . So we have , where is an irreducible -module, for and . Then

by Lemmas 2.13 and 2.14. Now note that the second sum is zero by comparing highest weights (where we use the fact that if there is a non-trivial extension between two simple modules for a semisimple algebraic group then their weights are comparable, see, e.g., Reference 4, II.2.14). Thus, for some, and hence, all . So with a non-split extension of with as a -module and acting trivially on . Since normalises , it stabilises this decomposition of . But acts with Jordan blocks of size on , so it must act by a single Jordan block on .

If is simple, as in case (1) of Proposition 3.2, there is no such extension of the two irreducible -modules by Proposition 3.7, giving the desired contradiction. Hence we must be in the case that , , and the action on the is as in Proposition 3.2(2). Again, there is no such non-split extension by Proposition 3.8.

Thus for Theorem 1, as far as simple groups of classical type are concerned, it remains to consider groups of type .

4. The case of

In this section we consider the following situation: for , and is a (not necessarily connected) reductive subgroup of the form , with not a torus, for a regular unipotent element of . Recall from Lemma 2.12(c) that regular unipotent elements of have two Jordan blocks on , of sizes if , and sizes when .

4.1. Almost simple groups containing an element with a large Jordan block

Proposition 4.1.

Let and be an almost simple algebraic group acting irreducibly on and such that some unipotent element of has a Jordan block of size . Then .

Proof.

For a simple group of exceptional type, the irreducible representations of whose image contains a unipotent element with a single non-trivial Jordan block are classified in Reference 23, Thm 1.1, and only in its 6-dimensional representation comes up. But here no unipotent elements have blocks of size 5. (See e.g. Reference 6.) The only almost simple but not simple group of exceptional type is . Here, the regular unipotent elements of have order 32, but has no faithful irreducible representation of dimension less than 54.

Now assume that is simple, simply connected, of classical type but not equal to and let denote the dimension of its natural module. Here we may assume is not an odd-dimensional orthogonal group as we will consider the isogenous symplectic group. The unipotent elements of have order less than . On the other hand, writing , an element with a Jordan block of size has order at least , so we get . All irreducible representations of such dimensions are known (see e.g. Reference 12, Thms 4.4 and 5.1), and we find that either , or up to twists one of the following holds:

with and is the tensor product of two 2-dimensional modules, or

with and is the exterior square of the natural module, or

for and is the spin module, or

for and is a spin module.

Looking at the precise order of unipotent elements rules out all cases except spin modules for , and . In the latter two cases, the image of the representation lies in the 8-dimensional orthogonal group, where there are no unipotent elements with Jordan blocks of size 7, and for the case of , both 4-dimensional representations have image in the symplectic group where there are also no unipotent elements with a block of size 3. Finally, if , by the above remarks we may assume is the natural representation of (); we note that does not have unipotent elements of the required Jordan block sizes on in characteristic 2 (see Reference 9, Lemma 6.2).

It remains to consider almost simple of type or . Here, unipotent elements have order less than , so as before we conclude . Now note that the natural module for is not invariant under the graph automorphism, nor is its exterior square for . For the spin modules for do not afford representations of . Again with Reference 12 we arrive at the Lie algebra for , the exterior square of the natural module for , and modules of dimension . All of these embed into a general orthogonal group, but the latter does not have unipotent elements of the required Jordan type in its natural representation (see Reference 9, Lemma 6.2). So none of these leads to examples, completing the proof.

4.2. A reduction result

Proposition 4.2.

Assume . Let be reductive with , semisimple, and regular unipotent in . Assume that does not stabilise any non-zero totally singular subspace of . Then one of the following four mutually exclusive cases occurs:

(1)

is irreducible on ; more specifically, either , or and ;

(2)

, is even, is irreducible on stabilising a pair of complementary totally singular subspaces interchanged by ;

(3)

there is an orthogonal decomposition into -submodules , where and is irreducible and tensor indecomposable for ; or

(4)

, stabilises a 1-dimensional non-singular subspace of , acts as a subgroup of on with having a single Jordan block on , and there exists no -complement to in .

Proof.

First assume that is a decomposable -module. Then by the Jordan block shape of we must have with , and acts with a single Jordan block on both summands. If acts non-trivially on then by Theorem 3.9, acts irreducibly on . Since then must be non-degenerate, so we obtain an -invariant decomposition . Lemma 2.4 shows that is tensor indecomposable for , so we reach conclusion (3). On the other hand, if is trivial on then it must act faithfully on and hence and . In particular, acts by an inner automorphism on and thus contains a regular unipotent element of by Lemma 2.9, which is impossible as these have order at least 4.

So now assume that is an indecomposable -module. In particular, there is no -invariant non-degenerate non-trivial proper subspace of . Thus, if denotes a non-zero -invariant subspace of of minimal dimension, then either , that is, acts irreducibly on , or is non-singular of dimension 1 and . In the latter case is contained in the stabiliser of , isomorphic to , and is the natural module for . Let’s first discuss this situation. Now is regular unipotent in by Lemma 2.10, so it has a single Jordan block on . By Theorem 3.9, acts irreducibly on , and thus is a direct sum of non-isomorphic irreducible -modules by Lemma 2.5, all non-trivial as acts non-trivially on . Assume that . As has no trivial -composition factor, this decomposition is -invariant. By dimension reasons, the irreducible -module must be non-degenerate, but this was excluded before. Thus has no -complement in , and hence (4) holds.

Thus we are left to consider the case that acts irreducibly on . By Lemma 2.5 then is a direct sum of non-isomorphic irreducible -modules transitively permuted by . Then with and is a power of . First assume has Jordan blocks of sizes on . Induction from Lemma 2.3 shows that has Jordan blocks of size , one Jordan block of size  and one Jordan block of size 1 on . Since has the same Jordan blocks on each  by transitivity, the only compatible solution is . If has Jordan blocks of sizes on (and so in particular ) then Lemma 2.3 shows that either , or has blocks of size , and 2 blocks of sizes each. This forces . In conclusion either , or . Let us consider these two cases in turn.

If , that is, if is irreducible, then is simple by Lemma 2.7 as none of (1)–(3) there can occur here. If we are in case (1) of our statement. If not, then by Proposition 2.11 the only possibility for is again the one given in (1). For , using that must also be irreducible on the natural 4-dimensional module for (e.g. by Borel–Tits), Proposition 2.11 implies that we must have with or . But in neither of these cases does act irreducibly on the exterior square , so this case does not occur here.

Now consider the case where so that the Jordan blocks of on , , have sizes . We claim that is totally singular. For otherwise, is non-degenerate and thus so is which must be , and therefore is contained in the stabiliser in of the orthogonal decomposition , hence in . But by Reference 17, Thm B(ii)(a) there is no reductive maximal subgroup of containing this stabiliser and a regular unipotent element of .

We thus have that is totally singular as claimed; then so is its image under . Hence stabilises a decomposition of into a direct sum of maximal totally singular subspaces and thus . According to Lemma 2.12(d) when is odd, the regular unipotent elements of the stabiliser of such a decomposition have a single Jordan block, so cannot lie in . Thus , which does not contain elements with a Jordan block of size . On the other hand, when is even the stabiliser contains regular unipotent elements of . Lemma 2.7 now implies that is simple, since the case (3) with cannot occur here as is even. By Proposition 4.1 this gives the examples in (2).

In what follows we investigate further the case (4) of the preceding result.

Proposition 4.3.

In the situation of Proposition , the following hold:

(a)

, with pairwise isomorphic factors , (with ) or , permuted transitively by ; and

(b)

there is a decomposition into -modules , irreducible for and transitively permuted by , and on which acts by a single Jordan block, with when or , respectively when .

Proof.

We keep the notation from Proposition 4.2(4). Recall that here . As acts by a single Jordan block on , this is an irreducible -module by Theorem 3.9. So by Proposition 3.2 there is a decomposition with transitively permuting the semisimple factors .

We first show that we are not in case (2) of Proposition 3.2. Suppose the contrary. Then with , , and , where is an -module which is a twist of a tensor product of two natural modules for and the are transitively permuted by . Let denote the full preimage of in , so . The Künneth formula Reference 20, Lemma 3.3.6 shows that , whence , with . So there is a similar decomposition for all of the , leading to a decomposition of the -module , contradicting Proposition 4.2(4).

Hence we are in case (1) of Proposition 3.2 and all are simple. Let us write . Now , where the are irreducible tensor indecomposable -modules, transitively permuted by , and acts by a single Jordan block on each. Hence the possibilities for are as listed in Table 1. Moreover, arguing as in the preceding paragraph we see that . Now by Reference 16 all pairs have except possibly for , or with as claimed, so by Proposition 3.2 we have and we get (a) and (b).

Proposition 4.4.

Let be reductive with , semisimple, and regular unipotent in . Then does not stabilise a maximal totally singular subspace of .

Proof.

Arguing by contradiction, assume that stabilises some maximal totally singular . Then its stabilizer has Levi factor . By Lemma 2.8, for the image of in contains a regular unipotent element of , and acts irreducibly on by Theorem 3.9 and hence on its dual . Note that acts non-trivially on .

Let us first consider the case when . Assume that acts with different orders on and on . Then a non-trivial power of , which we may take to be an involution, is in the kernel of the representation on . Letting denote the projection into the Levi factor of , we thus have lies in the normal subgroup , and has image , so in fact , that is, centralises . As has two Jordan blocks , the element has Jordan blocks with , on . (Here denotes a Jordan block of size .) Then by Reference 9, Thm 3.1, setting , we have that and the action of this group on is via two copies of the natural module for and one copy of the natural module for . As , the dimensions of the -composition factors on (two factors of dimension ) must be obtained as a refinement of the dimensions coming from the action of , which is not possible. So has the same order on as on . As has a Jordan block of size on , and blocks of sizes on , we must have . So we have proved that is odd when .

Now let be arbitrary again and write for the decomposition of into pairwise non-isomorphic irreducible -modules as in Remark 3.3, with transitively permuting the summands. Note that since has to be odd when we cannot be in case (1) of Proposition 3.2. As and are dual to each other, then with dual to . Let be the -submodule of with composition factors and (if is not completely reducible as an -module, this exists by Lemma 2.13 together with Remark 3.3). Then the transitive action of yields . From the block structure of , as in the proof of Proposition 4.2, this implies that or . But the latter cannot occur as is odd when . So , and is an extension of an irreducible -module by its dual, on both of which has a single Jordan block. Assume that this extension splits. If is self-dual, so is homogeneous, then it is an irreducible -module by Lemma 2.6, a contradiction. Else, the decomposition is -invariant, contrary to the block structure of .

So is a non-trivial extension, and since by Lemma 2.15(a) there are no self-extensions for simple groups, cannot be a self-dual -module. Moreover, still by Proposition 3.2, is now either simple or with . The second possibility is ruled out as is not self-dual. The possible non-self-dual for simple are listed in Table 1, and we find that with . But according to Reference 16, Cor. 1.1.1, there is no non-trivial extension of a twist of the natural module of with its dual.

The following proposition treats the special case arising out of Proposition 4.3 when . By our Lemma 2.9, contains a regular unipotent element. This case should have been treated in Reference 22 but the argument there is incomplete in precisely this setting. So we have included a proof here.

Proposition 4.5.

Let be a simple algebraic group with , , defined over a field of characteristic . Assume that stabilises a non-zero totally singular subspace of . Then does not contain a regular unipotent element of .

Proof.

Assume is regular unipotent. Choose a maximal totally singular -invariant subspace . Then the stabilizer has Levi factor and the projection of into the second factor does not lie in a proper parabolic subgroup.

By Lemma 2.8, for the image of in contains a regular unipotent element of . Thus, if , the projection onto the factor of is injective when restricted to . But the order of a regular unipotent element in is strictly less than the order of unless possibly when is maximal totally singular, and this latter case does not occur by Proposition 4.4. Hence we need only consider the case where . Let be the natural projection. Since does not lie in a proper parabolic subgroup of , by Reference 10, Lemma 2.2, one of:

(i)

, with all non-degenerate, inequivalent and irreducible -modules; or

(ii)

stabilizes a nonsingular 1-space of .

In the first case, when , since contains a regular unipotent element of , Proposition 2.11 implies that either , and the action of on is via a spin module, or . In both cases, there are no non-trivial extensions between and the trivial -module and we deduce that (and hence ) lies in the Levi factor, a contradiction.

In the first case, with , the Jordan block structure of then implies that and we may assume . But then the projection of to is trivial as the latter is a torus, contradicting the Jordan block structure of .

So we now have that lies in the stabilizer of a nonsingular 1-space of ; let be such a subspace, so that stabilizes the flag .

Now Proposition 4.2(4) gives that has one block on and so by Table 1 we are left with the following irreducible actions on :

or ;

, .

In the first case, we deduce that (the full stabilizer in of a non-singular -space), and so is a -dimensional tilting module for . In particular, there is no extension of this module by a trivial and we find that lies in a Levi factor of . This rules out the case where or .

In the second case, we have , and the action of on is as an -dimensional indecomposable tilting module (see Reference 8, Lemma 9.1.1) and as above there is no non-trivial extension with the trivial module. So lies in a proper Levi factor of , and hence cannot contain a regular unipotent element. This final contradiction completes the proof.

Proposition 4.6.

Let be reductive with , semisimple, and regular unipotent in . Assume that lies in a proper parabolic subgroup of . Then, with an -invariant totally singular subspace of of maximal possible dimension, we have: , , and acts non-trivially on and is not completely reducible on .

Proof.

Let be as in the statement. So lies in a proper parabolic subgroup of with Levi complement . By Lemma 2.8 the image of in the Levi factor is again regular unipotent, in particular acts by a single Jordan block on . Also note that acts non-trivially on , as otherwise it would act trivially on as well and hence on all of . By Proposition 4.4, cannot be maximal totally singular.

So, we now assume that , and thus is reducible on . By Theorem 3.9 this implies that cannot have a single Jordan block on . Thus, has exactly two Jordan blocks on , one of which has size at most . We set .

Note that does not stabilise any non-zero totally singular subspace of by the choice of . This then implies that acts non-trivially on since , lying in a Borel subgroup, has some totally singular fixed points. In particular . In fact, if then the image of in and hence itself is either or . If then acts on it by an inner automorphism, so contains a regular unipotent element of by Lemma 2.9, contradicting the main result of Reference 22. Similarly, if is the full Levi factor and thus , we conclude by the same argument. Hence we have . If is not completely reducible on , then we arrive at the conclusion.

So now assume that is a completely reducible -module. We will show that this leads to a contradiction. Write for the decomposition of into its -homogeneous components. Then permutes these components and can have at most two orbits on , as it has two Jordan blocks on .

Case 1.

We first discuss the case where is transitive on . Arguing precisely as in the proof of Proposition 4.2 the Jordan block shape of forces either , or . We consider these two cases in turn.

If , that is, if is homogeneous, then by Lemma 2.6 we find that , as is reducible on , contradicting .

If , one checks that the Jordan blocks of on , , have sizes . As is -invariant and homogeneous as an -module, it must be irreducible for by Lemma 2.6. Now interchanges and , so is irreducible for , a contradiction.

Case 2.

So now assume that has two orbits on . Let denote the subspaces spanned by these orbits, with . Then acts with a single Jordan block on each of them.

Consider first the case that acts non-trivially on . Then, since has a single Jordan block on we obtain by Theorem 3.9 that is an irreducible -module. Then either or . In the first case must be isomorphic to (as ), contradicting that . If on the other hand then is isomorphic to a submodule of . Thus the orthogonal group of dimension contains a unipotent element with a Jordan block of size at least . By the knowledge of possible Jordan block shapes (see Reference 9, Lemma 6.2) this is not possible as .

Finally, in Case 2 it remains to discuss the situation where acts trivially on . Then it must act irreducibly and faithfully on and hence and . In particular, acts by an inner automorphism on and thus contains a regular unipotent element of by Lemma 2.9, which is impossible by order considerations.

4.3. Proof of Theorem 1 for ,

Theorem 4.7.

Let , where , be a reductive subgroup, with a regular unipotent element of . Assume that . Then does not lie in any proper parabolic subgroup of .

Proof.

It suffices to prove the claim for , and hence we may and will assume that is semisimple. Moreover, by Remark 2.2 we may assume that is the product over a single -orbit of simple components. Assume that lies in a proper parabolic subgroup of . Then there is an -invariant flag with totally singular and dual to as an -module, and non-degenerate. We choose such that is maximal possible (and so is minimal). Hence, we are in the setting of Proposition 4.6.

By the choice of and since , we have that acts as in (1)–(4) of Proposition 4.2 on . Also, by Lemma 2.8, acts with a single Jordan block on as well as on , and with two Jordan blocks on .

Case 1.

First assume that acts non-trivially on and thus that . By Proposition 4.6, acts non-trivially on as well, which has dimension at least 6.

Case 1a.

We first discuss the case where . Let , , be the projections of into the two factors of the Levi subgroup , respectively. Since is irreducible on by Theorem 3.9, it cannot lie in a proper parabolic subgroup of , and by the choice of , neither is contained in a proper parabolic subgroup of . Write and so . We know that is a single Jordan block, and has a block of size and one of size 2, by Lemma 2.8. Since by assumption, has order smaller than , so some power with lies in ; we choose minimal with this property. As before, we see that must centralise . But then as well, as otherwise is a non-trivial unipotent element of centralised by , whence by Borel–Tits, lies in a proper parabolic subgroup of , which is not the case. Note that no smaller power of lies in as otherwise that element would (as before) centralise , forcing to lie in a proper parabolic of , again a contradiction. So have the same order .

Recall that has a single block of size on and blocks of sizes and on . Also, has two blocks on , one of size 1 or 2. But the first possibility is ruled out as and so .

We now show that has order on , that is, order twice as large as its order on (and on ). Let be minimal such that , so is also minimal so that . Hence we have and . In particular, . Since also , the order of in its action on is . In particular, acts as an involution on , with and (as ). As in a previous proof, the centraliser of in has composition factors of dimensions on , and as centralises , the -composition factors on must be obtained as a refinement of this.

Note that since and , there are at least three composition factors and so cannot act irreducibly on , ruling out configurations (1) and (2) of Proposition 4.2. Considering the cases in (3) and (4), we see that has composition factor dimensions on among

As and , the first case yields and . In the second case, the natural module for must remain irreducible for , as else there are five composition factors, and so either , or and .

If , then has one block of size 2 and the remaining blocks of size 1 on . By Lemma 2.3 this can only happen if for some , so has order on and by the previous analysis, its order on and on is (so as above). But this implies , a contradiction.

Hence we have and and has exactly two blocks of size 1 on (coming from the block of on ), while the one block of size produces only blocks of size  for . So is a power of , say . Thus, has order on and the order of on is , so and , forcing .

We claim that neither of the possible actions of on (as in Proposition 4.2(3) and (4)) is consistent with this. First suppose we have the configuration of Proposition 4.2(3), where , is irreducible on and . Let be the -submodules of such that , . On each of these acts reducibly and so has at least two Jordan blocks. Hence we have and , whence as has a single Jordan block on , which gives , contradiction.

So finally we are left to consider the case where acts on as in Proposition 4.2(4). Here we have , with , and . Recall that has order on and thus the same order on the codimension 1 subspace , which is again twice the order of on and on . So again acts as an involution on and the -composition factors are of dimensions . This is only consistent with the analysis if . This is the final contradiction settling Case 1a.

Case 1b.

So now we have . As , we may apply Proposition 4.2 to the action of on , and as we are in either case (1) or (3). If we are in case (1), acts as on , so acts by an inner automorphism on a component of and we are done by Lemma 2.9 and Reference 22.

In case (3) of Proposition 4.2, we have , with preimages in . Now acts non-trivially on both , normalised by , so has two Jordan blocks on each , by Theorem 3.9. Counting fixed points on as in Case 1a, we reach a contradiction.

Case 2.

Now assume that acts trivially on . Let be the -invariant subspace of of codimension 1. Note that is non-degenerate and acts as a regular unipotent element of by Lemma 2.8 and the image of lies in a proper parabolic subgroup of this orthogonal group. So it suffices to derive a contradiction in that situation, whence henceforth we assume . Again, as , we may apply the conclusion of Proposition 4.2 to the image of in (using our assumption that ).

Case 2a.

In case (1) of Proposition 4.2 again we have as in Case 1b, a situation that was handled in Reference 22.

Case 2b.

In case (2) of Proposition 4.2 we have , with as . Note that is then a completely reducible -module since there are no extensions between the natural and the trivial module for , so is isomorphic to an -submodule of . Assume that . Then, by dimension reasons, this intersection must be one of the two non-isomorphic irreducible -summands. But then has a non-degenerate -invariant form and thus is a self-dual -module, which it is not. Thus, is a non-degenerate -submodule of , and is its 2-dimensional orthogonal complement. Since are both sums of homogeneous -components of , the decomposition is -invariant, making an irreducible -module. Thus is a 1-dimensional totally singular subspace, but the fixed points of the non-trivial unipotent elements of are non-singular.

Case 2c.

In case (3) of Proposition 4.2, we have an -stable decomposition with irreducible on . Write for the full preimage of in , , both -submodules of . By Theorem 3.9, as is reducible for , we deduce that has two blocks on . If has more than one Jordan block on as well, then counting fixed points as in Case 1a we obtain a contradiction. So has a single Jordan block on , whence and .

Now first assume that so that . Using that we have , and a dimension count then shows the sum is direct. Hence , contradicting the Jordan block structure of on . Similarly, when and hence , consider . Then , , and these two intersect in . By assumption, has one fixed point on , but on it has a two-dimensional fixed point space, giving a contradiction.

Case 2d.

In case (4) of Proposition 4.2, we have and has -invariant subspaces with non-singular of dimension 1, such that acts with one Jordan block on . By Proposition 4.3 we may decompose into a product of simple groups all isomorphic to either , or , with , where we set in the case of . Furthermore, is a power of 2, with since otherwise we are done by Lemma 2.9 and Proposition 4.5. As has Jordan blocks of sizes on , has two Jordan blocks of size and two of size 1.

Note that acts trivially on the 2-dimensional full preimage of in . By our assumption, cannot be totally singular. Let be a 1-dimensional non-singular subspace. Then lies in the stabiliser isomorphic to , so in . We claim that has Jordan block sizes on . Indeed, as it has block sizes on , the only other possibility would be (note that all odd block sizes must occur an even number of times). But by Reference 9, Thm 6.6 the centraliser of such an element has reductive part of its centraliser containing an , while the reductive part of the centraliser of in is just a torus, a contradiction.

Now the -composition factors of are the and two trivial modules. Thus, by self-duality, has a submodule of codimension 1, and this is the sum of submodules of dimension at most (namely either the or extensions of some by a trivial module). But has block sizes times) on , so at least one block of size on . By Theorem 3.9, this contradicts the fact that has no irreducible -submodules of that dimension.

5. Exceptional types

In this section we consider algebraic groups defined over of characteristic . See Remark 2.1.

We will make extensive use of the known data on unipotent elements in simple algebraic groups of exceptional type, including element orders and power maps given in Reference 6 and structure of centralisers described in Reference 9. We follow the notation in Reference 6 for the labelling of unipotent classes. In particular, if the class of is denoted by some Dynkin type, then is a regular element in a Levi subgroup of that type.

In the course of our proof we will require precise knowledge on the existence and conjugacy classes of complements to in , for certain unipotent elements , as in the next result.

Lemma 5.1.

Let be unipotent and let be connected reductive, where

Then there exists a connected reductive group such that and lies in a conjugate of .

Proof.

Throughout we write . The existence of a complement to in follows from Reference 9, 17.6. We now turn to the proof of the remaining assertions.

Consider first , with , , a unipotent element of type and , where is a long root -subgroup of (see Reference 9, Tab. 22.1.2). By Reference 7, Thm 5 there exist two classes of such -subgroups in , coming from the two non-conjugate Levi factors of . By Reference 7, Cor., p.2, there exists a unique class of complements to in . For both classes of -subgroups, each non-trivial -composition factor of occurs as a composition factor of , for some , where is the natural -dimensional -module. (See Reference 7, Tab. 8.2.) There exists a unique class of -subgroups of , and restricting each of the given irreducible -modules to such an -subgroup we find that the composition factors of on have highest weights among , (the fundamental dominant weights of ), for , and the zero weight. Using Reference 4, II.2.14 we have for all -composition factors of and hence by Reference 20, Prop. 3.2.6, there exists a unique class of complements to in . Thus there exists with as claimed.

In the other two cases, we have , with , , and is a unipotent element of type either or , and , where , respectively , long root subgroups. The action of a long root on has composition factors the natural, dual or trivial module, and so in case , we have a unique class of complements to in , establishing the result.

In the case we must argue slightly differently because here there is a -dimensional -composition factor of with . Now lies in a parabolic subgroup of with . Moreover, considering the labelled diagram of the class of  (see Reference 9, Tab. 22.1.1) and applying Reference 9, Thm 17.4, we see that we may take to be a -parabolic subgroup of . There exists a composition series of as an -module, all of whose terms are -dimensional -modules and trivials. The subgroup is uniquely determined up to conjugacy in and, by Reference 8, Lemma 9.1.1, each such irreducible upon restriction to is the indecomposable tilting module with the 6-dimensional irreducible -module. By Reference 4, Prop. §E.1, . Hence, all complements to in are conjugate and by considering the action of on , we have the same statement for . Arguing as in the previous cases now yields the claim.

Theorem 5.2.

Let be a simple algebraic group of exceptional type and a reductive subgroup with a regular unipotent element of and . Then does not lie in a proper parabolic subgroup of .

Proof.

Let be as in the assertion and assume that lies in a proper parabolic subgroup of . Then we have by Reference 22, Thm 1.2. Also, by passing to we may assume that is semisimple. We will need to consider the image of in Levi factors, and for this throughout we write for the quotient of by its largest normal unipotent subgroup (which, being finite, is centralised by ). Note that maps isomorphically to a subgroup of on which the image of then acts faithfully.

By Remark 2.2, we may moreover assume that has a single orbit on the set of simple components of , and, by Lemma 2.9 it does not act as an inner automorphism on . In particular, . On the other hand, as lies in a proper parabolic subgroup of , . This already rules out the case . Furthermore, regular unipotent elements of are also regular in under the natural embedding, and proper parabolic subgroups of lie in such of , hence it suffices to prove our result for or . The orders of regular unipotent elements in these groups for small primes are given in Table 2.

Case 1.

We first consider the case that acts by an inner automorphism on and does not contain elements of order , so in particular .

Then by Lemma 3.5, is centralised by a unipotent element of this order. We discuss this situation by comparing the list of centralisers of unipotent elements Reference 9, §22 and the list of unipotent element orders Reference 6, Tab. 5–9.

Let first and consider the case , where we have and centralises . By Reference 9, Tab. 22.1.3 and Reference 6, Tab. 5 unipotent elements of order 8 centralising a group of semisimple rank at least 2 lie in class , with reductive part of the centraliser of type . Since we conclude that . Now acts as an inner automorphism of and so centralises , which lies in the class , by Reference 6, Tab. D. Moreover, using Reference 9, Tab. 22.1.3, we see that the full connected centraliser of has a reductive complement to , a -subgroup of generated by long root elements of . Now and we consider the possible embedding of in . By Reference 10, Lemma 2.2, must lie in a proper parabolic subgroup of and by rank considerations we find that lies in an -parabolic subgroup of . As is generated by long root subgroups of , the Levi factor is also generated by long root subgroups of . Now arguing as in Lemma 5.1, we find that is a long root -subgroup of , that is, a Levi factor of . (The Levi factor of acts on with composition factors the natural, dual or trivial module for .) But now the centraliser of is an and so normalizes an subgroup of . But there is no such example in Reference 17, Thm A. For we have and , and again by Reference 9, Tab. 22.1.3 and Reference 6, Tab. 5 there is no possibility. For we have , contrary to our assumption.

For and we have but all centralisers of such elements have semisimple rank at most 1 by Reference 9, §22. When and so , the unipotent classes , , , and need to be discussed. Here the semisimple parts of the centralisers have type , , , , , respectively. As is contained in one of those, and has a single orbit on its set of simple components, must be of type . Now contains acting as an inner element on , and so centralises . But by Reference 6, Tab. D, lies in class , with semisimple part of its centraliser a -subgroup generated by long root subgroups, by Reference 9, §22. By Lemma 5.1, must be a long root . There are two classes of Levi subgroups in . Using Borel–de Siebenthal one sees that one is centralised by an , the other by a . Thus in any case, contains a subgroup of of maximal semisimple rank, so the normaliser of and all of its overgroups are reductive. But by Reference 17, Thm A there are no such subgroups containing a regular unipotent element. For we have , contrary to our assumption.

Finally, assume . For we have . Only the 17 unipotent classes

of contain elements of order 16. Of these, only the first three cases have a semisimple part of the centraliser of rank at least 2, of type , and respectively. So or , with acting by the graph automorphism. In the second case, centralises and lies in class , but the latter has centraliser of rank 1. So in fact . By Reference 9, Tab. 22.1.1, the remaining possible subgroups , of are generated by long root subgroups, and by Lemma 5.1, is contained in one of them. Now all subgroups of type of these are again generated by long root subgroups, hence so is . Thus, the centraliser of is of type , and so acts on an , whose normaliser is a maximal subgroup of . But since this does not appear in Reference 17, Thm A, its normaliser does not contain regular unipotent elements.

If then , but no unipotent element of order bigger than 9 has a centraliser of semisimple rank at least 3. If then , but none of the seven unipotent classes having centraliser of semisimple rank at least 5 contains elements of order 25. Finally, for we have , which is not allowed here.

Case 2.

We next consider the case that acts by an inner automorphism on , and that does contain an element of order .

When and , then is a semisimple subgroup with an element of order 8 having a non-trivial graph automorphism transitively permuting the simple factors. Therefore, is one of . By assumption lies in a proper parabolic subgroup of with Levi factor , and thus contains one of the above groups, with the image of inducing a non-trivial graph automorphism of order . By rank considerations, only , and might occur. The minimal dimension of a representation of on which acts non-trivially is 10, so this cannot occur inside . This representation embeds into , but not into by the block structures given in Lemma 2.12. Also, the smallest faithful representation of has dimension , too large for any proper Levi subgroup. So in fact we must have inside a -parabolic. Using Reference 20, 3.2.6 one can check that this embedding is into a Levi factor, and so is a Levi subgroup of . By Reference 9, Tab. 22.1.3, no non-trivial unipotent element of has a in its centraliser, so is a torus, and then in fact it must be the centre of a Levi subgroup of type . As acts on , Proposition 7.8 below implies that must centralise . But lies in the class and has centraliser of rank by Reference 9, Tab. 22.1.3, a contradiction.

For and with , we have that contains elements of order 9 and has an outer automorphism of order 3, so . But the smallest faithful representation of has dimension 24, which is too large for containment in any proper parabolic subgroup of . For where the only option is that . But again by Borel–de Siebenthal no proper parabolic subgroup has a Levi factor containing a group with .

For with the semisimple group has an element of order 16 and a non-trivial graph automorphism, whence . But clearly no Levi factor of a proper parabolic subgroup of can contain with or . For with the only possibilities with a graph automorphism of order 3 are , and . All could only lie in a proper parabolic subgroup of type . But has no maximal rank subgroups or by Borel–de Siebenthal. When , Reference 7, Thm 5 shows that is a Levi factor of . Again by Borel–de Siebenthal there is a subgroup centralising , so is reductive. But by Reference 17, Thm A, there is no positive-dimensional maximal reductive subgroup of containing a regular unipotent element. For again the only possibility is . The only proper parabolic subgroups whose Levi factor might contain with are those of type . The list in Reference 17, Thm B shows that there is no maximal reductive subgroup of containing a regular unipotent element of and such an .

For and we have . The only semisimple groups of rank at most 7 with a unipotent element of order 16 and an even order graph automorphism are , and . Now for or the element of order 16 acts as an inner element of order 8. Thus is centralised by the element of order 16, which is not possible. Assume and induces a graph automorphism on . There is only one class of subgroups in by Reference 7, Thm 5 and hence is a Levi factor of . Again, normalises the centraliser of such an , hence a subgroup , and as above this is not possible by Reference 17, Thm A. If then , and there is no possible case. When or , then must have at least  simple components, whence . Now for , the group does not contain elements of order 25, so we have and . The only proper parabolic subgroup of with a Levi factor containing with is of type . By Reference 11, Thm 4 and Tab. 17 and 18, such an lies in a Levi factor . Now the centraliser of that contains the centralising the -Levi subgroup. So again the normaliser of in has maximal semisimple rank, and Reference 17, Thm A shows that this cannot contain regular unipotent elements.

Case 3.

Finally, consider the case that is not inner. Then either has at least components, or there are components and on each of them induces a graph automorphism of order . Either possibility forces , so . Furthermore, must act by an inner automorphism on . When then the possibilities are or . In the first case, by Lemma 3.5 there exists an element of order 4 centralising an , which is not possible by Reference 9, §22. In the case and acts as an outer automorphism of order 4. The only Levi factor possibly containing a subgroup with is of type , but contains elements of order 16 while does not have such elements. When then ; none of the groups , and contains elements of that order, so by Lemma 3.5 there is an element of of order 8 centralising such a subgroup, which is not the case by Reference 6 and Reference 9. Finally, when then again , and the candidates for are , and . Assume , then acts by an inner automorphism so centralises . But lies in class (there is a misprint in Reference 6, Tab. D), and its centraliser does not contain an . Similarly, if then , in class , centralises , which is not possible. The same argument rules out . This completes our case distinction and thus the proof.

Theorem 1 now follows by combining Theorems 3.9, 4.7 and 5.2.

6. Regular unipotent elements in almost simple groups

We now extend our main result to the case of regular unipotent elements in cosets of simple groups in almost simple groups.

Example 6.1.

The regular unipotent elements in a coset of an almost simple group of “exceptional type” can be realized as follows:

(a) The group occurs as a subgroup of in a natural way (see e.g. Reference 15, Ex. 13.9). Now according to Reference 6, Tab. 4 the only unipotent class of for containing elements of order 27 is the class of regular unipotent elements. Also, the regular unipotent elements in an outer coset of the disconnected group have order 27 (see Reference 13, Tab. 8). Thus, they are regular unipotent elements of .

(b) Similarly, the disconnected group occurs inside the normaliser of a Levi subgroup of type inside . Again by Reference 6, Tab. 7 the only unipotent class of for containing elements of order 32 is the class of regular unipotent elements, and since regular unipotent elements in the outer coset of have order 32 (see Reference 14, Tab. 10), they must be regular unipotent elements of .

We then obtain the following consequence of Theorem 1:

Corollary 6.2.

Let be almost simple of type or with , or of type with , and be a reductive subgroup with a regular unipotent element of and . Then does not lie in any proper subgroup of such that is a parabolic subgroup of .

Proof.

In each case, we embed in a simple algebraic group ; namely, and embed into via their natural representation, embeds into (see remarks before Lemma 2.12), and , , embed into , respectively under the embeddings given in Example 6.1. Applying Lemma 2.12 and Example 6.1, we have that the embedding sends regular unipotent elements in an outer coset of to regular unipotent elements of . Now, if lies in a proper subgroup of with a parabolic subgroup of with , then , and by the Borel–Tits theorem, the latter lies in a proper parabolic subgroup of . Thus, in all cases our claim for the almost simple group follows from Theorem 1 for the simple group .

Note that the type of subgroups allowed for in the preceding statement are those given by the most general possible definition of “parabolic subgroups of an almost simple group”.

7. Regular unipotent elements in normalisers of tori

Here, we show that if one removes the hypothesis that in Theorem 1, the conclusion is no longer valid. More generally, for a simple group we investigate the structure of torus normalisers in that contain a regular unipotent element and lie in some proper parabolic subgroup of .

7.1. Torus normalisers in

Proposition 7.1.

Let where is a torus and is unipotent with a single Jordan block. Then all weight spaces of on have the same dimension . Moreover is contained in a proper parabolic subgroup of if and only if .

Proof.

Since normalises , the weight spaces of on are permuted by . Moreover this action must be transitive as otherwise would have at least two Jordan blocks on . Thus, they all have the same dimension, and if they are 1-dimensional, is an irreducible -module and so does not lie in any proper parabolic subgroup of .

Now assume the common dimension of the weight spaces is and set . Since has -power order, the number of weight spaces, , is a -power. It follows by Lemma 2.3 that acts with a single Jordan block (of size ) on each weight space. In particular, centralises and thus lies in a proper parabolic subgroup of by the Borel–Tits theorem (Reference 15, Rem. 17.16).

Groups as in the previous result do in fact exist:

Proposition 7.2.

Let be an integer, where . There exists a -dimensional torus , where , with -dimensional weight spaces on , normalised by a unipotent element with a single Jordan block. Moreover, lies in a proper parabolic subgroup of if and only if .

Proof.

Decompose into a direct sum of subspaces of dimension . Let be the permutation matrix for a permutation sending an ordered basis of to an ordered basis of for , where . Then has order . For let be the torus of scalar matrices and a unipotent element with a single Jordan block. Set . As permutes the transitively and we have that is the wreath product of with , and we can write elements of as for and some . Then the element has th power which has Jordan blocks of size on . But then must have a single Jordan block on by Lemma 2.3. Now with the subgroup is as in Proposition 7.1 and thus the claim follows.

7.2. Torus normalisers in and

We next discuss those classical groups in which regular unipotent elements have a single Jordan block on the natural module, that is, the types and (see Lemma 2.12).

For this, note that weight spaces for non-zero weights of a torus in and are totally isotropic and totally singular, respectively, and weight spaces for weights with , are orthogonal to each other. To see this for , let be the quadratic form and the associated bilinear form on . Let be a weight of with weight space . Then for , we have , for all . Since there exists with we find . So non-zero weight spaces are indeed totally singular. Further, for and two weights of and , , we have for all , and if then . The argument for is completely analogous.

Lemma 7.3.

Let or with and where is a torus and is regular unipotent in . Then .

Proof.

Assume . Write for the natural module of and let be its -weight space decomposition. Note that we have since . Now permutes the and hence their corresponding weights. As acts as a single Jordan block on by Lemma 2.12, this action must be transitive, so is a power of . Since there is at least one non-zero weight . As we are in or , then is also a weight, so both and lie in one -orbit, contradicting that .

In the case , by the exceptional isogeny between and we need not consider type .

Proposition 7.4.

Let , or and with a torus and having a single Jordan block on . If is the -weight space decomposition then the are totally isotropic or totally singular, respectively, and permuted transitively by . Moreover, up to renumbering, there is an orthogonal decomposition .

Proof.

As has a single Jordan block, it permutes the transitively and so is a 2-power. All are totally isotropic or totally singular, respectively, by the remarks before Lemma 7.3, and orthogonal to all other weight spaces that do not have opposite weight. Further, if is a weight of then so is , and thus for a suitable numbering, and have opposite weights, for . Thus we obtain the claimed orthogonal decomposition.

Example 7.5.

The situation nailed down in Proposition 7.4 does give rise to examples within proper parabolic subgroups. To see this, let , with , where with odd. By Lemma 2.12, for any odd the stabiliser in of a maximal totally singular subspace contains a unipotent element with a single Jordan block. This normalises , and centralises . Now embed

( factors). The normaliser of in contains an element cyclically permuting the factors. Set . By construction, has Jordan blocks of size , so by Lemma 2.3, has a single Jordan block on , it normalises , and centralises . Thus, lies in a proper parabolic subgroup of . Since , this also provides examples in .

7.3. Torus normalisers in

Here we consider tori in normalised by a regular unipotent element.

Proposition 7.6.

Let with , a torus and regular unipotent in . Then and if is the -weight space decomposition then up to renumbering the , we have one of:

(1)

, is even, interchanges and , and acts with Jordan blocks of sizes on both and ;

(2)

permutes transitively (so for some ) and is the -weight space, with ; or

(3)

acts transitively on and on , so for some , and and are -dimensional weight spaces for opposed weights.

Proof.

Let be the decomposition of into non-zero -weight spaces. Note that we have since . From the block structure of it follows that has at most two orbits on the set of . In addition, the sum of the weight spaces in one of the orbits is of dimension at most . Since we are in , if is a weight of on , then so is . Now first assume that is odd. Then and can only lie in the same -orbit if . So has two orbits on the set of weight spaces, one of length  and the other of length 1. There is a non-zero weight in one of the orbits; the weight space of then lies in the other orbit. This forces , contrary to our assumption.

Thus we have . First assume permutes the transitively. Then stabilises each , and has same block sizes on each of them. Since is a 2-power, the blocks of on then have sizes , whence and so . Since and are both totally singular, is contained in the stabiliser of a decomposition of into a sum of two maximal totally singular subspaces. If is odd, then this stabiliser in fixes each (see Reference 5, Lemma 2.5.8). Thus is even, interchanges and and has Jordan blocks as claimed in (1).

Next assume that permutes transitively. Then without loss of generality . If is not the 0-weight space, then the opposite weight space must be one of the other , so , and , contradicting our assumption. So we arrive at (2).

Finally, assume that permutes transitively. Then and the corresponding weights are opposed and interchanged by , which is (3)

Example 7.7.

We show that the cases in Proposition 7.6 do give rise to examples within proper parabolic subgroups. So let .

(1) Let be even, be the 1-dimensional central torus of inside the stabiliser in of a pair of maximal totally singular subspaces. Thus acts by scalars on both . Then is normalised by the outer elements of interchanging . Now by Lemma 2.12, a regular unipotent element in the outer coset of has Jordan blocks of sizes , hence is regular unipotent in . Then lies in the centraliser of the non-trivial unipotent element (non-trivial as soon as ), thus inside a proper parabolic subgroup. This is an example of (1) in Proposition 7.6.

(2) Let be the stabiliser of an orthogonal decomposition of , where with . Then by Reference 17, Thm B(ii)(a) there is a subgroup , with a maximal torus and a regular unipotent element of . We number the weights of on such that acts as the permutation on these. For , the group is an example for case (2). On the other hand by taking the direct product of with a subgroup of as constructed in Example 7.5, and intersecting with we find an example for (3), and as in part (1) we see that both lie inside proper parabolic subgroups.

(3) The example for in Proposition 7.2 falls into case (3); this can be seen from the weight spaces on the two modules, as the natural module for is the wedge square of the natural module for .

We are not aware of examples of torus normalisers in disconnected groups containing outer regular unipotent elements and lying in a proper parabolic subgroup.

7.4. Torus normalisers in simple exceptional groups

Finally, we investigate the case of exceptional groups.

Proposition 7.8.

Let be a torus and of prime-power order . Then .

Proof.

We have with . If has order then it must have an eigenvalue that is a primitive th root of unity. But then all Galois conjugates of are also eigenvalues of , and there are of these.

Remark 7.9.

Let be a torus in a connected reductive group and a unipotent element acting non-trivially on . Then divides the order of the Weyl group of . Indeed, by assumption is non-trivial. As is a Levi subgroup of and , the claim follows with Reference 15, Cor. 12.11.

Proposition 7.10.

Let be simple of exceptional type and with a non-trivial torus and a regular unipotent element of . Then one of the following holds:

(1)

, , ; or

(2)

, , .

Proof.

The regular unipotent element induces a non-trivial automorphism of , so by the previous remark, divides the order of the Weyl group of .

Combining the -power map on unipotent classes Reference 6, Tab. D and E and the structure of centralisers Reference 9, §22 we have compiled in Table 3 a list of the dimensions of maximal tori in the centralisers for and .

Now first consider . Then , so has order at most 4 when , respectively 3 when , by Proposition 7.8. Hence , respectively , must centralise , which by Table 3 implies and . But in that case, has order at most 2, so centralises and we reach a contradiction to Table 3.

When , then and by Proposition 7.8, has order at most 8. Again by Table 3 this gives that has order 8 and , contradicting the bound in Proposition 7.8.

For with we have when respectively. For , using Proposition 7.8 and Table 3 we find that and of order 8 is the only possibility. Here lies in class by Reference 6, Tab. D, and its centraliser has rank 4. Let be an -Levi subgroup of containing ; it has connected centre of dimension 4, so this must be the torus in . Now normalises , so it also normalises , and thus . If acts by an inner automorphism on , then by Lemma 3.5 there is an element of order 16 centralising , but the only elements of of that order are regular, a contradiction. Therefore, it acts by a graph automorphism on the and is inner and hence some element of order centralises . Again by Reference 9 and Reference 6 there is no element of order 8 in with such a centraliser. So this does not occur. Next, for the case using Proposition 7.8 and Table 3 as above, only with of order 3 remains. So , in class by Reference 6, Tab. D, centralises , and we reach case (1) of the statement. The case is not possible by Table 3.

For with Proposition 7.8 and Table 3 and arguing as above we are left with the case that and either and has order 8, or and , or and . The last case occurs in the conclusion, so we need to exclude the former two. If and , then centralises and lies in class . Let be a Levi subgroup of this type containing . It has centre of dimension 2, so this is in fact . Now normalises and hence also . Now , of order 16, acts as an inner element on this, and by Lemma 3.5 and using Reference 6 and Reference 9 we arrive at a contradiction. The case where is similar.

Finally for , the same line of argument as for the other groups shows that no new configurations occur.

Example 7.11.

Both cases in Proposition 7.10 do actually lead to examples.

(a) Let with . By Reference 17, Thm A, there is a maximal subgroup of containing a regular unipotent element , with a 2-dimensional torus. As centralises , the subgroup of then lies in a proper parabolic subgroup and so yields an example for the situation in Proposition 7.10(a).

(b) Let with . According to Reference 17, Thm A there is a maximal subgroup in containing a regular unipotent element , with a 1-dimensional torus. Then yields an example for the situation in Proposition 7.10(b).

Acknowledgments

This work was motivated in part by a question which Jay Taylor raised after a talk by the second author in Pisa. In addition, we acknowledge having had several useful conversations on cohomology with Steve Donkin, Jacques Thévenaz and Adam Thomas, and thank Thomas, Mikko Korhonen and David Craven for their careful reading of and comments on an earlier version, and David for spotting a gap in a proof.

Note added in proof

After becoming aware of our results, M. Bate, B. Martin and G. Röhrle in a recent preprint entitled “Overgroups of regular unipotent elements in reductive groups” have proposed a short, case-free proof of our Theorem 1 using the machinery of -complete irreducibility.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. Theorem 1.
  3. 2. Preliminary results
    1. 2.1. Jordan forms and tensor products
    2. Lemma 2.3.
    3. Lemma 2.4.
    4. Lemma 2.5.
    5. Lemma 2.6.
    6. Lemma 2.7.
    7. 2.2. On subgroups containing regular unipotent elements
    8. Lemma 2.8.
    9. Lemma 2.9.
    10. Lemma 2.10.
    11. Proposition 2.11.
    12. 2.3. Jordan forms and orders of regular unipotent elements
    13. Lemma 2.12.
    14. 2.4. Some results on extensions
    15. Lemma 2.13.
    16. Lemma 2.14.
    17. Lemma 2.15.
  4. 3. The case of
    1. 3.1. The completely reducible case
    2. Lemma 3.1.
    3. Proposition 3.2.
    4. 3.2. The not completely reducible case
    5. Lemma 3.4.
    6. Lemma 3.5.
    7. Lemma 3.6.
    8. Proposition 3.7.
    9. Proposition 3.8.
    10. Theorem 3.9.
  5. 4. The case of
    1. 4.1. Almost simple groups containing an element with a large Jordan block
    2. Proposition 4.1.
    3. 4.2. A reduction result
    4. Proposition 4.2.
    5. Proposition 4.3.
    6. Proposition 4.4.
    7. Proposition 4.5.
    8. Proposition 4.6.
    9. 4.3. Proof of Theorem 1 for ,
    10. Theorem 4.7.
  6. 5. Exceptional types
    1. Lemma 5.1.
    2. Theorem 5.2.
  7. 6. Regular unipotent elements in almost simple groups
    1. Example 6.1.
    2. Corollary 6.2.
  8. 7. Regular unipotent elements in normalisers of tori
    1. 7.1. Torus normalisers in
    2. Proposition 7.1.
    3. Proposition 7.2.
    4. 7.2. Torus normalisers in and
    5. Lemma 7.3.
    6. Proposition 7.4.
    7. Example 7.5.
    8. 7.3. Torus normalisers in
    9. Proposition 7.6.
    10. Example 7.7.
    11. 7.4. Torus normalisers in simple exceptional groups
    12. Proposition 7.8.
    13. Proposition 7.10.
    14. Example 7.11.
  9. Acknowledgments
  10. Note added in proof

Figures

Table 1.

Simple modules with regular unipotent elements. The last column records (an upper bound for) the order of a regular unipotent element .

Table 2.

Orders of regular unipotent elements

Table 3.

Ranks of centralisers ,

Mathematical Fragments

Theorem 1.

Let be a simple linear algebraic group over an algebraically closed field, a closed reductive subgroup containing a regular unipotent element of . If , then lies in no proper parabolic subgroup of .

Remark 2.1.

In the situation of Theorem 1, assume that . As is finite, some power of a regular unipotent element will lie in . In characteristic 0 any power of a regular unipotent element is again regular unipotent, so here we are thus reduced to studying the connected reductive subgroup satisfying the same assumptions. In that case, the conclusion of Theorem 1 was established in Reference 22, Thm 1.2. Hence, in proving Theorem 1 we may assume whenever convenient. Furthermore, we will assume without loss of generality that .

Remark 2.2.

Let be a reductive subgroup of a connected reductive group such that is regular unipotent in and . Let be one of the simple components of and set . Then , so acts transitively on the set of simple components of , and if lies in a proper parabolic subgroup of , then so does . Thus, when proving Theorem 1 we may as well assume that the simple components of are permuted transitively by .

Lemma 2.3.

Assume that and let be unipotent with a single Jordan block. Write with . Then has Jordan blocks, of size  and the other of size .

Lemma 2.4.

Let be a unipotent element with a single Jordan block of size , or with two Jordan blocks of sizes or . If preserves the factors in a non-trivial tensor product decomposition of then and has two Jordan blocks on . If these are of sizes .

Lemma 2.5.

Let be groups with finite cyclic and be a finite-dimensional irreducible -module. Then , where is any irreducible -submodule of and runs over a system of coset representatives of the stabiliser of in . Moreover, the are pairwise non-isomorphic -modules.

Lemma 2.6.

Let be connected reductive with non-trivial derived subgroup and assume that is completely reducible and homogeneous. If is normalised by a unipotent element with a single Jordan block of size , or with two Jordan blocks of sizes or then either is an irreducible -module, or , preserves a non-trivial tensor product decomposition of , and has two Jordan blocks on , of sizes if .

Lemma 2.7.

Assume and let be a reductive subgroup and a unipotent element with a single Jordan block of size , or with two Jordan blocks of sizes or . If acts irreducibly on , then either is simple, or one of the following holds:

(1)

, preserves a non-trivial tensor decomposition of  and has two Jordan blocks on . If these are of size ;

(2)

, as an -module with , , permutes both sets of factors transitively and has a single Jordan block on . Moreover, has a single Jordan block on each ; or

(3)

, as an -module with , has Jordan blocks of sizes on and has a single Jordan block on each .

Here, in and , does not preserve the stated tensor product decomposition of .

Lemma 2.8.

Let be connected reductive, a parabolic subgroup with Levi complement and assume that is regular unipotent in . Then the image of is regular unipotent in and hence in each simple factor of .

Lemma 2.9.

Let be simple and be a connected reductive subgroup normalised by a regular unipotent element of . Assume that is a central product with such that acts by an inner automorphism on . Then contains a regular unipotent element of .

Lemma 2.10.

Let be simple and a connected reductive subgroup containing a regular unipotent element of . Then is regular in .

Proposition 2.11.

Let be a reductive subgroup of the simple classical group , or with simple and irreducible on , where when is of orthogonal type. If contains a regular unipotent element of , then acts tensor indecomposably on .

Furthermore, either , or and the highest weight of on are as in Table 1 (up to Frobenius twists and taking duals) and has a single Jordan block.

Lemma 2.12.

Let be almost simple of classical type with regular unipotent in . Then in the natural representation of :

(a)

has a single Jordan block for , (when ), , and for when ;

(b)

has two Jordan blocks of sizes when for ;

(c)

has two Jordan blocks of sizes when , respectively of sizes when for ;

(d)

has a single Jordan block of size when for with odd; and

(e)

has two Jordan blocks of sizes when for with even.

Lemma 2.13.

Let be a group, and with , , be -modules. Then

Lemma 2.14.

Let be a field, be groups and two finite-dimensional -modules on which acts trivially. Assume that . Then

Lemma 2.15.
(a)

Let be a semisimple algebraic group. Then there are no non-trivial self-extensions between irreducible -modules.

(b)

Let be a semisimple group acting on , with acting trivially on . Then

Lemma 3.1.

Let be a reductive subgroup of the form for a regular unipotent element of , such that is not a torus. Assume that is completely reducible. Then is an irreducible -module.

Proposition 3.2.

Assume . Let be a reductive subgroup of the form for a regular unipotent element of , such that is semisimple. Assume that acts irreducibly on . Then

for submodules , transitively permuted by , on which acts with a single Jordan block; and

for distinct semisimple normal subgroups transitively permuted by , with , where acts trivially on for , and in addition one of the following holds:

(1)

the are isomorphic simple algebraic groups and is an irreducible tensor indecomposable -module, for ; if then , for some , and ; or

(2)

, , for all , with permuting the factors transitively, and each , as a -module, is a tensor product of irreducible -dimensional -modules.

Remark 3.3.

In the situation of Proposition 3.2, for set if , and if . Then is an irreducible -module and acts trivially on for . So is the trivial -homogeneous component of and the decomposition is stabilised by . Moreover, acts by a single Jordan block on each .

Lemma 3.4.

Let be a reductive subgroup of the form for a regular unipotent element of , with . Then is an irreducible -module. That is, the socle series of as -module is a composition series as -module.

Lemma 3.5.

Let be groups with . Assume that has order and acts by an inner automorphism on . If does not contain elements of order then for some element of infinite order or of order a multiple of .

Lemma 3.6.

Let be a group, and assume that all composition factors of are mutually non-isomorphic. Then contains no non-trivial unipotent elements.

Proposition 3.7.

Let with a simple algebraic group and regular unipotent in . If has at most two -composition factors then acts irreducibly on .

Proposition 3.8.

Let with and regular unipotent in . If has at most two -composition factors, one of which is the -fold tensor product of isomorphic -dimensional -modules, then is an irreducible -module.

Theorem 3.9.

Let be simple of type or and be a reductive subgroup of the form for a regular unipotent element of with . Then does not lie in any proper parabolic subgroup of .

Proposition 4.1.

Let and be an almost simple algebraic group acting irreducibly on and such that some unipotent element of has a Jordan block of size . Then .

Proposition 4.2.

Assume . Let be reductive with , semisimple, and regular unipotent in . Assume that does not stabilise any non-zero totally singular subspace of . Then one of the following four mutually exclusive cases occurs:

(1)

is irreducible on ; more specifically, either , or and ;

(2)

, is even, is irreducible on stabilising a pair of complementary totally singular subspaces interchanged by ;

(3)

there is an orthogonal decomposition into -submodules , where and is irreducible and tensor indecomposable for ; or

(4)

, stabilises a 1-dimensional non-singular subspace of , acts as a subgroup of on with having a single Jordan block on , and there exists no -complement to in .

Proposition 4.3.

In the situation of Proposition , the following hold:

(a)

, with pairwise isomorphic factors , (with ) or , permuted transitively by ; and

(b)

there is a decomposition into -modules , irreducible for and transitively permuted by , and on which acts by a single Jordan block, with when or , respectively when .

Proposition 4.4.

Let be reductive with , semisimple, and regular unipotent in . Then does not stabilise a maximal totally singular subspace of .

Proposition 4.5.

Let be a simple algebraic group with , , defined over a field of characteristic . Assume that stabilises a non-zero totally singular subspace of . Then does not contain a regular unipotent element of .

Proposition 4.6.

Let be reductive with , semisimple, and regular unipotent in . Assume that lies in a proper parabolic subgroup of . Then, with an -invariant totally singular subspace of of maximal possible dimension, we have: , , and acts non-trivially on and is not completely reducible on .

Theorem 4.7.

Let , where , be a reductive subgroup, with a regular unipotent element of . Assume that . Then does not lie in any proper parabolic subgroup of .

Lemma 5.1.

Let be unipotent and let be connected reductive, where

Then there exists a connected reductive group such that and lies in a conjugate of .

Theorem 5.2.

Let be a simple algebraic group of exceptional type and a reductive subgroup with a regular unipotent element of and . Then does not lie in a proper parabolic subgroup of .

Example 6.1.

The regular unipotent elements in a coset of an almost simple group of “exceptional type” can be realized as follows:

(a) The group occurs as a subgroup of in a natural way (see e.g. Reference 15, Ex. 13.9). Now according to Reference 6, Tab. 4 the only unipotent class of for containing elements of order 27 is the class of regular unipotent elements. Also, the regular unipotent elements in an outer coset of the disconnected group have order 27 (see Reference 13, Tab. 8). Thus, they are regular unipotent elements of .

(b) Similarly, the disconnected group occurs inside the normaliser of a Levi subgroup of type inside . Again by Reference 6, Tab. 7 the only unipotent class of for containing elements of order 32 is the class of regular unipotent elements, and since regular unipotent elements in the outer coset of have order 32 (see Reference 14, Tab. 10), they must be regular unipotent elements of .

Corollary 6.2.

Let be almost simple of type or with , or of type with , and be a reductive subgroup with a regular unipotent element of and . Then does not lie in any proper subgroup of such that is a parabolic subgroup of .

Proposition 7.1.

Let where is a torus and is unipotent with a single Jordan block. Then all weight spaces of on have the same dimension . Moreover is contained in a proper parabolic subgroup of if and only if .

Proposition 7.2.

Let be an integer, where . There exists a -dimensional torus , where , with -dimensional weight spaces on , normalised by a unipotent element with a single Jordan block. Moreover, lies in a proper parabolic subgroup of if and only if .

Lemma 7.3.

Let or with and where is a torus and is regular unipotent in . Then .

Proposition 7.4.

Let , or and with a torus and having a single Jordan block on . If is the -weight space decomposition then the are totally isotropic or totally singular, respectively, and permuted transitively by . Moreover, up to renumbering, there is an orthogonal decomposition .

Example 7.5.

The situation nailed down in Proposition 7.4 does give rise to examples within proper parabolic subgroups. To see this, let , with , where with odd. By Lemma 2.12, for any odd the stabiliser in of a maximal totally singular subspace contains a unipotent element with a single Jordan block. This normalises , and centralises . Now embed

( factors). The normaliser of in contains an element cyclically permuting the factors. Set . By construction, has Jordan blocks of size , so by Lemma 2.3, has a single Jordan block on , it normalises , and centralises . Thus, lies in a proper parabolic subgroup of . Since , this also provides examples in .

Proposition 7.6.

Let with , a torus and regular unipotent in . Then and if is the -weight space decomposition then up to renumbering the , we have one of:

(1)

, is even, interchanges and , and acts with Jordan blocks of sizes on both and ;

(2)

permutes transitively (so for some ) and is the -weight space, with ; or

(3)

acts transitively on and on , so for some , and and are -dimensional weight spaces for opposed weights.

Example 7.7.

We show that the cases in Proposition 7.6 do give rise to examples within proper parabolic subgroups. So let .

(1) Let be even, be the 1-dimensional central torus of inside the stabiliser in of a pair of maximal totally singular subspaces. Thus acts by scalars on both . Then is normalised by the outer elements of interchanging . Now by Lemma 2.12, a regular unipotent element in the outer coset of has Jordan blocks of sizes , hence is regular unipotent in . Then lies in the centraliser of the non-trivial unipotent element (non-trivial as soon as ), thus inside a proper parabolic subgroup. This is an example of (1) in Proposition 7.6.

(2) Let be the stabiliser of an orthogonal decomposition of , where with . Then by Reference 17, Thm B(ii)(a) there is a subgroup , with a maximal torus and a regular unipotent element of . We number the weights of on such that acts as the permutation on these. For , the group is an example for case (2). On the other hand by taking the direct product of with a subgroup of as constructed in Example 7.5, and intersecting with we find an example for (3), and as in part (1) we see that both lie inside proper parabolic subgroups.

(3) The example for in Proposition 7.2 falls into case (3); this can be seen from the weight spaces on the two modules, as the natural module for is the wedge square of the natural module for .

Proposition 7.8.

Let be a torus and of prime-power order . Then .

Proposition 7.10.

Let be simple of exceptional type and with a non-trivial torus and a regular unipotent element of . Then one of the following holds:

(1)

, , ; or

(2)

, , .

Example 7.11.

Both cases in Proposition 7.10 do actually lead to examples.

(a) Let with . By Reference 17, Thm A, there is a maximal subgroup of containing a regular unipotent element , with a 2-dimensional torus. As centralises , the subgroup of then lies in a proper parabolic subgroup and so yields an example for the situation in Proposition 7.10(a).

(b) Let with . According to Reference 17, Thm A there is a maximal subgroup in containing a regular unipotent element , with a 1-dimensional torus. Then yields an example for the situation in Proposition 7.10(b).

References

Reference [1]
William W. Adams and Marc A. Rieffel, Adjoint functors and derived functors with an application to the cohomology of semigroups, J. Algebra 7 (1967), 25–34, DOI 10.1016/0021-8693(67)90065-8. MR218423,
Show rawAMSref \bib{AR67}{article}{ author={Adams, William W.}, author={Rieffel, Marc A.}, title={Adjoint functors and derived functors with an application to the cohomology of semigroups}, journal={J. Algebra}, volume={7}, date={1967}, pages={25--34}, issn={0021-8693}, review={\MR {218423}}, doi={10.1016/0021-8693(67)90065-8}, }
Reference [2]
Timothy C. Burness, Soumaïa Ghandour, Claude Marion, and Donna M. Testerman, Irreducible almost simple subgroups of classical algebraic groups, Mem. Amer. Math. Soc. 236 (2015), no. 1114, vi+110, DOI 10.1090/memo/1114. MR3364837,
Show rawAMSref \bib{BGMT}{article}{ author={Burness, Timothy C.}, author={Ghandour, Souma\"{\i }a}, author={Marion, Claude}, author={Testerman, Donna M.}, title={Irreducible almost simple subgroups of classical algebraic groups}, journal={Mem. Amer. Math. Soc.}, volume={236}, date={2015}, number={1114}, pages={vi+110}, issn={0065-9266}, isbn={978-1-4704-1046-9}, review={\MR {3364837}}, doi={10.1090/memo/1114}, }
Reference [3]
Timothy C. Burness and Donna M. Testerman, -type subgroups containing regular unipotent elements, Forum Math. Sigma 7 (2019), Paper No. e12, 61, DOI 10.1017/fms.2019.12. MR3942158,
Show rawAMSref \bib{BT}{article}{ author={Burness, Timothy C.}, author={Testerman, Donna M.}, title={$A_1$-type subgroups containing regular unipotent elements}, journal={Forum Math. Sigma}, volume={7}, date={2019}, pages={Paper No. e12, 61}, review={\MR {3942158}}, doi={10.1017/fms.2019.12}, }
Reference [4]
Jens Carsten Jantzen, Representations of algebraic groups, 2nd ed., Mathematical Surveys and Monographs, vol. 107, American Mathematical Society, Providence, RI, 2003. MR2015057,
Show rawAMSref \bib{Ja03}{book}{ author={Jantzen, Jens Carsten}, title={Representations of algebraic groups}, series={Mathematical Surveys and Monographs}, volume={107}, edition={2}, publisher={American Mathematical Society, Providence, RI}, date={2003}, pages={xiv+576}, isbn={0-8218-3527-0}, review={\MR {2015057}}, }
Reference [5]
Peter Kleidman and Martin Liebeck, The subgroup structure of the finite classical groups, London Mathematical Society Lecture Note Series, vol. 129, Cambridge University Press, Cambridge, 1990, DOI 10.1017/CBO9780511629235. MR1057341,
Show rawAMSref \bib{KL90}{book}{ author={Kleidman, Peter}, author={Liebeck, Martin}, title={The subgroup structure of the finite classical groups}, series={London Mathematical Society Lecture Note Series}, volume={129}, publisher={Cambridge University Press, Cambridge}, date={1990}, pages={x+303}, isbn={0-521-35949-X}, review={\MR {1057341}}, doi={10.1017/CBO9780511629235}, }
Reference [6]
R. Lawther, Jordan block sizes of unipotent elements in exceptional algebraic groups, Comm. Algebra 23 (1995), no. 11, 4125–4156, DOI 10.1080/00927879508825454. MR1351124,
Show rawAMSref \bib{La95}{article}{ author={Lawther, R.}, title={Jordan block sizes of unipotent elements in exceptional algebraic groups}, journal={Comm. Algebra}, volume={23}, date={1995}, number={11}, pages={4125--4156}, issn={0092-7872}, review={\MR {1351124}}, doi={10.1080/00927879508825454}, }
Reference [7]
Martin W. Liebeck and Gary M. Seitz, Reductive subgroups of exceptional algebraic groups, Mem. Amer. Math. Soc. 121 (1996), no. 580, vi+111, DOI 10.1090/memo/0580. MR1329942,
Show rawAMSref \bib{LS96}{article}{ author={Liebeck, Martin W.}, author={Seitz, Gary M.}, title={Reductive subgroups of exceptional algebraic groups}, journal={Mem. Amer. Math. Soc.}, volume={121}, date={1996}, number={580}, pages={vi+111}, issn={0065-9266}, review={\MR {1329942}}, doi={10.1090/memo/0580}, }
Reference [8]
Martin W. Liebeck and Gary M. Seitz, The maximal subgroups of positive dimension in exceptional algebraic groups, Mem. Amer. Math. Soc. 169 (2004), no. 802, vi+227, DOI 10.1090/memo/0802. MR2044850,
Show rawAMSref \bib{LS04}{article}{ author={Liebeck, Martin W.}, author={Seitz, Gary M.}, title={The maximal subgroups of positive dimension in exceptional algebraic groups}, journal={Mem. Amer. Math. Soc.}, volume={169}, date={2004}, number={802}, pages={vi+227}, issn={0065-9266}, review={\MR {2044850}}, doi={10.1090/memo/0802}, }
Reference [9]
Martin W. Liebeck and Gary M. Seitz, Unipotent and nilpotent classes in simple algebraic groups and Lie algebras, Mathematical Surveys and Monographs, vol. 180, American Mathematical Society, Providence, RI, 2012, DOI 10.1090/surv/180. MR2883501,
Show rawAMSref \bib{LS12}{book}{ author={Liebeck, Martin W.}, author={Seitz, Gary M.}, title={Unipotent and nilpotent classes in simple algebraic groups and Lie algebras}, series={Mathematical Surveys and Monographs}, volume={180}, publisher={American Mathematical Society, Providence, RI}, date={2012}, pages={xii+380}, isbn={978-0-8218-6920-8}, review={\MR {2883501}}, doi={10.1090/surv/180}, }
Reference [10]
Martin W. Liebeck and Donna M. Testerman, Irreducible subgroups of algebraic groups, Q. J. Math. 55 (2004), no. 1, 47–55, DOI 10.1093/qjmath/55.1.47. MR2043006,
Show rawAMSref \bib{LT04}{article}{ author={Liebeck, Martin W.}, author={Testerman, Donna M.}, title={Irreducible subgroups of algebraic groups}, journal={Q. J. Math.}, volume={55}, date={2004}, number={1}, pages={47--55}, issn={0033-5606}, review={\MR {2043006}}, doi={10.1093/qjmath/55.1.47}, }
Reference [11]
Alastair J. Litterick and Adam R. Thomas, Complete reducibility in good characteristic, Trans. Amer. Math. Soc. 370 (2018), no. 8, 5279–5340, DOI 10.1090/tran/7085. MR3803140,
Show rawAMSref \bib{LT18}{article}{ author={Litterick, Alastair J.}, author={Thomas, Adam R.}, title={Complete reducibility in good characteristic}, journal={Trans. Amer. Math. Soc.}, volume={370}, date={2018}, number={8}, pages={5279--5340}, issn={0002-9947}, review={\MR {3803140}}, doi={10.1090/tran/7085}, }
Reference [12]
Frank Lübeck, Small degree representations of finite Chevalley groups in defining characteristic, LMS J. Comput. Math. 4 (2001), 135–169, DOI 10.1112/S1461157000000838. MR1901354,
Show rawAMSref \bib{Lue}{article}{ author={L\"{u}beck, Frank}, title={Small degree representations of finite Chevalley groups in defining characteristic}, journal={LMS J. Comput. Math.}, volume={4}, date={2001}, pages={135--169}, review={\MR {1901354}}, doi={10.1112/S1461157000000838}, }
Reference [13]
Gunter Malle, Generalized Deligne-Lusztig characters, J. Algebra 159 (1993), no. 1, 64–97, DOI 10.1006/jabr.1993.1147. MR1231204,
Show rawAMSref \bib{Ma93}{article}{ author={Malle, Gunter}, title={Generalized Deligne-Lusztig characters}, journal={J. Algebra}, volume={159}, date={1993}, number={1}, pages={64--97}, issn={0021-8693}, review={\MR {1231204}}, doi={10.1006/jabr.1993.1147}, }
Reference [14]
Gunter Malle, Green functions for groups of types and in characteristic , Comm. Algebra 21 (1993), no. 3, 747–798, DOI 10.1080/00927879308824595. MR1204754,
Show rawAMSref \bib{Ma93b}{article}{ author={Malle, Gunter}, title={Green functions for groups of types $E_6$ and $F_4$ in characteristic $2$}, journal={Comm. Algebra}, volume={21}, date={1993}, number={3}, pages={747--798}, issn={0092-7872}, review={\MR {1204754}}, doi={10.1080/00927879308824595}, }
Reference [15]
Gunter Malle and Donna Testerman, Linear algebraic groups and finite groups of Lie type, Cambridge Studies in Advanced Mathematics, vol. 133, Cambridge University Press, Cambridge, 2011, DOI 10.1017/CBO9780511994777. MR2850737,
Show rawAMSref \bib{MT}{book}{ author={Malle, Gunter}, author={Testerman, Donna}, title={Linear algebraic groups and finite groups of Lie type}, series={Cambridge Studies in Advanced Mathematics}, volume={133}, publisher={Cambridge University Press, Cambridge}, date={2011}, pages={xiv+309}, isbn={978-1-107-00854-0}, review={\MR {2850737}}, doi={10.1017/CBO9780511994777}, }
Reference [16]
George J. McNinch, Dimensional criteria for semisimplicity of representations, Proc. London Math. Soc. (3) 76 (1998), no. 1, 95–149, DOI 10.1112/S0024611598000045. MR1476899,
Show rawAMSref \bib{McN}{article}{ author={McNinch, George J.}, title={Dimensional criteria for semisimplicity of representations}, journal={Proc. London Math. Soc. (3)}, volume={76}, date={1998}, number={1}, pages={95--149}, issn={0024-6115}, review={\MR {1476899}}, doi={10.1112/S0024611598000045}, }
Reference [17]
Jan Saxl and Gary M. Seitz, Subgroups of algebraic groups containing regular unipotent elements, J. London Math. Soc. (2) 55 (1997), no. 2, 370–386, DOI 10.1112/S0024610797004808. MR1438641,
Show rawAMSref \bib{SS97}{article}{ author={Saxl, Jan}, author={Seitz, Gary M.}, title={Subgroups of algebraic groups containing regular unipotent elements}, journal={J. London Math. Soc. (2)}, volume={55}, date={1997}, number={2}, pages={370--386}, issn={0024-6107}, review={\MR {1438641}}, doi={10.1112/S0024610797004808}, }
Reference [18]
Nicolas Spaltenstein, Classes unipotentes et sous-groupes de Borel (French), Lecture Notes in Mathematics, vol. 946, Springer-Verlag, Berlin-New York, 1982. MR672610,
Show rawAMSref \bib{Sp82}{book}{ author={Spaltenstein, Nicolas}, title={Classes unipotentes et sous-groupes de Borel}, language={French}, series={Lecture Notes in Mathematics}, volume={946}, publisher={Springer-Verlag, Berlin-New York}, date={1982}, pages={ix+259}, isbn={3-540-11585-4}, review={\MR {672610}}, }
Reference [19]
T. A. Springer and R. Steinberg, Conjugacy classes, Seminar on Algebraic Groups and Related Finite Groups (The Institute for Advanced Study, Princeton, N.J., 1968/69), Lecture Notes in Mathematics, Vol. 131, Springer, Berlin, 1970, pp. 167–266. MR0268192,
Show rawAMSref \bib{SpSt}{article}{ author={Springer, T. A.}, author={Steinberg, R.}, title={Conjugacy classes}, conference={ title={Seminar on Algebraic Groups and Related Finite Groups}, address={The Institute for Advanced Study, Princeton, N.J.}, date={1968/69}, }, book={ series={Lecture Notes in Mathematics, Vol. 131}, publisher={Springer, Berlin}, }, date={1970}, pages={167--266}, review={\MR {0268192}}, }
Reference [20]
David I. Stewart, The reductive subgroups of , Mem. Amer. Math. Soc. 223 (2013), no. 1049, vi+88, DOI 10.1090/S0065-9266-2012-00668-X. MR3075783,
Show rawAMSref \bib{Stew13}{article}{ author={Stewart, David I.}, title={The reductive subgroups of $F_4$}, journal={Mem. Amer. Math. Soc.}, volume={223}, date={2013}, number={1049}, pages={vi+88}, issn={0065-9266}, isbn={978-0-8218-8332-7}, review={\MR {3075783}}, doi={10.1090/S0065-9266-2012-00668-X}, }
Reference [21]
Irina D. Suprunenko, Irreducible representations of simple algebraic groups containing matrices with big Jordan blocks, Proc. London Math. Soc. (3) 71 (1995), no. 2, 281–332, DOI 10.1112/plms/s3-71.2.281. MR1337469,
Show rawAMSref \bib{Sup95}{article}{ author={Suprunenko, Irina D.}, title={Irreducible representations of simple algebraic groups containing matrices with big Jordan blocks}, journal={Proc. London Math. Soc. (3)}, volume={71}, date={1995}, number={2}, pages={281--332}, issn={0024-6115}, review={\MR {1337469}}, doi={10.1112/plms/s3-71.2.281}, }
Reference [22]
Donna Testerman and Alexandre Zalesski, Irreducibility in algebraic groups and regular unipotent elements, Proc. Amer. Math. Soc. 141 (2013), no. 1, 13–28, DOI 10.1090/S0002-9939-2012-11898-2. MR2988707,
Show rawAMSref \bib{TZ13}{article}{ author={Testerman, Donna}, author={Zalesski, Alexandre}, title={Irreducibility in algebraic groups and regular unipotent elements}, journal={Proc. Amer. Math. Soc.}, volume={141}, date={2013}, number={1}, pages={13--28}, issn={0002-9939}, review={\MR {2988707}}, doi={10.1090/S0002-9939-2012-11898-2}, }
Reference [23]
Donna M. Testerman and Alexandre E. Zalesski, Irreducible representations of simple algebraic groups in which a unipotent element is represented by a matrix with a single non-trivial Jordan block, J. Group Theory 21 (2018), no. 1, 1–20, DOI 10.1515/jgth-2017-0019. MR3739341,
Show rawAMSref \bib{TZ18}{article}{ author={Testerman, Donna M.}, author={Zalesski, Alexandre E.}, title={Irreducible representations of simple algebraic groups in which a unipotent element is represented by a matrix with a single non-trivial Jordan block}, journal={J. Group Theory}, volume={21}, date={2018}, number={1}, pages={1--20}, issn={1433-5883}, review={\MR {3739341}}, doi={10.1515/jgth-2017-0019}, }
Reference [24]
Charles A. Weibel, An introduction to homological algebra, Cambridge Studies in Advanced Mathematics, vol. 38, Cambridge University Press, Cambridge, 1994, DOI 10.1017/CBO9781139644136. MR1269324,
Show rawAMSref \bib{Wei}{book}{ author={Weibel, Charles A.}, title={An introduction to homological algebra}, series={Cambridge Studies in Advanced Mathematics}, volume={38}, publisher={Cambridge University Press, Cambridge}, date={1994}, pages={xiv+450}, isbn={0-521-43500-5}, isbn={0-521-55987-1}, review={\MR {1269324}}, doi={10.1017/CBO9781139644136}, }

Article Information

MSC 2020
Primary: 20G05 (Representation theory for linear algebraic groups), 20G07 (Structure theory for linear algebraic groups), 20E28 (Maximal subgroups)
Keywords
  • Regular unipotent elements
  • disconnected subgroups
  • reductive subgroups
  • almost simple linear algebraic groups
Author Information
Gunter Malle
FB Mathematik, TU Kaiserslautern, Postfach 3049, 67653 Kaiserslautern, Germany.
malle@mathematik.uni-kl.de
MathSciNet
Donna M. Testerman
Institut de Mathématiques, Station 8, École Polytechnique Fédérale de Lausanne, CH-1015 Lausanne, Switzerland
donna.testerman@epfl.ch
MathSciNet
Additional Notes

Donna Testerman is the corresponding author.

Work on this article was begun while the authors were visiting the Mathematical Sciences Research Institute in Berkeley, California in Spring 2018 for the programme “Group Representation Theory and Applications” supported by the National Science Foundation under Grant No. DMS-1440140. The second author was supported by the Fonds National Suisse de la Recherche Scientifique grant number 200021-175571. We thank the Isaac Newton Institute for the Mathematical Sciences, where this work was completed, for support and hospitality during the programme “Groups, Representations and Applications: New Perspectives”. This work was supported by: EPSRC grant number EP/R014604/1.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 8, Issue 25, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2021 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/72
  • MathSciNet Review: 4312324
  • Show rawAMSref \bib{4312324}{article}{ author={Malle, Gunter}, author={Testerman, Donna}, title={Overgroups of regular unipotent elements in simple algebraic groups}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={8}, number={25}, date={2021}, pages={788-822}, issn={2330-0000}, review={4312324}, doi={10.1090/btran/72}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.