Elementary abelian subgroups: from algebraic groups to finite groups
Abstract.
We describe a new approach for classifying conjugacy classes of elementary abelian subgroups in simple algebraic groups over an algebraically closed field, and understanding the normaliser and centraliser structure of these. For toral subgroups, we give an effective classification algorithm. For non-toral elementary abelian subgroups, we focus on algebraic groups of exceptional type with a view to future applications, and in this case we provide tables explicitly describing the subgroups and their local structure. We then describe how to transfer results to the corresponding finite groups of Lie type using the Lang-Steinberg Theorem; this will be used in forthcoming work to complete the classification of elementary abelian -subgroups for torsion primes in finite groups of exceptional Lie type. Such classification results are important for determining the maximal -local subgroups and -radical subgroups, both of which play a crucial role in modular representation theory.
2020 Mathematics Subject Classification:
Primary 20G07; Secondary 20G41, 20-08, 20D06, 22E401. Introduction
Many open conjectures in representation theory, like the McKay, Dade, and Alperin Weight conjectures, have reductions to finite quasi-simple groups. For example, Navarro & Tiep [35] have shown that the Alperin Weight Conjecture is true for every finite group if every finite simple group satisfies a stronger condition, namely, AWC goodness. (For another example, see the recent work of Feng, Li & Zhang [19].) Verifying this latter property requires a detailed study of the finite simple groups and their covering groups. An important role in this study is played by so-called -radical subgroups and their -local structure—that is, subgroups which satisfy , the largest normal -subgroup of the normaliser —together with their normalisers and centralisers . The recent articles of Malle & Kessar, e.g. [27], give an excellent survey of the current state of these conjectures and the role of -local structure. In fact, radical subgroups are relevant in many areas of modular representation theory: for instance, defect groups of blocks are radical, the subgroup of a weight is radical, and the first nontrivial subgroup in any radical chain is radical. If the radical subgroups of are known, then the essential rank of the Frobenius category ( a Sylow subgroup) can be determined, cf. [2]. Radical subgroups are also used in the study of so-called -local geometries, cf. [28]. The classification of radical subgroups therefore is an important open problem; classifications are known for the symmetric, classical, and sporadic groups, as well as for some exceptional groups of Lie type, see [6] and the references therein.
One approach to classify -radical subgroups of a finite group is via its -local subgroups: A subgroup is -local if it is the normaliser of a nontrivial -subgroup of . It is maximal -local if is maximal with respect to inclusion among all -local subgroups of . It is local maximal if it is -local for some prime and maximal among all subgroups of . If has a nontrivial normal -subgroup, then the only maximal -local subgroup is itself. Following the notation of previous works, we say is maximal-proper -local if is -local and maximal with respect to inclusion among all proper subgroups of that are -local. Thus if then the maximal-proper -local subgroups are exactly the maximal -local subgroups. Now, if is -radical with , then is -local and for every characteristic subgroup . In particular, is contained in some maximal-proper -local , so that and is -radical in . This shows that every radical -subgroup of with is radical in some maximal-proper -local subgroup of . Since , one can show that every maximal-proper -local subgroup can be realised as the normaliser of an elementary abelian -subgroup. This approach has been used successfully in recent work [3, 4, 5, 6, 9, 10], leading to various classifications of -local and -radical subgroups in finite exceptional groups of Lie type. Similar to -radical groups, -local groups also play an important role in group theory. For example, large parts of the classification of the finite simple groups are based on the analysis of -local subgroups; one reason for this is that the fusion of -elements is controlled by normalisers of -subgroups, by the Alperin-(Goldschmidt) Fusion Theorem. We note that Cohen et al. [16] classified local maximal subgroups of exceptional groups of Lie type. However, not every maximal-proper -local subgroup is local maximal, and the details obtained in the classification of maximal-proper -local subgroups have proved useful for the classification of radical subgroups.
All of this highlights the importance of knowing the elementary abelian -subgroups of a finite group of Lie type; Griess [22, Section 1] lists more references supporting this. With the knowledge of the elementary abelian subgroups, one can attempt a classification of the maximal-proper -local and -radical subgroups. Recent efforts in this direction have focused directly on the finite groups involved (see [3, 4, 6, 9, 10] for groups of type , , ). In contrast, this paper takes an alternative approach beginning with an ambient algebraic group over an algebraically closed field. In this case, the rich geometric and Lie-theoretic structure of the group drastically simplifies many arguments, and indeed there are a number of existing results in this case [11, 22, 24, 41] which we will build upon here. Supposing one has complete subgroup structure information for a given algebraic group in positive characteristic, the Lang-Steinberg Theorem (cf. [32, §21.2]) then gives a powerful technique for transferring results to the corresponding finite groups of Lie type. This both streamlines arguments, and avoids some of the lengthy ad-hoc calculations required in previous papers.
We note that the results of this paper have already been used successfully in [5, 6] to classify the maximal -local and -radical subgroups in the finite groups of type . It is also used in [8] for work on Donovan’s conjecture for blocks with extra-special defect groups , in [1] to classify weight subgroups of quasi-isolated -blocks of , and in the (ongoing) PhD thesis of Fu to classify elementary abelian subgroups in finite classical groups. In forthcoming work, we apply the results of this paper to classify the elementary abelian -subgroups in the finite exceptional groups of Lie type for small (torsion) primes. In particular, our results are also a significant step toward a classification of the maximal-proper -local subgroups in the finite groups of Lie type; this research will be published in a separate paper.
1.1. Main results and structure of the paper
Let be a prime and let be a simple algebraic group over an algebraically closed field of characteristic . If , then we choose a Steinberg endomorphism of , with corresponding fixed-point subgroup , a finite group of Lie type. If , then the -radical subgroups of are known by the Borel–Tits Theorem [21, Corollary 3.1.5], hence we assume throughout that . Thus each elementary abelian -subgroup of consists of semisimple elements, and therefore normalises a conjugate of a fixed maximal torus . Our strategy is to consider separately the case of toral subgroups (those with a conjugate contained in ) and non-toral subgroups (those which are not toral). While we consider the toral elementary subgroups for all simple algebraic groups, we restrict ourselves to exceptional groups for the non-toral ones; classification of non-toral elementary abelian groups in the classical case is more difficult (and is sometimes related to the classification of certain codes, see [22, Table 1]). Our main results are the following:
-
(1)
A constructive method to classify the toral elementary abelian -subgroups (up to conjugacy) in a simple algebraic group and to determine information on their local structure, see Section 5. Our method translates to a practical algorithm that we have implemented for the computer algebra system Magma [13]; our implementation is available under the link provided in [7].
- (2)
-
(3)
A description of how (1) and (2) can be used to classify the elementary abelian -subgroups (up to conjugacy) and their local structure in the finite groups , see Section 4.
Remark 1.1.
In fact, our method for classifying subgroups in (1) and (2) works for any finite subgroup of order coprime to the characteristic , and if the subgroups are moreover abelian, then the arguments relating to local structure also generalise. However, our results on torality depend implicitly on the fact that the subgroup in question is an abelian -group; for this reason and the reasons above, we keep our focus on the elementary abelian case.
First consider toral subgroups. For a maximal torus of , it is well known that the Weyl group controls conjugacy and determines the normaliser structure of subgroups of . By viewing as the group of points of a suitable group scheme, it follows that the classification of toral elementary abelian -subgroups of is independent of the characteristic as long as , which we assume throughout. Importantly, this shows that all of these calculations can be performed in a suitable finite group of Lie type, which gives rise to a practical computational approach toward a classification. We give full details in Section 5.
Turning to non-toral elementary abelian subgroups of an exceptional simple algebraic group in Section 6, the number of classes of such subgroups is bounded by an absolute constant, and in many cases these have been described elsewhere in the literature: For example, the maximal non-toral elementary abelian subgroups in complex groups are described by Griess [22]. Complete information on non-toral subgroups for in groups over is given by Andersen et al. [11]. For , much information is provided by Yu [41] for adjoint compact groups. Again, all these results carry over into any characteristic different from . We present explicit tables (including new information on the local structure) and we use the opportunity to correct some typographic errors in the ancillary data of [41].
Assuming , and using knowledge of the elementary abelian subgroups and their local structure in , in Section 4 we use a consequence of the Lang-Steinberg Theorem to see how the -classes of subgroups split into subgroups of the finite groups . If some -conjugate of an elementary abelian subgroup lies in , then the -subgroup classes arising from correspond to -classes in of elements contained in , and these are known by our previous calculations. The normaliser and centraliser structure also follow in short order.
2. Notation
Throughout, unless stated otherwise, is a simple algebraic group, defined over an algebraically closed field of characteristic . When discussing elementary abelian -groups, we always assume , restating this assumption when necessary. We treat as a Zariski-closed subgroup of some general linear group over the algebraically closed field, except in Section 3.1 where we briefly mention the scheme-theoretic background necessary to transfer results between groups over fields of different characteristics. We add subscripts “sc” and “ad” to denote the simply-connected and adjoint group, respectively, of a given type.
For a prime power, we denote by the finite field with elements. We denote by a fixed maximal torus of , and define the Weyl group of as ; this does not depend on the choice of since all maximal tori in a linear algebraic group are conjugate [32, Corollary 6.5]. Recall that has an associated root datum, consisting of a root system , dual root system , character lattice and co-character lattice , with a natural pairing . There are natural notions of homomorphism and isomorphism of root data, and is determined up to isomorphism by its root datum and the field , see [25, II.1.13].
We follow the convention of Malle and Testermann [32, Section 21.2] and call an endomorphism of a linear algebraic group a Steinberg morphism if some power is a Frobenius morphism with respect to some -structure, that is, is induced by the -power map . A Steinberg morphism is an automorphism of abstract groups, but not necessarily of algebraic groups. As usual, if is a set and is a function , then we denote by the fixed-point set of . If is a Steinberg morphism of a linear algebraic group with a Frobenius morphism, then are both finite groups. We note that this implies is a Steinberg morphism as defined in Gorenstein et al. [21, Definition 1.15.1]. Lastly, we mention a celebrated result of Steinberg which shows that if is simple (the case of most interest here), then every non-trivial endomorphism of is either an automorphism of algebraic groups or a Steinberg morphism, and the latter occurs if and only if is finite, see [32, Theorem 21.5].
For an algebraic group , we denote by the connected component of the identity element. If is a finite group then denotes the largest normal -subgroup of , and if is a -group, then , with an integer, denotes the subgroup generated by elements of order dividing .
Let be groups and be positive integers. Let be a prime. We denote by the direct product of and . We sometimes write for the cyclic group of size , and we denote by the direct product of copies of . This leads to an ambiguity if is a prime power; we avoid this by writing only for the direct product of groups of order , and for the cyclic group of order . An extension, split extension, and central extension of by are respectively denoted , , and ; here is the normal subgroup. We read as , and similarly for other products of groups. For group elements we denote conjugation as usual by and . The notation denotes an extraspecial group; there are two isomorphism types of these, which are denoted by and . The alternating and symmetric group on points are denoted by and , respectively. The dihedral and generalised quaternion group of order are and . We write for the general linear group of degree over the field with elements, and , , , and for the classical groups (special linear, unitary, symplectic, and orthogonal group). Other notation is introduced at the appropriate places.
3. Preliminaries
3.1. Independence from the characteristic
The following result is of vital importance for establishing our main results. In the present form this is due to Larsen [23, Appendix A], but see also [11, p. 137], [17, Section 3], [22, Appendix A], and [20]. It tells us that the classification of elementary abelian -subgroups of is independent of the defining characteristic , as long as it is different from ; in particular, it shows that the classification of elementary abelian -subgroups of a semisimple algebraic group coincides with that of the corresponding complex Lie group . The proof makes use of deep algebraic geometry, interpreting as a group scheme over a ring of Witt vectors, which is an integral domain of characteristic that maps onto an algebraically closed field of positive characteristic.
Proposition 3.1 ([23, Theorem 1.22]).
Let be a prime number and a semisimple split group scheme; let be a finite group of order coprime to . For any algebraically closed fields and of characteristic and , respectively, the sets and have the same cardinality, where the quotients are taken with respect to the natural conjugation action of the groups on the homomorphism sets. Furthermore, the process of reduction modulo induces a bijection.
The ‘process of reduction modulo ’ here entails showing that every finite subgroup of has a conjugate contained in , where is a suitable ring of Witt vectors that maps onto an algebraically closed subfield of , and then considering the induced map ; we refer to [23, Appendix A] and [20] for more details.
Recall, for instance from [11, Proposition 8.4], that a complex algebraic group has a maximal compact subgroup in the Hausdorff topology, say , which is unique up to conjugacy and strongly controls fusion: if for subsets and , then there is some such that for all . In particular, if , then , so . Moreover, if is a finite subgroup, then embeds into the (finite) automorphism group , and since has finite index in , it has finite index in ; thus, we have , and therefore . Based on all this, the next lemma shows that we can view the -local structure of inside .
Lemma 3.2.
Let be a reductive complex algebraic group with a maximal compact subgroup , and let be elementary abelian. Then the following hold:
-
(a)
The inclusion induces a bijection on conjugacy classes of elementary abelian subgroups of .
-
(b)
The group is a maximal compact subgroup of , with the same root datum.
-
(c)
We have
Proof.
By uniqueness of , each compact subgroup of has a -conjugate contained in . Uniqueness of this -conjugate up to -conjugacy follows since strongly controls fusion in ; this proves (a). Next, by [11, Proposition 8.4(3)] the group is a maximal compact subgroup of ; this is reductive, and the inclusion sends a maximal torus of into one of , matching up the corresponding root data, giving (b). For (c), note that each coset of in contains a translate of , which is a maximal compact subgroup of . The union of these translates is a finite extension of a compact group, hence is compact, and therefore meets each such coset. Identical reasoning applies for the centraliser of . (A more detailed justification for parts (b) and (c) can be obtained using the complexification functor from compact Lie groups to complex reductive groups, as detailed, e.g., in [26, Theorems 1 and 2].) ∎
Thus the classification of elementary abelian subgroups and their local structure in a complex reductive group can be viewed inside compact real Lie groups. We will refer in particular to [11, 22, 41].
Proposition 3.1 can already be applied to finite subgroups of and for an elementary abelian subgroup , but stronger results are available since and can themselves be viewed as the points of a group scheme. The following lemma summarises a series of results of Friedlander & Mislin [20] which make use of this viewpoint. We note that the results quoted from [20] require an elementary abelian -subgroup with , but their proofs should also hold more generally.
Lemma 3.3.
Let and be algebraically closed fields of characteristic and , respectively, and let be a prime. Let and respectively be elementary abelian -subgroups of and which correspond under the bijection of Proposition 3.1. Then reduction-mod- allows us to identify the root data of the reductive groups and , and also induces isomorphisms
Moreover, the order of the component group is a power of .
Proof.
We first claim that without loss of generality we can take and , the algebraic closure of the field of elements. Since , all representations of a finite elementary abelian -group in characteristic and characteristic are semisimple, see Maschke’s Theorem. This shows that and are linearly reductive in the sense of [12, §2], thus and are -completely reducible [12, Lemma 2.6], and our first claim is then [12, Corollary 5.5].
Now we interpret as the -points of a finite subgroup scheme of , defined over an appropriate ring of Witt vectors , where comes equipped with a fixed embedding into and a surjection onto . By hypothesis, the image of in is conjugate to , so we may replace by this image without loss of generality. Then and are the -points and -points of the scheme-theoretic normaliser of in , which we denote , and similarly and are the points of the scheme-theoretic centraliser . Then [20, Corollary 4.3 and Theorem 4.4] tell us that and are generalised reductive groups over (see [20, Definition 2.1]), and applying [20, Proposition 3.1(i)] to each of these, we deduce that the quotient by the connected component is the same over and . This gives the second and third isomorphisms above. The first isomorphism is derived in the proof of [20, Theorem 4.4].
Lastly, since the scheme-theoretic centraliser is a generalised reductive group, by definition the quotient of by its identity component is a finite étale group scheme which, by [20, Corollary 4.3], has order coprime to ; note that in [20] the quotient by the identity component is denoted . We can repeat this argument for any , replacing and as appropriate, and so the order is coprime to every prime except possibly ; in other words, its order is a -power. ∎
3.2. Torality
As hinted at in the introduction, the behaviour of elementary abelian subgroups admits a stark dichotomy depending on whether the subgroups in question are toral or not. In general, toral subgroups are much more well-behaved, and non-toral subgroups are comparatively rare, as illustrated by the following result of Steinberg. To state it, recall that a prime is called a torsion prime for a reductive group if some subsystem subgroup of has -torsion in its fundamental group, see [39, Definition 2.1]. Explicitly, if is simple then torsion primes are: for of type ; for of type , , , ; for of type , , ; for of type ; the exact possibilities depend on the isogeny type of the group and are listed in [39, Lemma 2.5] and [32, Table 9.2]. If is not simple, then is torsion for precisely when it is torsion for some simple factor.
Theorem 3.4 ([39, Theorem 2.28]).
Let be a reductive algebraic group in characteristic and let be a prime. The following conditions are equivalent:
-
(a)
The prime is not a torsion prime for .
-
(b)
The centraliser is connected for every elementary abelian -subgroup .
-
(c)
Every elementary abelian -subgroup of is toral.
We close this subsection with the following useful lemma.
Lemma 3.5.
Let be a reductive algebraic group in characteristic and let be an elementary abelian -subgroup, where .
-
(a)
The subgroup is toral in if and only if .
-
(b)
Let be central. Then is toral in if and only if is toral in .
-
(c)
Suppose that and view the Lie algebra of as a -module with character . Then
Proof.
Suppose is contained in some maximal torus of . Since is connected and abelian, we have and so . Conversely, suppose and recall that is reductive, see [32, Theorem 14.2]. Since the centre of a connected reductive group is contained in every maximal torus, is contained in every maximal torus of , in particular is contained in some torus of . This proves (a). Parts (b) and (c) are proved in [11, Theorem 8.2]. ∎
3.3. Semisimple elements of small order
In our discussion of non-toral elementary abelian subgroups we will require the information provided in Table 1 below, on certain conjugacy classes of semisimple elements in exceptional groups . We give the centraliser of each such element , as well as the trace (Brauer character) on the Lie algebra of ; for we also give the trace on a non-trivial module of least dimension. By the results of Section 3.1, this information remains the same for the corresponding algebraic group in any characteristic coprime to the element order. These results can be obtained using the algorithm of Moody and Patera [34], but they have already appeared throughout the literature, in particular [18, Table 2], [15, Tables 4, 6], [34, Table 10], and [21, Tables 4.3.1, 4.3.2, and 4.7.1]. We have chosen our class labels to be consistent with these references.
If is a prime, then stands for a conjugacy class of elements of order , labelled by X. In class labels, numbers in brackets indicate that these powers form distinct conjugacy classes – for instance squares of elements in the class form a distinct class (not otherwise listed), whose traces will be the complex conjugates of the traces shown. The notation and indicates that the squares of these elements lie in class .
In the description of the centraliser, indicates a -dimensional torus. The element is a fixed cube root of unity. The isogeny types of the groups in the centralisers can be determined from the action of the groups on and the minimal module, listed for instance in [29, §8]: the correct group is the smallest group (in the sense of taking quotients) which acts faithfully on some composition factor of if is adjoint, or of the minimal module if is simply connected of type or .
Class | ||||
2A | ||||
2A | ||||
2B | ||||
3A | ||||
3C | ||||
3D | ||||
2A | ||||
2B | ||||
3A | ||||
3B[2] | ||||
3C | ||||
3D | ||||
3E[2] | ||||
2A | ||||
2B | ||||
2C | ||||
3A | ||||
3B | ||||
3C | ||||
3D | ||||
3E | ||||
4A[A] | ||||
4H[A] | ||||
2A | ||||
2B | ||||
3A | ||||
3B | ||||
3C | ||||
3D | ||||
5C |
We will also refer to the inclusions of classes in Table 2, induced by the inclusions of simply connected groups . Some of this information is given in the above references, but it can also be determined directly from the character values and the known action of each group on the Lie algebra and minimal module of the next.
class | class | class | class | |||
---|---|---|---|---|---|---|
2A | 2A | 2B | 2A | |||
2B | 2B | 2C | 2B | |||
2A | 2A |
4. From algebraic groups to finite groups
Let be a connected reductive algebraic group over the algebraically closed field of characteristic , with maximal torus and Weyl group . In the following we suppose that a Steinberg endomorphism of is given, that is -stable, and that we know, up to -conjugacy, the elementary abelian subgroups of with some representative in , as well as their local structure in . Given this information, we show how to derive a classification of the elementary abelian subgroups of the finite group , together with their local structure. The main tool for establishing this result is an application of the Lang-Steinberg Theorem. Note that the results of this section hold for all finite subgroups; we will return to the particular case of elementary abelian -subgroups in subsequent sections when applying these.
4.1. The correspondence
If , then the -class of a coset in is the subset . Moreover, recall that if is finite.
Proposition 4.1.
Let be a finite subgroup such that and are conjugate in . The following hold:
-
(a)
There exists a -conjugate of which is -stable.
-
(b)
Suppose has a -conjugate in . Replacing by this conjugate, there is a bijection between -classes of subgroups of which are -conjugate to , and -classes in contained in : the -class of corresponds to the -class of subgroups with representative , where is chosen with .
Proof.
To begin, clearly acts on transitively. Also acts on since is -conjugate to , and this action is compatible with the -action since for . Note that is closed since it is finite, hence so are and . The Lang-Steinberg Theorem then implies the existence of an -stable conjugate of , see [32, Theorem 21.11], so (a) holds, and replacing by such -stable conjugate there is a bijection between
Now suppose is fixed point-wise by , so , and abbreviate . As shown in the proof of [32, Theorem 21.11], the -class of corresponds to the -class of subgroups with representative , where satisfies . For such an element , and for any , we have , for some lift of to . Therefore acts on as the map acts on . In particular, if and only if , so that the -class of consists of elements in ; thus (b) holds. ∎
We note that Proposition 4.1 works for both toral and non-toral elementary abelian subgroups, and for both twisted and untwisted Steinberg morphisms. Importantly, Proposition 4.1 allows us to determine all -classes of elementary abelian subgroups if we know the -classes of these subgroups in with representatives in (whenever they exist). One remaining problem is to find the latter, and Corollary 4.2 below will be useful for that. Clearly, if and there is such that , then and so and are conjugate. However, the converse is not always true. For instance, let where has characteristic , let be the subgroup of diagonal matrices, and let be a Steinberg morphism inducing on . Then and has no elementary abelian -subgroups of rank for , even though contains an -stable elementary abelian -subgroup of rank ; the action of on is then non-trivial. Despite this subtlety, it remains true that for any given finite subgroup of , there is some such that contains a conjugate of this subgroup, in which case the results of this section tell us exactly how many such classes arise, as well as their normaliser structure.
Corollary 4.2.
Let be a finite subgroup.
-
(a)
There is a -conjugate of in if and only if and are -conjugate and there is an -stable -conjugate of such that the restriction is induced by some .
-
(b)
If , then there is a -conjugate of in if and only if and are -conjugate.
Proof.
-
(a)
Suppose for some . Then is -stable and is -conjugate to . The proof of Proposition 4.1 shows that acts on -stable -conjugates of as elements of act on . Conversely, suppose and are -conjugate (so -stable -conjugates exist by Proposition 4.1(a)) and that there is such a -conjugate such that is induced by , that is, for some lift of . If satisfies , then is fixed point-wise by : if with , then , so , as claimed.
-
(b)
This follows readily from the previous proof: using the notation of (a), the assumption on implies that for every -stable conjugate of . ∎
4.2. Local structure
We continue with the previous notation; moreover, we denote by the map given by . The following propositions allow us to determine the local structure of the -conjugates of in from the structure of and .
Proposition 4.3.
Let be a finite subgroup. For let be the -conjugate of as defined in Proposition 4.1. If is any lift of , then
where, as usual, are the fixed points of in . Furthermore, is independent of the choice of . If acts as an inner automorphism of , then .
Proof.
Write . By construction, for some with and . Clearly, . If , then if and only if for , hence
Since is closed and -stable, so is . Now [21, Theorem 2.1.2(d)] proves that the restriction of to is a Steinberg endomorphism (in the sense of [21]), that is, is surjective (hence a bijection since is an abstract group automorphism, hence injective) and is finite. In particular, is a Steinberg endomorphism in the sense of [32] and as considered here. It follows from [21, Proposition 1.1.4(c)] that , so is -stable, and now [32, Exercise 30.1] shows that is a bijection. In conclusion, induces a Steinberg automorphism of . If with is another lift of , then the restrictions of and to differ by an inner automorphism of , and [32, Corollary 21.8] shows that and are isomorphic, as claimed. In particular, if induces an inner automorphism on , then the map equals for some , and . ∎
Proposition 4.4.
Let be a finite subgroup, let , and let be the -conjugate of under the correspondence in Proposition 4.1. Then
Proof.
Write and . By construction, we have for some such that . Let . If , then
Thus, we have an isomorphism , which also induces isomorphisms , , and . Since is connected, [21, Theorem 2.1.2(f)] shows that and are -conjugate, hence both finite. Since is also bijective on , it is a Steinberg morphism (in the sense of [21]). Recall that and are closed and mapped under into themselves. Now [21, Theorem 2.1.2(d)] shows that induces Steinberg morphisms on and on , respectively. Applying [21, Theorem 2.1.2(d)] to and to , respectively, we see that maps and onto themselves and that the natural maps and are surjective. The claim follows. ∎
4.3. The -action on
The previous results require us to study the action of on the normaliser quotient.
Proposition 4.5.
Let be a finite subgroup. The map induced by is given by , where is a -cocycle , that is, satisfies
for all . In particular, induces the identity map on .
Proof.
By hypothesis, if and then , thus and the first statement holds. Writing , we have , from which the cocycle condition on follows. ∎
While acts trivially on , it does not necessarily act trivially on , even if the action on is trivial:
Example 4.6.
Let and , with induced by . Let be a Sylow -subgroup and note that and . If is a primitive element of , then satisfies . This shows that there is only one -class of -orbits in . In particular, acts as the identity on , see Proposition 4.5, and on , but acts non-trivially on .
Corollary 4.7.
Let be finite. If acts trivially on , then there is a bijection between the -classes in and the -classes of subgroups of that are -conjugate to ; this bijection maps an element to , as defined in the correspondence in Proposition 4.1; moreover
and
4.4. Cohomology
Recall from [36, I.§5] that if is a group which acts on a group by automorphisms, then the cohomology set is the set of all -cocycles, modulo the equivalence relation where if there exists such that for all . When is abelian, this coincides with the usual definition of via right-derived functors of Hom. In general, is not a group, but a pointed set, the distinguished point given by the class of the map for all .
The next result shows that for certain and given in the algebraic group, we find a -conjugate with the same normaliser quotient in .
Corollary 4.8.
Let be a finite subgroup with . There exists a subgroup which is -conjugate to , such that
Moreover, the -classes of are precisely the conjugacy classes of , and for the -stable conjugate corresponding to the -class of we have
Proof.
The hypothesis implies that every -cocycle is of the form for some . Thus, by Proposition 4.5, there is such that , . Now define as the -stable -conjugate of corresponding to the -class of , so that the first claim follows from Proposition 4.4. Recall that we can define for some with . Since for all and for some , it follows that every is fixed by . Together with Proposition 4.4, this implies the second claim. ∎
5. Toral subgroups of simple algebraic groups
Recall the notation we have fixed in Section 2, with a simple algebraic group in characteristic . Using [22, (2.13)(iii)], we can assume that a given elementary abelian -subgroup (with is contained in the normaliser of our fixed maximal torus of . We study separately the cases that is toral or non-toral: in the former case we assume that ; in the latter case, has non-trivial image in the Weyl group .
Our goal in this section is to describe a practical algorithm for enumerating such toral elementary abelian -groups and determining their normaliser structure. We begin with Proposition 5.1 (as stated below) which simplifies our calculations: Part (a) enables us to deduce all relevant information working solely within . Part (b) then allows us to describe the normaliser of a toral subgroup by splitting it into canonical sub-quotients which can be studied separately and recombined. Following this proposition, in Section 5.1 we then describe our algorithm for enumerating elementary abelian subgroups and determining their normaliser structure. Finally, in Section 5.2 we describe a process for extending this classification of subgroups of to a classification of subgroups in the corresponding finite groups of Lie type, that is, in .
Much of the argument below applies equally well to arbitrary abelian subgroups (not just elementary abelian), but we concentrate on elementary abelian subgroups for two reasons: Firstly, methods from linear algebra can be brought to bear, allowing for a very efficient classification algorithm (see Section 5.1 and our implementation [7]); secondly, as explained in the introduction, elementary abelian subgroups have many interesting applications, for example, in the classification of maximal -local and -radical subgroups.
In the following, recall that indicates a (possibly non-split) extension with normal subgroup and quotient .
Proposition 5.1.
Let be a simple algebraic group, with maximal torus and Weyl group . If are finite subgroups, then the following hold.
-
(a)
If with , then for some and ; in particular, and are conjugate in .
-
(b)
We can decompose , with isomorphisms
where is the Weyl group of the reductive group .
Proof.
-
(a)
Note that . If , then is semisimple, and [32, Theorem 14.2] shows that is a connected reductive group containing and . Since is finite, induction proves that contains and ; these are therefore -conjugate, hence for some , that is, .
-
(b)
In the following, write ; clearly, . It follows from Part (a) that , and so
Note that
These isomorphisms allow us to identify as a subgroup of , and we deduce . ∎
5.1. Computation of toral elementary abelian subgroups
Continuing with the previous notation, we now explain how to classify, up to conjugacy, the toral elementary abelian -subgroups of algorithmically, by working in a suitably-chosen finite group of Lie type. For technical reasons, we first assume that is odd, and treat the case in a moment.
Recall that is a fixed maximal torus of with corresponding set of roots . First, by Dirichlet’s Theorem we can choose a prime-power such that divides ; by Proposition 3.1, we may assume that the characteristic divides . Second, we can choose a Steinberg morphism of which induces the -power map on . This gives rise to the finite group , containing the fixed-point subgroup which is a product of cyclic groups of order . The subgroup of is defined as ; it contains a representative of each toral elementary abelian -subgroup of .
Recall from Proposition 5.1(a) that two toral subgroups of are conjugate in if and only if they are conjugate under the action of the Weyl group . In we therefore consider the extended Weyl group of , see [38, Notation 2.1]. Originally defined by Tits [40, § 4.6] when is simply connected, is the subgroup of generated by elements , where are root elements of and ranges over the simple roots of . By construction , and is an elementary abelian subgroup of order , where is the rank of ; the corresponding quotient is the Weyl group . Specifically, normalises , and . Our choice of and [14, Section 12.1 and p. 61] imply that , and therefore the action of on can be viewed inside the finite group . In our setting, may not be simply connected but the same definition still produces a subgroup normalising and , and mapping surjectively to under the quotient map . We compute as the image of defined by its action on , and we determine all toral elementary abelian -subgroups of , up to conjugacy, by computing the -conjugacy classes of elementary abelian subgroups contained in . In the following, let be such an elementary abelian subgroup; we now describe how to determine the structure of within .
By Proposition 5.1(b), to describe the structure of we need to determine and the finite groups and . As in the proof of Proposition 5.1(a), it follows from [32, Theorem 14.2] and a finite inductive argument that is generated by and those root subgroups which commute with . To find which root subgroups we need, it suffices to check which of the finitely many elements commute with ; note that each lies in by our choice of . This allows us to compute the root datum of . Since , adding a suitable toral subgroup determines most of the structure of ; the remaining work is to determine the isogeny types of the occurring simple factors in . It now remains to understand and . The action of on restricts to an action on the characteristic subgroup , giving a surjective homomorphism . The only algebraic endomorphisms of the multiplicative group are power maps, and it follows that the full group of algebraic automorphisms of is isomorphic to . Moreover, the restriction to is precisely the homomorphism given by reducing matrix entries modulo . Restricting this map to the finite subgroup gives us the above map . A result of Minkowski [33], see also [37, Lemma 1], tells us that the kernel of is torsion-free for all integers , thus the induced map is injective. This proves that we can compute the entire structure of inside ; in conclusion, the structure of is determined.
When , the above argument goes through similarly, with the exception that Minkowski’s Lemma fails to apply to the map ; indeed, the kernel can contain elements of order in this case. For this reason, instead of working with , we pick so that , and let be the characteristic subgroup of generated by elements of order dividing . Then the -orbits on elementary abelian -subgroups of coincide with the -orbits on elementary abelian -subgroups of , where is the image of in . We can now apply Minkowski’s Lemma as above to deduce that the map is an isomorphism. This proves that for each elementary abelian -subgroup , the structure of is again determined by working in the finite group .
A Magma implementation of this algorithm is available at [7].
5.2. Translation to finite groups
We use the notation of Section 4, that is, is a connected reductive algebraic group over the algebraically closed field of characteristic , with maximal torus and Weyl group . The prime is different from , and is a (possibly twisted) Steinberg endomorphism such that is -stable. The aim of this section is to describe how to classify, up to -conjugacy, the toral elementary abelian subgroups of , together with their local structure in , assuming that this information is known for . The approach is based on the following lemma (and results of Section 4), and summarised in the subsequent remark.
Lemma 5.2.
If is toral, then
and acts trivially on .
Proof.
Let be the natural projection. Its restriction to has kernel ; thus, the first claim follows if we show that this restriction is surjective. For this recall that the extended Weyl group satisfies , see Section 5.1. Moreover, Proposition 5.1 shows that if , then for some and . Since and , we can assume that and ; this shows that is surjective. The claim about the centralisers follows analogously. ∎
Remark 5.3.
For a simple algebraic group , to classify all toral elementary abelian -subgroups of up to conjugacy, we now have the following recipe.
- (a)
-
(b)
For each such representative , determine the -classes in .
- (c)
In Step (b), we note that if , then the -class of contains only one -class, and and . If , then is a torsion prime by Theorem 3.4, and some work is required in Step (c).
6. Non-toral subgroups of exceptional algebraic groups
We now turn our attention to non-toral elementary abelian subgroups. In this case, the analogue of Proposition 5.1 fails, that is, the normaliser structure of such a subgroup is not controlled by the normaliser of a maximal torus, and thus more ad-hoc calculations are required. Nevertheless the results of Section 3.1 still apply, allowing us to transfer many known results from the characteristic setting.
Let be an exceptional simple algebraic group, still over an algebraically closed field of characteristic , different from a fixed prime . Each non-toral elementary abelian -subgroup of is contained in a maximal such subgroup, which have been classified up to conjugacy by Griess [22] for groups over . Also for complex groups, when is odd, a complete description of non-toral elementary abelian -subgroups and their normaliser structure is given by [11, Section 8]. When and is adjoint, much information regarding the collection of all elementary abelian -subgroups is given in [41]. Comparing this with the information on toral subgroups provided by our algorithm, this allows us to derive a complete list of non-toral subgroups, and it then remains to determine properties of these subgroups.
We summarise these results in the subsequent tables. In these tables, we also give the distribution of elements among the conjugacy classes of the group . For an elementary abelian -subgroup of , and for conjugacy classes of order- elements of labelled , , etc. in Table 1, we write to indicate that , , and so on. So for instance in Table 3, the group labelled has two elements in class and twenty-four in class (together accounting for all non-identity elements).
6.1. Statement of results for odd
In view of Proposition 3.1 and Lemma 3.3, the next proposition follows from [22, §1] and [11, §8]. The class distribution of the non-toral subgroup in is derived in the proof of [22, Theorem 7.4].
Proposition 6.1 (Griess; Andersen et al.).
Let be an odd prime and let be a simple algebraic group of exceptional type over an algebraically closed field of characteristic . If is a non-toral elementary abelian -subgroup, then and appears in Table 3. Each line of the table corresponds to a unique -conjugacy class; supplementing comments are given in Remark 6.2.
\cdashline2-5 | ||||
---|---|---|---|---|
\cdashline2-5 | ||||
\cdashline2-5 | ||||
\cdashline2-5 | ||||
\cdashline2-5 | ||||
\cdashline2-5 | ||||
Remark 6.2.
These remarks supplement Proposition 6.1.
- (a)
-
(b)
For adjoint of type , in [11, Theorem 8.10] the distribution is given of the pre-image of in the simply connected cover of . Since multiplication by an element of the centre permutes , , and elements whose square is in (while preserving the class of elements in and ), it is straightforward to derive the above distribution of from that of , and vice-versa. The notation denotes the unique conjugacy class which is the image of the classes and (and also the inverse of elements in ). The inverse of the class is denoted .
-
(c)
For of type , the classification is independent of the isogeny type, because the simply connected group has centre of order , and now the Schur-Zassenhaus Theorem implies that for each elementary abelian -subgroup of the adjoint group lifts uniquely to an isomorphic copy in the simply connected group. In the table, the group is the kernel of the isogeny from the simply connected cover of , that is, if , and if and .
6.2. Statement of results for
We now consider elementary abelian -subgroups. Again, by Proposition 3.1 and Lemma 3.3 we are able to use results concerning complex Lie groups and compact real Lie groups, in particular [22] and [41]. However, in this case complete information is not given, and we supplement existing results with arguments for , as well as our own normaliser structure calculations. Since the lists of subgroups for and are much larger than the other cases, we separate these out for readability. For the same reason, we have outsourced some more technical considerations to the appendix.
Proposition 6.3.
Let be a simple algebraic group of exceptional type over an algebraically closed field of characteristic . If is a non-toral elementary abelian -subgroup, then appears in Table 5, Table 6, or Table 7. Each line of a table corresponds to a unique -conjugacy class of subgroups; supplementing comments are given in Remark 6.4.
Remark 6.4.
These remarks supplement Proposition 6.3.
-
(a)
For of type , the classification is independent of the isogeny type, because the simply connected group has centre of order or , and the same argument as in Remark 6.2(c) applies. In the table, the group is the kernel of the isogeny from the simply connected cover of ; thus, if , and for and .
-
(b)
For adjoint of type , the notation denotes the unique conjugacy class which is the image of the classes and . We abuse notation and write and for the involutions coming from the corresponding classes in the simply connected cover of .
- (c)
The remainder of this section is devoted to a proof of Proposition 6.3. Using the algorithm of Section 5.1, we find that the number of classes of toral elementary abelian -subgroups is as given in Table 4.
Number | |||||
---|---|---|---|---|---|
Total |
\cdashline2-5 | ||||
---|---|---|---|---|
\cdashline2-5 | ||||
\cdashline2-5 | ||||
\cdashline2-5 | ||||
\cdashline2-5 | ||||
\cdashline2-5 | ||||
\cdashline2-5 | ||||
Name | ||||||
Name | ||||||
It follows from our description in Section 5.1 that the number of toral subgroups does not depend on the isogeny type (although the class distributions and normaliser structure of the subgroups do vary). On the other hand, [41, Theorem 1.1] tells us that for a group of type , , , , and we respectively have exactly , , , , and classes of elementary abelian -subgroups in . If is adjoint, then , unless has type in which case (by inspecting [41, §3.4]) we see that contains classes of elementary abelian -subgroups; using the notation of [41], these are the classes in ‘Class 3’ and ‘Class 4’.
Thus for adjoint groups of type , , , , and there are respectively , , , , and classes of non-toral elementary abelian -subgroups. So if we verify the information stated in the tables, it becomes evident that all the listed subgroups are non-conjugate, and counting then shows that these are all non-toral subgroups with elementary abelian image in . For the simply connected group , which is not isomorphic to its adjoint version as the characteristic is not , we provide our own detailed calculations, showing explicitly that the given classes of subgroups above constitute all non-toral subgroups.
We now split the proof into several cases.
6.3. Proof of Proposition 6.3 for .
This follows from [22, Table I].
6.4. Proof of Proposition 6.3 for .
The collection of all elementary abelian -subgroups of is described in [41, Proposition 5.2]. These twelve subgroups are labelled for and , and it follows directly from the proof of [loc. cit.] that the class distribution of is . Using our algorithm from Section 5.1, we find that of these possible class distributions are the distribution of a toral elementary abelian -subgroup; the exceptions are , , and . We conclude that has exactly three conjugacy classes of non-toral subgroups (cf. also [22, Table I and Theorem 7.3]).
The structure of for each subgroup occurring is given in [41, Proposition 5.5]; it remains to consider their centralisers. Using Lemma 3.5, the subgroups of order , , and have centralisers of dimension , , and , respectively. By [31, Theorem 1], the group has a maximal connected subgroup of type , where the factor is adjoint as this maximal subgroup is not contained in an involution centraliser in . The non-toral subgroup of this factor, having all involutions conjugate, must be non-toral in . Thus is a representative of the -class of non-toral subgroups . Now centralises a subgroup , and we deduce that . Moreover, normalises ; since the latter has no outer automorphisms, . From the maximality of the subgroup of type it follows that , hence . Since is the unique non-toral subgroup and , we have .
If is non-toral of order , then is the centraliser of a non-central involution of , and is therefore . Finally, when is non-toral of order we have .
6.5. Proof of Proposition 6.3 for .
Recall that is the kernel of the isogeny from the simply connected cover of . From [22, Table I] it follows that, up to conjugacy, the three non-toral subgroups of the simply connected group of type are precisely the images of the non-toral subgroups of under the inclusion : in the terminology of [22], this is because these are each precisely the subgroups of complexity at least 3. Their class distributions in Table 5 follow from their distributions in . Using Lemma 3.5, the centralisers of , , and have dimension , , and , respectively. By [22, Remark 8.3], the group has a maximal subgroup of type . The unique non-toral is non-toral in (since all its involutions are conjugate), and extends to a non-toral contained in . The known dimension tells us that is precisely the factor of type . We claim that its centre is . It follows from [31, Tables 10.1 and 10.2] that this factor has a natural -dimensional module (highest weight ‘’) as a composition factor of the -dimensional module, hence this factor has a non-trivial centre when is simply connected; on the other hand, all composition factors of this subgroup on have highest weight ‘’ or ‘’, hence occur as sub-quotients of . Thus the centre of this subgroup acts trivially on , and so lies in . Since is maximal, the normaliser of in is , hence . Since is a direct product and , we have
Since , we observe that is a -dimensional reductive subgroup of , and is therefore a Levi subgroup isomorphic to . The centraliser normalises this , hence is equal to . Next, is reductive and -dimensional, hence is a subtorus of the subgroup above. The centraliser is contained in and normalises this torus, hence is the product of with an elementary abelian subgroup of . Every such subgroup is toral in by [32, Corollary 14.17], hence .
Finally, the groups are given in [41, Proposition 6.10], but we correct errors for and . Note that the three non-toral groups are , , and as defined in [41, p. 272], see also Case above. Since and , we have . Note that contains involutions and , so is a proper subgroup of . But is a maximal (parabolic) subgroup of , so . Similarly, , since is maximal.
6.6. Proof of Proposition 6.3 for .
By [22, Table I] there exists a unique maximal non-toral subgroup of with
We now determine a set of generators for which allow us to easily determine the class of an element of , and to concretely realise as the subgroup of of all matrices of the form
(6.1) |
where , , and denotes an arbitrary matrix.
First, write for the non-toral subgroup arising from . This remains non-toral in by [22, Remark 8.4], has distribution in , and we have shown above that it lies in the factor of a subgroup of type , maximal among connected subgroups of . Since the centraliser in of the factor contains a subgroup , by inspecting [29, Table 8.2] it follows that the connected centraliser of this factor is simple of type . Using Lemma 3.5, we deduce that has dimension , and therefore is simple of type . Inspecting [29, Tables 8.2 and 8.6] also shows that the centre of this subgroup of type equals since it acts trivially on . Now the subgroup of of type is a maximal subgroup of , and so each factor is the full centraliser of the other. Hence . It follows that
The above also shows that . Since and , we obtain
We now use this decomposition to describe the -classes of elements of and the action of . From [31, Table 10.2], the subgroup acts on the -dimensional -module with two composition factors; one is a tensor product of the natural - and -dimensional modules for and , denoted in [31] by their highest weights ‘’ and ‘’. The other is a -dimensional irreducible -module, (highest weight ‘’). Straightforward weight calculations show that the alternating third power of is isomorphic to as an -module. Hence the eigenvalues of an element on can be determined directly from those on . In particular, letting , , and denote the characters of the three respective modules, if and , then ; see Section 3.3 for the definition of . For , let be a generator of the centre of the -th factor in , and write . Each is -conjugate to a diagonal matrix , hence has trace on and on . The products have trace on and on , and has trace on and on . Therefore if and , we find
which allows us to determine the distribution of via Table 1 as . Now, acts on , and hence on , via its subgroup . The subgroup is then the unique subgroup of containing all -involutions not lying in the -orbit , and is therefore -invariant. Thus if then preserves the direct-product decomposition , and taking the elements shown as an ordered vector space basis for , we realise as the set of matrices as in (6.1) above.
Now define , and note that , so is non-toral by Lemma 3.5 and has distribution by Table 1. A direct calculation with (6.1) shows that has eleven orbits on subgroups of containing a conjugate of or : in addition to the subgroups , , and , there are subgroups
Each such subgroup is non-toral, and their distributions (given in Table 5) show that the subgroups are pairwise non-conjugate in . There are further -orbits on subgroups of of rank or more, and each of these has an -conjugate contained in the subgroup . We claim that this subgroup, and therefore all of its subgroups, are toral. By construction, is contained in a subgroup and is a central subgroup of , which is in turn contained in a simple subgroup of type , and these subgroups and commute. Moreover, is toral in , and is toral in by Lemma 3.5. The claim follows.
Centralisers. We now determine centralisers. Each subgroup above is generated by either or together with some involutions in . The -centraliser of a non-central involution is , and it is then straightforward to see that the -centraliser of an elementary abelian -subgroup of is either itself, or , or . This determines the centraliser of every non-toral subgroup of .
Normalisers. It remains to determine the normaliser quotient for each subgroup . Firstly, the non-toral subgroups of remain non-toral in under the inclusion as they each contain . These three subgroups are each contained in a subgroup by construction, and then [41, Propositions 5.4 and 5.5] show that every possible class-preserving automorphism of such a subgroup is already induced by conjugation by an element of , hence the subgroups , , and with distributions , and have the same normaliser quotient as in , and in . For , any automorphism induced by conjugation in must preserve the subgroup as this is generated by all -involutions in . The group of automorphisms of centralises , and the stabiliser of in induces all such automorphisms on . It follows that is the stabiliser of a -dimensional subspace.
A direct calculation shows that if or , then contains a unique -pure subgroup of order , namely , and so . Since , , and are involutions in lying in distinct -conjugacy classes, and since is centralised by the action of , it follows that . The group contains a unique subgroup with distribution , and this in turn contains a unique involution in class ; thinking of as a subgroup of , it is contained in the set of matrices of the form
with . The -stabiliser of induces all such automorphisms, giving the stated normaliser structure . Similarly, contains a unique subgroup with distribution and preserves the direct-sum decomposition . Moreover, contains a unique involution in class . Hence can be identified with a set of matrices of the form
with . Again, induces all such automorphisms, hence we obtain . If , then contains a unique subgroup with distribution , and is a conjugate of . Thus, is contained in . The subgroup centralises and thus , while the subgroup contains an element swapping and (taking to be the central element of without loss of generality), and acting trivially on the subgroup of . Hence . We again find that induces all such automorphisms on , so . Finally, let . This contains a unique subgroup with distribution , and in turn contains a unique subgroup with distribution . Thus is contained in the set of all matrices of the form
with . Once again, induces all such automorphisms on , so is determined.
6.7. Proof of Proposition 6.3 for
Let and . The three classes of involutions in are denoted by , , and , corresponding to the classes in the simply connected cover of which map onto them (see Section 3.3). In [41], Yu has identified 78 classes of elementary abelian -subgroups of and determined their normaliser quotients. However, it remains to determine their centraliser structure and whether or not these subgroups are toral. In the course of determining this information, we have found some typographic errors in the ancillary data of [41], particularly in [41, Remark 7.27], although we agree with the eventual classification itself. For this reason, we now carry out a number of our own calculations, verifying much of the information in [41].
The toral elementary abelian -subgroups of can be determined with our algorithm in Section 5.1. We find that there are conjugacy classes of such toral subgroups. It follows from [22] that has two maximal non-toral subgroups up to conjugacy, of orders and , respectively. Between them, these contain representatives of each class of non-toral subgroups; since the subgroup of order is obtained by adjoining a Chevalley involution to a maximal toral subgroup of order , it also contains a representative of every class of toral subgroups.
By [22, Table 1], the maximal non-toral subgroups and are each self-centralising, and have normaliser quotients as follows
Because and are finite, and their actions on and are known (cf. [41, Proposition 7.27] and Appendix A.2), we can explicitly construct these groups as matrix groups of degree and , respectively. In particular, we can calculate the orbits of these groups on the subgroups of and , using Magma [13]. Our computation also returns the complete subgroup lattice for these groups. Appendix A.3 comments on our construction. Recall that every non-toral subgroup of appears as a subgroup of at least one of the maximal and . The (known) distributions of and allow us to identify the distributions of the subgroups in each orbit. From this, together with the information on toral subgroups already calculated, we can determine for most subgroups occurring whether the subgroup is toral or non-toral. A few cases require further arguments, and for these we defer to information from [41]. In the end, we are able to identify conjugacy classes of non-toral elementary abelian -subgroups of ; these are listed in Table 6. We label these groups according to the classification in [41], see Appendix A.1 for a description of this classification. Our investigations yield the following corollary.
Corollary 6.5.
The complete Hasse diagram of non-toral -subgroups in is given in Figure 1.
The structure of the normaliser quotients of the groups in Table 6 has been determined in [41, Proposition 7.26], see also Appendix A.2 for more details. In the remainder of this section, we discuss the centralisers of the non-toral subgroups; for this the inclusions implied in the Hasse diagram (Figure 1) will be useful. In most cases, the centraliser of with is determined in where for some non-toral ; the column labelled “” in Table 6 lists this subgroup, see Case (2) in the proof of Proposition 6.6.
Proposition 6.6.
If is as given in Table 6, then is as given in that table.
Proof.
We consider the following sets of groups as defined in Table 6:
and make a case distinction.
-
(Case 1)
Suppose . The centraliser of each Klein four-subgroup of a compact simple real Lie group of type is given in [24, Table 6], and in our three cases , this is the direct product of the non-toral subgroup in question with a maximal connected subgroup (in the real topology) of the centraliser in the complex algebraic group; cf. [21, Table 4.3.3]. (Note that [24, Table 6] lists the centraliser of each Klein-four subgroup of , and the latter is a maximal compact subgroup of since is adjoint.) The centraliser structure for the subgroups of rank in Table 6 follows.
To consider and , we need some preliminary results. Let , so that where . By [21, Tables 4.3.1–4.3.3] there is an involution such that
where the central product is over . Note that , and where generates . It follows also from [21, Tables 4.3.1–4.3.3] that has centraliser , where the central product is over . This shows that . Since is generated by , it follows that is contained in , so lies in . A direct computation shows that there is a toral with such that and , and . Let for some involution . Since , it follows that and .
By [21, Tables 4.3.1–4.3.3], there is a projective involution such that , where with being the characteristic of . Choose a subgroup isomorphic to with . By [21, Tables 4.3.1–4.3.3], we may suppose , and so ; this shows that is a projective involution as well. By [21, Tables 4.3.1-4.3.3], we may suppose , so
Now write with and . It follows from [21, Table 4.3.1] that acts via inversion on ; note that normalizes . This implies that centralises , so centralises . By [21, Table 4.3.1], the element induces on ; here and denote a graph automorphism and inversion, respectively. It follows that acts via inversion on and induces a graph automorphism on . Note that acts on both and . Moreover, we have , and it follows that centralises . If , then centralises at least two distinct involutions in , which is impossible; this proves that . Let , so that . Note that is toral in and . A direct computation shows that and it yields that
where acts via inversion on and induces a graph automorphism on .
Now consider for some ; in particular, , and induces a graph automorphism on . By [21, Tables 4.3.1–4.3.3], there are two graph automorphisms, and , such that and , and and . Note that each group is non-toral. Comparing dimensions, we can therefore suppose that
in particular, we can assume that is a subgroup of and . If , then it follows from above that , so and , as claimed. If , then and each . Write and note that
This yields , and it then follows that and , as claimed. Similarly, we deduce that
Let such that . Since acts trivially on , it follows that , and so has order or . Replacing by , if necessary, we may suppose that . Now , and so . Since , it follows that and therefore . This shows that , as claimed.
-
(Case 2)
Suppose . It follows from Figure 1 that for some non-toral elementary abelian subgroup containing some subgroup , and an involution in . In particular,
for some and and , see Table 6. Thus, is a reductive group and acts on . The centralisers of involutions of can be determined by [21, Tables 4.3.1–4.3.3]; note that and . We use this information to determine ; the subgroups we use are listed in Table 6.
-
(Case 2a)
First suppose . By the choice of , it follows from [21, Tables 4.3.1–4.3.3] that is uniquely determined, up to conjugacy in , by the dimension , and so can be determined by [21, Tables 4.3.1–4.3.3]. For example, if , then has dimension . By Figure 1, we have with and . It follows from [21, Tables 4.3.1–4.3.3] that contains a unique involution such that . In fact, , thus
The centraliser structure of the other groups can be obtained similarly; we give a few details for with :
-
We have with ; since is simply connected, the centraliser is connected, hence .
-
We have and . Consider the element and let be a permutation matrix corresponding to , so that , , and swaps the two factors of . If , then the outside 2 of centralises , and so does and ; hence .
-
We have with . Let be in with and let be the permutation matrix corresponding to . Then and . Note that the graph automorphism in also gives ; if and , then .
-
We have with . Write with and , and note that , so that with . Similarly, and . Let and such that and , so that inverts and induces inverse-transpose on . Thus, inverts and induces a graph automorphism on . Now is an involution with . The element centralises , inverts , and induces on .
Note that with ; choose with and . For note that acts on but the graph automorphism acts like an inner automorphism of ; thus we can replace by .
-
-
(Case 2b)
Suppose . The methods described in Section 5.1 assist us with determining . For example, if , then is the only subgroup of size with . We take , so , and a direct computation shows that has an elementary abelian subgroup such that . Thus, and by the uniqueness of . Note that the action of is also explicitly determined. The centraliser structure of the other can be obtained similarly. For example, the group can be defined via a subgroup and an involution with , where swaps the two factors . For the group we can choose with each , and the outside 2 of centralises . If for some , then also centralises , so .
-
(Case 2c)
Suppose . First, consider , so that and has dimension . In addition, , where induces a graph automorphism on . Let and , so and, by [21, Tables 4.3.1–4.3.3], we have , where , and inverts and interchanges the two factors of . Note that we may take to be the permutation matrix corresponding to , so that , and hence . Since stabilises each , we have that induces an inner automorphism on ; by replacing by for some , we may suppose that centralises each . Note that we may suppose , hence , thus where the outside 2 inverts . Since , we may suppose , and hence with
where commutes with , inverts , and interchanges the two factors of .
Now let , so that , and let with given three lines above. By Figure 1 and Table 6, we have for some involution and . In particular, it follows from Table 6 that and are the only subgroups of size whose centralisers have dimension 3. Thus, contains exactly two involutions whose centralisers have dimension 3; we now construct two such centralisers. In the following we identity .
Note that and , where , thus
and so . Let , where . So , , and with acting via inversion (see [21, Tables 4.3.1–4.3.3]). Note that and , so
In particular, . If , then stabilises , and so it also stabilises . Thus is a subgroup of the parabolic subgroup . This is impossible, since , see Table 6. In conclusion, , and so .
-
(Case 3)
Let , so that , and let with . Let be acting via inversion on , so that . Now Table 6 implies that . ∎
6.8. Proof of Proposition 6.3 for .
Yu [41, §8] has classified all elementary abelian -subgroups of , see Appendix B.1 for details and group labels; the non-toral groups are listed in Table 7. The structure of the corresponding normaliser quotients is also determined in [41], see Appendix B.2 for details. In the following we determine the structure of the centralisers.
Preliminary subgroups. By [21, Tables 4.3.1–4.3.3], there is an involution whose centraliser is , where
Note that for a suitable Steinberg morphism ; by [4, Lemma 6.3], the group has subgroups such that
We fix this notation in the following and write for , so that each .
We first determine . The proof of [4, Lemma 6.2] yields . Note that and . It follows from [21, Tables 4.3.1–4.3.3] that , and acts on . Thus
By [21, Tables 4.3.1–4.3.3], we have ; moreover, we deduce from that ; similarly, follows. Lastly, note that and acts on . Since , it follows from [21, Tables 4.3.1–4.3.3] that .
Now let
such that with . Each has type and satisfies
In particular, each is non-toral, and the dimensions listed in Table 7 determine
A correspondence. We now show a correspondence between non-toral subgroups of and non-toral subgroups of . First, consider an elementary abelian subgroup
where and . Note that is non-toral, and by Table 6. Recall from the previous paragraph that and define
so that . If , then and . If denotes the projection from onto , then and . In particular, is a non-toral elementary abelian subgroup of and
Conversely, if and is an elementary abelian subgroup. Then is non-toral, is elementary abelian, and is non-toral with .
The following lemma will be useful.
Lemma 6.7.
With the previous notation, .
Proof.
Since , we have and , so . Note that , so and
Moreover, since . Now shows that . ∎
Centralisers. There are subgroups with names , , etc. in and in , respectively; in the following, we write “” and “” to indicate which subgroup is meant.
Let be as in the previous paragraph, so that is a non-toral subgroup of . Note that , and is listed in Table 7 and is given in Table 9. In most of the cases, is uniquely determined by , and so we can identify , therefore determining . For example, if , then , so and . Now Table 9 implies that with , hence
Similar proofs work for the subgroups of in Table 7 that are not in
In the following we suppose that with , but cannot be determined uniquely by the dimension of ; this includes the groups in .
Note that if , then its normaliser structure and [22, Table 1] show that is maximal non-toral and . According to Table 9, the only other non-toral with centraliser of dimension 0 are , both with centralisers isomorphic to . However, has no as subgroups; this determines the centralisers of .
Now let . Recall that with ; in particular, and so . We now use Lemma 6.7 to determine and so .
For example, if , then , and Figure 9 shows that is one of . As subgroups of , we know that and . By Table 7, a Sylow -subgroup of has order . If , then , which is impossible by Lemma 6.7; thus . Note that , so is determined. If , then , and so ; this determines . In a similar way, Lemma 6.7 can be used to identify the subgroups for the pairs and . Lemma 6.7 also helps to identify for : if , then Table 10 shows that centralises an involution ; if , then
which is impossible. Now set ; Lemma 6.7 shows that . Table 10 yields . Since , we deduce that is not a subgroup of ; hence .
Let ; here is one of , but has no non-toral as a subgroup. This determines , hence . Similarly, let ; here is one of ; however, have no as subgroup. This determines and .
Lastly, consider for all ; it follows from Table 7 that has distribution . It is shown in [30, Section 2] that is a maximal subgroup of , and with . Thus, is non-toral and we determine , hence .
The structure of the centralisers has the following corollary.
Corollary 6.8.
If and , and and for , then if and only if .
In conclusion, we have proved the following proposition, and the proof of Proposition 6.3 is complete.
Proposition 6.9.
If is as given in Table 7, then is as given in that table.
Appendix A Details for and
This section complements the results in Section 6.7.
A.1. The families of subgroups in [41]
Yu [41] has classified the subgroups of , up to conjugacy, into the following families of subgroups. Families (1a)–(1d) below are the subgroups which contain an element from class , hence lie in the centraliser . Families (2a) and (2b) are those subgroups which contain an element of but no element of , hence lie in a subgroup but not . The final two families (3a) and (3b) are those subgroups containing only involutions from class . For an elementary abelian -subgroup , let ; by [41, Lemma 7.3], this is a subgroup of , and has rank at most . A direct calculation with our algorithm from Section 5.1 shows that has rank at most for toral subgroups. We can now state the families of subgroups.
-
(1a)
. These subgroups are described in [41, §5, §6, Lemma 7.6; Propositions 7.8(1) and 7.9]. They are precisely the subgroups containing a conjugate of , and are therefore all non-toral; this gives classes of non-toral subgroups. The rank of such a group is , and has rank .
-
(1b)
. These are described in [41, §6, Lemma 7.6; Propositions 7.8(2) and 7.9]. By definition these are subgroups of the form , where with for certain elementary abelian subgroups . Since lies in the factor, such a subgroup is toral if and only if is toral in . From the classification for , it follows that the non-toral subgroups here are precisely the three groups with . The rank of a subgroup in each case is , and has rank .
-
(1c)
. These are described in [41, Propositions 2.24, 2.29, 7.8(3), and 7.9; Definition 2.27; Lemma 7.6]. Such a subgroup has rank and has rank , hence all of these subgroups are non-toral. We take this opportunity to point out an error in the stated definition in [41, p. 273]; correcting this error, the definition of in [41, p. 278] is
where the elements are also defined in [41, p. 273]. With this corrected definition the remaining results in [41] regarding these groups are correct as stated.111We thank Dr. Yu for clarifying this.
-
(1d)
. These subgroups all have the form where with and . Each subgroup is toral in as the non-toral subgroups of lie in the class also labelled in [41, p. 272]. Thus all subgroups in this class are toral.
-
(2a)
. This family is described in [41, Proposition 7.14]. They are the elementary abelian subgroups with with and , and . The rank of such a subgroup is , and it follows that where . Using the algorithm of Section 5.1, we see that there are four toral such subgroups, with respective distributions , , , and . From the description of the groups and their generators given in [41, p. 280], the subgroups are visibly toral, and thus the remaining six subgroups are non-toral.
-
(2b)
. These are also described in [41, Proposition 7.14], and are precisely the elementary abelian subgroups with with and , and . It follows at once that all four such subgroups are non-toral, the rank of such a subgroup is , and its distribution is . We have .
-
(3a)
. The containment relations stated in [41, Remark 7.27] show that these subgroups are always toral.
-
(3b)
. These are described in [41, Proposition 7.22]. These are precisely the subgroups where is one of the subgroups above. Our algorithm from Section 5.1 finds exactly toral -pure subgroups, which are necessarily the toral -pure subgroups above. Hence the three subgroups are all non-toral. We have and , see [41, p. 284].
This provides us with classes of non-toral subgroups, as expected. From their definition, we see that the subgroups are ordered by if and only if and , and similarly for the . It is easy to see when the various subgroups are contained in one another, from their definition. The remaining distributions can be calculated from [41, Proposition 7.9], combined with the knowledge of . Specifically, if has rank and has rank , then where and is the defect , as defined in [41, Definition 7.2] and calculated in [41, Proposition 7.9]. Hence and .
A.2. Normaliser quotients
The structure of for the various subgroups is stated in [41, Proposition 7.26]. We now unravel some of the notation there. The notation means the group of block-upper-triangular matrices in with blocks of size and (cf. [41, Proposition 5.5]); thus
Also, means in our notation, and the groups are understood as the following matrix groups of degree :
with () in its natural orthogonal representation, and indicates the -dimensional symplectic module for , considered as a subspace of co-dimension in the space of -dimensional row vectors. This gives the structure for the normaliser quotients of the non-toral subgroups in these families, see Table 8. Here , etc. indicate that all matrices in these groups occur in the block, and indicates that all matrices of the appropriate size occur in the block. For clarity, when dealing with the groups we write to indicate the precise size of the blocks occurring. If some dimension is , then the corresponding rows and columns are omitted entirely.
A.3. A direct computation of all non-toral subgroups
Recall from Section 6.7 that we can compute all elementary abelian subgroups of within the finite normalisers of the maximal non-toral and . In Tables 9(A) and (B) we list the elementary abelian -subgroups of the maximal non-toral and , respectively. Note that some groups occur twice because the computations have been done up to conjugacy in and , respectively. We comment on the identification of the subgroups, using the notation introduced in [41]. Those subgroups named “toral” have been identified because they have a distribution unique to a toral subgroups, or they are computed as a subgroup of such a toral subgroup. Subgroups with distributions different to those of toral subgroups must be non-toral, and we use the notation of [41] (together with information on their distribution) to identify them. Below are the ad-hoc arguments which have been used in this identification. The complete Hasse diagram of non-toral subgroups is given in Figure 1; inclusions in that diagram are computed up to conjugacy in the normalisers.
Remark A.1.
-
(a)
The Klein four-subgroups of and their distributions are given in [41, Table 3]. Our algorithm shows that all but three of these are toral; the non-toral ones are the subgroups , , , which we denote , , in Table 6. Note that these are also precisely the Klein four-subgroups of of such that has rank .
-
(b)
The group can be identified because there is no other non-toral subgroup in or with distribution 2BC7. Similarly is the only un-identified subgroup with distribution 2BC15, and is the only un-identified subgroup with distribution 2BC31.
-
(c)
Group (33) in Tables 9(A) and group (27) in Table 9(B) are both subgroups of a copy of a non-toral . If they are both toral, then there is no unidentified subgroup left with distribution 2BC7A8, which means the non-toral is missing. Hence contains a copy of , which is non-toral, hence (33) in Table 9(A) and (27) in Table 9(B) have this type.
- (d)
-
(e)
If group (25) in Table 9(B) was a copy of then, because it is contained in in the -lattice, we would also see in the -lattice; since we only see there, we conclude that group (25) and its subgroup (13) are toral.
- (f)
-
(g)
Since group (41) in Table 9(B) contains a copy of it must be , as the only other subgroup with the same distribution is , which does not contain such a subgroup. Since (42) does not contain it must be .
- (h)
-
(i)
It remains to identify group (48) in Table (A), and groups (44) and (45) in Table (B). All three have distribution and contain , and the only non-toral subgroups with this distribution that have not been found elsewhere are and . The groups (48) from (A) and (44) from (B) are each contained in a copy of , while group (45) from (B) is not; hence the first two are -conjugate, and not conjugate to the third. Let be one of the subgroups contained in , and let a subgroup from the other class. From the discussion in Case 2c) of Proposition 6.6 we have for , and thus for some involution . Table 6 now implies that , hence , and so by [21, Tables 4.3.1–4.3.3]. It follows that and .
(A) subgroups of | |||
---|---|---|---|
Dist | Cent dim | name | |
(1) | 2BC1 | 69 | toral |
(2) | 2BC1 | 69 | toral |
(3) | 4H1 | 79 | toral |
(4) | 4A1 | 63 | toral |
(5) | 2BC3 | 37 | toral |
(6) | 2BC3 | 37 | toral |
(7) | 2BC14H2 | 47 | toral |
(8) | 2BC14A2 | 31 | toral |
(9) | 2BC3 | 37 | toral |
(10) | 2BC14A14H1 | 39 | toral |
(11) | 2BC14A2 | 31 | toral |
(12) | 4H3 | 52 | |
(13) | 4A24H1 | 36 | |
(14) | 4A3 | 28 | |
(15) | 2BC7 | 21 | toral |
(16) | 2BC7 | 21 | toral |
(17) | 2BC34A24H2 | 23 | toral |
(18) | 2BC34A4 | 15 | toral |
(19) | 2BC7 | 21 | |
(20) | 2BC34A34H1 | 19 | toral |
(21) | 2BC34A4 | 15 | |
(22) | 2BC14A34H3 | 24 | |
(23) | 2BC14A54H1 | 16 | |
(24) | 2BC14A6 | 12 | |
(25) | 2BC34A4 | 15 | toral |
(26) | 2BC14A44H2 | 20 | |
(27) | 2BC14A6 | 12 | |
(28) | 2BC34H4 | 31 | toral |
(29) | 2BC14H6 | 36 | |
(30) | 2BC15 | 13 | toral |
(31) | 2BC15 | 13 | |
(32) | 2BC74A64H2 | 11 | toral |
(33) | 2BC74A8 | 7 | |
(34) | 2BC74A74H1 | 9 | |
(35) | 2BC34A94H3 | 10 | |
(36) | 2BC34A114H1 | 6 | |
(37) | 2BC74A8 | 7 | toral |
(38) | 2BC34A104H2 | 8 | |
(39) | 2BC34A12 | 4 | |
(40) | 2BC34A12 | 4 | |
(41) | 2BC74A44H4 | 15 | toral |
(42) | 2BC34A64H6 | 16 | |
(43) | 2BC34A84H4 | 12 | |
(44) | 2BC34H12 | 28 | |
(45) | 2BC31 | 9 | |
(46) | 2BC154A144H2 | 5 | |
(47) | 2BC74A214H3 | 3 | |
(48) | 2BC154A16 | 3 | |
(49) | 2BC74A224H2 | 2 | |
(50) | 2BC74A24 | 0 | |
(51) | 2BC154A124H4 | 7 | toral |
(52) | 2BC74A184H6 | 6 | |
(53) | 2BC74A204H4 | 4 | |
(54) | 2BC74A124H12 | 12 | |
(55) | 2BC314A284H4 | 3 | |
(56) | 2BC154A424H6 | 1 | |
(57) | 2BC154A444H4 | 0 | |
(58) | 2BC154A364H12 | 4 | |
(59) | 2BC314A844H12 | 0 |
(B) subgroups of | |||
Dist | Cent dim | name | |
(1) | 4A1 | 63 | toral |
(2) | 2BC1 | 69 | toral |
(3) | 4A1 | 63 | toral |
(4) | 4H1 | 79 | toral |
(5) | 2BC14A2 | 31 | toral |
(6) | 4A3 | 28 | |
(7) | 4A24H1 | 36 | |
(8) | 2BC3 | 37 | toral |
(9) | 2BC14A2 | 31 | toral |
(10) | 2BC14H2 | 47 | toral |
(11) | 2BC3 | 37 | toral |
(12) | 2BC14A14H1 | 39 | toral |
(13) | 2BC34A4 | 15 | (toral) |
(14) | 2BC14A6 | 12 | |
(15) | 2BC14A44H2 | 20 | |
(16) | 2BC34A4 | 15 | |
(17) | 2BC14A54H1 | 16 | |
(18) | 2BC7 | 21 | toral |
(19) | 2BC34A4 | 15 | toral |
(20) | 2BC7 | 21 | toral |
(21) | 2BC34A24H2 | 23 | toral |
(22) | 2BC34H4 | 31 | toral |
(23) | 2BC34A34H1 | 19 | toral |
(24) | 2BC34A14H3 | 27 | toral |
(25) | 2BC74A8 | 7 | (toral) |
(26) | 2BC34A12 | 4 | |
(27) | 2BC74A8 | 7 | |
(28) | 2BC34104H2 | 8 | |
(29) | 2BC34A84H4 | 12 | |
(30) | 2BC34114H1 | 6 | |
(31) | 2BC34A94H3 | 10 | |
(32) | 2BC74A8 | 7 | toral |
(33) | 2BC74A64H2 | 11 | toral |
(34) | 2BC74A44H4 | 15 | toral |
(35) | 2BC74A44H4 | 15 | toral |
(36) | 2BC74A24H6 | 19 | toral |
(37) | 2BC15 | 13 | toral |
(38) | 2BC15 | 13 | toral |
(39) | 2BC74A24 | 0 | |
(40) | 2BC74A224H2 | 2 | |
(41) | 2BC74A204H4 | 4 | |
(42) | 2BC74A204H4 | 4 | |
(43) | 2BC74A184H6 | 6 | |
(44) | 2BC154A16 | 3 | |
(45) | 2BC154A16 | 3 | |
(46) | 2BC154A124H4 | 7 | toral |
(47) | 2BC154A104H6 | 9 | toral |
(48) | 2BC154A84H8 | 11 | toral |
(49) | 2BC154A64H10 | 13 | toral |
(50) | 2BC31 | 9 | toral |
(51) | 2BC154A444H4 | 0 | |
(52) | 2BC154A424H6 | 1 | |
(53) | 2BC154A404H8 | 2 | |
(54) | 2BC154A384H10 | 3 | |
(55) | 2BC314A32 | 1 | |
(56) | 2BC314A204H12 | 7 | toral |
(57) | 2BC314A164H16 | 9 | toral |
(58) | 2BC63 | 7 | toral |
(59) | 2BC314A844H12 | 0 | |
(60) | 2BC314A804H16 | 1 | |
(61) | 2BC634A64 | 0 | |
(62) | 2BC634A364H28 | 7 | toral |
(63) | 2BC634A1644H28 | 0 |
Appendix B Details for and
This section complements the results in Section 6.8.
B.1. The families of subgroups in [41]
Yu [41, §8] classified all elementary abelian -subgroups of , separating these into six disjoint families depending on some parameters, as follows:
-
(1)
. Such a subgroup has rank . From [41, Proposition 8.8], these have distribution where and . Since there is no toral subgroup with distribution , we conclude that is non-toral, hence all these subgroups are non-toral as they all contain .
-
(2)
. Such a subgroup has rank . Its distribution is given by [41, Proposition 8.8]. The subgroup turns out to be the unique subgroup with distribution , which matches a toral subgroup returned by our algorithm. Thus, and all its subgroup are toral.
-
(3)
. Such a subgroup has rank . According to the discussion after [41, Definition 8.15], these subgroups have distribution where and . Note that this does not agree with [41, Proposition 8.8(3)], which is missing an exponent ‘’ in a given formula. In particular, has distribution , which it shares with no toral subgroup. Hence is non-toral, as are all the subgroups .
-
(4)
. Such a subgroup has rank . By [41, Proposition 8.8(4)] this has distribution where and . In particular is the unique subgroup of rank with distribution , hence is toral, as are all of its subgroups .
-
(5)
. Such a subgroup has rank . By [22, (9.2)], the groups with are toral, and , the unique -pure subgroup of rank , is non-toral.
-
(6)
. Such a subgroup has rank . Its distribution follows from [41, Definition 8.15], however this contains several typographic errors so we give this description again for completeness. We have decompositions
where is a -pure subgroup of rank , is a -pure subgroup of rank , and and are subgroups of rank and respectively, each containing a unique element from class . Then [41, Definition 8.15] states that an element lies in class precisely when in the decomposition arising from the above. In particular . From this and from inspecting the distributions of the remaining toral subgroups, we find that is toral when , and is non-toral when .
This gives classes of subgroups in all; the non-toral ones are listed in Table 7.
B.2. Normaliser quotients
For each elementary abelian as above, the quotient is determined in [41, Proposition 8.17]. Adopting the same notation for matrix groups as in the case, the normaliser structure is as in Table 10. Here a wreath product denotes an action on a direct sum of two copies of the natural -dimensional module. The group is the group in [41, p. 259]. From the description there it is not hard to show that
where each group is acting in its natural orthogonal representation.
References
- [1] Jianbei An, Weight -subgroups of the Chevalley groups , 2023, (in preparation).
- [2] Jianbei An and Heiko Dietrich, The AWC-goodness and essential rank of sporadic simple groups, J. Algebra 356 (2012), 325–354. MR 2891136
- [3] by same author, Radical 3-subgroups of with even, J. Algebra 398 (2014), 542–568. MR 3123785
- [4] by same author, Maximal 2-local subgroups of , J. Algebra 445 (2016), 503–536. MR 3418067
- [5] Jianbei An, Heiko Dietrich, and Shih-Chang Huang, Classification of some 3-subgroups of the finite groups of Lie type , J. Pure Appl. Algebra 222 (2018), no. 12, 4020–4039. MR 3818291
- [6] by same author, Radical 3-subgroups of the finite groups of Lie type , J. Pure Appl. Algebra 222 (2018), no. 12, 4040–4067. MR 3818292
- [7] Jianbei An, Heiko Dietrich, and Alastair Litterick, Magma code for computing toral elementary abelian subgroups, 2023, see github.com/heikodietrich/toralelabsubgroups.
- [8] Jianbei An and Charles Eaton, Morita equivalence classes of blocks with extraspecial defect groups , 2023, (in preparation).
- [9] Jianbei An and Shih-chang Huang, Radical 3-subgroups and essential 3-rank of , J. Algebra 376 (2013), 320–340. MR 3003729
- [10] Jianbei An and Shih-Chang Huang, Maximal 2-local subgroups of and , J. Pure Appl. Algebra 219 (2015), no. 1, 101–120. MR 3240826
- [11] K.K.S. Andersen, J. Grodal, J.M. Møller, and A. Viruel, The classification of -compact groups for odd, Ann. of Math. (2) 167 (2008), no. 1, 95–210. MR 2373153
- [12] Michael Bate, Benjamin Martin, and Gerhard Röhrle, A geometric approach to complete reducibility, Invent. Math. 161 (2005), no. 1, 177–218. MR 2178661
- [13] Wieb Bosma, John Cannon, and Catherine Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput. 24 (1997), no. 3-4, 235–265, Computational algebra and number theory (London, 1993). MR 1484478
- [14] Roger W. Carter, Simple groups of Lie type, Wiley Classics Library, John Wiley & Sons, Inc., New York, 1989, Reprint of the 1972 original, A Wiley-Interscience Publication. MR 1013112
- [15] Arjeh M. Cohen and Robert L. Griess, Jr., On finite simple subgroups of the complex Lie group of type , The Arcata Conference on Representations of Finite Groups (Arcata, Calif., 1986), Proc. Sympos. Pure Math., vol. 47, Amer. Math. Soc., Providence, RI, 1987, pp. 367–405. MR 933426
- [16] Arjeh M. Cohen, Martin W. Liebeck, Jan Saxl, and Gary M. Seitz, The local maximal subgroups of exceptional groups of Lie type, finite and algebraic, Proc. London Math. Soc. (3) 64 (1992), no. 1, 21–48. MR 1132853
- [17] Arjeh M. Cohen and David B. Wales, Finite simple subgroups of semisimple complex Lie groups—a survey, Groups of Lie type and their geometries (Como, 1993), London Math. Soc. Lecture Note Ser., vol. 207, Cambridge Univ. Press, Cambridge, 1995, pp. 77–96. MR 1320515
- [18] by same author, Finite subgroups of and , Proc. London Math. Soc. (3) 74 (1997), no. 1, 105–150. MR 1416728
- [19] Zhicheng Feng, Zhenye Li, and Jiping Zhang, Jordan decomposition for weights and the blockwise Alperin weight conjecture, Adv. Math. 408 (2022), no. part A, Paper No. 108594, 37. MR 4457347
- [20] Eric M. Friedlander and Guido Mislin, Conjugacy classes of finite solvable subgroups in Lie groups, Ann. Sci. École Norm. Sup. (4) 21 (1988), no. 2, 179–191. MR 956765
- [21] Daniel Gorenstein, Richard Lyons, and Ronald Solomon, The classification of the finite simple groups. Number 3. Part I. Chapter A, Mathematical Surveys and Monographs, vol. 40, American Mathematical Society, Providence, RI, 1998, Almost simple -groups. MR 1490581
- [22] Robert L. Griess, Jr., Elementary abelian -subgroups of algebraic groups, Geom. Dedicata 39 (1991), no. 3, 253–305. MR 1123145
- [23] Robert L. Griess, Jr. and A. J. E. Ryba, Embeddings of and in , Duke Math. J. 94 (1998), no. 1, 181–211, With appendices by Michael Larsen and J.-P. Serre. MR 1635916
- [24] Jing-Song Huang and Jun Yu, Klein four-subgroups of Lie algebra automorphisms, Pacific J. Math. 262 (2013), no. 2, 397–420. MR 3069067
- [25] Jens Carsten Jantzen, Representations of algebraic groups, second ed., Mathematical Surveys and Monographs, vol. 107, American Mathematical Society, Providence, RI, 2003. MR 2015057
- [26] John Jones, Dmitriy Rumynin, and Adam Thomas, Compact lie groups and complex reductive groups, https://arxiv.org/abs/2109.13702, preprint.
- [27] Radha Kessar and Gunter Malle, Local-global conjectures and blocks of simple groups, Groups St Andrews 2017 in Birmingham, London Math. Soc. Lecture Note Ser., vol. 455, Cambridge Univ. Press, Cambridge, 2019, pp. 70–105. MR 3931409
- [28] Masaaki Kitazume and Satoshi Yoshiara, The radical subgroups of the Fischer simple groups, J. Algebra 255 (2002), no. 1, 22–58. MR 1935034
- [29] Martin W. Liebeck and Gary M. Seitz, Reductive subgroups of exceptional algebraic groups, Mem. Amer. Math. Soc. 121 (1996), no. 580, vi+111. MR 1329942
- [30] by same author, A survey of maximal subgroups of exceptional groups of Lie type, Groups, combinatorics & geometry (Durham, 2001), World Sci. Publ., River Edge, NJ, 2003, pp. 139–146. MR 1994964
- [31] by same author, The maximal subgroups of positive dimension in exceptional algebraic groups, Mem. Amer. Math. Soc. 169 (2004), no. 802, vi+227. MR 2044850
- [32] 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. MR 2850737
- [33] Hermann Minkowski, Zur Theorie der positiven quadratischen Formen, J. Reine Angew. Math. 101 (1887), 196–202. MR 1580123
- [34] R. V. Moody and J. Patera, Characters of elements of finite order in Lie groups, SIAM J. Algebraic Discrete Methods 5 (1984), no. 3, 359–383. MR 752042
- [35] Gabriel Navarro and Pham Huu Tiep, A reduction theorem for the Alperin weight conjecture, Invent. Math. 184 (2011), no. 3, 529–565. MR 2800694
- [36] Jean-Pierre Serre, Galois cohomology, English ed., Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2002, Translated from the French by Patrick Ion and revised by the author. MR 1867431
- [37] by same author, Bounds for the orders of the finite subgroups of , Group representation theory, EPFL Press, Lausanne, 2007, pp. 405–450. MR 2336645
- [38] Britta Späth, Sylow -tori of classical groups and the McKay conjecture. I, J. Algebra 323 (2010), no. 9, 2469–2493. MR 2602390
- [39] Robert Steinberg, Torsion in reductive groups, Advances in Math. 15 (1975), 63–92. MR 354892
- [40] J. Tits, Normalisateurs de tores. I. Groupes de Coxeter étendus, J. Algebra 4 (1966), 96–116. MR 206117
- [41] Jun Yu, Elementary abelian 2-subgroups of compact Lie groups, Geom. Dedicata 167 (2013), 245–293. MR 3128780