Malle, Testerman - Linear Algebraic Groups and Finite Groups of Lie Type-Cambridge University Press (2011)
Malle, Testerman - Linear Algebraic Groups and Finite Groups of Lie Type-Cambridge University Press (2011)
Editorial Board
B . B O L L O B Á S , W . F U L T O N , A . K A T O K , F . K I R W A N ,
P. SARNAK, B. SIMON, B. TOTARO
Originating from a summer school taught by the authors, this concise treatment
includes many of the main results in the area. An introductory chapter describes the
fundamental results on linear algebraic groups, culminating in the classification of
semisimple groups. The second chapter introduces more specialized topics in the
subgroup structure of semisimple groups, and describes the classification of the
maximal subgroups of the simple algebraic groups. The authors then systematically
develop the subgroup structure of finite groups of Lie type as a consequence of the
structural results on algebraic groups. This approach will help students to understand
the relationship between these two classes of groups.
The book covers many topics that are central to the subject, but missing from
existing textbooks. The authors provide numerous instructive exercises and examples
for those who are learning the subject as well as more advanced topics for research
students working in related areas.
Editorial Board:
B. Bollobás, W. Fulton, A. Katok, F. Kirwan, P. Sarnak, B. Simon, B. Totaro
All the titles listed below can be obtained from good booksellers or from Cambridge University Press. For a
complete series listing visit: http://www.cambridge.org/mathematics
Already published
85 J. Carlson, S. Müller-Stach & C. Peters Period mappings and period domains
86 J. J. Duistermaat & J. A. C. Kolk Multidimensional real analysis, I
87 J. J. Duistermaat & J. A. C. Kolk Multidimensional real analysis, II
89 M. C. Golumbic & A. N. Trenk Tolerance graphs
90 L. H. Harper Global methods for combinatorial isoperimetric problems
91 I. Moerdijk & J. Mrčun Introduction to foliations and Lie groupoids
92 J. Kollár, K. E. Smith & A. Corti Rational and nearly rational varieties
93 D. Applebaum Lévy processes and stochastic calculus (1st Edition)
94 B. Conrad Modular forms and the Ramanujan conjecture
95 M. Schechter An introduction to nonlinear analysis
96 R. Carter Lie algebras of finite and affine type
97 H. L. Montgomery & R. C. Vaughan Multiplicative number theory, I
98 I. Chavel Riemannian geometry (2nd Edition)
99 D. Goldfeld Automorphic forms and L-functions for the group GL(n,R)
100 M. B. Marcus & J. Rosen Markov processes, Gaussian processes, and local times
101 P. Gille & T. Szamuely Central simple algebras and Galois cohomology
102 J. Bertoin Random fragmentation and coagulation processes
103 E. Frenkel Langlands correspondence for loop groups
104 A. Ambrosetti & A. Malchiodi Nonlinear analysis and semilinear elliptic problems
105 T. Tao & V. H. Vu Additive combinatorics
106 E. B. Davies Linear operators and their spectra
107 K. Kodaira Complex analysis
108 T. Ceccherini-Silberstein, F. Scarabotti & F. Tolli Harmonic analysis on finite groups
109 H. Geiges An introduction to contact topology
110 J. Faraut Analysis on Lie groups: An Introduction
111 E. Park Complex topological K-theory
112 D. W. Stroock Partial differential equations for probabilists
113 A. Kirillov, Jr An introduction to Lie groups and Lie algebras
114 F. Gesztesy et al. Soliton equations and their algebro-geometric solutions, II
115 E. de Faria & W. de Melo Mathematical tools for one-dimensional dynamics
116 D. Applebaum Lévy processes and stochastic calculus (2nd Edition)
117 T. Szamuely Galois groups and fundamental groups
118 G. W. Anderson, A. Guionnet & O. Zeitouni An introduction to random matrices
119 C. Perez-Garcia & W. H. Schikhof Locally convex spaces over non-Archimedean valued fields
120 P. K. Friz & N. B. Victoir Multidimensional stochastic processes as rough paths
121 T. Ceccherini-Silberstein, F. Scarabotti & F. Tolli Representation theory of the symmetric groups
122 S. Kalikow & R. McCutcheon An outline of ergodic theory
123 G. F. Lawler & V. Limic Random walk: A modern introduction
124 K. Lux & H. Pahlings Representations of groups
125 K. S. Kedlaya p-adic differential equations
126 R. Beals & R. Wong Special functions
127 E. de Faria & W. de Melo Mathematical aspects of quantum field theory
128 A. Terras Zeta functions of graphs
129 D. Goldfeld & J. Hundley Automorphic representations and L-functions for the general linear group, I
130 D. Goldfeld & J. Hundley Automorphic representations and L-functions for the general linear group, II
131 D. A. Craven The theory of fusion systems
132 J. Väänänen Models and games
Linear Algebraic Groups and
Finite Groups of Lie Type
GUNTER MALLE
University of Kaiserslautern, Germany
DONNA TESTERMAN
École Polytechnique Fédérale de Lausanne, Switzerland
cambridge university press
Cambridge, New York, Melbourne, Madrid, Cape Town,
Singapore, São Paulo, Delhi, Tokyo, Mexico City
Cambridge University Press
The Edinburgh Building, Cambridge CB2 8RU, UK
Published in the United States of America by Cambridge University Press, New York
www.cambridge.org
Information on this title: www.cambridge.org/9781107008540
C G. Malle and D. Testerman 2011
A catalogue record for this publication is available from the British Library
Preface page ix
List of tables xiii
Notation xiv
These notes grew out of a summer school on “Finite Groups and Related
Geometrical Structures” held in Venice from September 5th to September
15th 2007. The aim of the course was to introduce an audience consisting
mainly of PhD students and postdoctoral researchers working in finite group
theory and neighboring areas to results on the subgroup structure of linear
algebraic groups and the related finite groups of Lie type.
As will be seen in Part I, a linear algebraic group is an affine variety which
is equipped with a group structure in such a way that the binary group op-
eration and inversion are continuous maps. A connected (irreducible) linear
algebraic group has a maximal solvable connected normal subgroup such
that the quotient group is a central product of simple algebraic groups, a so-
called semisimple algebraic group. Thus, one is led to the study of semisimple
groups and connected solvable groups. A connected solvable linear algebraic
group is the semidirect product of the normal subgroup consisting of its
unipotent elements with an abelian (diagonalizable) subgroup (for example,
think of the group of invertible upper triangular matrices). While one cannot
expect to classify unipotent groups, remarkably enough this is possible for
the semisimple quotient.
The structure theory of semisimple groups was developed in the middle of
the last century and culminated in the classification of the semisimple linear
algebraic groups defined over an algebraically closed field, a result essentially
due to Chevalley, first made available via the Séminaire sur la classification
des groupes de Lie algébriques at the Ecole Normale Supérieure in Paris,
during the period 1956–1958 ([15]). Analogously to the work of Cartan and
Killing on the classification of the complex semisimple Lie algebras, Cheval-
ley showed that the semisimple groups are determined up to isomorphism
by a set of combinatorial data, based principally upon a root system (as for
the semisimple Lie algebras) and a dual root system. Moreover, the set of
x Preface
possible combinatorial data does not depend on the characteristic of the un-
derlying field. Part I of this text is devoted to developing the tools necessary
for describing this classification. We have followed the development in two
very good texts ([32] and [66]) on linear algebraic groups, and we often refer
to these books for the proofs which we have omitted. Our aim is to give the
reader a feel for the group-theoretic ingredients of this classification, without
going into the details of the underlying algebraic geometric foundations, and
then to move on to the material of Parts II and III, which should perhaps
be seen as the distinguishing feature of this text.
The 20- to 30-year period following the classification was a productive time
in “semisimple” theory, during which many actors, notably Borel, Bruhat,
Springer, Steinberg and Tits, played a role in the further study of these
groups. The conjugacy classes, endomorphisms, representations, and sub-
group structure were among the topics of consideration, with the principal
aim of reducing their classification and the description of their structure to
combinatorial data related to the root system and the Weyl group of the
ambient group. Part II of this book treats some of these subjects. In partic-
ular, we describe the Tits BN -pair for a semisimple linear algebraic group
and obtain a Levi decomposition for the associated parabolic subgroups;
we discuss conjugacy classes of semisimple elements and their centralizers;
we describe the parametrization of the irreducible representations of these
groups and their automorphism groups. We leave the discussion of the gen-
eral endomorphisms of simple algebraic groups to Part III, where these are
used to construct the finite groups of Lie type.
In the last chapters of Part II, we turn to more recent developments in the
theory of semisimple algebraic groups, where we describe the classification of
their maximal positive-dimensional subgroups. These results can be seen as
an extension of the fundamental work of Dynkin on the maximal subalgebras
of the semisimple complex Lie algebras. It is the subject of several very long
research articles by Liebeck, Seitz and others and was completed in 2004.
In the course of the classification of finite simple groups, attention turned
to the analogues of algebraic groups over finite fields. These so-called fi-
nite groups of Lie type were eventually shown to comprise, together with
the alternating groups, almost all the finite simple groups. Steinberg found
a unified approach to constructing not only the well-known finite classical
groups, but also their twisted “Steinberg variations”, as well as further seem-
ingly sporadic examples, the Suzuki and Ree groups, as fixed point subgroups
of certain endomorphisms of simple linear algebraic groups defined over Fp .
There does not yet seem to be a generally accepted terminology for such
endomorphisms, and we call them “Steinberg endomorphisms” in this text.
Preface xi
Part III is devoted to the definition and study of these finite groups. We
first classify the endomorphisms whose fixed point subgroups are finite, fol-
lowing the work of Steinberg, and hence are able to describe the full set of
finite groups “of Lie type” thus obtained. The theorem of Lang–Steinberg
then provides the necessary machinery for applying the results of Parts I
and II to the study of these fixed point subgroups. For example, we give the
proof of the existence of a BN -pair for these groups, which allows one to
deduce a formula for their order. Furthermore, we study their Sylow sub-
groups and touch on some other aspects of the subgroup structure. Finally,
we return to the question of the maximal subgroups, this time sketching a
proof of Aschbacher’s reduction theorem for the maximal subgroups of the
finite classical groups and indicating how it has been applied (by Kleidman,
Liebeck and others) and how it must still be applied if one hopes to deter-
mine the maximal subgroups of the finite classical groups. We conclude with
a discussion of what is known about the maximal subgroups of the excep-
tional finite groups of Lie type, including work of Liebeck, Saxl, Seitz and
others. We then come full-circle and sketch the proof of a result which en-
ables one to lift certain embeddings of finite groups of Lie type to embeddings
of algebraic groups, where one can apply the more complete information of
Part II.
of Lie type, without needing to understand every detail of the proof of the
classification of semisimple groups. In particular, Parts II and III should give
a good overview of much of what is known about the subgroup structure of
these groups and to a lesser extent their conjugacy classes and representation
theory. The numerous exercises are intended to supplement and illustrate the
theory, and should help the book fulfill its objective of serving as the basis
for a first-year graduate level course.
In this part we introduce the main objects of study, linear algebraic groups
over algebraically closed fields.
We assume that the reader is familiar with basic concepts and results from
commutative algebra and algebraic geometry. More specifically, the reader
should know about affine and projective varieties, their associated coordinate
ring, their dimension, the Zariski topology, and basic properties thereof.
In Chapter 1 we define our main objects of study. The examples which
will guide us throughout the text are certain subgroups or quotient groups
of the isometry group of a finite-dimensional vector space equipped with a
bilinear or quadratic form. We state the important result which says that
any linear algebraic group is a closed subgroup of some group of invertible
matrices over our fixed field, which is nearly obvious for all of our examples
(the proof will be given in Chapter 5). In Chapter 2, we show that the
Jordan decomposition of a matrix results in a uniquely determined Jordan
decomposition of elements in a linear algebraic group. This in turn gives
us the notion of semisimple and unipotent elements in these groups. We
establish the important result that any group consisting entirely of unipotent
elements is conjugate to a subgroup of the upper unitriangular matrices.
Chapter 3 is devoted to the structure theory of commutative linear alge-
braic groups. In particular, Theorem 3.1 focuses attention on groups consist-
ing entirely of unipotent elements or of semisimple elements. Theorem 3.2
classifies the connected one-dimensional linear algebraic groups. While one
can say something about the structure of connected commutative groups
consisting entirely of unipotent elements, these will not play a role in this
text. Hence we turn at this point to commutative groups consisting entirely
of semisimple elements and introduce the notion of a torus, its character
2
group and its cocharacter group. These will play a crucial role in the classifi-
cation of semisimple groups. We turn in Chapter 4 to the structure theory of
connected solvable groups, for which the prototype is the group of upper tri-
angular invertible matrices. Indeed, the Lie–Kolchin theorem (Theorem 4.1
and Corollary 4.2) shows that any such group is isomorphic to a closed sub-
group of the group of upper triangular matrices. The importance of closed
connected solvable subgroups will become apparent in Chapter 6.
But before defining these so-called Borel subgroups, we must extend our
theory to cover group actions and in particular quotient groups; this is the
content of Chapter 5. The results on homogeneous spaces prepare the terrain
for establishing the main result of Chapter 6, the Borel fixed point theorem,
Theorem 6.1, some of whose many applications we discuss. We can also
finally define the radical of a linear algebraic group and establish its con-
nection with Borel subgroups. In these two chapters, we omit some essential
geometric arguments and notions. In particular, we do not prove results on
complete varieties but restrict ourselves only to projective varieties.
The last three chapters of this part are devoted to introducing the combi-
natorial data which classifies semisimple algebraic groups and to establish-
ing structural results and the classification theorem. The most important
ingredient of the data is a root system, which is obtained via the adjoint
representation of the group, acting on its tangent space; this theory is de-
scribed in Chapter 7. Theorem 8.17 is the main structural result on reductive
groups and we study in detail the case of the group SL2 in order to sketch a
proof of this result. The final chapter describes the classification of semisim-
ple algebraic groups in terms of the data mentioned above. We conclude by
explaining where our standard examples appear in this classification.
1
Basic concepts
X@
ϕ
/Y
@
@ f
ϕ∗ (f ) @
k
µ : G × G −→ G, i : G −→ G,
(g, h) −→ gh, g −→ g −1 ,
Example 1.2 The base field k provides two natural examples of algebraic
groups:
(1) The additive group G = (k, +) of k is defined by the zero ideal I = (0)
in k[T ], and addition is given by a polynomial; hence G is an algebraic
group, with coordinate ring k[G] = k[T ]. The group G is called the
additive group, noted Ga .
(2) The multiplicative group G = (k × , ·) of k can be identified with the set
of pairs {(x, y) ∈ k 2 | xy = 1} (where multiplication is componentwise,
again given by polynomials), which is the algebraic set defined by the
ideal I = (XY − 1) k[X, Y ]. So here k[G] = k[X, Y ]/(XY − 1) ∼ =
k[X, X −1 ]. The group G is called the multiplicative group and noted
Gm .
It is not immediately obvious from the above definition that the general
linear group
GLn := {A ∈ k n×n | det A = 0}
Example 1.4 (1) If G ≤ GLn is a closed subgroup then the natural em-
bedding G → GLn is a morphism of linear algebraic groups.
(2) The determinant map det : GLn → Gm , A → det A, is a group homo-
morphism and clearly also a morphism of varieties, so a morphism of
algebraic groups.
For the proof of the above statement, we will make use of the following
property of morphisms of varieties, which will also be used in subsequent
chapters (see [66, Thm. 1.9.5] or [26, Cor. 2.2.8]):
1.3 Connectedness
We now recall a topological notion which will play a crucial role in the study
of linear algebraic groups.
Proposition 1.9 The following are equivalent for an affine algebraic va-
riety X:
(i) X is irreducible.
(ii) Every non-empty open subset of X is dense.
(iii) Any two non-empty open subsets of X intersect non-trivially.
(iv) The vanishing ideal I of X is a prime ideal.
(v) k[X] is an integral domain.
Furthermore, we need the following basic properties (see [32, Prop. 1.3A,
1.3B and 1.4] and also Exercise 10.1):
10 Basic concepts
The maximal irreducible subsets in the preceding statement are called the
irreducible components of X. Note that by (a) irreducible components are
necessarily closed.
Note that any irreducible set is connected; the converse is not true in
general. See Exercise 10.2.
(1) Ga and Gm are connected by Proposition 1.9(v) since k[Ga ] = k[T ] and
k[Gm ] = k[T, T −1 ] are integral domains.
(2) GLn is connected since k[GLn ] = k[Tij ]det(Tij ) is an integral domain,
being a localization of the polynomial ring k[Tij ].
The next result gives a first example of how the Zariski topology on a linear
algebraic group allows one to deduce group theoretic structural results.
(a) The irreducible components of G are pairwise disjoint, so they are the
connected components of G.
(b) The irreducible component G◦ containing 1 ∈ G is a closed normal sub-
group of finite index in G.
(c) Any closed subgroup of G of finite index contains G◦ .
1.4 Dimension
Another fundamental invariant of algebraic varieties is their dimension.
For an irreducible variety X, the coordinate ring k[X] is an integral domain
by Proposition 1.9(v). Let k(X) be the field of fractions of k[X]. We define
the dimension of X by dim(X) := trdegk (k(X)), the transcendence degree of
k(X) over k. Equivalently, the dimension of X equals the maximal length of
descending chains of prime ideals in k[X]. If X is a reducible affine algebraic
variety, then according to Proposition 1.10(d) it can be decomposed as a
finite union X = X1 ∪ · · · ∪ Xt of its irreducible components Xi , and we set
dim(X) := max{dim(Xi ) | 1 ≤ i ≤ t}.
For G a linear algebraic group, dim(G) = dim(G◦ ) since G is the union
of the finitely many cosets gG◦ of the irreducible subgroup G◦ , by Propo-
sition 1.13, and dim(G) = dim(gG◦ ) = dim(G◦ ) by the above definition. In
particular, dim(G) = 0 if and only if G is a finite algebraic group.
Dimension behaves well with respect to morphisms in the following sense
(see [26, Cor. 2.2.9] or [32, Thm. 4.3]):
In particular, for short exact sequences of linear algebraic groups this gives:
(a) For any embedding ρ of G into some GL(V ) and for any g ∈ G, there
2.1 Decomposition of endomorphisms 17
See Definition 3.3 and Proposition 3.9 for the remaining step in the case
that G = Gs .
to
(γ1 + γ2 )(x) := γ1 (x)γ2 (x) for γ1 , γ2 ∈ Y (G), x ∈ Gm .
χ ◦ γ : Gm → Gm , t → χ(γ(t)),
for all t ∈ Gm and all i. Thus di = 0 for all i, that is, γ = 0. Hence
Y ∼= Hom(X, Z). Similarly one argues that X ∼
= Hom(Y, Z).
Definition 3.7 Let T be a torus. For a closed subgroup H ≤ T we define
a closed subgroup of T .
X(T )/H ⊥ ∼
= X(H).
(b) For any subgroup X1 ≤ X(T ), X1⊥⊥ /X1 is a finite p-group, where p =
char(k); in particular, X1⊥⊥ = X1 if X(T )/X1 has no p-torsion.
The following can be shown using properties of characters (see [32, §16]
or [66, 3.2.7]):
This yields the missing part in the proof of Theorem 3.2 in case G = Gs :
embed G into some GLn and diagonalize.
The following crucial result shows the “rigidity” of tori: their normaliz-
ers are only a finite bit larger than their centralizers. It will later allow us
to define the Weyl group, a finite group which plays an important role in
controlling the structure of semisimple groups (see [66, Thm. 3.2.9] for a
proof):
and NG (T ) ∼
= NGLm (Dm ) × GLn−m where NGLm (Dm ) = M is the set of all
monomial m × m-matrices. As Dm has finite index in M , M ◦ = Dm . Also,
GLn−m is connected, so we get
NG (T )◦ ∼
= Dm × GLn−m ∼
= CG (T ) = CG (T )◦ ,
and NG (T )/CG (T ) ∼
= M/Dm ∼
= Sm is a symmetric group, hence indeed
finite.
4
Connected solvable groups
|Gu : G◦u | = |Gu /G : G◦u /G | = |ϕ(Gu ) : ϕ(G◦u )| = |G̃u : ϕ(G◦u )|.
4.2 Structure of connected solvable groups 29
Hence ϕ(G◦u ) is a closed subgroup of finite index in the connected group G̃u ,
so of index 1. Hence, Gu is also connected.
The proof of part (b) is more difficult and requires results about actions
of diagonalizable groups (see [32, Thm. 19.3]).
Finally, the assertion on the normalizer follows from the Frattini argument:
if g = ut ∈ NG (T ), with u ∈ Gu and t ∈ T , then uT u−1 = T . But, for s ∈ T ,
usu−1 = (u su−1 s−1 )s lies in T if and only if u su−1 s−1 = 1, that is, u
centralizes s.
Note the similarity of the second part with the Schur–Zassenhaus The-
orem for finite groups ([2, §18.1]). Indeed, G is already known to be an
extension of the unipotent group Gu by a torus π(G), all of whose elements
are semisimple. The theorem asserts that a complement to Gu exists and
that furthermore all complements are conjugate. If k is the algebraic clo-
sure of a finite field Fp , then unipotent elements have p-power order, while
semisimple elements have order prime to p. Thus Gu is the analogue of a
normal Sylow p-subgroup of G, and any maximal torus is a p -complement.
In that situation, the assertion in (b) is just as in the Schur–Zassenhaus
Theorem.
If G is not connected, anything may happen. For example, take G =
NGL2 (D2 ), an extension of D2 by an element of order 2, and assume that
char(k) = 2. Then, although G is solvable, Gu is not a subgroup, in particular
not normal, and the maximal torus D2 is not self-normalizing.
Corollary 4.5 Let G be a connected, solvable linear algebraic group. Then
any semisimple element of G lies in a maximal torus and any unipotent
element of G lies in a connected unipotent subgroup of G.
The result for unipotent elements follows from Theorem 4.4(a); for semi-
simple elements it follows from the omitted proof of Theorem 4.4(b).
Finally, we need a result on centralizers of tori in connected solvable groups
which will later on be generalized to arbitrary connected groups (see Theo-
rem 6.19) (for the proof see [66, Cor. 6.3.6]):
Proposition 4.6 Let G be connected solvable, S ≤ G a torus. Then CG (S)
is connected.
5
G-spaces and quotients
One aspect of the theory of linear algebraic groups which has been missing
up to now is that of a quotient group. We need to first see how to give the
structure of variety to a quotient and it will become clear that we cannot
limit ourselves to affine varieties. Thus, we begin by recalling some basic
aspects of the general theory of varieties and morphisms.
for U ⊆ X open, defines a sheaf on X. One can show that OX (X) = k[X],
the usual k-algebra of regular functions on X.
n
Now consider the covering Pn = i=0 Ui , where Ui consists of the points
in Pn with non-zero ith homogeneous coordinate. Then Ui can be identified
with affine n-space k n via
x0 xi−1 xi+1 xn
(x0 , x1 , . . . , xn ) → ,..., , ..., .
xi xi xi xi
Moreover, the induced topology on the sets Ui is precisely the topology of
the affine space k n . For each x ∈ Ui , we have the ring of functions Ox as
above, using the identification of Ui with k n . Then for U ⊆ Pn , set O(U ) :=
$
x∈U Ox . The sheaf of functions for an arbitrary projective variety is defined
by restriction of the given sheaf on Pn . It is then clear that the only globally
defined functions on Pn are the constant functions. This is more generally
true for an arbitrary projective variety.
The dimension of an irreducible projective variety is the dimension of any
affine open subset. For a general projective variety, one simply takes the
maximum dimension of the irreducible components of the variety. (See [66,
Prop. 1.2.4, §1.8.1].)
Now a morphism of varieties is a continuous map ϕ : X → Y between two
(affine or projective) varieties such that for all V ⊆ Y open and U = ϕ−1 (V )
and all f ∈ OY (V ), we have f ◦ ϕ ∈ OX (U ).
Henceforth, we will include projective varieties when we consider a general
variety, though we continue to limit ourselves to linear algebraic groups,
which are always affine varieties. We point out that Propositions 1.6, 1.19,
and 1.22 remain valid in the setting of projective varieties.
Important examples of projective varieties are provided by partial flag
varieties of an n-dimensional vector space V over k. Recall that a flag in
V is a chain of subspaces 0 V1 · · · Vr = V . One can equip the
set of flags having a fixed sequence of dimensions (dim V1 , . . . , dim Vr ) with
the structure of a projective variety, the partial flag variety. The case where
dim Vi = i for all i is said to be the flag variety of V . (See for example [26,
§3.3] for details.)
by Proposition 1.22 has smaller dimension. (As pointed out above, the proof
of the latter result goes through in that case as well.)
Let’s point out two important consequences. The first is the characteriza-
tion of linear algebraic groups already stated in Theorem 1.7:
Corollary. Any linear algebraic group can be embedded as a closed subgroup
into GLn for some n.
In the case of a normal subgroup, even more is true (see [4, Thm. 6.8]):
the sum of the Vχ is direct there are only finitely many χ for which Vχ = 0.
We may then replace V by this sum and assume henceforth that V is the
direct sum of Vχ , with H acting as scalar multiplications on each Vχ .
Let W = {x ∈ End(V ) | x(Vχ ) ⊆ Vχ for all χ ∈ X(H)}, which is naturally
isomorphic to χ End(Vχ ). Now GL(V ) acts via conjugation on End(V ) and
ϕ(G) ≤ GL(V ) stabilizes W . Thus we have a group homomorphism ψ : G →
GL(W ), which is a rational representation. We claim that ker ψ = H. Indeed,
if g ∈ H then ϕ(g) acts as a scalar on each Vχ and so conjugating by ϕ(g) is
the identity on W . Conversely, if ψ(g) = 1, then ϕ(g) stabilizes each Vχ and
commutes with End(Vχ ), which implies that ϕ(g) acts as a scalar on each
Vχ . In particular ϕ(g) stabilizes the line L and by Theorem 5.5, g ∈ H. We
now have a group isomorphism G/H ∼ = im(ψ). To see that this is indeed an
isomorphism of algebraic varieties, refer to [4, Thm. 6.8].
Remarks 5.8 (a) In the situation of Proposition 5.7 let π : G → G/H be
the canonical surjection. One checks that π ∗ : k[G/H] → k[G] is injective.
Hence k[G/H] may be viewed as a subalgebra of k[G]. One can show that
k[G/H] ∼
= {f ∈ k[G] | ρx (f ) = f for all x ∈ H},
that is, k[G/H] is precisely the algebra of H-invariant functions on G.
(b) It follows from Proposition 5.7 that the quotient G/G considered in
the proof of Theorem 4.4 is a linear algebraic group.
Example 5.9 Let G = GLn , H = Z(G) = {tIn | t ∈ k × }. Then the
projective general linear group PGLn := GLn /Z(GLn ) is a linear algebraic
group of dimension
dim(PGLn ) = dim(GLn ) − dim(Z(GLn )) = n2 − 1
(see Exercise 10.17 for a direct proof).
Similarly one obtains the projective conformal symplectic group
PCSp2n := CSp2n /Z(CSp2n )
as well as the projective conformal orthogonal group and its connected com-
ponent of the identity
PCO2n := CO2n /Z(CO2n ), PCO◦2n = CO◦2n /Z(CO◦2n )
as linear algebraic groups.
6
Borel subgroups
Remarks 6.2 (a) The conclusion of the theorem in fact holds more gener-
ally when X is just assumed to be a complete variety (see [32, Thm. 21.2]).
(b) The Lie–Kolchin Theorem (Theorem 4.1) asserts that for G a con-
nected, solvable subgroup of GL(V ), there exists 0 = v ∈ V such that
g.v ∈ v for all g ∈ G, that is, G, acting on P(V ), has a fixed point. Borel’s
fixed point theorem can hence be used to give a short proof of this assertion.
Clearly, Borel subgroups exist: just take a closed, connected, solvable sub-
group of maximal possible dimension. In fact, there is not much choice (see
[32, §21.3]):
Proof of (a) assuming (b) Firstly, we consider the case G = GLn . The
group Tn of upper triangular matrices is a closed, connected, solvable sub-
group of G. If H is any closed, connected, solvable subgroup of G, then
Corollary 4.2 gives H ≤ g −1 Tn g for some g ∈ G. Therefore, Tn is maximal
among closed, connected, solvable subgroups of G, hence a Borel subgroup,
and any Borel subgroup is conjugate to Tn .
For the general case we can replace G by G◦ as Borel subgroups of G◦ are
Borel subgroups of G. Let H be a Borel subgroup of G. Assuming G/B is
projective for a fixed Borel subgroup B, H acts on the projective space G/B
via gB → hgB. By Theorem 6.1, there exists a fixed point, that is, there
exists g ∈ G such that HgB = gB. Therefore, g −1 Hg ≤ B. By maximality
of H, we conclude that H = gBg −1 .
38 Borel subgroups
Corollary 6.5 Let G be a linear algebraic group. Then all maximal tori of
G are conjugate.
Note that by Corollary 6.5, as for connected solvable groups, maximal tori
are the tori of maximal dimension so the rank is well-defined.
Example 6.7 Let’s find Borel subgroups, maximal tori and the rank of
classical groups.
(1) For G = GLn , Tn is a Borel subgroup (see the proof of Theorem 6.4),
and Dn is a maximal torus of Tn , hence of G. Therefore, rk(GLn ) =
dim(Dn ) = n.
(2) For G = SLn , if we repeat the argument given for GLn , we observe that
Tn ∩ SLn is a Borel subgroup and Dn ∩ SLn is a maximal torus of SLn .
The exact sequence
det
1 −→ Dn ∩ SLn −→ Dn −→ Gm −→ 1
But it is easily checked from the definition of GO2n in Section 1.2 that
D2n ∩ GO2n = T , so T = T1 is a maximal torus and rk(SO2n ) = n.
Analogously one shows that rk(Sp2n ) = rk(SO2n+1 ) = n (see Exer-
cise 10.19).
A1 A2
(4) Let B := SO2n ∩T2n . A short calculation shows that ∈ SO2n
0 A3
1 Kn A3 = Kn , and A3 Kn A2 = −A2 Kn A3 , that is, A3 =
if and only if Atr tr tr
−tr tr −1
Kn A1 Kn , and A3 Kn A2 = Kn A1 A2 is skew-symmetric. (Recall that,
by our definition, SO2n = GO◦2n in any characteristic.) In particular, B
is the product of
In Kn S
S ∈ k n×n
skew-symmetric
0 In
with {diag(A, Kn A−tr Kn ) | A ∈ Tn } ∼
= Tn , of dimension
% & % &
n
2 +
n+1
= n2 , if n ≥ 2
dim(B) = %n+1& 2
2 = n2 , if n = 1.
We claim that B is a Borel subgroup of SO2n . Let B1 ≥ B be a Borel
subgroup of SO2n . Then B1 lies in a Borel subgroup of GL2n , so it sta-
bilizes a complete flag on V = k 2n . The only one-dimensional subspace
invariant under B is v1 , generated by the first standard basis vector.
Thus, this must also be fixed by B1 . But then B1 also fixes v1 ⊥ , a sub-
space of codimension 1. Now applying induction to the action of B1 , B
on v1 ⊥ /v1 one sees that B1 can only fix the standard flag, hence is
contained in T2n . It follows that B1 = B, as claimed.
Analogously one shows that Sp2n ∩ T2n , respectively SO2n+1 ∩ T2n+1
is a Borel subgroup of Sp2n , respectively SO2n+1 (see Exercise 10.19).
with di := dim Vχi and χi (t)di = 1. But there exist only finitely many such
matrices.
(c) By Proposition 1.18, [G, G] is connected. As R([G, G]) is a charac-
teristic normal subgroup of [G, G], R([G, G]) ≤ R(G). Thus R([G, G]) ≤
(R(G) ∩ [G, G])◦ = 1 by part (b), hence [G, G] is semisimple.
We also have that G = R(G)[G, G], but the proof of this statement requires
considerably more work (see Corollary 8.22 later).
7
The Lie algebra of a linear algebraic group
We now give a second, more geometric construction for the Lie algebra of
an algebraic group G.
For an affine variety X we define the tangent space of X at x ∈ X by
Tx (X) := {δ : k[X] → k linear | δ(f g) = f (x)δ(g)+δ(f )g(x) for f, g ∈ k[X]}
(the k-vector space of point derivations at x). In case X is a linear algebraic
group G, then as G acts homogeneously on itself by left translation, the
tangent space at any group element g ∈ G is naturally isomorphic to T1 (G).
Thus, we will consider the tangent space T1 (G). Now
Θ : Lie(G) → T1 (G), Θ(D)(f ) := D(f )(1),
is a k-linear map. In fact, a straightforward calculation shows this to be an
isomorphism of vector spaces, see [32, Thm. 9.1]. By this, the Lie algebra
structure on Lie(G) can be transported to T1 (G). With this alternative in-
terpretation one can establish the following (see [66, Thm. 4.3.7, Cor. 4.4.6,
§4.1.7] for (a) and (b), [32, §5.1] for (c)):
while
"
λg (DX Tij ) = λg ( Til Xlj )
l
"
= Xlj λg (Til )
l
" " ""
−1 −1
= Xlj gim Tml = Xlj gim Tml .
l m l m
Furthermore, the tangent space approach allows one to complete the func-
torial connection between linear algebraic groups and their Lie algebras.
(a) dx (ψ ◦ ϕ) = dϕ(x) ψ ◦ dx ϕ.
7.1 Derivations and differentials 47
The proof is given in [32, §5.4, Thm. 9.1] and [66, Cor. 5.3.3].
The proof of the first part is a direct calculation, see [32, Lemma 9.4], and
uses the identification of Lie(H) with T1 (H). For the second part see [32,
Cor. 10.4A and Sec. 11.5].
Ad : G −→ GL(Lie(G)), x → Ad x,
See [32, Prop. 10.3, Thm. 10.4] for (a) and (b). The statement of (c)
will follow from Theorem 8.17, see Exercise 10.32. Thus, if G is connected
reductive, Ad is almost faithful: only the center of G is lost.
The following is easily verified: if H ≤ G is closed, then Lie(H) (as a Lie-
subalgebra of Lie(G) according to Theorem 7.9) is Ad G (H)-invariant and
the restriction of the adjoint representation of G to H, acting on Lie(H), is
just the adjoint representation of H, i.e., Ad G |H = Ad H on Lie(H).
Example 7.13 Let G = GLn , so Lie(G) ∼ = gln by Example 7.5 and for
X ∈ gln we have the derivation DX ∈ Lie(G) given by DX (Tij ) = l Til Xlj .
Via the vector space isomorphism Θ defined above, we can associate a point
derivation δX to X, with δX (Tij ) = ( l Til Xlj )(In ) = Xij .
For g = (gij ) ∈ G, we evaluate (Intg )∗ (Tij ): For x = (xij ) ∈ G,
"
−1
(Intg )∗ (Tij )(x) = Tij (gxg −1 ) = gil gmj xlm ,
l,m
−1
so (Intg )∗ (Tij ) = l,m gil gmj Tlm . Thus Ad g = dIntg : Lie(G) → Lie(G)
50 The Lie algebra of a linear algebraic group
satisfies
(Ad g)(δX )(Tij ) = δX ((Intg )∗ Tij )
"
−1
= δX ( gil gmj Tlm )
l,m
"
−1
= gil gmj δX (Tlm )
l,m
"
−1
= gil gmj Xlm = (gXg −1 )ij .
l,m
It turns out that the best way to investigate reductive algebraic groups is
via their adjoint action on the Lie algebra. We start by decomposing the Lie
algebra according to the action of a maximal torus. Then, as a preparation
for the case of an arbitrary semisimple group we consider the smallest non-
solvable groups in some detail.
Φ(G) := {χ ∈ X(T ) | χ = 0, gχ = 0}
(1) Let G = GLn , with Lie algebra gln (see Example 7.5). By Example 7.13
the adjoint action of G on gln is by conjugation. Let T = Dn . Then
Lie(T ) is the set of diagonal matrices in gln by Exercise 10.26. Let Eij ∈
gln be the n × n-matrix whose (l, m) entry is δil δjm . A straightforward
calculation gives
of order |Φ(G)| = n(n − 1). Note that all gα , α ∈ Φ(G), are one-
dimensional. The Weyl group is
These actions can easily be seen to be faithful, and moreover they are com-
patible with the pairing , : X(T ) × Y (T ) → Z defined in Example 3.5
(see Exercise 10.30):
We now claim that W stabilizes the set of roots Φ(G) ⊆ X(T ). Indeed, for
n ∈ NG (T ), α ∈ Φ(G) and v ∈ gα , (Ad n)(v) is again a common eigenvector
8.2 Semisimple groups of rank 1 53
We are thus led to consider carefully the structure of the rank 1 group
PGL2 = GL2 /{aI2 | a ∈ k × }. For this, we investigate the action of PGL2
on its Lie algebra Lie(PGL2 ). Write X̄ for the image of X ⊆ GL2 under
the natural projection π : GL2 → PGL2 . By dimension considerations one
sees that T2 is a Borel subgroup of PGL2 and T := D2 is a maximal torus
contained in T2 . A second Borel subgroup containing T is the image of the
lower triangular matrices. By Example 7.10 we have
Lie(PGL2 ) ∼
= gl2 /{aI2 | a ∈ k}.
This is the unipotent radical of our chosen Borel subgroup T2 and is therefore
normalized by T . Fix an isomorphism
1 c
u : Ga −→ U2 , u(c) = .
0 1
and the definition of n. The assertion of (b) now follows by comparing di-
mensions.
Finally, Lie(U ∩ nU n−1 ) ⊆ Lie(U ) ∩ Lie(nU n−1 ) = gα ∩ g−α = 0, so the
group U ∩nU n−1 is finite by Theorem 7.4(b), unipotent and normalized by T .
Hence this group lies in CG (T ) = T by Exercise 10.4. So U ∩nU n−1 = 1.
The above considerations finally enable one to establish the following clas-
sification of rank 1 semisimple groups ([66, Thm. 7.2.4]):
Theorem 8.8 Any semisimple group of rank 1 is isomorphic to SL2 or
PGL2 .
Example 8.9 For later use, let’s have a closer look at the case of SL2 . By
Example 8.2(2) we get the same root space decomposition
sl2 = (sl2 )0 + (sl2 )α + (sl2 )−α
= D2 ∩ SL
of Lie(SL2 ) = sl2 with respect to T 2 as for the action of PGL2 ,
c
where now α ∈ X(T ) is defined by α = c2 . Note that α = 2X(T )
c−1
in this case.
Remark 8.10 In our calculations for PGL2 above we also found two roots
±β and a Weyl group of order two interchanging these two. That is, the
groups SL2 and PGL2 have the same Weyl group and “isomorphic root
systems”.
We finish this section with the following extension of our results on groups
of rank 1:
Proposition 8.11 Let G be connected reductive of semisimple rank 1,
T ≤ G a maximal torus, Z = Z(G) and G = [G, G]. Then:
(a) CG (T ) = T and thus Z ≤ T .
(b) G is semisimple of rank 1 with maximal torus T ∩ G , and G = G Z.
Proof By Proposition 6.20, the radical R(G) = Z(G)◦ is a torus, so con-
tained in T , G is semisimple and G ∩ R(G) is finite. As G/R(G) is assumed
to be semisimple of rank 1 we have dim(G/R(G)) = 3 by Proposition 8.6,
and G/R(G) is isomorphic to SL2 or PGL2 by Theorem 8.8. Now observe
that a maximal torus of SL2 or PGL2 is self-centralizing. So T and CG (T )
have the same image in G/R(G) under the canonical epimorphism, and hence
T ≤ CG (T ) ≤ T R(G) = T .
Since the product map R(G) × G → G has finite kernel we have
dim G ≤ dim G − dim R(G) = dim G/R(G) = 3.
8.3 Structure of connected reductive groups 57
Note that one inclusion is given by Proposition 6.16. For a reductive group,
in particular, we obtain:
Proof Part (a) follows directly from the preceding theorem, applied to a
maximal torus T of G containing S, and Theorem 6.19. The second assertion
then follows from (a) with S = T , Proposition 6.18 and Example 6.17(2).
We now study general reductive algebraic groups via their adjoint action
on the Lie algebra. Let G be a non-solvable connected reductive algebraic
group. We keep the notations T , g = Lie(G), Φ and W as in Section 8.2.
Let α ∈ Φ; then Tα := (ker α)◦ ≤ T is a subtorus of T by Proposition 3.9,
of codimension 1, and we let Cα := CG (Tα ) be its centralizer.
Proof Part (a) and thus (b) follows from Propositions 7.14 and 8.16. Now let
uα : Ga → Gα be the morphism u defined in the proof of Proposition 8.7,
relative to the torus T1 . Then for all t ∈ T , c ∈ k, we have tuα (c)t−1 =
uα (α(t)c) as required, since T = T1 ·ker(α). The argument that im(duα ) = gα
is as in the proof of loc. cit. Let Uα := im(uα ). Suppose that u : Ga → G is
a homomorphism as in (c), then clearly V := im(u ) ≤ Cα . Since
(R1) Φ is finite, 0 ∈
/ Φ, Φ = E;
(R2) if c ∈ R is such that α, cα ∈ Φ, then c = ±1;
(R3) for each α ∈ Φ there exists a reflection sα ∈ GL(E) along α stabilizing
Φ;
(R4) (crystallographic condition) for α, β ∈ Φ, sα .β − β is an integral mul-
tiple of α.
64 The classification of semisimple algebraic groups
1 2 3 n 1 2 n−1 n
An Bn >
n≥1 n≥2
n−1
1 2 n−2 1 2 n−1 n
Dn Cn <
n≥4 n≥3
n 1 3 4 5 6
E6
1 2 1 2 3 4
G2 < F4 > 2
1 3 4 5 6 7 1 3 4 5 6 7 8
E7 E8
2 2
A root system or Dynkin diagram is said to be simply laced if all roots have
the same length, or equivalently, if all mα,β ∈ {0, 1}. Thus the indecompos-
able, simply laced Dynkin diagrams are those of types An , Dn , E6 , E7 , E8 .
Finite reflection groups arising from root systems only satisfying (R1)–
(R3), but not necessarily the crystallographic condition (R4), are the so-
called finite Coxeter groups. Interestingly, in addition to the Weyl groups
arising from Theorem 9.6 above, only the two-dimensional dihedral groups
and two further indecomposable cases, denoted H3 and H4 , in dimension 3,
respectively 4, occur. For more information on this and on the indecompos-
able root systems and their Weyl groups, see for example [34, §2.7–2.11].
base of Φ is given by
∆ := {χi,i+1 | 1 ≤ i ≤ n − 1}.
Indeed,
χij = χi,i+1 + χi+1,i+2 + . . . + χj−1,j for i < j,
from which we see that the root system of SLn is of type An−1 .
Example 9.9 Let’s compute the coroots for SL2 and PGL2 .
(1) G = SL2 . The character group is generated by χ, where χ t
t−1 = t.
We have seen in Example 8.9 that Φ = {±α}, where α t t−1 = t2 . So
ZΦ = 2χ and X = Zχ. The coroot α∨ is given by α∨ : t → t t−1 ,
thus ZΦ∨ = Y .
(2) G = PGL2 . By Section 8.2 we have Φ = {±β}, where β t
1 = t; in
9.2 The classification theorem of Chevalley 69
It follows easily that whenever (X, Φ, Y, Φ∨ ) is a root datum then the Weyl
groups of Φ and of Φ∨ are isomorphic via sα → sα∨ (see Exercise 10.34).
The preceding definition is justified by the following result:
Y := {γ ∈ Y | γ(Gm ) ≤ T } = Y (T )
and
X := {χ|T | χ ∈ X} ∼
= X/Ann(Y ) = X(T ),
from a simply connected group Gsc and to an adjoint group Gad , each with
root system Φ, with ker(π1 ) ∼
= Λ(G)p , ker(π2 ) ∼
= (Λ(Gad )/Λ(G))p , where
p = char(k), and such that dπi is an isomorphism for i = 1, 2.
Proof for the case that char(k) is prime to |Λ(Φ)| Let T be a maximal to-
rus of G with corresponding root system Φ. Then Z := Z(G) ≤ CG (T ) = T ,
$
so Z = α ker(α). Since Z lies in the kernel of the adjoint representation, G
acts via G/Z on Lie(G), so the roots of G and G/Z are the same under the
natural inclusion X(T /Z) ≤ X(T ) induced by the surjection T → T /Z (see
Exercise 10.13). In particular, for any subgroup S ≤ Z, we have inclusions
ZΦ ≤ X(T /Z) ≤ X(T /S) ≤ X(T ). In fact, as gcd(char(k), |Λ|) = 1, Propo-
sition 3.8 implies that X(T )/ZΦ ∼ = Z. Hence, X(T /Z)/ZΦ = 0 and G/Z is
of adjoint type; so we take π2 : G → G/Z to be the natural surjection.
On the other hand, starting with Gsc and Λ1 ≤ X(Tsc )/ZΦ the funda-
mental group of G, then with S = Λ⊥ 1 ≤ Z(Gsc ) the quotient Gsc /S has
root datum with fundamental group Λ1 . With π1 : Gsc → Gsc /S we then
have G ∼ = Gsc /S. The assertion that dπi is an isomorphism follows from
Theorem 7.9(b).
For the general case, the claim follows from the isogeny theorem [66,
Thm. 9.6.5] which asserts that any morphism of root data induces an isogeny
of corresponding semisimple groups.
Table 9.2 gives a list of the possible isogeny types for simple groups and
an identification with various classical groups (see e.g. [73, p.45]).
72 The classification of semisimple algebraic groups
For more information on the various types of classical groups, see for
example the books by Dieudonné [21], Grove [30], or Goodman and Wallach
[27] (for k of characteristic 0).
Remark 9.17 Steinberg has shown that Gsc is the universal perfect central
extension, in the category of abstract groups, of any semisimple group with
9.2 The classification theorem of Chevalley 73
root system Φ [73, Thm. 10]. There is no such universal central extension
in the category of algebraic groups since there exist bijective morphisms of
simple groups which are not isomorphisms. These isogenies will play a cru-
cial role in Part III. Over the field k = C of complex numbers the groups Gsc
are in fact simply connected in the complex topology, and for G of arbitrary
type, Λ(G) is the topological fundamental group [73, Thm. 13]. Finally we
note that the group Gad is isomorphic to its image under the adjoint repre-
sentation, which explains the denomination. (It is not however the case that
the image of the adjoint representation of any group is necessarily of adjoint
type. See Exercise 10.37.)
There is another universal property of groups of simply connected type
which will be needed later on:
Proposition 9.18 Let G be semisimple of simply connected type. Then
whenever π : H1 → H2 is an isogeny of semisimple groups with dπ an
isomorphism, any isogeny ϕ : G → H2 lifts to an isogeny ψ : G → H1 such
that ϕ = π ◦ ψ.
Proof Let G̃ := {(g, h) ∈ G × H1 | ϕ(g) = π(h)}, a closed subgroup of
G × H1 . The projection pr1 : G̃ → G onto the first factor is a surjective
morphism with finite central kernel 1 × C, where C = ker(π), hence its
restriction to G1 := G̃◦ is an isogeny.
pr2
G1 ⊆ G × H1 /H
r8 1
r
ψ
r
pr1
r π
r r
G / H2
ϕ
[Hint: For (c) first show that H H̄ ⊆ H̄. For (d), first show that if U, V ⊆ G are
dense open subsets, then U V = G.]
[Hint: Look at the last non-trivial term in the descending central series and use
Proposition 1.18.]
Exercise 10.16 The purpose of this exercise is to point out that the struc-
ture of connected unipotent groups is difficult to classify.
(a) Show that
⎧⎛ ⎞ ⎫
⎨ 1 a b ⎬
H1 = ⎝0 1 a⎠ a, b ∈ k
⎩ ⎭
0 0 1
is a closed connected two-dimensional commutative unipotent algebraic
group.
(b) Show that H1 ∼ = Ga × Ga as algebraic groups if and only if k is a field
of characteristic different from 2.
(c) Show that
⎧⎛ ⎞ ⎫
⎪ 1 t u s ⎪
⎪
⎨⎜ ⎪
⎬
0 1 0 u ⎟
H2 = ⎜ ⎝
⎟ s, t, u ∈ k ≤ Sp4
⎠
⎪
⎪ 0 0 1 −t ⎪
⎪
⎩ ⎭
0 0 0 1
is a three-dimensional connected unipotent algebraic group.
(d) Determine [H2 , H2 ]; conclude that H2 is abelian if and only if char(k) =
2.
[Hint: For (b) note that if char(k) = 2, H1 has elements of order 4.]
Exercise 10.17 The goal of this exercise is to see that PGLn , the projec-
tive linear group in dimension n, is a linear algebraic group.
Set PGLn = GLn /Z, where Z = {cIn | c ∈ k × }, as abstract group. Let V
be the n-dimensional vector space over k on which GLn naturally acts and
V ∗ = Hom(V, k) the dual space. Consider the action of GLn on V ⊗ V ∗ . This
defines a group homomorphism ρ : GLn → GLn2 .
(a) Show that ρ is a morphism of algebraic groups and hence its image is a
closed subgroup H of GLn2 .
(b) Show that ker ρ = Z.
Hence we conclude that PGLn is isomorphic, as an abstract group, to the
closed subgroup H of GLn2 and thus can be given the structure of a linear
algebraic group.
Exercises for Part I 77
(a) Show that the fixed point set X g := {x ∈ X | g.x = x} is closed for any
g ∈ G. Conclude that X G := {x ∈ X | g.x = x for all g ∈ G} is closed.
(b) Let Y and Z be closed subsets of X and set TranG (Y, Z) := {x ∈ G |
x.Y ⊆ Z} (the transporter of Y into Z). Show that TranG (Y, Z) is a
closed subset of G.
(c) Show that for H a closed subgroup of G, NG (H) is a closed subgroup.
[Hint: For (a), consider the morphism ψ : X → X × X, x → (x, g.x), and use that
the diagonal in X × X is closed. For (b), consider the maps ϕy : G → X given by
x → x.y. Then TranG (Y, Z) = y∈Y (ϕ−1 y (Z)).]
Exercise 10.20 Show that a connected linear algebraic group with nilpo-
tent Borel subgroup is solvable. Conclude that any two-dimensional con-
nected group is solvable.
[Hint: Consider a counterexample of minimal dimension and use Exercise 10.6 and
Proposition 6.8.]
(a) Let G be a linear algebraic group. Show that the radical R(G) is equal
$
to ( B B)◦ , where B runs over all Borel subgroups of G.
(b) Show that Sp2n is semisimple.
(a) Let G be a connected linear algebraic group whose elements are semisim-
ple. Show that G is a torus.
(b) Give an example of an algebraic group (closed subgroup of GLn ) which
consists of semisimple elements but which is not conjugate to a subgroup
of Dn , the group of diagonal matrices.
(so S is a torus of P ). Find NP (S), NP (S)◦ , CP (S), and CP (S)◦ and verify
that NP (S)/CP (S) is finite.
Exercise 10.25 Let G be a linear algebraic group. Show that the inversion
morphism i : G → G, g → g −1 , on G has differential di(X) = −X, for
X ∈ Lie(G).
[Hint: Apply Example 7.8(1) and Proposition 7.7(a) to the morphism µ ◦ (i, id) :
G × G → G.]
Exercise 10.28 (a) Let G be a linear algebraic group. Show that Z(G) ≤
ker(Ad ).
(b) Determine ker(Ad ) for each of the following groups: GLn , Un , Tn .
Exercises for Part I 79
(c) This example shows that ker(Ad ) may be larger than Z(G). Let k be a
⎧⎛ ⎞ ⎫
⎨ a 0 0 ⎬
field of characteristic p > 0. Let G = ⎝0 ap b ⎠ a ∈ k × , b ∈ k .
⎩ ⎭
0 0 1
Show that Z(G) ker(Ad ) G.
(a) In the case n = 2, find the roots and root subspaces of Lie(G). For each
root α, exhibit a one-dimensional closed subgroup Uα ≤ G whose Lie
algebra is the corresponding root subspace.
(b) Generalize to the case of arbitrary n. Show that the root system of Sp2n
is of type Cn . Conclude that dim Sp2n = 2n2 + n.
Proof of (b)–(d) Parts (b) and (c) are just Proposition 9.4(c) together with
Proposition 9.2. For (d) note that Uα , U−α contains a preimage of the sim-
ple reflection sα , for α ∈ ∆, so H := Uα | α ∈ ±∆ contains preimages of all
w ∈ W by (c). Hence Uβ ≤ H for all β ∈ Φ by Proposition 9.4(c) and The-
orem 8.17(e). Then (d) follows from Theorem 8.17(g). See [66, Prop. 8.2.1]
for a proof of (a).
Remark 11.3 Recall from Theorem 8.17(f) and Proposition 8.15 that the
subgroup Uα , U−α is non-solvable, so it follows from Theorem 11.1(a) that
Uα ⊆ B for α ∈ Φ− .
Uij := In + Eij .
We investigate the structure of B and Ru (B) a bit further; for this, let’s
choose and fix isomorphisms uα : Ga → Uα , for all α ∈ Φ.
11.1 On the structure of B 85
does not involve uβ any more, but a non-trivial element from Uγ occurs in
the expression. So it is a non-trivial product of smaller length, contradicting
our minimal choice.
The following consequence will be used in the investigation of maximal
rank subgroups:
Corollary 11.6 Let G be connected reductive with maximal torus T and
H ≤ G a connected reductive subgroup normalized by T . Then
H = T ∩ H, Uα | Uα ≤ H.
Proof Note that HT is a reductive subgroup of G. Let BH be a Borel
subgroup of HT containing T . It lies in a Borel subgroup B of G, hence
so does U := Ru (BH ). Application of Proposition 11.5 shows that U =
α∈M Uα for some subset M ⊆ Φ. But then HT = T, Uα | α ∈ ±M by
Theorems 8.17(g) and 11.1. As HT is reductive, we have U ≤ [HT, HT ] ≤ H
and the claim follows.
Example 11.7 (Root system and root subgroups of SO2n ) Let G = SO2n
with Borel subgroup B = G ∩ T2n and maximal torus T = G ∩ D2n (see
Example 6.7). Here,
Uij := I2n + Eij − E2n−j+1,2n−i+1 , 1 ≤ i < j ≤ 2n − i,
are one-dimensional connected subgroups of the unipotent radical Ru (B) =
86 BN-pairs and Bruhat decomposition
where the product is over all integers m, n > 0 such that mα + nβ ∈ Φ, taken
according to the chosen ordering.
The cmn
αβ above are called structure constants. Note that all unipotent
elements of a connected reductive group G are contained in its derived sub-
group, which is semisimple, and also that any semisimple group is a quotient
of a group of simply connected type by a subgroup consisting of semisimple
elements. So it suffices to prove the above statement for semisimple groups
of simply connected type. The above, together with more precise versions of
Theorem 8.17(c) and (e), yields the very important Steinberg presentation
of a semisimple group G, see [73, §6] or [13, Thm. 12.1.1].
From now on, we fix a total ordering on Φ and, for each α ∈ Φ, an
isomorphism uα : Ga → Uα , c → uα (c), for which Theorem 11.8 holds.
Example 11.9 We determine the commutator relations for SL3 and Sp4 .
(1) For the group G = SL3 , with B and T as in Example 9.8, by Exam-
ple 11.4 the root groups of G with respect to T are
Φ An (n ≥ 2) Dn (n ≥ 5) D4 E6
ΓG Z2 Z2 S3 Z2
Proof for type An−1 For An−1 , n ≥ 3, there is a unique non-trivial graph
automorphism of the Dynkin diagram, which sends αi to αn−i , for 1 ≤ i ≤
n−1 (with the simple roots labeled as in Table 9.1). An easy calculation with
the root subgroups in Example 11.4 shows that
theproduct of the transpose-
0 . .1
inverse automorphism with conjugation by . · diag(1, −1, 1, −1, . . .) is
1 0
an automorphism of SLn with the required property. Factoring out the center
we obtain a corresponding automorphism on the adjoint group PGLn .
The automorphisms in Theorem 11.12 are called graph automorphisms
of G. For SO2n , of type Dn , elements of GO2n \ SO2n induce a non-trivial
graph automorphism of order 2 (see Exercise 20.1). The exceptional graph
automorphisms for D4 , of order 3, and for E6 , of order 2, are realized inside
suitable larger groups, see Examples 12.12 and 13.9.
Remark 11.13 There exist further abstract group automorphisms of a
simple algebraic group G, but which are not invertible as morphisms. For
example, any field automorphism of the underlying field k extends to an
automorphism of G. Also, if G is of type B2 or F4 in characteristic 2, or of
type G2 in characteristic 3, and k is perfect, there exists a bijective endo-
morphism which comes from the non-trivial symmetry of the corresponding
Coxeter diagram (see [73, Thm. 29 and Cor.]). These will be important in
Part III.
for any w ∈ W (note that by Theorem 11.8 this is a subgroup of G). With
92 BN-pairs and Bruhat decomposition
this we have the following extension of Theorem 8.17(g) (see [32, Thm. 28.3
and 28.4]):
Theorem 11.17 (Bruhat decomposition) Let G be a group with a BN-pair.
Then
)
G= B ẇB
w∈W
by assumption. So v = v s = w.
On the other hand, an easy induction on (w) shows from (BN4) that the
union of the B ẇB, where w runs over W , is closed under multiplication.
Since it contains B and N , it is all of G by (BN1).
The proof of the uniqueness assertion uses in an essential way that G is
reductive.
Corollary 11.18 In the notation of Theorem 11.17, there exists a unique
element w0 ∈ W such that w0 (∆) = −∆. Moreover w02 = 1 and B ẇ0 ∩B = T .
Proof The first part follows directly from the simple transitivity of W on
the set of bases, see Proposition 9.4(b). Now assume that g ∈ B ẇ0 ∩ B, so
g = ẇ0 utẇ0 = t u for some u, u ∈ U = Ru (B), t, t ∈ T . Thus
ẇ0−1 t u = uẇ0 tẇ0 .
As Uw−0 = U by the defining property of w0 , the uniqueness statement in
Theorem 11.17 shows that u = 1 = u, whence g = t ∈ T . The other
inclusion is clear since ẇ0 normalizes T .
11.2 Bruhat decomposition 93
It can be shown that the map π above is even an isomorphism onto its
image, see [32, 28.5].
12
Structure of parabolic subgroups, I
Proof (a) By Lemma 11.14 we have B ẇB ·B ṡB ⊆ PI for all w ∈ WI and all
s ∈ I. Hence, by an easy induction, PI ≤ G. Now assume that g ∈ NG (PI ).
Then B, B g are two Borel subgroups of PI , hence conjugate by some p ∈ PI
by Theorem 6.4(a). Then gp ∈ NG (B) = B, by Theorem 6.12, so g ∈ PI .
In (b), if PJ = PIg with g ∈ G then B, B g are Borel subgroups of PJ ,
whence g ∈ PJ as before, so PI = PJ . Next, if PI = PJ for I, J ⊆ S then
by (a) we must have WI = WJ . This implies that WI = WI∪J , so we may
assume that I ⊆ J. But the elements of S are reflections on XR , so the fixed
space of WI on XR has codimension at most |I|. On the other hand, by
Corollary 8.22 and the proof of Theorem 8.21, the fixed space of W = WS
on XR has codimension rkss (G), which equals |∆| by Theorem 11.1(b), so
we must in fact have equality. Thus WI = WJ implies that I = J.
For (c), write Φ± ±
I := ΦI ∩ Φ . Since B ⊆ PI we have Uα ⊆ PI for all
α ∈ Φ+ . Since ΦI is a root system by Proposition 12.1, the longest element
− −
w ∈ WI satisfies w(Φ+ I ) = ΦI (see Corollary 11.18). Thus, for β ∈ ΦI
−1
there exists α ∈ ΦI such that wα = β. Then Uβ = Uwα = ẇUα ẇ ⊆ PI ,
+
which proves the inclusion “⊇”. For the converse note that if α ∈ ∆ is the
simple root corresponding to s ∈ S, then we may choose ṡ ∈ U±α . So
T, Uα | α ∈ ΦI contains preimages of all s ∈ I, hence of all w ∈ WI ,
which in view of the decomposition in (a) gives the other inclusion. The
connectedness of PI now follows by Proposition 1.16.
Finally, all overgroups of B are of the form PI for some I ⊆ S by Exer-
cise 20.3.
Note that all statements except for (c) and the connectedness of PI remain
valid for arbitrary BN-pairs, see Exercise 20.3.
12.1 Parabolic subgroups 97
More generally, parabolic subgroups in GLn and in SLn are just the sta-
bilizers of flags of subspaces in their natural n-dimensional representation
space. In fact, the parabolic subgroups of all classical groups (groups with
root system of type An , Bn , Cn or Dn ) have an interpretation in terms of
the natural module for the group. We give a precise statement and proof in
the next section (see Proposition 12.13).
The BN-pair setting also allows one to give an easy proof that simple
algebraic groups (as defined in Theorem 8.21) are simple modulo center as
abstract groups; for this recall from the discussion before Theorem 9.6 that
the Weyl group of a simple algebraic group is irreducible.
For later use let’s also point out the following (see Exercise 20.4):
P = StabCl(V ) (0 ⊂ V1 ⊂ . . . ⊂ Vr ⊂ V ),
Proof To simplify the exposition, throughout this proof we will use the
terms “singular” and “totally singular” for isotropic vectors and subspaces
in the case where the space is equipped with a symplectic form or the zero
form.
Given a flag F of totally singular subspaces, refine this to obtain a maximal
flag inside a maximal totally singular subspace of V . By Witt’s lemma (see [2,
§20]), all maximal singular subspaces, and moreover all maximal flags inside
a given maximal singular subspace are Isom(V )-conjugate. By the explicit
constructions in Example 6.7, there exists a Borel subgroup of G stabilizing
a maximal flag of a maximal singular subspace of V . So the stabilizer of F
contains a Borel subgroup and hence is parabolic by Proposition 12.2.
In addition to the above remarks, we see that the explicit constructions of
Borel subgroups in Example 6.7 also show that the Borel subgroups of a sim-
ple classical group Cl(V ) act indecomposably on V . Indeed, one first checks
that except when V is an odd-dimensional orthogonal space and char(k) = 2,
the unipotent radical of the given Borel subgroup B has a one-dimensional
fixed point space on V ; as Cl(V ) acts irreducibly this also follows easily from
the theory of highest weights, see Corollary 15.10. In the exceptional case,
one quotients out by the radical of the bilinear form and passes to the sym-
plectic group, where B acts indecomposably. Then the result follows from
the fact that the radical of the form has no B-invariant complement in V .
Now let P ≤ Cl(V ) be a parabolic subgroup with Levi decomposition
P = QL, containing a fixed Borel subgroup B. We first show that if P
stabilizes a flag of subspaces on V with Q acting as the identity on the
quotient spaces, then P is the full stabilizer in G of this flag. Let F be the
flag 0 = V0 ⊂ V1 ⊂ . . . ⊂ Vr+1 = V with P ≤ StabG (F) and with Q
acting as the identity on each of the quotients Vi /Vi−1 . By Proposition 12.2,
StabG (F) is a parabolic subgroup, containing B, say P̂ . By Proposition 12.6,
Ru (P̂ ) ≤ Q. Now let U ≤ P̂ be the subgroup of elements which act as the
identity on all quotients in F. Then U is a closed unipotent, normal subgroup
of P̂ , and so U ◦ ≤ Ru (P̂ ). But Q ≤ U and so Q ≤ U ◦ ≤ Ru (P̂ ) ≤ Q, whence
Q = Ru (P̂ ). By Corollary 12.8, this implies P = P̂ .
To complete the proof, we must find a flag of totally singular subspaces
of V stabilized by P such that Q acts as the identity on the quotients. By
Proposition 2.9, Q has a non-zero fixed point on any representation space,
so P stabilizes the flag 0 = V0 ⊂ V1 ⊂ . . . ⊂ Vr+1 = V , where Vi /Vi−1 :=
(V /Vi−1 )Q , for i ≥ 1. If V is equipped with the trivial form (so G = SL(V )),
this flag satisfies the criteria.
Now suppose V is equipped with a non-degenerate form Q, and assume
for now that char(k) = 2 if V is equipped with a quadratic form. Set W :=
12.2 Levi decomposition 103
Note that by Theorem 11.8 the integers cmn αβ only depend on the root
system, not on the isogeny type of a corresponding semisimple group G.
It is clear from (C1) that closed and p-closed subsets are root systems in
their own right, in the appropriate Euclidean spaces.
See Table B.1 for the list of highest roots in indecomposable root systems.
The extended Dynkin diagrams for the indecomposable root systems are
shown in Table 13.1 (see [34, §4.7]). It is clear from this list that the highest
root is always a long root, if Φ has two root lengths. This can also be shown
without using the classification of root systems, see Corollary B.7.
1 2 3 n 0 1 2 n−1 n
Ãn C̃n > <
n≥2 n≥2
n−1
0
0 2 3 n−2
D̃n
0 n≥4
2 3 n−1 n 1 n
B̃n >
n≥3 1 2 0
G̃2 <
1
0 1 3 4 5 6 7
0 1 2 3 4 Ẽ7
F̃4 >
2
1 3 4 5 6
Ẽ6
1 3 4 5 6 7 8 0
Ẽ8
2
2
0
The following result (see Theorem B.18), which was first shown in [5, §7]
in the context of compact Lie groups exhibits the role played by the extended
Dynkin diagrams in the study of subsystem subgroups:
Theorem 13.12 (Borel–de Siebenthal) Let Φ be an indecomposable root
system with base ∆ and highest root α0 = α∈∆ nα α with respect to ∆.
Then the maximal closed subsystems of Φ up to conjugation by W are those
with bases:
(1) ∆ \ {α} ∪ {−α0 }, for α ∈ ∆ with nα a prime, and
(2) ∆ \ {α} for α ∈ ∆ with nα = 1.
It is thus easy to write down explicitly the possibilities for maximal closed
subsystems in any of the various types of indecomposable root systems, see
Exercise 20.9.
13.2 The algorithm of Borel and de Siebenthal 109
1 3 4 5 6 7 8 0
Ẽ8
Removal of the first node leads to the Dynkin diagram of type D8 , so the
E8 -root system contains a closed subsystem of type D8 . Since the highest
root for E8 is given by α0 = 2α1 + 3α2 + 4α3 + 6α4 + 5α5 + 4α6 + 3α7 + 2α8
(see Table B.1), this is in fact maximal in E8 by Theorem 13.12.
0 7 0 3
2 3 4 5 6 2
D̃8 D̃4
1 8 1 4
the D4 subgroups and hence E8 has a subsystem subgroup with root system
of type (A1 )8 . In particular we have inclusions of subsystem subgroups of
E8 as follows:
(A1 )8 ≤ D4 D4 ≤ D8 ≤ E8 .
Here, by abuse of notation, we identify the group with its root system.
We next discuss the extent to which closed subsystems account for all
subsystem subgroups.
Example 13.16 Let G = Sp4 , with root system of type C2 , with base
{α, β}, where α is short. Here U±β ×U±(2α+β) is a subsystem subgroup of
type A1 A1 corresponding to the subdiagram of the extended Dynkin diagram
of G where the middle node is removed, see Table 13.1.
Now consider the short roots, α, α+β. If p = 2 we have [Uα , Uα+β ] = 1 and
indeed the group generated by the root subgroups corresponding to the short
13.2 The algorithm of Borel and de Siebenthal 111
Proof The inclusion “⊇” is clear in both (a) and (b) (see Theorem 8.17(c)
for the action of s on Uα ). Now let g ∈ CG (s). By the Bruhat decomposition
in Theorem 11.17, g can be uniquely expressed as a product g = u1 ẇtu2
with u1 ∈ Uw− , w ∈ W , t ∈ T and u2 ∈ U = Ru (B). Now
CG (s)◦ lies inside a proper Levi subgroup. Such semisimple elements are
called isolated .
conjugacy classes. See Springer and Steinberg [67], Steinberg [75], and Spal-
tenstein [64] for further information.
We conclude our investigation of conjugacy classes by introducing an im-
portant type of element, which in a sense comprises most elements in a
connected group (see Corollary 14.10 below).
Definition 14.8 An element x of a linear algebraic group G is called
regular if dim CG (x) is smallest possible among all elements of G.
Proposition 14.9 Let G be connected, x ∈ G. Then dim CG (x) ≥ rk(G).
Proof Let B be a Borel subgroup of G containing x (according to Theo-
rem 6.10), with unipotent radical U = Ru (B). As B/U is abelian by Theo-
rem 4.4(a), the conjugacy class [x]B of x in B is contained in the coset xU ,
whence of dimension at most dim U = dim B − dim T , where T denotes a
maximal torus of B, hence of G. Consider the surjective morphism B → [x]B ,
g → gxg −1 , with fibers the cosets of CB (x), of dimension dim CB (x). This
shows that dim[x]B = dim B − dim CB (x) by Proposition 1.19, so
dim B − dim CB (x) = dim[x]B ≤ dim B − dim T = dim B − rk(G),
whence dim CG (x) ≥ dim CB (x) ≥ rk(G) as claimed.
Regular elements play an important role in the study of reductive groups.
For semisimple elements, we have the following characterization:
Corollary 14.10 Let G be connected reductive with maximal torus T and
root system Φ. For s ∈ T the following are equivalent:
(i) s is regular;
(ii) α(s) = 1 for all α ∈ Φ;
(iii) CG (s)◦ = T .
Moreover, regular semisimple elements are dense in G.
Proof By Theorem 14.2(a) the conditions (ii) and (iii) are equivalent. As
dim T = rk(G), any element with CG (s)◦ = T is necessarily regular by
Proposition 14.9, so (ii), (iii) imply (i). Now ker(α) is a (closed) subgroup
of T of codimension 1 for any α ∈ Φ, and |Φ| is finite, so the complement
of ∪α ker(α) is non-empty open in T and thus dense by Proposition 1.9.
In particular, again by Theorem 14.2(a), the bound in Proposition 14.9 is
attained and so (i) implies (iii).
Now let B denote a Borel subgroup of G containing T , with unipotent
radical U . Let Treg denote the open set of regular semisimple elements of
G in T . We claim that Treg · U consists of regular semisimple elements of
116 Centralizers and conjugacy classes
G, and thus regular semisimple elements (of G) are dense in B. Indeed, let
su ∈ Treg U with Jordan decomposition tv, where t is semisimple and v ∈ U .
As U acts transitively by conjugation on the set of maximal tori of B, the
semisimple part t of su is conjugate to s, hence also regular in G. By the
first part, CB (t)◦ is a torus, so v = 1 by Proposition 14.7, and hence su = t
is regular semisimple.
Since G is the union of conjugates of B (see Theorem 6.10), regular
semisimple elements are also dense in G.
For example in GLn , regular elements are those whose characteristic poly-
nomial and minimal polynomial agree. This shows that GLn and SLn also
do contain regular unipotent elements. The proof of this fact for other types
of semisimple groups is much more delicate, see [71, Thm. 4.6].
Proposition 14.13 The bad primes, respectively torsion primes, for the
indecomposable root systems are as given in Table 14.1. The bad primes and
torsion primes for a decomposable root system are those of its indecomposable
components.
14.2 Connectedness of centralizers 117
Proof This can be verified from the criteria involving coefficients of highest
roots in Corollary B.28 and Proposition B.32; see also Exercise B.4.
The fundamental groups for the various simple groups can be read off
from Table 9.2. Comparison with that list shows that for simple groups G
the torsion primes for G are those of its root system Φ, except possibly for
prime divisors of n + 1 for type An , and for 2 for type Cn .
We can now characterize an important situation in which centralizers of
semisimple elements are always connected (see [14, Thm. 3.5.6] for (a), [74,
Thm. 0.1] for (b)):
Note that part (b) generalizes Proposition 12.14: by Proposition 12.6 any
Levi subgroup L of G is the centralizer of its connected center Z(L)◦ ; by
Exercise 20.11 we can choose a semisimple element s ∈ Z(L)◦ of order prime
to all torsion primes with L = CG (s).
Corollary 14.17 Let G be connected reductive with simply connected de-
rived subgroup [G, G], and s1 , . . . , sr ∈ G mutually commuting semisimple
elements of finite order. Let n be the number of si whose order is not prime
to all torsion primes of G. If n ≤ 2 then there exists a maximal torus of G
containing all si .
Proof Assume that s1 , . . . , sr−2 have order not divisible by any torsion
prime of G, and for i = 1, . . . , r put Gi := CGi−1 (si ), with G0 := G. By
Theorem 14.16(a) and (b) applied inductively we see that Gi is connected
and [Gi , Gi ] is simply connected for i = 1, . . . , r − 2. Furthermore, Gr−1 is
still connected and contains sr . Now any maximal torus of Gr−1 containing
sr (see Corollary 6.11(a)) is as claimed.
Corollary 14.18 Let G be a semisimple group, r = char(k) a prime which
is not a torsion prime for G. Then the maximal rank of an elementary abelian
r-subgroup of G equals rk(G).
Proof Let R be an elementary abelian r-subgroup of G and consider the
natural isogeny π : Gsc → G from a simply connected group of the same
type (see Proposition 9.15). Then ker(π) is a subgroup of the fundamental
group of G, so of order prime to r. Hence, the preimage π −1 (R) ≤ Gsc has
a Sylow r-subgroup R̃ isomorphic to R. Since r is not a torsion prime, by
Corollary 14.17 there exists a maximal torus T ≤ Gsc with R̃ ≤ T . Since T ∼=
(k × )l , with l = rk(G), we see that R ∼
= R̃ has r-rank at most l. Conversely,
T clearly contains an elementary abelian r-subgroup of rank l.
Example 14.19 We illustrate Theorem 14.16 and its corollaries.
(1) For G = GLn or SLn the derived group G = SLn (by Exercise 10.36)
is simply connected, and indeed, in both groups any set of commuting
semisimple elements can be simultaneously diagonalized, hence embed-
ded into a maximal torus.
(2) The assumption on torsion primes in Theorem 14.16(b) is necessary:
Let G = G2 with char(k) = 2, a group of simply connected type,
with root system Φ with respect to a maximal torus T ≤ G. Then
the closed subsystem Ψ1 of type A1 A1 defined in Example 14.12 satis-
fies |ZΦ/ZΨ1 | = 2. Let H denote the subsystem subgroup of G of type
A1 A1 corresponding to Ψ1 and containing T (see Theorem 13.6). As G
14.2 Connectedness of centralizers 119
Cases where the conclusion of Corollary 14.18 fails for torsion primes lead
to interesting maximal subgroups, see Theorem 29.3 below.
For semisimple groups, Theorem 14.16 can be generalized as follows: the
extent to which centralizers of semisimple elements can fail to be connected
is controlled by the fundamental group.
0
120 Centralizers and conjugacy classes
Proof By Proposition 2.9, the fixed point space V ρ(Ru (G)) of the unipotent
radical Ru (G) of G is a non-trivial G-invariant subspace of V , so all of V by
irreducibility. That is, Ru (G) ≤ ker(ρ).
So ρ induces an irreducible representation ρ̄ : Ḡ → GL(V ) of the con-
nected reductive group Ḡ := G/Ru (G). Moreover, as G acts irreducibly on
122 Representations of algebraic groups
Thus, for example, the weights in the adjoint representation of G are just 0
and the roots of G. The multiplicity of a non-zero weight (that is, a root) is 1
and the multiplicity of the 0 weight is the rank of G by Theorem 8.17(a). The
weights of the natural representation of GLn with respect to the maximal
torus Dn of diagonal matrices are just the n coordinate functions, again with
one-dimensional weight spaces.
We consider the action of certain elements and subgroups on weight spaces
of V , starting with NG (T ). As seen in Section 8.1, the Weyl group W acts
on X(T ): for w ∈ W and χ ∈ X(T ) we have w.χ(t) = χ(ẇ−1 tẇ), where
ẇ ∈ NG (T ) is an arbitrary preimage of w under the natural map NG (T ) →
NG (T )/T = W . This leads to the following generalization of Proposition 8.4:
so ẇv ∈ Vw.χ . Hence ẇ(Vχ ) ⊆ Vw.χ and ẇ−1 (Vw.χ ) ⊆ Vχ , and the result
follows.
Recall from Theorem 8.17(g) that a connected reductive group G is gener-
ated by a maximal torus and its root subgroups Uα , for α ∈ Φ. We describe
the action of the root groups Uα on the Vχ .
Lemma 15.4 Let G, T , Φ be as above, ρ : G → GL(V ) a rational repre-
sentation. Let α ∈ Φ and λ ∈ X(T ) with T -weight space Vλ . Then for all
v ∈ Vλ we have
"
Uα .v ⊆ v + Vλ+mα ,
m∈N
so Uα Vλ ⊆ m∈N0 Vλ+mα .
Proof Fix an isomorphism uα : Ga → Uα , c → uα (c). Since G acts ra-
tionally on V , uα (c).v is given by a polynomial function in c with respect
to some basis for V and coordinates with respect to that basis. Collecting
n
terms with equal c-power, we find vi ∈ V such that uα (c).v = i=0 ci vi , for
some n ≥ 0. For t ∈ T we have on the one hand
"
n
tuα (c)t−1 .v = λ(t−1 )tuα (c).v = λ(t−1 ) ci t.vi ,
i=0
Since this is true for all c ∈ k, the coefficients at equal powers of c have
to agree: λ(t−1 )t.vi = α(t)i vi , that is, t.vi = (λ + iα)(t)vi for all i. Hence
vi ∈ Vλ+iα for all i, and setting c = 0 we also find that v0 = v.
As a consequence, Uα , U−α preserves the submodule m∈Z Vλ+mα of V .
Example 15.5 Let G be connected reductive with root system Φ. For
α ∈ Φ let nα ∈ Uα , U−α as defined in Section 8.4. Applying the above
result to the adjoint action of G on its Lie algebra g, we see that nα preserves
the sum of weight spaces m gβ+mα for all β ∈ Φ, so sα .β = β + mα for
some m ∈ Z. This shows that the root system axiom (R4) is satisfied for Φ,
as claimed in Proposition 9.2; that is, we have sα .β − β an integral multiple
of α.
We continue our study of rational representations ρ : G → GL(V ) of a
connected reductive group G, where V = 0. By Lie–Kolchin (Theorem 4.1),
124 Representations of algebraic groups
there exists v + ∈ V \ {0} such that v + is invariant under the image ρ(B)
of the Borel subgroup B of G. Any such vector v + is called a maximal vector
of V (with respect to B). As v + is stabilized by any maximal torus T of
B, v + ∈ Vλ for some λ ∈ X(T ).
Proposition 15.6 In the notation introduced above, set V = Gv + . Then
the weights of V are of the form
"
λ− mα α with all mα ∈ N0 .
α∈∆
Moreover, ẇ0−1 f is of weight w0−1 (−λ) = −w0 (λ). The result now follows
from Corollary 15.10.
An : Dn :
E6 : 1 3 4 5 6
2
= L(λ)∗ ∼
For example, for type E6 and λ = λ1 + λ6 , L(λ) ∼ = L(−w0 (λ)).
Note that there are only finitely many p-restricted weights for a fixed
semisimple group G and a fixed prime p.
λ = µ0 + pµ1 + · · · + pa µa ,
L(λ) ∼
a
= L(µ0 ) ⊗ L(µ1 )(p) ⊗ · · · ⊗ L(µa )(p )
as kG-modules.
(1) Let G = SLn and let X be the natural n-dimensional module for G, of
highest weight λ1 (as mentioned in Example 15.21(2)). Then by Propo-
sition 15.12 the highest weight in V = X ⊗ X is 2λ1 , occurring with
multiplicity 1. In particular V possesses a kG-composition factor with
highest weight 2λ1 . If p > 2, this weight is p-restricted and it can be
shown that the corresponding highest weight module is precisely the
symmetric square S 2 (X) of X and V = S 2 (X) ⊕ ∧2 (X). If however
p = 2, the irreducible module with highest weight 2λ1 is simply the
“twist” of the natural module: L(2λ1 ) ∼= L(λ1 )(2) = X (2) by Proposi-
tion 16.6. Nevertheless, ∧ (X) is an irreducible kG-module and V has
2
(2) Consider the kSL3 -module with highest weight λ = 2λ1 +3λ2 , where λi is
the fundamental dominant weight corresponding to the simple root αi . If
p > 3, then λ is p-restricted and dim L(λ) = 39 if p = 5 and dim L(λ) =
42 if p = 5. (See [54].) If however p = 3, the p-adic expansion of λ is
λ = µ0 + 3µ1 where µ0 = 2λ1 and µ1 = λ2 , so L(λ) ∼ = L(2λ1 ) ⊗ L(λ2 )(3)
in this case by Theorem 16.12. Thus, by the previous example and by
Proposition 16.1 and Remark 16.2, L(λ) ∼ = S 2 (X)⊗X ∗ (3) , where X is the
natural three-dimensional module for SL3 . Hence, dim L(λ) = 6 · 3 = 18.
If p = 2, then Theorem 16.12 gives L(λ) ∼ = L(λ2 ) ⊗ L(λ1 + λ2 )(2) . Using
Example 15.19, one can show that L(λ1 + λ2 ) is isomorphic to the Lie
algebra of SL3 , and so dim L(λ) = 3 · 8 = 24.
17
Structure of parabolic subgroups, II
j ≥ 1, set
Qj = Uβ | β ∈ Φ+ of level ≥ j ≤ Q.
Example 17.2 Let G = SLn , the simply connected simple group of type
An−1 , with root system base {α1 , . . . , αn−1 } and corresponding set of posi-
tive roots Φ+ (see Example 9.8).
Let P be the standard parabolic subgroup corresponding to the subset
I = {sα1 } of S with unipotent radical
. "
n−1 "
n−1 /
Q = Uβ β = ai αi ∈ Φ+ , ai = 0
i=1 i=2
x ∈ Uα , xUβ x−1 lies in the product of root subgroups for roots of level at
least j. Since Qj is a product of root subgroups for T , Qj is also normalized
by T and (a) follows.
For (b), let’s choose a total order on the positive roots so that ht(α) >
ht(β) implies that α > β. For each α ∈ Φ, fix an isomorphism uα : Ga →
Uα . Then by Theorem 11.1(a) any element u of Qj may be written as a
product β uβ (cβ )y over β ∈ Φ+ of level j, with y ∈ Qj+1 and the cβ
uniquely determined by u. Thus, as Qj /Qj+1 is abelian by the commutator
formula, the map (cβ )β → β uβ (cβ )Qj+1 defines an isomorphism β Ga ∼ =
Qj /Qj+1 , with β as before.
As Qj P for all j, L acts on the quotient Qj /Qj+1 . It is immediate to
check that the rule cuβ (t)Qj+1 := uβ (ct)Qj+1 , for β ∈ Φ+ \ ΦI of level j and
c ∈ k, defines a k-vector space structure on Qj /Qj+1 which commutes with
the action of L.
142 Structure of parabolic subgroups, II
Now for a shape S of level j let VS := ( β of shape S Uβ )Qj+1 /Qj+1 . From
the commutator relations it is then easy to verify:
About the proof Essentially this follows easily from an inductive argument
and the commutator relations, Theorem 11.8. The exceptions occur when
certain structure constants cmn
αβ appearing in the commutator formula are
divisible by char(k).
In fact one has very good control over the kL-modules which occur in the
quotients Qj /Qj+1 .
(a) There exists a unique root βS of shape S having maximal height with
respect to this property.
(b) For each shape S, VS is an irreducible kL-module of highest weight βS .
(c) If S = S , then VS ∼
= VS as kL-modules.
(d) Qj /Qj+1 is the direct sum of the VS , for the shapes S of level j, hence
a completely reducible kL-module.
Sketch of proof The assertion of (a) is just Exercise A.7. As all roots of
shape S and of the same length are conjugate under the Weyl group WI of
L, by Lemma A.31, this already shows that VS is irreducible when there is
only one root length. By avoiding the small characteristic configurations ex-
cluded in the theorem, it is straightforward to show this in general. Part (b)
then follows from the classification of highest weight modules given in Theo-
rem 15.17(a) since clearly any non-identity element in the root group UβS is
17.1 Internal modules 143
a maximal vector of weight βS for the Levi factor L. Part (c) is now imme-
diate by Theorem 15.17(a). We have already shown in Proposition 17.4(b)
that Qj /Qj+1 decomposes as a direct sum of the VS , whence (d).
Remark 17.7 Under the assumptions of Theorem 17.6, Azad, Barry and
Seitz also show that the Levi complement L has finitely many orbits on each
VS . In the proof, they apply Richardson’s result [59, Thm. E] to see that
there exist a finite number of L-orbits on Q = Q1 and so on Q1 /Q2 , and
therefore on the VS occurring in Q1 /Q2 as well. For the general case, one
realizes Qj /Qj+1 as the commutator quotient of the unipotent radical of a
parabolic subgroup of a certain reductive subgroup of G. See [3, Thm. 2(f)]
for the details.
(1) Let G = SLn , the simply connected group of type An−1 , with the
parabolic subgroup P as in Example 17.2. The Levi complement has
the form L = U±α1 , T = [L, L]Z(L)◦ , where Z(L)◦ is an (n − 2)-
dimensional torus and [L, L] is a rank 1 semisimple group, so isomorphic
to SL2 or PGL2 by Theorem 8.8 (in fact isomorphic to SL2 by Proposi-
tion 12.14). Let’s consider some of the levels worked out in Example 17.2.
At level 1, for shape S1 = (1, 0, . . . , 0) we have VS1 = Uα2 · Uα1 +α2 .
This is easily seen to be the natural two-dimensional module for [L, L].
For each i > 1, VSi = Uαi+1 is the trivial module for the shape Si =
(0, . . . , 0, 1, 0, . . . , 0).
At level i for i ≥ 2, let S = (1 , . . . , n−2 ), j = 0 or 1 for all j and
n−2
j=1 j = i be a shape of level i. Then if 1 = 1, the corresponding
simple module VS is the natural two-dimensional module for [L, L] and
if 1 = 0, then VS is the trivial module.
(2) For a more elaborate example, let G = SO2n (see Definition 1.15) and B
the Borel subgroup of G consisting of upper triangular matrices exhibited
in Example 6.7(4). Recall the basis {v1 , . . . , v2n } of the orthogonal space
on which G naturally acts, such that the quadratic form with respect to
this basis is given by f (x1 , . . . , x2n ) = x1 x2n + x2 x2n−1 + · · · + xn xn+1 .
Then set V1 = v1 , . . . , vn , a totally singular subspace of V , and let
P = StabG (V1 ), a parabolic subgroup (see Proposition 12.13). Then one
144 Structure of parabolic subgroups, II
3
1 2
4
= QLH /Q ≤ P/Q ∼
LH ∼ = L.
Now L acts on each of the quotients Qj /Qj+1 in Q (defined in Section 17.1),
with LH stabilizing the image of QH in this quotient. Using Theorem 17.6,
we obtain a composition series for QH as kLH -module and this must be
compatible with the restrictions of the kL-modules Qj /Qj+1 to LH . This
can be used to provide precise information about the embeddings LH → L
and QH ≤ Q.
17.2 The theorem of Borel and Tits 147
The above corollary also has the following important consequence for the
study of maximal subgroups of a reductive group.
ψ : V1 ⊗ V2 → V, v ⊗ ϕ → ϕ(v),
defines an isomorphism, since indeed both sides have the same dimension,
and surjectivity is easily verified. Endowing V2 with the trivial H-action, ψ
even becomes an isomorphism of KH-modules, as
for all h ∈ H, v ∈ V1 , ϕ ∈ V2 .
Now clearly 1 ⊗ GL(V2 ) ≤ CGL(V ) (H). Conversely, if g ∈ CGL(V ) (H)
then g acts on the embeddings of V1 into V , so on V2 via g : V2 → V2 ,
(gϕ)(v) = gϕ(v). This action is faithful, since an element g in the kernel
satisfies gϕ(v) = ϕ(v) for all ϕ ∈ V2 , v ∈ V1 . Thus CGL(V ) (H) embeds into
1 ⊗ GL(V2 ). The last statement follows as before by replacing H acting on
V1 by GL(V2 ) acting on V2 .
(b) The form f is symmetric if and only if f1 , f2 are either both symmetric
or both skew-symmetric.
(c) If char(k) = 2 then there is a unique quadratic form Q on V , with
associated bilinear form f , such that Q(v1 ⊗ v2 ) = 0 for all vi ∈ Vi , and
Q is preserved by Sp(V1 ) ⊗ Sp(V2 ).
The first weak reduction result for closed connected subgroups of classical
groups is now the following:
Step 3: H is finite.
Assume that H is infinite, so H ◦ = 1. By Step 2 we have that V |H ◦ is
irreducible. As in the proof of Proposition 18.4, we see that H ◦ is reductive,
indeed semisimple as Z(H ◦ ) must lie in Z(G). Thus H ◦ = H1 · · · Hr , a
commuting product of simple algebraic groups. If r > 1, then V |H ◦ is a tensor
product of irreducible Hi -modules by Corollary 18.2 and H ≤ M ∈ C4 . So
we must have that r = 1 and H ◦ = H1 acts irreducibly on V , and indeed
tensor indecomposably (else H ≤ M ∈ C4 ). So finally we have H as in (2) of
the theorem, contradicting our assumption on H.
Step 4: H is finite with no components and H ≤ Z(G).
By Step 3, H is finite. As H is not contained in a subgroup lying in C(G)
and does not satisfy the conclusion (2) of the result, the above arguments
show that any non-trivial normal subgroup of H acts irreducibly on V , and
so H ≤ Z(G). Let E := E1 ◦ · · · ◦ Et be the product of the components of H,
a characteristic subgroup of H. Suppose E = 1. Arguing as in Step 3 above,
replacing H ◦ by E, we see that either H lies in a member of C4 , or t = 1
and E = E1 acts irreducibly on V . In the latter case CG (E) ≤ Z(G), so
HZ(G)/Z(G) ≤ Aut(EZ(G)/Z(G)) and so HZ(G)/Z(G) is almost simple
and H ∞ ≥ E acts irreducibly on V as in conclusion (2). Thus we have E = 1.
Step 5: Final contradiction.
So finally we have that H is finite and has no components. Set H̄ :=
HZ(G)/Z(G) = 1. Then all minimal normal subgroups of H̄ are elemen-
tary abelian, that is, direct products of cyclic groups of equal prime order.
Fix one such prime r, and let Q be a Sylow r-subgroup of the preimage in
H of the product of all minimal normal r-subgroups of H̄, a characteristic
subgroup of H. Let R denote its subgroup generated by all elements of order
r, respectively of order 4 if r = 2. Then R is a normal r-subgroup of H.
Then as above, V |R is irreducible. In particular, r = p = char(k) by
Proposition 2.9. In this case the irreducibility of V |R forces every character-
istic abelian subgroup of R to be cyclic. Then R has precisely the structure
of the r-groups described in C5 , by [2, 23.9]. Hence H ≤ M ∈ C5 , a final
contradiction.
The condition (i)* implies that the character group of a maximal torus of
Y coincides with the lattice of abstract weights Ω, and thus in particular, by
Theorem 15.17(b), there exist irreducible highest weight modules V for any
dominant weight from Ω. Finally, the condition (v) corresponds to the fact
that we are interested in the maximality of NG (X) in the smallest classical
group containing it.
The main results of Dynkin [23, 24] (which cover the case when char(k) =
0) and Seitz [60] and Testerman [80] (which cover the case when char(k) =
p > 0) are summarized in the following theorem. For the purposes of this
theorem, we will take p = ∞ when char(k) = 0. Before stating the result, let
us introduce some additional notation. Let {ωi | 1 ≤ i ≤ rk(X)}, respectively
{λj | 1 ≤ j ≤ rk(Y )}, be a set of fundamental dominant weights with respect
to a fixed choice of maximal torus and Borel subgroup of X, respectively Y ,
where we label Dynkin diagrams as in Table 9.1.
(1) the triple (ρ(X), ρ(Y ), V ) (or (ρ(X), ρ(Y ), V ∗ )) belongs to one of the
infinite series given in Table 18.1; or
(2) the triple is among a specified finite list of additional triples, each cor-
responding to a fixed embedding of simple algebraic groups and a kY -
module V of fixed p-restricted highest weight.
X<Y V |X V |Y conditions
Cn < A2n−1 , n ≥ 2 aω1 aλ1 a≥2
Cn < A2n−1 , n ≥ 2 aωj + bωj+1 , aλj + bλj+1 a+b=p−1>1
j<n a = 0 if j = n − 1
Bn < A2n , n ≥ 3 ωj , 2 ≤ j < n λj p=2
Bn < A2n , n ≥ 2 2ωn λn p=2
Dn < A2n−1 , n ≥ 4 ωj , 2 ≤ j < n − 1 λj p=2
Dn < A2n−1 , n ≥ 4 ωn−1 + ωn λn−1 p=2
An < A(n2 +n−2)/2 , n ≥ 3 ω1 + ω3 λ2 p=2
An < A(n2 +3n)/2 , n ≥ 2 2ω1 + ω2 λ2 p=2
Bn < Dn+1 , n ≥ 3 aωn aλn , aλn+1 a>0
Bn < Dn+1 , n ≥ 3 aωi + bωn aλi + bλj a + b + n ≡ i (p)
1≤i<n j = n, n + 1 a=0=b
Bn−j Bj < Dn+1 , n ≥ 2 ωn−j + ωn λn , λn+1
X → Bn−j Bj < Dn+1 , ωn−j + ωn λn , λn+1 p ≥ 5 if
n≥2 some πi (X) = A1
Bn A1 < Dn+3 , n ≥ 2 ω
nj+ 3ωn+1 λn+2 , λn+3 p≥5
X → B n1 · · · B nj < Dm , i=1 ωn1 +...+ni λm−1 , λm p=2
m = n1 + . . . + nj + 1 n−1 n−1 n−1
Dn < C n ji=1 ai ωi + an−1 ωn i=1 ai λ i p = 2, i=2 ai = 0
X → B n1 · · · B nj < B m , i=1 ωn1 +...+ni λm p=2
m = n1 + . . . + nj
groups F4 and E6 . The existence was later established in [81] and hence
there are no remaining question marks in the above result.
(f) The above result in case Y is of type Am and char(k) > 0 was estab-
lished independently by Suprunenko in [77].
(g) Note that there are essentially three types of examples:
(I) Positive characteristic analogues of characteristic 0 configurations. For
example, in characteristic 0, the naturally embedded X = Sp2n in
Y = SL2n = SL(W ) acts irreducibly on the symmetric powers S a (W )
for all a; in characteristic p > 0, X acts irreducibly on S a (W ) for all
1 < a < p. This is the first family of examples in the table.
(II) For a fixed embedding X < Y , families which exist only in positive
characteristic, for each positive characteristic p. For example, taking
X and Y as in (I), the second family of examples in the table has no
characteristic 0 analogue.
(III) Examples which occur for a fixed prime characteristic. For example,
the adjoint representation of the group SL3 induces a seven-dimen-
sional irreducible representation when p = 3. (See Theorem 15.20(1).)
Moreover, SL3 preserves a non-degenerate orthogonal form on the
corresponding module (see [11, Table 2]). Hence, in characteristic 3,
we have an embedding X := PGL3 < Y := SO7 = SO(W ). It
turns out that the symmetric square S 2 (W ) is irreducible for Y when
char(k) = 0, but has a 21-dimensional quotient when char(k) = 3.
One can show that the restriction of this quotient module to the sub-
group X is an irreducible kX-module, hence giving an example. This
particular example, which is unique with respect to one further con-
dition (see Proposition 18.9 below) is among the isolated examples
mentioned in the statement of the theorem.
The relatively short length of this list of triples justifies the following
statement: If ρ : H → SL(V ) is a rational irreducible, tensor indecomposable
representation of a simple algebraic group H, then, most of the time, ρ(H)
is maximal among closed connected subgroups of G = Cl(V ), the smallest
classical group containing ρ(H). In particular, the following result is deduced
directly from the classification given by Theorem 18.7 (see [60, Cor.4]).
Corollary 18.9 Let H, ρ and G be as above. If Y is a closed connected sub-
group of G with ρ(H) < Y < G, then with just two exceptions, Y is maximal
among closed connected subgroups of G. The exceptions are as follows:
(1) when p = 3, there is an embedding of irreducible subgroups A2 < G2 <
B3 < SO27 , and
160 Maximal subgroups of classical type simple algebraic groups
Remark 18.10 Note that the exception of (2) above is not listed in [60,
Cor. 4]. Moreover, the precise statement concerning the type of the classical
group Cl(V ) is not at this time a published result. The three groups do fix
a symplectic form on V ; see Remark 16.2.
We conclude this section with a discussion of how the results of this chapter
are applied to determine the triples (X, Y, V ). As a first step we classify the
triples where X is the group SL2 or PGL2 . We will apply the following
lemma.
Lemma 18.11 Let H = SL2 , with maximal torus T , and let V be a ra-
tional irreducible kH-module. Then the T -weight spaces of V are all one-
dimensional.
"
t "
t
qi (ai λ − mi α) = µ − mi qi α,
i=1 i=1
for some 0 ≤ mi ≤ ai . Two such weights are equal only if the coefficients of
q1 , . . . , qt are the same. Thus, there is a one-dimensional space of vectors of
weight µ − (m1 q1 + · · · + mt qt )α, for each choice of mi with 0 ≤ mi ≤ ai ,
which establishes the result.
18.2 Maximal subgroups of the classical algebraic groups 161
For the remainder of the section we assume the triple (X, Y, V ) satisfies
the conditions (i)*, (ii), (iii)*, (iv) and (v) above. Fix a maximal torus TX
of X and a maximal torus TY of Y with TX ≤ TY . Thus, when X = SL2 or
PGL2 , Lemma 18.11 shows that V must have one-dimensional weight spaces
with respect to TY as well. This leads to a short list of possibilities for λ,
the highest weight of V ; for example, if Y has type Bn , λ = λ1 or λn , or
if Y has type E6 , then λ = λi for i = 1 or 6. At this point, case-by-case
considerations lead to two isolated examples, one in G2 and another in D5 ,
as well as the configurations in Table 18.1 where X is diagonally embedded
in a central product of a certain number of copies of A1 , when p = 2. See
[60, Thm. 7.1] for the proof.
Now one is in a position to proceed by induction on rk(X). In order to
do so, we will want to consider the action of a Levi subgroup of X on the
irreducible module V . Choose a parabolic subgroup PX of X; by Corol-
lary 17.12, there exists a parabolic subgroup PY of Y , with PX ≤ PY and
QX := Ru (PX ) ≤ QY := Ru (PY ). Let LX , respectively LY , be Levi sub-
groups of PX , respectively PY . The following result plays a crucial role in
applying the induction hypothesis:
Proposition 18.12 With notation as above, we have that the fixed point
spaces V QX and V QY are equal.
Remarks 18.14 (a) The above considerations, based upon the reduction
theorem 18.6, do not treat the odd-dimensional orthogonal groups defined
over fields of characteristic 2. Indeed, while these are isomorphic as abstract
groups to the symplectic groups acting on a space of dimension one less,
Theorem 9.13 and the identification of the different simple algebraic groups
with certain classical groups (Table 9.2) shows that the groups SO2n+1 and
Sp2n are not isomorphic as algebraic groups. (See Exercise 20.25 for further
details.) Nevertheless the natural map (homomorphism of abstract groups)
ϕ : SO2n+1 → Sp2n is indeed a bijective morphism of algebraic groups. Then
the fact that ϕ is continuous, together with Proposition 1.5, shows that it
suffices to determine the maximal closed positive-dimensional subgroups of
Sp2n in order to determine those of SO2n+1 .
(b) The classification of the maximal positive-dimensional closed sub-
groups of the simple algebraic groups of types An , Bn , Cn and Dn fol-
lows from the above considerations. Let Gsc be a simply connected simple
algebraic group with the given root system. Then there exists an isogeny
18.2 Maximal subgroups of the classical algebraic groups 165
π : Gsc → Cl(V ), for one of the classical groups Cl(V ). The maximal closed
positive-dimensional subgroups of Gsc are then inverse images of those of
Cl(V ). Moreover, the maximal closed positive-dimensional subgroups of any
homomorphic image ϕ(Gsc ) of Gsc are images of the corresponding sub-
groups of Gsc , since by Proposition 1.5 images of closed subgroups under
morphisms are closed.
19
Maximal subgroups of exceptional type
algebraic groups
We now consider the case of the exceptional type algebraic groups. The
classification of the maximal closed connected subgroups was obtained by
Dynkin [23] in the case where char(k) = 0. The case of positive characteristic
is covered by three lengthy articles of Seitz [61] and Liebeck–Seitz [48, 53].
G X
G2 A1 Ã1 , A2 , Ã2 (p = 3)
F4 B4 , A1 C3 (p = 2), C4 (p = 2), A2 Ã2
E6 A1 A5 , A2 A2 A2
E7 A1 D6 , A7 , A2 A5
E8 D8 , A1 E7 , A8 , A2 E6 , A4 A4
G X simple
G2 A1 (p ≥ 7)
F4 A1 (p ≥ 13), G2 (p = 7)
E6 A2 (p = 2, 3), G2 (p = 7), C4 (p = 2), F4
E7 A1 (2 classes, p ≥ 17, 19 respectively), A2 (p ≥ 5)
E8 A1 (3 classes, p ≥ 23, 29, 31 respectively), B2 (p ≥ 5)
G X semisimple, non-simple
F4 A1 G2 (p = 2)
E6 A2 G2
E7 A1 A1 (p = 2, 3), A1 G2 (p = 2), A1 F4 , G2 C3
E8 A1 A2 (p = 2, 3), G2 F4
G X M/X
G2 A1 Ã1 , A2 , Ã2 (p = 3) 1, Z2 , Z2
F4 (p = 2) B4 , D4 , A1 C3 , A2 Ã2 1, S3 , 1, Z2
F4 (p = 2) B4 , C4 , D4 , D̃4 , A2 Ã2 1, 1, S3 , S3 , Z2
E6 A1 A5 , (A2 )3 , D4 T2 , T6 1, S3 , S3 , W (E6 )
E7 A1 D6 , A7 , A2 A5 , (A1 )3 D4 , 1, Z2 , Z2 , S3 ,
7
(A1 ) , E6 T1 , T7 PSL3 (2), Z2 , W (E7 )
E8 D8 , A1 E7 , A8 , A2 E6 , 1, 1, Z2 , Z2 ,
2
(A4 ) , (D4 ) ,2
Z 4 , S3 × Z 2 ,
(A2 )4 , (A1 )8 , T8 GL2 (3), AGL3 (2), W (E8 )
For the proof of (b), one must use the fact that the centralizer in G of a
semisimple, or nilpotent, element in Lie(G) contains a maximal torus of G,
or has a non-trivial unipotent radical. See [61, 1.3] for details.
With these lemmas in place, and after fixing some additional notation,
we introduce the main tool used in the study of the embedding X < G.
Let T be a maximal torus of G, TX a maximal torus of X with TX ≤ T
and ΦX the corresponding root system of X. Let BX be a Borel subgroup
of X containing TX and let ∆X be the corresponding base of ΦX . For α ∈
ΦX , let α∨ : Gm → TX be the coroot as in Lemma 8.19. For c ∈ k × ,
set γ(c) = α∨ (c), the product taken over all roots α ∈ Φ+ X . Then im(γ)
is a one-dimensional torus in X and we shall first consider the action of
this torus on Lie(X). Let Lie(X) = Lie(TX ) ⊕ α∈ΦX Lie(X)α and choose
vα ∈ Lie(X)α \ {0}.
Proof For (a), we simply note that im(γ) ≤ TX and TX acts trivially on
itself via conjugation, hence TX fixes pointwise its Lie algebra. For (b), let
β ∈ ∆X , c ∈ k × . Then we note that γ(c)vβ = ( α∨ (c))vβ = cr vβ where
∨
∨
α∈Φ+ β, α . Setting ρ = 2 β, 2ρ. By
1
r = α∈Φ+ α , we have r =
X X
Exercise A.1 {α∨ | α ∈ Φ+ X } forms a positive system in the root system
∨
ΦX . Now by Exercise 20.26 we know that ρ is the sum of the fundamental
dominant weights relative to the abstract root system Φ∨ X and ΦX = ΦX .
∨∨
Finally we indicate one of the key tools in the analysis which follows:
About the proof For β ∈ Φ, the root system of G with respect to T , let
rβ ∈ Z such that γ(c)v = crβ v for all v ∈ Lie(G)β ; that is, rβ = β, γ. One
first shows that all TX -weights on Lie(G) are integral linear combinations
of elements of ΦX . See [61, 2.3] for this proof. Then apply Lemma 19.7 to
see that the rβ must be even integers. Choosing an appropriate ordering on
X(T ) one finds a base ∆ of Φ such that rβ ≥ 0 for all β ∈ ∆ and so for all
19.2 Indications on the proof 171
(a) Show how to reduce the proof of Theorem 11.12 on the existence of
graph automorphisms to the case of simple groups of simply connected
type.
(b) Verify the details of the proof for type SLn , n ≥ 3.
(c) Show that a suitable element of GO2n induces a non-trivial graph auto-
morphism of SO2n , n ≥ 2.
[Hint: For (a), reason by induction on (w) and apply Theorem 11.17; for (b)
apply (a). The result of (c) is a corollary of (a) and (b).]
(a) Determine the proper closed subsystems of Φ of rank equal to the rank
of Φ.
(b) Let G = Sp2n , a simple algebraic group of type Cn , acting on the nat-
ural 2n-dimensional module V preserving a symplectic bilinear form.
For n1 , . . . , nt ∈ N with ni = n, there exist subspaces Vi ≤ V with
dim Vi = 2ni and V = V1 ⊥ · · · ⊥ Vt . Hence the group Sp2n1 ×· · ·×Sp2nt
is naturally embedded in G; let H denote the corresponding closed sub-
group of G.
(i) Show that H is a maximal rank subgroup of G, for all characteris-
tics char(k).
(ii) Find a closed subsystem Ψ of the root system of G such that H is
the subsystem subgroup associated to this subsystem.
[Hint: (a) You should find no example for An , n − 1 for type Bn , (n − 1)/2 for
type Cn and (n − 2)/2 for type Dn .]
[Hint: For (a), use the uniqueness of expression in the Bruhat decomposition.]
Exercises for Part II 175
[Hint: (c) Use Proposition 14.6. (d) By Proposition B.12 every Wa -orbit on Y ⊗
Qp ⊂ Y ⊗ R contains a point x from the fundamental alcove A. Then for s = τ (x),
for α ∈ Φ we have α(s) = 1 if and only if sα , x = 0, that is, if and only if sα∨
fixes x, that is, x lies on the wall of A corresponding to α∨ .]
Exercise 20.15 Show that the torsion primes of a reductive group G are
precisely the torsion primes of its root system Φ together with the prime
divisors of the order of its fundamental group.
[Hint: First show that Λ(G) ∼ = Y /ZΦ∨ by applying Hom(−, Z) to the exact se-
quence defining Λ(G). Now, if G ≤ G is a subsystem group with root system Ψ,
the torsion of Y /ZΨ∨ is the torsion of Y /ZΦ∨ together with that of ZΦ∨ /ZΨ∨ .]
(b) If s is of finite order, the exponent of CG (s)/CG (s)◦ divides the order
of s.
(c) If s is of finite order prime to | ker(π)| then CG (s) is connected.
[Hint: Fix a preimage ŝ ∈ Gsc of s. For g ∈ CG (s) with preimage ĝ, ĝŝĝ −1 = zŝ
for some z ∈ ker(π). Then g → z = [ĝ, ŝ] is a well-defined group homomorphism
with the required properties. For (b), use that [ĝ, ŝ] is central, so [ĝ, ŝ]n = [ĝ, ŝn ]
for all n.]
∆ the corresponding set of simple roots in the root system Φ. Let λ ∈ X(T )
be a dominant weight.
[Hint: For (a), let v + ∈ L(λ) be a maximal vector with respect to the Borel
subgroup B. Then B ≤ StabG (v + ), so StabG (v + ) is a parabolic subgroup of
G. Now use the structure of parabolic subgroups (or see Corollary A.29). For (b),
use Lemma 15.3 and Proposition 15.8.]
Exercise 20.20 Let char(k) = 2 and let G = SL2 . Fix a maximal torus T
of G, a Borel subgroup B containing T and a simple root corresponding to
the choice of B. Let λ ∈ X(T ) be the unique fundamental dominant weight.
Let F : k → k be the 2-power map.
(a) Show that the adjoint representation of G equips Lie(G) with the struc-
ture of an indecomposable kG-module, having a composition series 0 ⊂
V1 ⊂ Lie(G) = V , with V1 ∼
= L(0) and V /V1 ∼= L(λ)(2) .
(b) Show that V2 , as in Exercise 20.18, is an indecomposable kG-module
with a composition series 0 ⊂ W1 ⊂ V2 , such that W1 ∼ = L(λ)(2) and
∼
V2 /W1 = L(0).
In this part we introduce and study finite analogues of the simple linear alge-
braic groups which were the subject of the first two parts. The construction
of the various classical groups over algebraically closed fields in Section 1.2
generalizes in a straightforward way to give versions over arbitrary base
fields. It is much less obvious how to obtain versions of the simple excep-
tional groups. This was first achieved by Chevalley [15] who showed how to
construct analogues of all simple algebraic groups as automorphism groups
of simple Lie algebras over any base field. Still this approach falls short of
producing all versions in which we will be interested. For example, over an
algebraically closed field there is just one class of non-degenerate orthogo-
nal forms up to similarity in any dimension, while over arbitrary fields there
may be many, with corresponding non-isomorphic isometry groups. Also, the
isometry groups of unitary forms do not arise in Chevalley’s setup.
Shortly after Chevalley’s construction, Steinberg [68] presented a varia-
tion of this by considering fixed points of field automorphisms composed
with algebraic group automorphisms. This allows one to recover the uni-
tary groups, for example. But this wasn’t yet the end of the story. In 1960
M. Suzuki discovered an infinite series of finite simple permutation groups
which at that time did not seem to have any relation with algebraic groups.
It was recognized in the same year by Steinberg how his construction could
be generalized to yield these as subgroups of the four-dimensional symplectic
group over an algebraically closed field of characteristic 2.
We will follow this approach of Steinberg, which now seems to be the best
way to recover via a general construction all of the versions described above:
given a field K, the various analogues of linear algebraic groups are obtained
from algebraic groups over the algebraic closure k of K by means of a Galois
180
via the coefficients, then leaves I invariant. Thus Fq also acts on the set V
of common zeros of I in k n , by
Fq : V → V, (v1 , . . . , vn ) −→ (v1q , . . . , vnq ).
This induced map Fq is called the Frobenius morphism of V with respect to
the Fq -structure given by I. As in the example above, we write
V Fq := V (Fq ) := {v ∈ V | Fq (v) = v}
for the Fq -fixed points of V . Note that, as Fq is induced by an element of
the Galois group of k/Fq , it is a bijective map.
Thus, if a closed subgroup G ≤ GLn is defined by equations over Fq , this
gives rise to a Frobenius morphism Fq : G → G, (aij ) → (aqij ), with respect
to this Fq -structure, which clearly is a morphism of algebraic groups, with
finite fixed point group G(Fq ) = GFq ≤ GLF n = GLn (q). Note, however,
q
Here, the fixed point group GUn (q) := GF is the so-called general unitary
n = GUn (q) ∩ SLn (q ), the special
group over Fq2 . We also set SUn (q) := SLF 2
unitary group. (Note that, despite the suggestive notation, GUn (q) cannot
be obtained as a subgroup of GLn (q). Some justification for this notation will
be given in Example 22.11.) The above definition shows that it is the group
of invertible n × n-matrices over Fq2 leaving invariant the non-degenerate
sesquilinear form
."n "
n / "n
, : Fnq2 × Fnq2 → Fq2 , x i ei , y i ei = xi yiq ,
i=1 i=1 i=1
Examples 21.1 and 21.2 have already shown two instances of Steinberg
endomorphisms of GLn . Both of them fall into the second case of the funda-
mental dichotomy for endomorphisms of simple algebraic groups proved by
Steinberg ([72, Thm. 10.13]):
Proof Clearly the natural map π : G → G/H induces a surjection from the
set of F -conjugacy classes of G to those of G/H. Now let g1 , g2 ∈ G have
F -conjugate images, so g2 H = F (x)g1 x−1 H for some x ∈ G. Then g2 =
F (x)g1 hx−1 for some h ∈ H. Now the Lang–Steinberg Theorem, applied to
the endomorphism F = g1−1 F : H → H, g → g1−1 F (g)g1 , of the connected
group H yields that h = F (y)y −1 = g1−1 F (y)g1 y −1 for some y ∈ H, whence
g2 = F (x)F (y)g1 y −1 x−1 = F (xy)g1 (xy)−1 are F -conjugate, as claimed.
186 Steinberg endomorphisms
Proof By Theorems 6.4 and 4.4 the group G acts transitively by conjugation
on
Note that here GF ∼= GUn (q) is again isomorphic to the general unitary
group by Corollary 21.8, since we have changed the Frobenius map from
Example 21.2 only by the inner element w.
22
Classification of finite groups of Lie type
We study the Steinberg endomorphisms more closely via their action on the
character group and on the Weyl group and present their classification which
in turn gives a classification of the finite groups of Lie type. As for semisimple
algebraic groups, this classification can be given in combinatorial terms via
so-called complete root data. We then comment on the construction of the
various finite groups of Lie type.
Proof Let’s first consider the case where T = Dn ≤ GLn and F = Fq is the
standard Frobenius map on GLn . Then
F (χ)(t) = χ(F (t)) = χ(tq ) = χ(t)q = (qχ)(t) for all χ ∈ X, t ∈ T,
so F |X = q idX as claimed. Moreover, T F ∼ = (F× n
q ) (see Example 21.14(1)).
In the general situation, as F is a Steinberg endomorphism, there exists
an embedding T → GLn such that some power F of F is the restriction
to T of a standard Frobenius on GLn . Let T1 be a maximal torus of GLn
containing the torus T . This is conjugate to the maximal torus Dn , so T1 =
Dgn for some g ∈ GLn . Now g has entries in some finite subfield of k, so
there is a positive power F δ of F which fixes g. Then conjugation by g
defines an isomorphism T1 ∼ = Dn which commutes with F δ . By the first
part, F δ acts by multiplication with some power r of p on X(Dn ), so on
X(T1 ). Restriction of characters from T1 to T is F -equivariant, with F -
stable kernel T ⊥ (see Proposition 3.8(a)), so F δ also acts by the scalar r on
X = X(T ) ∼ = X(T1 )/T ⊥ , proving (a).
For (b), again apply Proposition 3.8(a) to the closed subgroup T F ≤ T to
obtain that X(T F ) ∼= X/(T F )⊥ . Now we claim that
t ∈ TF ⇐⇒ F (t) = t ⇐⇒ (F − 1)χ(t) = 1 for all χ ∈ X.
The implication from left to right is clear. Conversely, if F (t) = t then clearly
there is χ ∈ X with χ(F (t)t−1 ) = 1. So we have T F = ((F − 1)X)⊥ . As
(F −1)X contains (F δ −1)X = (r −1)X, it is of index prime to p in X, so we
have (T F )⊥ = (F − 1)X by Proposition 3.8(b). Finally, as |T F | is prime to p
and k contains all p -roots of unity, X(T F ) = Hom(T F , k × ) ∼
= Hom(T F , C× ),
F
and the latter is isomorphic to T . Thus indeed
= X(T F ) ∼
TF ∼ = X/(T F )⊥ ∼
= X/(F − 1)X.
Write Φ ⊂ X for the root system of G, with positive system Φ+ with
respect to T and B. For α ∈ Φ let’s choose isomorphisms uα : Ga → Uα
onto the root subgroups (see Theorem 8.17(c)). We set XR := X ⊗Z R.
Proposition 22.2 Let G be a connected reductive algebraic group with
Steinberg endomorphism F : G → G, and T , B, X, Φ as above.
(a) There exists a permutation ρ of Φ+ and, for each α ∈ Φ+ , a positive
integral power qα > 1 of p = char(k) and aα ∈ k × such that F (ρ(α)) =
qα α and F (uα (c)) = uρ(α) (aα cqα ) for all c ∈ k.
(b) There exists δ ≥ 1 such that F δ |X = q δ idX and F = qφ on XR for
some positive fractional power q of p and some φ ∈ Aut(XR ) of order δ
inducing ρ−1 on Φ+ .
190 Classification of finite groups of Lie type
Proof Part (a) is just Lemma 11.10 applied to the endomorphism F . For
part (b), by Lemma 22.1 there exists some minimal δ ∈ N such that F δ |X =
1
r idX . Thus all eigenvalues of F on X ⊗Z C have absolute value q := r δ > 1, a
fractional power of p. Then with φ := 1q F : XR → XR we get φδ = q −δ F δ =
idXR , so φ has finite order and permutes the roots in the same way as ρ−1
by (a).
By Lemma 11.10, ρ stabilizes the set of simple roots ∆ ⊆ Φ+ . Hence, in
its action on the Weyl group F stabilizes the set of simple reflections of W
and thus both F and ρ induce graph automorphisms of the Coxeter diagram
of Φ.
Thus ρ(β) and w1 ρ(α) are proportional roots with positive proportionality
factor qα /qβ , so they are equal by axiom (R2) whence qα = qβ is constant
on roots of the same length.
Since G is simple, W is irreducible, so there is a unique W -invariant scalar
product on XR up to scalar multiples, which must hence be fixed by φ, up
to a scalar. If qα is constant on Φ+ , then by Proposition 22.2 we necessarily
22.1 Steinberg endomorphisms 191
Thus, finally, by Examples 22.6 and 22.8 the existence question in Theo-
rem 22.5 is only about the very twisted automorphisms of groups of excep-
tional type, which we will not discuss here. Note that there necessarily q is
a non-integral power of 2 or 3. See [73, §11] for this case.
22.2 The finite groups GF 193
Φ δ GF
sc GF
ad other isogeny types
An−1 1 SLn (q) PGLn (q) Zd/e .PSLn (q).Ze
e|d := (n, q − 1)
Bn 1 Spin2n+1 (q) SO2n+1 (q) −
Cn 1 Sp2n (q) PCSp2n (q) −
Dn 1 Spin+2n (q) (PCO◦2n )+ (q) SO+ +
2n (q), HSpin2n (q)
G2 1 G2 (q) −
F4 1 F4 (q) −
E6 1 (E6 )sc (q) (E6 )ad (q) −
E7 1 (E7 )sc (q) (E7 )ad (q) −
E8 1 E8 (q) −
An−1 2 SUn (q) PGUn (q) Zd/e .PSUn (q).Ze
e|d := (n, q + 1)
Dn 2 Spin−
2n (q) (PCO◦2n )− (q) SO−2n (q)
D4 3 3
D4 (q) −
E6 2 (2 E6 )sc (q) (2 E6 )ad (q) −
B2 2 2
B2 (q ) = Suz(q 2 ), q 2 = 22f +1
2
−
G2 2 2
G2 (q 2 ), q 2 = 32f +1 −
F4 2 2
F4 (q 2 ), q 2 = 22f +1 −
(1) Recall the orthogonal group GO2n with respect to the quadratic form
See the books of Dieudonné [21] or Grove [30], for example, for more
information on the various types of finite classical groups.
The various possibilities for the finite groups of Lie type GF can again
be encoded in a purely combinatorial way. The root datum (X, Φ, Y, Φ∨ )
of a semisimple group G is obtained by fixing a maximal torus T of G.
If F : G → G is a Steinberg endomorphism, we may assume T to be F -
stable. Then its Weyl group W = NG (T )/T is also F -stable, and this defines
a semidirect product W F . Any element wF in the coset W.F of W in
W F still stabilizes T , hence the root datum. By Corollary 21.8 it also
defines a G-conjugate, hence isomorphic, finite group of fixed points. So, by
Theorem 22.5, (G, F ) is determined up to isomorphism by the root datum
of G, the coset W φ and q, where φ ∈ Aut(XR ) stabilizes Φ ⊂ X and also
Φ∨ ⊂ Y , hence normalizes W , in the following sense: if (G, F ), (G , F ) both
correspond to (X, Φ, Y, Φ∨ , W φ) and the same q, there is an isomorphism
σ : G → G such that F ◦ σ = σ ◦ F .
The complete root datum G, together with q, determines (G, F ) and hence
the finite group GF up to isomorphism by Corollary 21.8. We write GF =:
G(q). Thus, G stands for a whole family of finite groups of Lie type, for
example {SLn (q)}, {SUn (q)} or {E8 (q)}, where q runs over all prime powers.
It is not true, though, that groups GF for different complete root data
are necessarily non-isomorphic. For example we have SL3 (2) ∼ = PGL3 (2)
(non-isomorphic but isogenous algebraic groups may lead to isomorphic fi-
nite groups), Sp2n (2f ) ∼
= SO2n+1 (2f ) (different root systems may lead to
isomorphic finite groups), or Sp4 (3) ∼ = SU4 (2) (here even the underlying
characteristic differs). But this happens not very often and all such “excep-
tional” isomorphisms are known, see Remark 24.9.
Example 22.11 (Complete root data for GLn (q) and GUn (q))
(1) Let G ≤ GLn be a connected reductive subgroup stable under the stan-
dard Frobenius map Fq on GLn , and assume that Fq acts trivially on the
Weyl group W of G. Then φ is trivial and the corresponding complete
root datum is G = (Γ, W ), where Γ is the root datum of G.
(2) Let G = GLn , Fq : G → G the standard Frobenius map, and σ : G → G,
(aij ) → ((aij )−tr )w , with w as in Example 21.14(2). Let F := Fq σ, so
GF = GUn (q). Then T := Dn is an F -stable maximal torus, and σ
acts on X(T ) as −w, hence F acts as −wq on X(T ). Thus, in this case
196 Classification of finite groups of Lie type
G = (Γ, −W ) is the complete root datum for the family {GUn (q)} (with
Γ the root datum of GLn and W its Weyl group).
This is sometimes expressed by saying that GUn (q) is obtained from
GLn (q) by replacing q by −q, or “GUn (q) is Ennola-dual to GLn (q)”.
23
Weyl group, root system and root subgroups
under the standard Frobenius map. Here, the last arrow is in general not
surjective, the image PSLn (q) has index gcd(n, q − 1) in PGLn (q), see Sec-
tion 24.1.
This phenomenon cannot occur if the kernel is connected, see Exercise 30.7:
W F = (NG (T )/T )F ∼
= NGF (T )/T F .
Lemma 23.3 Let W be the Weyl group of a root system with set of simple
reflections S ⊂ W and φ an automorphism of W stabilizing S.
The proof shows that the element sI can also be described as being the
longest element of the parabolic subgroup WI . The various possibilities for
W φ when W is irreducible and φ is non-trivial are displayed in Table 23.1.
It turns out that W φ is again a Coxeter group, and even a Weyl group
except for W of type F4 with φ acting non-trivially; in the latter case, W φ
is isomorphic to the dihedral group I2 (8) of order 16, which is a real, but
non-rational reflection group.
Φ A2n−1 A2n Dn D4 E6 B2 G2 F4
o(φ) 2 2 2 3 2 2 2 2
Wφ Cn Bn Bn−1 G2 F4 A1 A1 I2 (8)
23.1 The root system 199
We next introduce a root system for a finite group of Lie type. Let G be
a semisimple algebraic group with a Steinberg endomorphism F : G → G.
As we saw in Section 22.1, fixing an F -stable torus inside an F -stable Borel
subgroup B of G with associated root system Φ induces an action of F on the
set of corresponding root subgroups Uα , α ∈ Φ+ , and hence a permutation
ρ of Φ+ . As in Proposition 22.2 we write F |XR = qφ with φ ∈ Aut(XR ) of
finite order δ. Let XRφ denote the fixed space of φ on XR , i.e., the eigenspace
for the eigenvalue 1. Then the homomorphism
1" i
δ−1
π : XR −→ XRφ , χ→ φ (χ),
δ i=0
clearly is the identity on XRφ , hence a projection. Let S ⊂ W denote the set
of simple reflections of W .
Definition 23.5 For each equivalence class ω for the equivalence relation
∼ defined in Lemma 23.4 let αω denote the vector of maximal length among
the {π(α) | α ∈ ω}. The set ΦF := {αω | ω ∈ Ω} is called the root system of
GF . We let ∆F := {αω | ω ⊆ ∆} ⊆ ΦF .
second simple root is fixed, while the other three are permuted cyclically by
triality. Let’s set
1
µ := π(α1 ) = (α1 + α3 + α4 ), ν := π(α2 ) = α2 .
3
Then π(α1 +α2 ) = π(α2 +α3 ) = π(α2 +α4 ) = µ+ν, π(α1 +α2 +α3 ) = 2µ+ν,
π(α1 + α2 + α3 + α4 ) = 3µ + ν and π(α1 + 2α2 + α3 + α4 ) = 3µ + 2ν. Thus
here ΦF is a root system of type G2 , with base {µ, ν} (see Example 9.5).
q2 q2 q2 q3 q2 q2 q2 q
2 2
A2n > A2n−1 <
n≥1 n≥2
q q q q2
2
Dn >
n≥4
q3 q q q q2 q2
3 2
D4 < E6 >
q4 q6 q4 q2
2 2 2
B2 G2 F4 <
202 Weyl group, root system and root subgroups
where the product runs over the orbits µ ∈ Ω such that αµ = mαω + nαν for
some m, n > 0, in any fixed order.
24
A BN-pair for GF
We now derive a Bruhat decomposition for the finite groups of Lie type by
investigating the fixed points of a Steinberg endomorphism on the Bruhat
decomposition for algebraic groups. This leads to a first order formula for
groups of Lie type. A second, factored form can be derived from this using
some standard results on the invariant theory of finite reflection groups. It
turns out that the orders in a family of groups of Lie type belonging to the
same complete root datum are given by a polynomial in q.
The Bruhat decomposition also gives rise to a BN-pair structure for finite
groups of Lie type. Using general properties of BN-pairs this then allows
us to conclude that most groups of Lie type derived from simple algebraic
groups of simply connected type are quasi-simple, that is, simple modulo
their center. Furthermore, we are able to determine their automorphism
groups.
Similarly, a formula for |GF | in terms of the qα from Proposition 22.2 can
be deduced for arbitrary connected reductive groups G, see Exercise 30.9.
The polynomial w∈W x
(w)
∈ Z[x] is called the Poincaré polynomial
of the Weyl group W . In order to obtain a factorization, we need another
interpretation of this polynomial in terms of the invariant theory of W in
its reflection representation on V := X ⊗Z R. It is a fundamental property
of finite reflection groups, i.e., finite linear groups generated by complex
reflections, that their invariants form a polynomial ring:
See [9, V, Thm. 5.3] for (a), [9, V, §5.1, Cor.] for (b), and [9, V, Cor. 5.3
and Prop. 5.3] for (c).
The d1 , . . . , dn are called the degrees of the reflection group W . Note that
for W the Weyl group of a root system Φ we have N = |Φ|/2 = |Φ+ | since
by axiom (R2), Proposition A.21 and Exercise A.6 there are precisely two
roots for each reflection in W .
Returning to our algebraic group G, the Steinberg endomorphism F acts
on V as a scalar times an automorphism φ of finite order δ, which normalizes
W (see Proposition 22.2). Thus φ also acts naturally on the ring of invari-
ants S(V )W , and the algebraically independent homogeneous generators fi
may in fact be chosen to be eigenvectors of φ. We denote the corresponding
eigenvalues by i . Thus, the i are roots of unity of order dividing δ. Clearly
all i = 1 if F acts as a scalar on V . Using that the Poincaré polynomial
is the Hilbert series of the coinvariant algebra of S(V ) one obtains (see [72,
Thm. 2.1 and Cor. 2.9]):
Proposition 24.5 Let W ≤ GL(V ) be a finite reflection group on a real
vector space V of dimension n, φ ∈ NGL(V ) (W ) and fi , 1 ≤ i ≤ n, homo-
geneous generators of S(V )W of degrees di which are eigenvectors of φ with
eigenvalues i . Then the i are also the eigenvalues of φ on V and we have
" !n
xdi − i
x(w) = .
i=1
x − i
w∈W φ
This leads to the following factorization of the order formula from Propo-
sition 24.3 (see [72, Thm. 11.16]):
Corollary 24.6 Let G be connected reductive, F : G → G a Steinberg
endomorphism. Then
!
rk(G)
|GF | = q |Φ |
+
(q di − i ),
i=1
!
rk(G)
|G| := x|Φ |
+
(xdi − i ) ∈ Z[x]
i=1
the order polynomial of the complete root datum G. The preceding corollary
can then be rephrased as saying that |G(q)| = |G|(q), i.e., the order of
GF is the order polynomial evaluated at q. Note that deg |G| = dim G by
Theorems 8.17(b) and 24.4(c).
A further immediate consequence of the above formula is the fact that
|G(q)| is independent of the isogeny type of G, e.g., |SLn (q)| = |PGLn (q)|,
and |E6 (q)sc | = |E6 (q)ad |. We also get a divisibility property of order poly-
nomials, see Exercise 30.10:
n !
n
|GLn (q)| = q ( 2 ) (q i − 1).
i=1
(2) For G = SLn with the standard Frobenius map, the Weyl group acts
via the deleted permutation module on XR , hence the invariants are
generated by the elementary symmetric polynomials of degree at least 2.
We obtain
n !
n
|SLn (q)| = q ( 2 ) (q i − 1) = |PGLn (q)|.
i=2
(3) For the special orthogonal group G = SO8 , with root system of type
D4 the degrees of W are (d1 , d2 , d3 , d4 ) = (2, 4, 6, 4) and N = d1 + d2 +
208 A BN-pair for GF
The list of order polynomials for the various types is given in Table 24.1
(see for example [13, Prop. 10.2.5 and Thm. 14.3.2]).
GF |GF |
n
SLn (q) q ( 2 ) (q 2 − 1)(q 3 − 1) · · · (q n − 1)
n
SUn (q) q ( 2 ) (q 2 − 1)(q 3 + 1) · · · (q n − (−1)n )
2
SO2n+1 (q) q n (q 2 − 1)(q 4 − 1) · · · (q 2n − 1)
2
Sp2n (q) q n (q 2 − 1)(q 4 − 1) · · · (q 2n − 1)
2
SO+2n (q) q n −n (q 2 − 1)(q 4 − 1) · · · (q 2n−2 − 1)(q n − 1)
2
SO−2n (q) q n −n (q 2 − 1)(q 4 − 1) · · · (q 2n−2 − 1)(q n + 1)
G2 (q) q 6 (q 2 − 1)(q 6 − 1)
3
D4 (q) q 12 (q 2 − 1)(q 6 − 1)(q 8 + q 4 + 1)
F4 (q) q 24 (q 2 − 1)(q 6 − 1)(q 8 − 1)(q 12 − 1)
E6 (q) q 36 (q 2 − 1)(q 5 − 1)(q 6 − 1)(q 8 − 1)(q 9 − 1)(q 12 − 1)
2
E6 (q) q 36 (q 2 − 1)(q 5 + 1)(q 6 − 1)(q 8 − 1)(q 9 + 1)(q 12 − 1)
E7 (q) q 63 (q 2 − 1)(q 6 − 1)(q 8 − 1)(q 10 − 1)(q 12 − 1)(q 14 − 1)(q 18 − 1)
E8 (q) q 120 (q 2 − 1)(q 8 − 1)(q 12 − 1)(q 14 − 1)(q 18 − 1)(q 20 − 1)(q 24 − 1)
·(q 30 − 1)
2
B2 (q 2 ) q 4 (q 2 − 1)(q 4 + 1) (q 2 = 22f +1 )
2
G2 (q 2 ) q 6 (q 2 − 1)(q 6 + 1) (q 2 = 32f +1 )
2
F4 (q 2 ) q 24 (q 2 − 1)(q 6 + 1)(q 8 − 1)(q 12 + 1) (q 2 = 22f +1 )
Remark 24.9 It is easy to see from the formulas in Table 24.1 that most of
the time the defining prime occurs to a much higher power in |GF | than any
other prime. From this, and using Zsigmondy’s Theorem (see Theorem 28.3),
it is straightforward to check for coincidences of orders among the (generally
24.2 BN-pair, simplicity and automorphisms 209
simple, see Theorem 24.17) groups GF /Z(GF ) for G simple of simply con-
nected type. Apart from those coming from an isomorphism of the underlying
root system (like A3 ∼= D3 or B2 ∼= C2 , for example), or from the isogeny
between type Bn and Cn in characteristic 2 (viz. Sp2n (2f ) ∼
= SO2n+1 (2f ), see
Remark 18.14(a)) there are only the following three cases of isomorphisms
PSL2 (4) ∼
= PSL2 (5), PSL2 (7) ∼
= PSL3 (2), PSU4 (2) ∼
= PSp4 (3),
and the following infinite series of order coincidences between non-isomorphic
groups:
|PSp2n (q)| = |SO2n+1 (q) | for q odd, n ≥ 3
(see for example [44, §2.9]).
A double coset B wvB ˙ on the right-hand side is F -stable if and only wv is,
that is, if v ∈ W F . But by Lemma 23.3 the only F -stable elements in WI
are 1 and sI , whence taking F -fixed points we find (BN4).
Finally, let s = sI with preimage ṡ ∈ N F and set U := Us− . Then
!
ṡU ṡ = U−α | α ∈ Φ+ , s(α) ∈ Φ− ≤ Uα ≤ B w˙0 ,
α∈Φ−
210 A BN-pair for GF
GF Z(GF ) GF Z(GF )
SLn (q), n ≥ 2 (n, q − 1) 2
B2 (22f +1 ) 1
SUn (q), n ≥ 3 (n, q + 1) 2
G2 (32f +1 ) 1
Spin2n+1 (q), n ≥ 3 (2, q − 1) G2 (q) 1
Sp2n (q), n ≥ 2 (2, q − 1) 3
D4 (q) 1
Spin+2n (q), n ≥ 4 even (2, q − 1)2 2
F4 (22f +1 ) 1
Spin+2n (q), n ≥ 5 odd (4, q − 1) F4 (q) 1
Spin−2n (q), n ≥ 4 even (2, q − 1) E6 (q) (3, q − 1)
Spin−2n (q), n ≥ 5 odd (4, q + 1) 2
E6 (q) (3, q + 1)
E7 (q) (2, q − 1)
E8 (q) 1
with the right-hand side being solvable since B is, while GF is perfect, so
H = GF . The result follows.
This criterion leads one to ask under which conditions GF is perfect, and
how this can be verified. For this, let’s set G1 := GF
u , the normal (in fact,
characteristic) subgroup of GF generated by its unipotent elements.
Theorem 24.15 (Steinberg) Let G be a simply connected semisimple linear
algebraic group with Steinberg endomorphism F : G → G. Then GF = G1 ,
that is, GF is generated by its unipotent elements.
Proof Let B ≤ G be an F -stable Borel subgroup of G with F -stable
maximal torus T and unipotent radical U , and U − = U w˙0 the unipotent
radical of the opposite Borel subgroup. By Corollary 24.2 we have that
GF = U F , (U − )F , T F . Since T F normalizes both U F and (U − )F , we con-
clude that G1 = U F , (U − )F .
Thus it remains to show that T F ≤ U F , (U − )F . Now for α ∈ ∆ let
Gα = Uα , U−α and Tα = T ∩Gα , a one-dimensional torus (see Section 8.3).
Since G is simply connected, the coroots {α∨ : k × → Tα ≤ T | α ∈ ∆}
generate Y (T ), so T = α∈∆ Tα . Collecting according to F -orbits we obtain
the F -invariant decomposition T = I⊆S TI , where I runs over F -orbits in
S and TI := α∈∆I Tα , so T F = I⊆S TIF .
We are thus reduced to the case that S is a single F -orbit. Moreover, by
Exercise 30.2 we may assume that G is simple. But then Φ has root system
of type A1 , A2 , B2 or G2 . In the second case, GF = SU3 (q) and the claim
can be checked by direct computation, as in Exercise 30.11. In the other
three cases, W F = {1, w = wS }. Let t ∈ T F and t1 ∈ T with t21 = t. Then
ẇt1 ẇ−1 = t−1 −1
1 , so t1 ẇt1 = ẇt. Recall from Corollary 24.2 that we may
choose ẇ ∈ G1 . If char(k) = 2 then ẇ and ẇt are semisimple elements of GF
which are G-conjugate, so also GF -conjugate by Theorem 26.7(c) below. In
particular, as we chose ẇ ∈ G1 , ẇt ∈ G1 and so t ∈ G1 . If char(k) = 2 this
also holds, since then |T F | is odd, so t1 may already be chosen in T F .
Example 24.16 Let G = SL2 with Steinberg endomorphism F , so GF =
SL2 (q) for some prime power q by the classification of Steinberg endo-
morphisms. We claim that G1 = GF u is perfect for q > 3 and equal to
F
G . Indeed, by Theorem 24.15 SL2 (q) is generated by its root subgroups
F
U±α = {u±α (c) | c ∈ Fq }, where
1 c 1 0
uα (c) := , u−α (c) := ,
0 1 c 1
so GF = G1 . Since q > 3 there exists a ∈ F×
q \ {±1}. Set t := diag(a, a
−1
)∈
24.2 BN-pair, simplicity and automorphisms 213
GF . Then tu±α (c)t−1 = u±α (a±2 c), so by the choice of a the map
F
U±α −→ U±α
F
, u±α (c) → [t, u±α (c)] = u±α ((a±2 − 1)c),
is surjective. Thus U±αF
≤ [GF , GF ], and so GF is perfect. Note that SL2 (q)
is solvable for q ≤ 3.
Similarly using Example 23.10(2) one can check that SU3 (q) is perfect for
q > 2 (see Exercise 30.11).
This example generalizes to yield perfectness and hence simplicity modulo
center in the simply connected case.
Theorem 24.17 (Tits) Let G be a simply connected simple linear algebraic
group with Steinberg endomorphism F : G → G. Then, unless GF is one of
SL2 (2), SL2 (3), SU3 (2), Sp4 (2), G2 (2), 2 B2 (2), 2 G2 (3), 2 F4 (2),
G1 = GF is perfect, and thus GF /Z(GF ) is simple.
Proof when F is not very twisted and q > 3 If F is not very twisted, then
an orbit ω of F on the set of simple roots of G either consists of mutually
orthogonal roots, so is of type A1 × · · · × A1 , or it is of type A2 . In the first
case, the various U±α , for α ∈ ω, commute pairwise and are permuted
transitively by F , so
U±α | α ∈ ωF ∼
m
= U±β F
for m = |ω| and any β ∈ ω, by Exercise 30.2. But Gβ := U±β is a central
quotient of SL2 by Section 8.2 and Theorem 8.17(f), so an application of
Example 24.16 shows that (Gβ )1 is perfect, and similarly when ω is of type
A2 (see Exercise 30.11). Now by Proposition A.11 any root subgroup of
GF = G1 is conjugate to a root subgroup corresponding to a simple root.
So [G1 , G1 ] contains all root subgroups of G1 , hence a Sylow p-subgroup by
Corollary 24.11, and thus equals G1 . Now apply the previous theorem and
Corollary 24.14.
A slight variation of the above argument applies when q ≤ 3. The groups
of type 2 B2 and 2 G2 have to be treated differently. Then, the groups of type
2
F4 can again be handled as before, see [82].
Remark 24.18 The groups SL2 (2), SL2 (3), SU3 (2), 2 B2 (2) occurring in
the exceptions of Theorem 24.17 are solvable, the groups Sp4 (2) ∼ = S6 ,
G2 (2) ∼ = Aut(PSU3 (3)), 2 G2 (3) ∼= Aut(PSL2 (8)) are almost simple, and
2
F4 (2) contains a normal simple subgroup of index 2 which does not occur
elsewhere in the classification of finite simple groups, the so-called Tits group
(see [82, 4.3]).
214 A BN-pair for GF
1 −→ Z −→ H −→ H/Z −→ 1
(a) GF /π(GF ∼
sc ) = Z/L(Z) and
F ∼
(b) π(Gsc ) = Gsc /Z F .
F
F ∼
In particular, if GF F F F F
sc is perfect then π(Gsc ) = [G , G ] = Gsc /Z , and
F F
[Gad , Gad ] is simple.
Proof We apply Lemma 24.20 with H = Gsc . As Gsc is connected the map
L : Gsc → Gsc , g → F (g)g −1 , is surjective by Theorem 21.7 and we obtain
the exact sequence
π
1 −→ Z F −→ GF
sc −→ G −→ Z/L(Z) −→ 1,
F
In fact, these already generate the full automorphism group. Let’s first
consider the case of groups of rank 1:
Proposition 24.23 Any automorphism of S = PSL2 (pf ) is the product of
an inner, a diagonal and a field automorphism.
This is an easy exercise, see for example [37, Aufg. 11.15].
Theorem 24.24 (Steinberg) Let Gsc be simple of simply connected type,
F : Gsc → Gsc a split Frobenius map, G = Gad the adjoint type image
under the natural isogeny, and assume that GF sc is perfect (and hence S :=
[GF , GF ] is simple). Then any element of Aut(S) is the product of an inner,
a diagonal, a field and a graph automorphism.
Proof for the case that all roots of G have the same length Let T ≤ B be
a maximal torus inside a Borel subgroup of G with root system Φ and set
of simple roots ∆. As pointed out above, we may assume that F = Fpf
for the endomorphism Fp from Theorem 16.5, for some f ≥ 1, and T, B
are F -stable and F acts trivially on Φ. Let U = Ru (B) be the unipotent
radical of B. Then P := U F is a Sylow p-subgroup of GF with normalizer
B F by Corollary 24.11. By Proposition 24.21, GF /S consists of semisimple
elements, so P is a Sylow p-subgroup of S. Thus, by Sylow’s Theorem any
σ ∈ Aut(S) can be multiplied by an inner automorphism so as to stabilize P .
We claim that we may modify σ further so as to also stabilize P − :=
(U − )F = (U ẇ0 )F , a second Sylow p-subgroup of S. Indeed, σ(P − ) must be
S-conjugate to P , by an element g = uẇb ∈ S with w ∈ W F , u ∈ (Uw− )F
and b ∈ B F (see Theorem 24.1); as B normalizes U we have σ(P − ) =
uẇP ẇ−1 u−1 . Now
uẇP ẇ−1 u−1 ∩ P = σ(P − ) ∩ σ(P ) = σ(P − ∩ P ) ≤ σ(U − ∩ U ) = 1
by Corollary 11.18, whence ẇP ẇ−1 ∩ P = 1 by conjugation by u ∈ (Uw− )F ≤
U F = P . If α ∈ Φ+ is not made negative by w, then ẇUαF ẇ−1 ≤ P , so w
must send all positive roots to negative ones and thus w = w0 . So σ(P − ) =
uw˙0 P w˙0 −1 u−1 = uP − u−1 . Changing σ by the inner automorphism induced
by u (which stabilizes P ) then gives σ(P − ) = P − .
Since σ fixes B F , it permutes the minimal parabolic subgroups above B F ,
which are of the form B F ∪ B F s˙α B F for some α ∈ ∆ by Proposition 12.2
(and the subsequent remark). Now U − ∩ B ṡα B = U−α , so σ induces a
permutation on the set {U−α F
| α ∈ ∆}, and similarly a permutation ρ on
{Uα | α ∈ ∆}. Note that these two permutations must agree since Uα , U−β
F
We now turn to the subgroup structure of the finite groups of Lie type,
first discussing maximal tori and then some aspects of Sylow subgroups. Let
G be connected reductive with Steinberg endomorphism F , B an F -stable
Borel subgroup of G. Recall that p = char(k). By Corollary 24.11 (or by the
order formula in Corollary 24.6) U F = Ru (B)F is a Sylow p-subgroup of GF .
Some results on the structure of U F were obtained in Section 23.2; see also
Theorem 26.5 below. Here we will be concerned with Sylow subgroups of GF
for the other prime divisors of its order. Again, there is a close connection
to the structure of the algebraic group G.
!
rk(G)
!
|G| = x|Φ |
(xdi − i ) = x|Φ |
+ +
Φd (x)a(d)
i=1 d≥1
for suitable integers a(d) ≥ 0. Thus the irreducible factors of the generic
order |G| are the Φd (x) with a(d) > 0, so the Φd (q) can in some sense be
considered as the “generic primes” dividing |G|(q) = |GF |. In this section
we will substantiate this statement. In order to do this, we first need more
precise information on the classification of F -stable maximal tori and the
orders of their F -fixed points.
25.1 F -stable tori 219
Proof By Lemma 22.1 we have T F ∼ = X(T )/(F −1)X(T ). The first equality
is thus a direct consequence of the elementary divisor theorem. With F |X =
qφ we get det(F − 1) = det(φ) det(q − (φ)−1 ). Since φ has finite order, all its
eigenvalues are roots of unity. So | det(φ)| = 1, and moreover all eigenvalues
of q − (φ)−1 have real part at least q − 1 > 0, whence the real number
det(q − (φ)−1 ) is positive.
Writing n̄ for the image of n in NG (T1 )/T1 = W this is the case if and only
if n̄w = wφ(n̄), so if and only if n̄ lies in the centralizer CW (wφ), giving (a).
For (b) note that F sends gtg −1 ∈ gT1 g −1 = Tw to gwF (t)w−1 g −1 , so
it acts on Tw like wF acts on T1 . The claim is then just the assertion of
Lemma 22.1(b). Part (c) is Proposition 25.2.
(1) For G = GLn with Fq : G → G the standard Frobenius map with fixed
point group GFq = GLn (q) we let T = Dn denote the maximally split
torus of diagonal matrices. For g ∈ GLn with
⎛ ⎞
0 ··· 0 1
⎜ 1 0. ⎟
g −1 Fq (g) = w := ⎜
⎝ 0 ...
⎟
.. ⎠ ∈ GLn
0 0 1 0
F×
q n Gal(Fq n /Fq ) → GLn (q)
where [wφ] denotes the φ-class of w. Here, |NGF (Tw )| = |CW (wφ)| · |TwF |
by Proposition 25.3(a). Now by Proposition 25.3(c) and using that |[wφ]| =
222 Tori and Sylow subgroups
|GF | ! q N ! q di − i
l l
1 N +d1 +...+dl −l
= −1 −di = q = q 2N ,
q l i=1 1 − −1
i q −di q l
i=1
1 − i q
where we’ve also used Theorem 24.4(c) and Corollary 24.6. (Note that the
multiset {(di , i )} is invariant under complex conjugation, since φ is a real
endomorphism.)
SF ∼
= X/(F − 1)X ∼
= X/(qζ − 1)X ∼
= X/(q − ζ −1 )X
as R-modules.
So from now on we may assume that X is a projective R-module of rank 1,
hence can be identified with an ideal of R. Then multiplying by its inverse
we obtain X/(q − ζ −1 )X ∼= R/(q − ζ −1 )R, whose order equals the norm of
−1
the principal ideal (q − ζ )R, so that |S F | = Φd (q) in this case. If n ∈ N
with nX ⊆ (q − ζ −1 )X, then q − ζ −1 divides n in R. As Z[ζ] = Z[x]/(Φd )
this implies n ≡ q − x (mod Φd ), so n ≡ 0 (mod Φd (q)). Hence S F is cyclic
of order Φd (q).
More generally, maximally split tori always contain a Sylow 1-torus, and
conversely an F -stable maximal torus containing a Sylow 1-torus is maxi-
mally split. Tori always possess Sylow d-tori (see Exercise 30.16):
Lemma 25.9 Let (T, F ) be a torus with corresponding complete root datum
T = (X, ∅, Y, ∅, φ). Then S = (X , ∅, Y , ∅, φ), with Y = kerY (Φd (φ)) and
X = X/Ann(Y ), is the complete root datum of the Sylow d-torus of T .
We now show why the name Sylow tori is justified. For this we need a
further fundamental result from the theory of finite reflection groups (see
[65, 3.4 and 6.2]):
V (wφ, ζ) := {v ∈ V | wφ v = ζv}
denote the ζ-eigenspace of wφ. Let a(ζ) := maxw∈W dim V (wφ, ζ). Then we
have:
(a) For all w1 , w2 ∈ W with dim V (w1 φ, ζ) = dim V (w2 φ, ζ) = a(ζ) there
exists w ∈ W with w.V (w1 φ, ζ) = V (w2 φ, ζ).
(b) For any w ∈ W there exists w ∈ W with V (wφ, ζ) ≤ V (w φ, ζ) and
such that dim V (w φ, ζ) = a(ζ) is maximal.
This implies the following result (see [10, Thm. 3.4], and also Exercise 30.17
for the case of GLn ):
224 Tori and Sylow subgroups
so that r|q e −1, but r does not divide q d −1 for any positive integer d < e. We
require the following standard number theoretical lemma (see [56, Lemma
5.2], for example).
(a) There exists a unique d such that a(d) > 0 and r|Φd (q). Then any Sylow
d-torus of G contains a Sylow r-subgroup of GF .
(b) The Sylow r-subgroups of GF are homocyclic with a(d) cyclic factors of
order |Φd (q)|r .
so we cannot expect to embed them into tori. But it turns out that Sylow
r-subgroups do embed in the normalizer of a maximal torus.
Definition 25.15 An algebraic group S is supersolvable if there exists a
series 1 = S0 S1 · · · Sn = S of closed normal subgroups Si S such
that Si /Si−1 is either cyclic or a torus for all i = 1, . . . , n.
For example, finite p-groups, having non-trivial center, are supersolvable,
hence so are finite nilpotent groups.
Theorem 25.16 Let G be a connected reductive group, F : G → G a
Steinberg endomorphism, S ≤ G an F -stable supersolvable subgroup such
that each Si S is F -stable and consists of semisimple elements. Then there
exists an F -stable maximal torus T ≤ G with S ≤ NG (T ).
Proof We argue by induction on the length of the chain 1 = S0 · · · Sn =
S. If S = 1, then the existence of an F -stable maximal torus T is guaranteed
by Corollary 21.12. Secondly, if S = 1 but CG (S1 ) = G, then S1 ≤ Z(G),
so S1 lies in every maximal torus of G by Corollary 8.13(b). In this case we
may argue with a shorter chain in G/S1 by induction.
Otherwise, G1 := CG (S1 )◦ is a proper connected F -stable subgroup of G,
reductive by Corollary 8.13(a) (if S1 is a torus) respectively by Theorem 14.2
(if S1 is cyclic), and the maximal tori of G1 are maximal tori of G by the
same references. Now, by the previous case there exists an F -stable maximal
torus of G1 which is normalized by S. This is a maximal torus of G.
Corollary 25.17 Let r be a prime, r = p. Then any Sylow r-subgroup of
GF lies in NGF (T ) for some F -stable maximal torus T ≤ G.
Proof As the Sylow r-subgroups are nilpotent and consist of semisimple
elements, and finite nilpotent groups are supersolvable, this follows immedi-
ately from Theorem 25.16 by taking F -fixed points.
Example 25.18 Let G = GLn , F the standard Frobenius map with re-
spect to Fq . The normalizer in GF of the maximally split torus Dn of G
equals NGF (Dn ) = DF F
n .W = Dn .Sn (see Example 8.2(1)). It can be checked
from the order formula that for q ≡ 1 (mod 4) this contains a Sylow 2-
subgroup of GLn (q). On the other hand, if q ≡ 3 (mod 4) then the order
of NGF (Dn ) is not divisible by the full 2-part of the order of GLn (q), so in
particular it cannot contain a Sylow 2-subgroup. In this case, though, the
normalizer of a Sylow 2-torus contains a Sylow 2-subgroup.
We now show that centralizers of Sylow subgroups are strongly related to
centralizers of Sylow tori.
25.2 Sylow subgroups 227
Theorem 25.19 Let G be simple with a not very twisted Steinberg endo-
morphism F : G → G, let r be a prime divisor of |GF |, r = p. Then every
semisimple element g ∈ GF which centralizes a Sylow r-subgroup of GF lies
in an F -stable torus containing a Sylow d-torus of G, where d = er (q). In
particular, g centralizes a Sylow d-torus.
Proof The first claim follows exactly as in the special case I = ∅, that
is, PI = B, in Corollary 21.12. Indeed, PI is self-normalizing by Proposi-
tion 12.2(a), so there is a unique GF -conjugacy class of F -stable G-conju-
gates of PI by Theorem 21.11(b). Taking F -fixed points in the double coset
decomposition for PI in Proposition 12.2 we obtain that PIF = PIF .
For part (b) we use that the standard Levi complement LI of PI is gen-
erated by an F -stable subset of root groups and so is F -stable, and that all
Levi complements in PI are PI -conjugate (see Proposition 12.6). It is clear
by this same result that UIF ∩ LFI ≤ UI ∩ LI = {1}. Now let p ∈ PI . Then
F
That is, the standard parabolic subgroups of the BN-pair GF arise as the
F -fixed points of F -stable standard parabolic subgroups of G. Note however
that in the finite case it is no longer true in general that all complements
to UIF in PIF are conjugate to LFI . Over small finite fields, the central torus
of LI used in the proof of Proposition 12.6 may have a trivial fixed point
group, so that the argument given there fails in this case; see the following
example.
∼ ×
and F -stable Levi complement LI with F -fixed points LF I = Fq 2 ×
GUn−2 (q).
(2) On the other hand, for n ≥ 3 the maximal parabolic subgroup
∗ ∗
PJ :=
0 A A ∈ GLn−1
Example 26.8 How many semisimple conjugacy classes are there in the
general linear group GLn (q)? A conjugacy class is determined by its Jordan
canonical form, hence semisimple classes are parametrized by the character-
istic polynomial of their elements. Any monic polynomial over Fq of degree n
and with non-vanishing constant coefficient occurs as characteristic polyno-
mial, and there are precisely (q − 1)q n−1 of these, so this is the number of
semisimple conjugacy classes of GLn (q).
Sketch of proof The sum in question is the scalar product on the coset W.σ
of the class function detV (x idV − σw) with the trivial character (see [22, I.6]
for the elementary properties of such scalar products). Now note that the
characteristic polynomial of any z ∈ GL(V ) is given by
"
n
det V (x idV − z) = (−1)i χi (z) xn−i ,
i=0
where n = dim V and χi is the character of GL(V ) on the ith exterior power
∧i (V ). Using the assumptions on W one shows that W has trivial fixed space
on all ∧i (V ), i > 0, so only the term for i = 0 in the sum contributes to that
scalar product, yielding the value xn .
Theorem 26.10 (Steinberg) Let G be a connected reductive algebraic
group, G := [G, G], l := rk G , Z := Z(G)◦ , and F : G → G a Steinberg
endomorphism. Then:
(a) The number of F -stable semisimple conjugacy classes of G equals q l |Z F |.
(b) If the derived group G is simply connected, then this is also the number
of GF -classes of semisimple elements in GF .
Proof By Proposition 14.6 the F -stable semisimple conjugacy classes of G
are in bijection with the set (T /W )F of F -stable W -orbits on an F -stable
maximal torus T of G. Thus, we are counting t ∈ T satisfying F (t) = tw for
some w ∈ W , up to W -conjugation. Hence
1 " " 1 " " 1 " w−1 F
|(T /W )F | = 1= 1= |T |.
|W | |W | |W |
t∈T w∈W w∈W
t∈T w∈W
F (t)=tw F (t)=tw
−1 −1
Now |T w F | = |Z F | · |S w F | with Z = Z(G)◦ and S = T ∩ G by Ex-
−1
ercise 30.22, and by Proposition 25.2 we have |S w F | = detV (q − φ−1 w),
where V = X(S) ⊗Z R, giving
|Z F | "
|(T /W )F | = det V (q − φ−1 w).
|W |
w∈W
Now (a) follows from Lemma 26.9, (b) follows from this by Theorem 26.7(c).
w F
Hw NG (Hw )F /Hw
F
|Z(Hw
F
)| |Z(H̄w
F
)|
() A2 (q)3 S3 gcd(3, q − 1)2 gcd(3, q − 1)
(1 2) A2 (q 2 ) · 2A2 (q) Z2 gcd(3, q 2 − 1) gcd(3, q + 1)
(1 2 3) A2 (q 3 ) Z3 gcd(3, q − 1) gcd(3, q − 1)
The aim of this chapter is to study the subgroup structure of the finite clas-
sical groups. The classical groups are defined as certain subgroups of the
isometry groups of the zero form or of non-degenerate bilinear, sesquilinear
or quadratic forms f on finite-dimensional vector spaces over finite fields.
According to the classification of such forms and their isometry groups
(which can be found for example in [2, §21]), there exist the following non-
degenerate possibilities: f is symplectic on an even-dimensional vector space,
f is quadratic of maximal Witt index on an odd-dimensional vector space,
f is quadratic on a 2n-dimensional vector space, of Witt index n or n − 1, or
f is unitary on a vector space over a field with a subfield of index 2. Here,
the Witt index of a quadratic form is the maximal dimension of a totally
singular subspace.
The various possibilities for the isometry groups are collected in Table 27.1.
Definition 27.1 We shall refer to groups in the last column of the above
table as classical groups, and we shall denote them by Cl(V ).
27.1 The theorem of Liebeck and Seitz 237
This is clear from our definition of the classical algebraic groups in Sec-
tion 1.2, respectively from the construction of SO− 2n (q) in Example 22.9.
(In the case of orthogonal groups in characteristic 2 this can serve as the
definition of the finite special orthogonal groups.)
A similar construction is not possible for the classical groups SUn (q),
despite the fact that they also arise as fixed points of SLn under a Steinberg
endomorphism, as the concept of a sesquilinear form is tied to the field in
question. So it is not possible to obtain compatible Fq -structures on SLn and
the underlying natural n-dimensional k-vector space k n . The SUn (q)-case is
therefore more technical and we will not go into the proofs for this case here.
We want to apply techniques from algebraic groups to obtain results on
classical groups. For this, we need the following standard result (see for
example [66, §11.1]).
define
f˜ : V × V → k, f˜(v, w) := F −1 (f (F (v), F (w))).
We now state a refined version of the Reduction Theorem 18.6 for maximal
subgroups of classical algebraic groups from Part II which accommodates for
the presence of a Steinberg endomorphism, see [50, Thm. 1 ]. This will allow
us later to descend to the finite groups GF . The proof is very similar in that
one basically has to go through the proof of Theorem 18.6 once again, but
now paying attention to F -stability.
Recall the subgroup classes C1 , . . . , C5 from Section 18.1.
In the case that all Vi are one-dimensional, this leads to the normalizer of
the Singer cycle as in Example 25.4(1).
27.2 The theorem of Aschbacher 241
n = mt , or
Class C4 : twisted tensor product subgroups Clm (q t ).t, where n = mt with
F
t prime.
Here, a tensor product structure is preserved over Fq , but not over Fq , like
for example (modulo scalars) PSLm (q t ) < PSLmt (q), PSLm (q 2 ) < PSUm2 (q)
or PSpm (q t ) < PSpmt (q).
Next, if M ∈ C5 is F -stable, then it has a characteristic (and hence also
F -stable) normal r-subgroup R which is extraspecial or a central product of
such with Z4 . Here the F -fixed points lie in class
Class C5F : NGF (R) with NG (R) in C5 , as in Table 27.2 (see [50, p. 430] or
[44, Tab. 4.6.A]).
This completes our discussion of F -stable members of C1 , . . . , C5 . If H ≤
G is F -stable but not contained in any F -stable member of C1 ∪ . . . ∪ C5 ,
then by Theorem 27.4 it is an almost simple subgroup (modulo Z(G)), with
H ◦ acting irreducibly on V . Let’s single out two further subclasses in this
situation:
242 Maximal subgroups of finite classical groups
G NG (R)/Z(G)
SLrm r2m .Sp2m (r)
Sp2m 22m .GO−2m (2)
SO2m 22m .GO+2m (2)
2
Class C6F : subgroups of GF which are classical on V F (respectively on V F ,
for GF a unitary group).
Examples of this type are the normalizers of Spn (q) or SO(±) n (q) inside
SLn (q) and SUn (q), the normalizer of SUn (q 1/2 ) inside SLn (q), and the nor-
malizer of SOn (q) inside Spn (q) if p = 2.
Finally, for subfields Fq < Fq of prime index, we get:
Class C7F : subfield subgroups NGF (Cln (q )), with |Fq : Fq | prime, where
GF = Cln (q).
We can now state the principal reduction result on maximal subgroups of
finite classical groups, first proved by Aschbacher [1]; here we present the
approach by Liebeck and Seitz [50] using descent from the corresponding
algebraic groups.
Theorem 27.5 (Aschbacher, Liebeck–Seitz) Let G ≤ SL(V ) be a classical
group on the finite-dimensional k-vector space V , stable under a Steinberg
endomorphism F : SL(V ) → SL(V ), so that GF is a finite classical group on
2
V F (respectively V F ), and let H ≤ GF . Then one of the following holds:
(1) H lies in a member of C1F , . . . , C7F , or
(2) H lies in class S, that is, HZ(G)/Z(G) is almost simple, H ∞ acts
absolutely irreducibly and tensor indecomposably on V F (respectively
2
V F ), the representation cannot be realized over a proper subfield, and
if GF = SLn (q) then H fixes no non-degenerate bilinear or unitary form
on V .
Proof for GF not unitary Apply Theorem 27.4 and our previous discus-
sion. Since H is F -stable, either H ≤ M F for an F -stable member M ∈
C1F ∪ . . . ∪ C5F , or HZ(G)/Z(G) is almost simple and H ∞ acts absolutely
irreducibly (since otherwise it fixes a direct sum decomposition over Fq and
H lies in a member of C2 ). If H lies in no member of C7F , then it cannot be
F
Example 27.6 Assume that GF = SLn (q) for n a prime. Then only the
classes C1F , C5F , C6F , C7F and S can occur, since for the other classes n has to
be composite (see Proposition 28.1 for arbitrary n).
In fact Liebeck and Seitz [50] prove a more general version of the re-
duction theorem for normalizers of classical groups GF inside ΓL(V F ), the
extension of GL(V F ) by the group of field automorphisms, where additional
classes C1 and C7 can arise. We shall not attempt to describe this here.
F F
This general version then allows one to obtain a similar reduction theorem
for maximal subgroups of all isogeny types of classical groups. Indeed, for
our given classical group G = Cl(V ) let Gsc be the corresponding group of
simply connected type. By Proposition 9.15 there exists a natural isogeny
π : Gsc → G which restricts to a homomorphism GF sc → G
F
with central
F F
kernel and whose image contains [G , G ] by Proposition 24.21. Now note
that π induces a bijection between maximal subgroups of GF sc and those of
F F
its image π(Gsc ) in G . In this way, the results on maximal subgroups of
[GF , GF ] can be transferred to those of GF sc .
On the other hand, if Gad denotes the corresponding group of adjoint type,
then the natural isogeny G → Gad induces a homomorphism GF → GF ad
whose image contains [GF F F F F
ad , Gad ]. Now, Gad is an extension of [Gad , Gad ] by
certain (diagonal) automorphisms, so the above-mentioned general version
of the reduction theorem applies to the maximal subgroups of GF ad as well.
Remark 27.7 (Relationship between the classes CiF and Aschbacher’s and
Kleidman–Liebeck’s classes Ci ) The classes C1F , . . . , C7F of natural subgroups
defined above do not agree completely with the classes Ci introduced by Asch-
bacher in his original paper [1]. Moreover, the latter were already modified
in the book of Kleidman and Liebeck [44] in which the structure, conjugacy
and maximality of members in these classes were determined. The following
table indicates the relationship between the classes CiF here and in [50], and
the classes Ci in [44]:
[50] C1F , C1 C2F , C2 , C3F C4 C7F , C7 C4 , S
F F F F F
C4F C5F C6F
[44] C1 C2 , C3 C4 C7 C6 C8 C5 S
28
About the classes C1F , . . . , C7F and S
The theorem of Aschbacher in the previous chapter still leaves several im-
portant questions open:
Proposition 28.1 The maximal members in the classes C1F , . . . , C7F for
28.1 Structure and maximality of groups in CiF 245
GF = SLn (q), n ≥ 3, are as given in Table 28.1. Here, for the classes
C2F , C2 , C4F , C4 , the maximal subgroup is obtained as the normalizer in
F F
SLn (q) of the intersection of the corresponding entry in the table with SLn (q).
About the proof It is clear from the description in Table 18.2 of the maximal
members in the classes C1 , C2 , C4 , C6 that there exist F -stable representatives,
with F -fixed points in C1F , C2F , C4F , C4 , C6F as given in Table 28.1. Moreover,
F
In all classes and for all classical groups there are only three families of
examples in which maximal members of a class CiF lie in some group in
the class S. They all originate from the fact that in characteristic 2 the
246 About the classes C1F , . . . , C7F and S
GF class H K conditions
C4
F
SO+
2t
(q) Sp2 (q) St Sp2t (q) q ≥ 4 even, t ≥ 4
C4
F
SO+
2t
(q) Sp2 (q) St SO+
2t+2 (q) q ≥ 4 even, t ≥ 5 odd
F
SO+
4t
(q) C4 Sp4 (q) St Sp4t (q) q even
Example 28.2 Let GF = SO+ 2t (q), with t ≥ 4 and q ≥ 4 even. Then the
class C4 contains the tensor product subgroup H := Sp2 (q) St . On the
F
of Lie type and its factorization into cyclotomic polynomials, which in some
sense behaved like generic prime divisors of the order. There is one further
aspect to this factorization, given by the theorem of Zsigmondy (see [38, IX,
Thm. 8.3]):
Theorem 28.3 Let q > 1 be a prime power. Then there exists a prime
divisor of q n − 1 which does not divide q i − 1 for 1 ≤ i ≤ n − 1, unless one
of:
Zsigmondy primes can often be used to rule out the existence of certain
embeddings of quasi-simple groups of Lie type in the same characteristic, like
248 About the classes C1F , . . . , C7F and S
for example B8 (q) ≤ E8 (q) for all q (using (16, q)). To deal with embeddings
in different characteristic, it is useful to know the minimal dimensions of
faithful representations of groups of Lie type in cross characteristic. These
turn out to be exponential in the rank of the group, and thus much larger
than dimensions of representations in the defining characteristic. We give
one example of such a result:
Proposition 28.6 (Landazuri–Seitz) Let K be a field of characteristic
r = p.
by the order formula (see Example 24.8(2)), while the order of |GLd (2)| is
of the order of magnitude
2 n−1
−1)2 2n−2
2d = 2(q = 2q
by the bound for d in Proposition 28.6. So indeed these are extremely small
subgroups.
We close the chapter with some comments on our fourth question, the
maximality of the groups in class S. Here, one has to investigate the follow-
ing situation: we have embeddings H < K < GF = Cl(V )F (respectively in
2
SU(q 2 )), both H, K absolutely irreducible on V F (respectively V F ), with
HZ(G)/Z(G) and KZ(G)/Z(G) almost simple. By the classification of finite
simple groups, H, K are either sporadic, alternating, or of Lie type. More-
over, in the latter case, the characteristics of H, K may coincide or not, and
may or may not equal the characteristic of G. This leads to quite a number
of different possible configurations for (H, K, Cl(V )) to consider.
For example, if K is one of the 26 sporadic groups, the classification of
triples H < K < GF is essentially a question on maximal subgroups of
sporadic simple groups. In the situation where H is sporadic, one has to
determine all absolutely irreducible representations of sporadic groups and
their covering groups. Both are finite, but nevertheless very difficult problems
which have not yet been solved for all groups.
If K is an alternating group An , one has to decide which absolutely ir-
reducible representations of covering groups of An restrict irreducibly to
(almost simple) subgroups. The cases where H and K are of Lie type in dif-
ferent characteristics can essentially be dealt with using the Landazuri–Seitz
bounds.
This short overview should have convinced the reader that the methods
to study the groups in class S are of a representation theoretic nature; in
particular they lie outside the scope of this text.
29
Finite exceptional groups of Lie type
In this chapter we are concerned with the determination of the maximal sub-
groups of the finite exceptional groups of Lie type. According to Table 22.1
these are the members of the ten infinite series
2
B2 (22f +1 ), 2 G2 (32f +1 ), G2 (q), 3D4 (q), 2F4 (22f +1 ),
other isogeny types the discussion before Remark 27.7 applies; by Table 9.2
this is only relevant for groups of types E6 and E7 .) The first result here is
(see [48, Thm. 1]):
show that CG (s) ∼ = SL2 ◦SL2 , the subgroup constructed in Example 14.19(2).
By Example 14.19(3) both factors contain a quaternion subgroup of order 8.
Their central product in CG (s) is then extraspecial of order 32 and contains
a self-centralizing elementary abelian subgroup R of order 8, with the stated
normalizer. Similarly, for G = F4 and r = 3, respectively G = E8 and r = 5
there is s ∈ R \ {1} with centralizer SL3 ◦ SL3 , respectively SL5 ◦ SL5 , cor-
responding to maximal rank subsystems of type A2 A2 , respectively A4 A4 ,
inside which R can be constructed.
Note that there do exist maximal local subgroups which are not covered by
Theorem 29.3: they are normalizers of some positive-dimensional semisimple
subgroups, as for example in Theorem 29.1(4)!
To complete our investigation of maximal subgroups now assume that
H ≤ GF is not local, and does not normalize a proper connected subgroup
of positive dimension, that is, it is not covered by Theorem 29.1. Then its
generalized Fitting subgroup is a direct product F ∗ (H) = H1 × · · · × Ht
of non-abelian simple groups Hi . We aim to show that t = 1 (and hence
H is almost simple), so instead assume t > 1. For each i, let CC(Hi ) :=
CG (CG (Hi )) be the double centralizer of Hi .
A5 , A6 , or PSU4 (2). Finally one can show that the only possibility is G = E8
and F ∗ (H) = A5 ×A6 , see Borovik [8]. Otherwise, t = 1 and so H ≤ NG (H1 )
is almost simple.
We have thus sketched the proof of the following reduction theorem (see
[48, Thm. 2]):
a subgroup of a finite group of Lie type, does it arise as the group of fixed
points under a Steinberg endomorphism of some closed positive-dimensional
subgroup of the associated algebraic group? In particular, when considering
the configurations of Theorem 29.5(4), when H is again a finite group of
Lie type in characteristic p, one is led to ask whether H is the fixed point
subgroup of some maximal positive-dimensional closed connected subgroup?
Let’s make this question a bit more precise. Let G, H be simple algebraic
groups over k = Fp . Let F : G → G, F1 : H → H be Steinberg endomor-
phisms, so GF , H F1 are finite groups of Lie type. Let ϕ : H F1 → GF be a
group homomorphism. Then under what conditions can ϕ be lifted to an
appropriate morphism of algebraic groups, that is, when does there exist
ϕ : H → G, a morphism of algebraic groups, such that ϕ|H F1 = ϕ? The first
major result in this direction is due to Steinberg [70, Thm. 1.3]:
Theorem 29.6 (Steinberg) Let V be a finite-dimensional vector space over
Fq , q = pa . Let ϕ : H F1 → GL(V ) be an absolutely irreducible representation
of H F1 , i.e., H F1 acts irreducibly on V ⊗Fq k. Then there exists a rational
representation ϕ : H → GL(V ⊗Fq k) such that ϕ|H F1 = ϕ. In particular,
absolutely irreducible representations of finite groups of Lie type over fields
of characteristic p are also parametrized by highest weights.
Remark 29.7 In fact Steinberg also shows how to obtain a complete set
of absolutely irreducible representations for H F1 . In case F1 is a standard
Frobenius morphism, say a q-power map, then
{L(λ) | 0 ≤ λ, α∨ < q for all α ∈ ∆}
forms a complete set of non-isomorphic irreducible kH F1 -modules.
The following example shows that the condition of absolute irreducibility
is necessary.
Example 29.8 Let GF = SL2 (3) (isomorphic to a 2-fold central extension
of A4 ) with N GF the normal Sylow 2-subgroup, so GF /N is cyclic of
order 3, generated by c. Consider the representation
⎛ ⎞
1 1 1
GF /N → GL3 (F3 ), c → ⎝0 1 1⎠ .
0 0 1
This defines an indecomposable but reducible representation of GF .
We claim that there does not exist a rational representation of G :=
SL2 (F3 ) whose restriction to GF is the above described representation. Sup-
pose the contrary and let V be the underlying F3 G-module which affords the
256 Exceptional groups of Lie type
The goal of this section is to discuss results that generalize Theorem 29.6.
It is perhaps not surprising that one can obtain a similar result when GL(V )
is replaced by one of the classical groups SL(V ), Sp(V ) or SO(V ), making
some appropriate assumptions on the action of H F1 on the natural module
for the classical group. In fact, one can also say something for representations
inside groups of exceptional type. We just need to know that a semisimple
group is “recovered” by its action on its Lie algebra, as shown in [69, 4.2]:
with
! !
Q := qα and qw := qα
α∈Φ+ α∈Φ+ ,w(α)∈Φ−
with complete root data G, respectively H. Then the polynomial |H| divides
|G| in Z[x].
[Hint: You may want to use that both polynomials are monic and that the rational
function |G|/|H| takes integral values for infinitely many prime powers q.]
Exercise 30.11 Use Example 23.10(2) to show that SU3 (q) is perfect for
q>2
Exercise 30.12 (Steinberg) Let G be a group, Z ≤ G a central subgroup,
and F : G → G an automorphism normalizing Z. Then the exact sequence
π
1 −→ Z −→ G −→ G/Z −→ 1
induces a long exact sequence
π δ
1 −→ Z F −→ GF −→ (G/Z)F −→ (L(G) ∩ Z)/L(Z) −→ 1,
where L : G → G is defined by L(g) := g −1 F (g).
[Hint: Show that δ := ν ◦ L ◦ π −1 , with ν : Z → Z/L(Z) the natural map, is a
well-defined homomorphism. See for example [26, Prop. 4.2.3].]
(a) Classify the maximal tori in GUn (q) and work out their orders.
(b) Classify the maximal tori in Sp2n (q) and work out their orders.
Exercise 30.15 (Molien’s formula) Let V be a finite-dimensional complex
vector space, W ≤ GL(V ) finite and φ ∈ NGL(V ) (W ). Denote by V W the
invariants of W in V .
1
(a) Show that tr(φ, V W ) = |W | w∈W tr(φw, V ).
(b) For a graded complex W -module ⊕i≥0 Vi and g ∈ GL(⊕Vi ) define the
graded trace grtr(g, ⊕Vi ) := i≥0 tr(g, Vi )xn ∈ C[[x]]. Now the symmet-
ric algebra S(V ) of V is a graded W -module. For g ∈ GL(V ) show that
grtr(g, S(V )) = detV (1 − gx)−1 .
266 Exercises for Part III
Exercise 30.20 Show that the general unitary group GUn (q) has exactly
two orbits on the lines in its natural representation on Fnq2 , and determine
the stabilizers.
[Hint: Theorem 21.11 does not apply, since here the F -structures on GLn and
on V = kn are not compatible. You should find the groups constructed in Exam-
ple 26.3.]
The structure theory for linear algebraic groups shows that many questions
on these groups can be translated into questions on the associated root
system, which are then of a purely combinatorial nature.
In this appendix we collect some basic results on root systems as they are
pertinent to our study of algebraic groups and finite groups of Lie type. Most
of these are well-known (see for example [9, §VI] or [33, Chap. III]), except
possibly for the discussion of maximal subsystems in Sections B.3 and B.4.
We do not present the proof of the classification of indecomposable root
systems, since it is not used here and is very well-documented in the literature
(see for example [9, VI, §4] or [33, §11]), nor the properties of invariant rings
of finite groups generated by reflections (see [9, V, §5] or [47]).
for all v ∈ E.
Proof Suppose that this fails for some α, β. By Proposition A.1 we then
have sα .β = β − cα for some c > 0, and by (R3) this has to lie in Φ,
so either sα .β or −sα .β lies in Φ+ . If sα .β ∈ Φ+ then we can write it as
sα .β = γ∈∆ cγ γ with cγ ≥ 0. Now, if cβ ≥ 1 we get
"
0 = sα .β − (β − cα) = cα + (cβ − 1)β + cγ γ
γ =β
which is absurd since all summands on the right are non-negative, and the
first is strictly positive. On the other hand, if cβ < 1 then the above equation
implies that (1 − cβ )β is a non-negative linear combination of ∆ \ {β} in
contradiction to the minimal choice of ∆. Thus −sα .β ∈ Φ+ , but then a
similar argument leads to a contradiction.
In particular this shows that bases of root systems exist, and we can speak
of the base in a positive system, or the positive system containing a given
base. The elements of ∆ are then also called the simple roots with respect
to Φ+ .
We next aim to show that the Weyl group acts transitively on the set of
bases. For this we need:
show. Now assume that n > 0. Then the base ∆ of Φ+1 cannot be contained in
Φ2 , say α ∈ ∆ ∩ −Φ2 . Then Lemma A.8 shows that |sα .Φ+
+ +
1 ∩ −Φ2 | = n − 1.
+
The result now follows from the inductive hypothesis applied to sα .Φ+ +
1 , Φ2 .
272 Root systems
Proof The result is trivially true for the roots in ∆. Now let β ∈ Φ and
(β,α)
α ∈ ∆. If the claim holds for β, then so it holds as well for sα .β = β −2 (α,α) α
by (R4). Since any β ∈ Φ is of the form w.γ for some w ∈ W and γ ∈ ∆, and
w can be written as a product of reflections sα for α ∈ ∆ by Proposition A.11,
the claim follows by induction.
Note that by Proposition A.1 this notion does not depend on the choice
of a W -invariant scalar product on E.
α α
3α + 2β
B2 G2
α+β α+β 2α + β 3α + β
β 2α + β β
α α
The roots α ∈ Φ of maximal length are called long roots. If there are two
distinct root lengths, then the shorter ones are called short roots.
We now introduce a convenient description of root systems. By Proposi-
tion A.11 any root system can be recovered from a base. The base is encoded
in terms of a directed graph which we now describe. Its nodes are in bijec-
tion with the elements of a base ∆, and two different nodes corresponding
to α, β ∈ ∆ are joined by an edge of multiplicity n(α, β)n(β, α). The result-
ing graph is called the Coxeter diagram associated to Φ or W . It does not
determine Φ uniquely. So in addition, whenever α, β are of different lengths
and joined by at least one edge, then we put an arrow on this edge, pointing
towards the shorter of the two. This directed graph is called the Dynkin di-
agram of Φ. It can be shown that the base can essentially be recovered from
its Dynkin diagram (that is, up to changing lengths in different connected
components of the diagram), see [33, 11.1].
The indecomposable root systems can be classified, see Theorem 9.6 and
Table 9.1 for their Dynkin diagrams. We’ll not repeat this proof here, since
it is well documented in the literature, see for example [9, VI, §4] or [33,
Thm. 11.4].
the number of positive roots made negative by w ∈ W . It will turn out that
this coincides with the length (w) introduced above. For this we need:
whence the result. On the other hand, if w.α ∈ Φ− then this shows that
N (wsα ) = N (w) \ {w.α}.
ΦI := Φ ∩ Z∆I
Proof Let v ∈ E and choose w ∈ W such that |{β ∈ Φ+ | (w.v, β) < 0}| is
minimal. We claim that then w.v ∈ C̄. Otherwise, (w.v, α) < 0 for some α ∈
∆, but then (sα w.v, α) = (w.v, sα .α) = −(w.v, α) > 0, and by Lemma A.8
this contradicts our choice of w. Hence the W -orbit of any v ∈ E has a
representative in C̄.
To finish the proof it is enough to show that when v1 , v2 ∈ C̄ with w.v1 =
v2 for some w ∈ W then v1 = v2 . We argue by induction on (w), the case
(w) = 0 being trivial. If (w) > 0 then by Proposition A.21 there is some
simple root α ∈ ∆ made negative by w. By Lemma A.20 this means that
n(wsα ) = n(w) − 1. Now v1 , v2 ∈ C̄, so
Proof Clearly, the pointwise stabilizer of M is the same as that of its linear
span, so we may replace M by a basis of the latter. Now let v ∈ M . We may
assume that v ∈ C̄ by Theorem A.27 and then H = CW (v) is a parabolic
subgroup by the above observation. Now, by induction on |M |, the centralizer
in H of M \ {v} is a parabolic subgroup of H, hence of W , as claimed.
For the converse, note that for I ⊆ S the fixed space M := V WI of the
280 Root systems
we call (dγ )γ∈∆\∆I the shape of β. Note that WI preserves the set of roots
of a given shape. Let’s note the following property:
(a) Any two roots in Φ+ \ ΦI of the same shape and the same length are
conjugate under WI .
(b) There is a unique root of minimal height of a given shape.
root of the same length as β and γ. But then sδ (β) = β−n(β, δ)δ = β−δ = γ,
with sδ ∈ WI , proving (a). In general, either n(β, γ) = 1 or n(γ, β) = 1,
whence β −γ ∈ Φ, a contradiction if β, γ have the same (minimal) height.
Exercises
A.1 Let Φ be a root system with base ∆. Show that Φ∨ is a root system
with base ∆∨ := {α∨ | α ∈ ∆}.
A.2 Let Φ be a root system with dual root system Φ∨ . Then the Weyl
groups of Φ and of Φ∨ are isomorphic via sα → sα∨ .
A.3 Let Φ ⊂ E be an indecomposable root system with Weyl group W and
fix a W -invariant scalar product on E. Show the following: There are
at most two W -orbits of roots in Φ, and any two roots of the same
length are conjugate.
[Hint: First argue that given α, β ∈ Φ we can always find w ∈ W such
that α, w.β are not orthogonal, then apply Proposition A.17 and Corol-
lary A.18(a).]
A.4 Let Φ be an indecomposable root system with two different root lengths
and let p = (α, α)/(β, β) for α, β ∈ Φ with α long and β short.
pα∨ α long,
(a) The map α → extends to a homothety.
α∨ α short,
(b) A root γ = α∈∆ cα α is long if and only if p|cα for all short α ∈ ∆.
A.5 Let Φ be a root system, α, β ∈ Φ with β = −α. Show that there exists
a base ∆ of Φ with α ∈ ∆ such that β ∈ Φ+ (with respect to ∆).
A.6 Let Φ be a root system in E with Weyl group W . Show that the only
reflections in W are of the form sα for α ∈ Φ.
[Hint: You may want to use Corollary A.29.]
A.7 Let Φ be a root system with base ∆, I S a subset of the correspond-
ing set of simple reflections, and S a shape with respect to the subset
I. Then there exists a unique root of maximal height of shape S.
[Hint: Use Lemma A.31, and show that if β is of maximal height and of
shape S, then wI (β) is of minimal height and of shape S, where wI is the
longest element of the parabolic subgroup WI .]
Appendix B
Subsystems
Proof Let v be a maximal element in the orbit W.v. We claim that v ∈ C̄.
Indeed, if (v , α) < 0 for some α ∈ ∆ then sα .v = v − 2(v , α)/(α, α)α >
v , contradicting the maximality of v . By Theorem A.27 this implies the
uniqueness of v .
Definition B.6 The root α0 above is called the highest root of Φ with
respect to ∆ (since, clearly it has the largest height (see Definition A.10) of
all roots in Φ).
Proof For (a) let α ∈ Φ. Since C̄ is a fundamental domain for the action
of W by Theorem A.27 we may assume that α ∈ C̄. Now α0 − α ≥ 0 by
Proposition B.5, so (α0 −α, v) ≥ 0 for all v ∈ C̄. In particular (α0 −α, α0 ) ≥ 0
and (α0 − α, α) ≥ 0, which gives (α0 , α0 ) ≥ (α0 , α) ≥ (α, α).
In (b) we have n(α, α0 ) ∈ {0, ±1} by Table A.1. But since α0 ∈ C̄ this
integer is non-negative.
1 1 1 1 1 2 2 2
An Bn >
n≥1 n≥2
1
2 2 2 1 1 2 2
Cn < Dn
n≥3 n≥4
1 2 3 2 1 1
E6
2 3 4 2 3 2
2 F4 > G2 <
2 3 4 3 2 1 2 4 6 5 4 3 2
E7 E8
2 3
tv : E → E, u → u + v.
Definition B.9 The affine Weyl group Wa of a root system Φ is the group
of affine transformations of E generated by all sα,j , for α ∈ Φ, j ∈ Z.
Let L := ZΦ∨ , the lattice generated by the coroots. Then we have the
following structure result for Wa :
Proposition B.10 The affine Weyl group is the semidirect product of the
translation subgroup corresponding to L with W .
The proposition shows that closed subsets satisfy (R1)–(R4), and therefore
are automatically subsystems.
Proposition B.16 The closed subsystems of Φ are precisely the sets Φ∩H
for H a subgroup of ZΦ of finite index.
2 4 6 5 4 3 2
E8 ......
−α0
3
Proof By Proposition A.11 any base ∆ of Φ contains short and long roots.
Since Φ is indecomposable, there is a short root α and a long root β in ∆
with (α, β) = 0. Then by Table A.2, Φ∩(Zα+Zβ) is of type B2 or G2 , and in
both cases, all long roots are integral linear combinations of the short ones.
Since W acts transitively on the set of long roots (see Corollary A.18(b))
this completes the proof.
Proof First note that Ψ is of the same rank as Φ, since otherwise by Corol-
lary A.29 it is contained in the proper (closed) parabolic subsystem of the
parabolic subgroup pointwise fixing Ψ⊥ . Since Ψ is not closed, it contains
some short root by Example B.15(3). So there is an indecomposable compo-
nent Ψ1 of Ψ containing short roots. If Ψ1 is proper in Ψ, then
Ψ = Ψ1 (Ψ ∩ Ψ⊥ ⊥
1 ) ⊆ (Φ ∩ ZΨ1 ) (Φ ∩ Ψ1 ),
Note that the conditions in Proposition B.21 are necessary, but not always
B.4 Other subsystems 291
Φ An Bn Cn Dn E6 E7 E8 F4 G2
Φl An Dn An1 Dn E6 E7 E8 D4 A2
Φs − An1 Dn − − − − D4 A2
|Φ| n(n + 1) 2n2 2n2 2n(n − 1) 72 126 240 48 12
Since duality of root systems (see Definition A.3) interchanges long and
short roots, we see from the classification that type Cn is dual to type Bn ,
and all other indecomposable root systems are self-dual.
(1) If Φ is indecomposable with only one root length, then any subsystem
is automatically closed, by Example B.15(3).
(2) Let Φ be of type Bn . The indecomposable proper parabolic subsystems
Ψ1 containing short roots are of types Br , 1 ≤ r < n, and then Ψ1
(Φ ∩ Ψ⊥1 ) are maximal subsystems of type Br Bn−r . As for the situation
in Proposition B.21(2), Φs is of type An1 by Table B.2, without proper
subsystems of maximal rank. The only indecomposable root system of
rank n having fewer roots than Bn is An , but this does not contain a
maximal rank subsystem of type An1 . Thus, the only maximal non-closed
subsystems of Bn are of type Br Bn−r , 1 ≤ r < n.
(3) Let Φ be of type Cn . This root system is dual to that of type Bn , so
its maximal subsystems can be obtained as duals of those in Bn . The
duals of parabolic subsystems are parabolic, hence closed. According
to Theorem B.18 the closed maximal rank subsystems in type Bn are
Dr Bn−r with 2 ≤ r ≤ n, with duals Dr Cn−r in Cn . Finally the dual
of the maximal non-closed subsystem Br Bn−r from part (2) is Cr Cn−r ,
which is closed. So here the maximal non-closed subsystems are of type
Dr Cn−r , 2 ≤ r ≤ n.
(4) Let Φ be of type F4 . The indecomposable proper parabolic subsys-
tems containing short roots are of types A1 , A2 , C3 , which by Propo-
sition B.21(1) lead to subsystems A1 B3 , A2 A2 , C3 A1 , only the first of
292 Subsystems
A1 , Bn , Cn , Dn (n even), E7 , E8 , F4 , or G2 .
Proof This follows from the above description of subsystems with Proposi-
tion A.30.
n(Φ) = max{ni | 1 ≤ i ≤ l}
B.5 Bad primes and torsion primes 293
denote the largest coefficient of the highest root of Φ. Clearly, this is inde-
pendent of the chosen base.
Proposition B.26 Let Φ be indecomposable, Ψ ⊆ Φ a closed indecompos-
able subsystem. Then n(Ψ) ≤ n(Φ).
Proof Let ∆1 be a base of Ψ. Extend this to a basis of E and choose an
ordering on E with respect to which this basis is positive. Let Φ+ be the
positive system of Φ with respect to this ordering, with base ∆. Then all
roots in ∆1 are positive in this ordering, hence all elements of ∆1 are non-
negative integral combinations over ∆. Substituting these combinations into
the expression for the highest root of Ψ shows the claim.
Lemma B.27 Let Φ be indecomposable with ∆ = {α1 , . . . , αl } labeled such
that α1 , . . . , αr is a sequence in ∆ of minimal length r ≥ 0 subject to
(1) nr = n(Φ) is maximal among the ni , 1 ≤ i ≤ l, and
(2) (αi , αi+1 ) = 0 for i = 0, . . . , r − 1.
Then for i = 1, . . . , r we have (αi , αi ) = (α0 , α0 ) and ni = i + 1.
Proof If nr = 1 then r = 0 and there is nothing to prove. Now assume
that nr > 1 and let 0 ≤ j < r. Taking the scalar product of αj∨ with
l
α0 = i=1 ni αi we obtain
"
l
2= (α0 , α0∨ ) = ni (αi , α0∨ ) ≥ n1 (α1 , α0∨ ) and
i=1
"
l "
j+1
0 ≤ (α0 , αj∨ ) = ni (αi , αj∨ ) ≤ ni (αi , αj∨ ) for j = 1, . . . , r − 1
i=1 i=j−1
This characterization of bad primes could also easily be verified “by inspec-
tion” of the various indecomposable root systems, see Table B.1. Steinberg
[74, §1] has given the above proof relying on Lemma B.27 which does not
use the classification.
We now come to the second type of primes:
In particular, if Φ only contains roots of one fixed length, then α0∨ is the
highest root of Φ∨ . Else, by definition β ∈ Φ is long if and only if β ∨ ∈ Φ∨
is short, so the preceding result can be paraphrased as saying that α0∨ is
the highest short root of Φ∨ . A list of highest short roots in the case where
two different root lengths occur can readily be derived from Table B.1, see
Table B.3.
We let n∨ (Φ) := max{n∨ i | 1 ≤ i ≤ l}, the largest coefficient of the highest
B.5 Bad primes and torsion primes 295
1 2 3 2 2 1
F4 > G2 <
short root in the dual root system. Then we have the following analogue of
Proposition B.26:
Proposition B.31 Let Φ be indecomposable, Ψ ⊆ Φ a closed indecompos-
able subsystem. Then n∨ (Ψ) ≤ n∨ (Φ).
Proof As in the proof of Proposition B.26 we may choose compatible bases
in Ψ and Φ. Then, by induction we may assume that Ψ is a maximal subsys-
tem of Φ. Now let α̃0 denote the highest root of Ψ and write α̃0 = i ñi α̃i ,
where {α̃1 , . . . , α̃r } is the chosen base of Ψ. Expressing α̃i = j mij αj in
the base {α1 , . . . , αl } of Φ with non-negative integral coefficients, we have
"%" &
α̃0 = ñi mij αj
j i
and thus i ñi mij ≤ nj for all j by Proposition B.5. Now sums of long roots
are long, by Example B.15(3), so for any short root α̃i there is some j with
αj short and mij = 0. Thus,
max{ñi | α̃i short} ≤ max{nj | αj short},
and similarly for long roots. Now if α̃0 has the same length as α0 (hence is
long) then by Lemma B.30 we find
n∨ (Ψ) = max{ñ∨ ∨ ∨
i } ≤ max{nj } = n (Φ).
Thus we may assume α̃0 is short, whence all roots in Ψ are short, and
among the possible maximal closed subsystems described in Theorem B.18
this can only happen when Ψ = Z(∆ \ {αi }) ∩ Φ for some i, with αi long
and all other simple roots short. Let s denote the reflection in α̃0 . Then β :=
s.αi = αi − (αi , α̃0∨ )α̃0 is long, and since α̃0 is a positive linear combination
in ∆ \ {αi }, (αi , α̃0∨ ) < 0 by Lemma A.6. So the coefficients of α̃0∨ are
dominated by those of β ∨ , whose coefficients in turn are dominated by those
of α0∨ , because both β, α0 are long.
296 Subsystems
Exercises
B.1 Let Φ be a root system and α ∈ Φ. Show that the affine transformation
sα,j is a reflection in the hyperplane Hα,j := {v ∈ E | (v, α) = j} along
α∨ and that sα,j = tjα∨ sα .
B.2 Let Φ be an indecomposable root system, Ψ a subsystem consisting
only of long roots. Then Ψ is closed in Φ.
[Hint: Use the classification of two-dimensional root systems in Proposi-
tion A.17.]
B.3 Consider the root system Φ of type E8 (see Example B.19). By Theo-
rem B.18, the subsystems Ψi generated by ∆ \ {αi } ∪ {−α0 } are not
maximal when ni is not prime. In each such case find a (base of a)
maximal subsystem containing Ψi .
B.4 Let Φ be a root system, S the set of simple reflections corresponding
to some base of Φ.
(a) Show that ZΦ/ZΦI has no torsion, for any I ⊆ S.
(b) Conclude that Φ of type An has no bad primes nor torsion primes.
(c) Determine the torsion primes for Φ of type Cn .
Appendix C
Automorphisms of root systems
(a) For each F -orbit I ⊆ S we have WIF = wI , where wI is the longest
element of the parabolic subgroup WI .
(b) The group of fixed points W F is generated by {wI | I ⊆ S an F -orbit}.
Now suppose that α, β ∈ Φ with π(β) = aπ(α) for some a > 0. By the
previous argument, w.α ∈ Φ+ I for some w ∈ W
F
and some F -orbit I ⊆ S.
So π(w.α) is a non-negative linear combination of roots in ∆I . However
1" 1" i
δ−1 δ−1
wI .π(α) = wI .φi (α) = φ (wI .α) = −π(α).
δ i=0 δ i=0
"
δ−1 "
δ−1
i
0= (φ (α), v) = (α, v),
i=0 i=0
Definition C.4 For each equivalence class ω for the equivalence relation ∼
defined in Lemma C.2 we let αω denote the vector of maximal length among
the {π(α) | α ∈ ω}. We set ΦF := {αω | ω ∈ Ω} and ∆F := {αω | ω ⊆ ∆} ⊆
ΦF .
Exercises
C.1 (Automorphisms of root systems) Let Φ be a root system in the Eu-
clidean space V and σ ∈ GL(V ) stabilizing a positive system Φ+ ⊂ Φ.
Then σ induces a graph automorphism of the Dynkin diagram of Φ.
[Hint: First check that σ stabilizes the base ∆ in Φ+ . Then use the clas-
sification of root systems of rank 2 to see that both the number of edges
between two roots α, β ∈ ∆, as well as the direction of the arrow, is already
determined by the set (Zα + Zβ) ∩ Φ.]
C.2 Verify the structure of W F and ΦF for the cases occurring in Table 23.1.
References
[1] M. Aschbacher, On the maximal subgroups of the finite classical groups. In-
vent. Math. 76 (1984), 469–514.
[2] M. Aschbacher, Finite Group Theory. Second edition. Cambridge Studies in
Advanced Mathematics, 10. Cambridge University Press, Cambridge, 2000.
[3] H. B. Azad, M. Barry, G. M. Seitz, On the structure of parabolic subgroups.
Comm. Algebra 18 (1990), 551–562.
[4] A. Borel, Linear Algebraic Groups. Second edition. Graduate Texts in Math-
ematics, 126. Springer-Verlag, New York, 1991.
[5] A. Borel, J. de Siebenthal, Les sous-groupes fermés de rang maximum des
groupes de Lie clos. Comment. Math. Helv. 23 (1949), 200–221.
[6] A. Borel, J. Tits, Groupes réductifs. Inst. Hautes Études Sci. Publ. Math. 27
(1965), 55–150.
[7] A. Borel, J. Tits, Eléments unipotents et sous-groupes paraboliques de groupes
réductifs. I. Invent. Math. 12 (1971), 95–104.
[8] A. V. Borovik, The structure of finite subgroups of simple algebraic groups.
(Russian) Algebra i Logika 28 (1989), 249–279, 366; translation in Algebra
and Logic 28 (1989), 163–182 (1990).
[9] N. Bourbaki, Groupes et Algèbres de Lie. IV, V, VI, Hermann, Paris, 1968.
[10] M. Broué, G. Malle, Théorèmes de Sylow génériques pour les groupes réductifs
sur les corps finis. Math. Ann. 292 (1992), 241–262.
[11] J. Brundan, Double coset density in classical algebraic groups. Trans. Amer.
Math. Soc. 352 (2000), 1405–1436.
[12] M. Cabanes, Unicité du sous-groupe abélien distingué maximal dans certains
sous-groupes de Sylow. C. R. Acad. Sci. Paris Sér. I Math. 318 (1994), 889–
894.
[13] R. W. Carter, Simple Groups of Lie Type. John Wiley & Sons, London, 1972.
[14] R. W. Carter, Finite Groups of Lie Type—Conjugacy Classes and Complex
Characters. John Wiley & Sons, New York, 1985.
[15] C. Chevalley, Classification des Groupes Algébriques Semi-simples. Collected
Works. Vol. 3. Springer-Verlag, Berlin, 2005.
[16] E. Cline, B. Parshall, L. Scott, On the tensor product theorem for algebraic
groups. J. Algebra 63 (1980), 264–267.
302 References