Rep N Control
Rep N Control
Rep N Control
for the
Pure Mathematician
by
Peter Webb
Preface
This book is a teaching tool intended for students who are studying finite group
representation theory beyond the level of a first course in character theory. It started as
notes for courses given at the second-year graduate level at the University of Minnesota
and it is appropriate for a year-long course at this level.
Most of the students who attend such a course do not go on to be specialists in group
representation theory (for otherwise the audience would be much smaller!) but instead
go on to pursue interests in other areas of pure mathematics, which might perhaps be
combinatorics, topology, number theory or commutative algebra. These students need a
solid, well-balanced grounding in representation theory that enables them to apply the
theory to their own situation when the occasion demands it. It is probably irrelevant for
them to be presented with overly deep material whose main function - at least at present
- is to serve the internal workings of the subject. Thus topics have been selected and in
some instances standard results have been omitted (sometimes just the proofs) where they
appear to me not to have applications. This book is written primarily for those students.
Their needs have provided the main criterion for whether or not to include certain material,
and I have also tried to aim the level of exposition at them. The explanations are intended
to be full, but not overly lengthy. For the students who go on to be specialists in the area
there is no shortage of more advanced texts - but they may find it helpful to start with
this one!
One feature of the treatment is that wherever possible we develop the subject over
an arbitrary ring. The philosophy behind this is that representations over many different
rings do arise in practice in pure mathematics, and we need to be able to handle them.
These rings might be finite fields arising in the context of homology groups of spaces, or
they might be local rings arising from questions in commutative algebra, or they might be
rings of algebraic integers. It is my intention that the reader should come away from this
book no less confident in characteristic p than in characteristic 0. The questions that arise
in positive characteristic or over rings of integers are genuinely harder than over fields of
characteristic zero, but at least the reader should come away with some idea of what can
be done and how to do it, and what cannot be done in this context.
The exercises at the ends of sections are an important part of the book. The benefit
of learning actively by having to apply the theory to calculate with examples and solve
problems cannot be underestimated. I have tried to provide a good number of exercises:
some of them are easy and some are more challenging. In some instances I use the exercises
as a place to present extensions of results that appear in the text. Sometimes the exercises
indicate how the subject may be developed further.
It is supposed that the reader has already studied the material in a first course in
algebra and is familiar with the properties of groups, rings, field extensions and with linear
algebra. More specifically the reader should know about Sylow subgroups, solvable and
nilpotent groups, as well as the examples that are introduced in a first group theory course,
such as the dihedral, symmetric, alternating and quaternion groups. The reader should
also be familiar with tensor products, Noetherian properties of commutative rings, the
structure of modules over a principal ideal domain, and the first properties of ideals, as
well as Jordan and rational canonical forms for matrices.
Many people have read sections of this book, worked through the exercises and made
important comments and I wish to thank them all. These include Cihan Bahran, Dave
Benson, Daniel Hess, John Palmieri, Sverre Smaland many others.
Table of Contents
1
2
3
4
5
6
7
8
9
10
11
12
A
B
a matrix representation of G. In this situation the rank of the free R-module V is called
the degree of the representation. Sometimes by abuse of terminology the module V is also
called the representation, but it should more properly be called the representation module
or representation space (if R is a field).
To illustrate some of the possibilities that may arise we consider some examples.
(1.1) Examples. 1. For any group G and commutative ring R we can take V = R and
(g) = 1 for all g G, where 1 denotes the identify map R R. This representation is
called the trivial representation, and it is often denoted simply by its representation module
R. Although this representation turns out to be extremely important in the theory, it does
not at this point give much insight into the nature of a representation.
2. A representation on a space V = R of rank 1 is in general determined by specifying
a homomorphism G R . Here R is the group of units of R, and it is isomorphic to
GL(V ). For example, if G = hgi is cyclic of order n and k = C is the field of complex
numbers, there are n possible such homomorphisms, determined by g 7 e
2ri
n
where
which assigns to each permutation its sign, regarded as an element of the arbitrary ring R.
3. Let R = R, V = R2 and G = S3 . This group G is isomorphic to the group of
symmetries of an equilateral triangle. The symmetries are the three reflections in the lines
that bisect the equilateral triangle, together with three rotations.
Positioning the triangle at the origin of V and labelling the three vertices of the triangle
Peter Webb
1 0
0 1
1
1
0 1
1 0
0
1
1 1
0 1
1
1
0 1
1 1
1
0
where we have taken basis vectors in the directions of vertices 1 and 2, making an angle of
2
3 to each other. In fact these matrices define a representation of degree 2 over any ring
R, because although the representation was initially constructed over R the matrices have
integer entries, and these may be interpreted in every ring. No matter what the ring is,
the matrices always multiply together to give a copy of S3 .
4. Let R = Fp , V = R2 and let G = Cp = hgi be cyclic of order p generated by an
element g. We see that the assignment
r
(g ) =
1 0
r 1
is a representation. In this case the fact that we have a representation is very much
dependent on the choice of R as the field Fp : in any other characteristic it would not work,
because the matrix shown would no longer have order p.
We can think of representations in various ways. One of them is that a representation
is the specification of an action of a group on an R-module, as we now explain. Given
a representation : G GL(V ), an element v V and a group element g G we get
another module element (g)(v). Sometimes we write just g v or gv for this element. This
rule for multiplication satisfies
g (v + w) = g v + g w
(gh) v = g (h v)
1v =v
gG
ag g)(
hG
bh h) =
X X
(
ag bh )k.
kG gh=k
More concretely, we may exemplify the definition by listing some elements of QS3 . The
elements of S3 such as (1, 2) = 1 (1, 2) are also elements of QS3 (they appear as basis
elements), and () serves as the identity element of QS3 (as well as of S3 ). In general,
elements of QS3 may look like (1, 2) (2, 3) or 51 (1, 2, 3) + 6(1, 2) 71 (2, 3). Here is a
computation:
((1, 2, 3) + (1, 2))((1, 2) (2, 3)) = (1, 3) + () (1, 2) (1, 2, 3).
Having defined the group algebra, we may now define a representation of G over
R to be a unital RG-module. The fact that this definition coincides with the previous
ones is the content of the next proposition. Throughout this text we may refer to group
representations as modules (for the group algebra).
(1.2) PROPOSITION. A representation of G over R has the structure of a unital
RG-module; conversely, every unital RG-module provides a representation of G over R.
Proof. Given a representation : G GL(V ) we define a module action of RG on V
P
P
by ( ag g)v =
ag (g)(v).
Given a RG-module V , the linear map (g) : v 7 gv is an automorphism of V and
(g1 )(g2 ) = (g1 g2 ) so : G GL(V ) is a representation.
We have defined the group algebra without saying what an algebra is! For the record,
an (associative) R-algebra is a ring A with a 1, equipped with a (unital) ring homomorphism
R A whose image lies in the center of A. The group algebra RG is indeed an example
of an R-algebra.
The group algebra gives another example of a representation, called the regular representation. In fact for any ring A we may regard A itself as a left A-module with the
action of A on itself given by multiplication of the elements. We denote this left A-module
by A A when we wish to emphasize the module structure, and this is the (left) regular
Peter Webb
Semisimple representations
We come now to our first non-trivial result, and one that is fundamental to the study
of representations over fields of characteristic zero, or characteristic not dividing the group
order. The result says that in this situation representations fall apart remarkably easily,
so that they can be understood by taking direct sums of smaller representations. We do
now require the ring R to be a field, and in this situation we will often use the symbols F
or k instead of R.
(1.4) THEOREM (Maschke). Let V be a representation of the finite group G over a
field F in which |G| is invertible. Let W be an invariant subspace of V . Then there exists
an invariant subspace W1 of V such that V = W W1 as representations.
Proof. Let : V W be any projection of V onto W as vector spaces, i.e. a linear
transformation such that (w) = w for all w W . Since F is a field, we may always
find such a projection by finding a vector space complement to W in V , and projecting
off the complementary factor. Then V = W Ker() as vector spaces, but Ker() is not
necessarily invariant under G. Consider the map
=
1 X
gg 1 : V V.
|G|
gG
1 X
g(g 1w)
|G|
gG
1 X 1
gg w
|G|
gG
1
|G|w
|G|
= w.
Peter Webb
1 X
g(g 1hv)
|G|
gG
1 X
h(h1 g)((h1 g)1 v)
|G|
gG
= h (v)
Peter Webb
10
Summary of Chapter 1
1. Representations of G over R are the same thing as RG-modules.
2. Semisimple modules may be characterized in several different ways. They are
modules that are the direct sum of simple modules, or equivalently the sum of simple
modules, or equivalently modules for which every submodule is a direct summand.
3. If F is a field in which G is invertible, F G-modules are semisimple.
4. The sum of all simple submodules of a module is the unique largest semisimple
submodule of that module: the socle.
Peter Webb
11
Assuming this, show that for each finite group G and each integer n there are only finitely
many isomorphism classes of representations of G on Zn .
6. (a) Write out a proof of Maschkes theorem in the case of representations over C
along the following lines.
Given a representation : G GL(V ) where V is a vector space over C, let ( , ) be any
positive definite Hermitian form on V . Define a new form ( , )1 on V by
(v, w)1 =
1 X
(gv, gw).
|G|
gG
Show that ( , )1 is a positive definite Hermitian form, preserved under the action of G,
i.e. (v, w)1 = (gv, gw)1 always.
If W is a subrepresentation of V , show that V = W W as representations.
(b) Show that any finite subgroup of GL(n, C) is conjugate to a subgroup of U (n, C)
(the unitary group, consisting of n n complex matrices A satisfying AAT = I). Show
that any finite subgroup of GL(n, R) is conjugate to a subgroup of O(n, R) (the orthogonal
group consisting of n n real matrices A satisfying AAT = I).
7. Let U = S1 Sr be an A-module that is the direct sum of finitely many simple
modules S1 , . . . , Sr . Show that if T is any simple submodule of U then T
= Si for some i.
8. Let : U V be a homomorphism of A-modules. Show that : (Soc U ) Soc V ,
and that if is an isomorphism
then restricts to an isomorphism Soc U Soc V .
2
2
9. Let G = hx, y x = y = 1 = [x, y]i be the Klein four-group, R = F2 , and consider
the two representations 1 and 2 specified on the generators of G by
1 1 0
1 0 1
1 (x) = 0 1 0 , 1 (y) = 0 1 0
0 0 1
0 0 1
and
2 (x) = 0
0
0 0
1 1,
0 1
2 (y) = 0
0
0 1
1 0.
0 1
1 1 0
1 1
1 (x) = 0 1 1
and 2 (x) = 0 1
0 0 1
0 0
neither representation is
3. Consider the two
1
0.
1
Calculate the socles of these two representations and show that neither representation is
semisimple. Show that the second representation is nevertheless the direct sum of two
non-zero subrepresentations.
12
11. (a) Using 1.6 show that if A is a ring for which the regular representation A A is
semisimple, then every finitely generated A-module is semisimple.
(b) Extend the result of part (a), using Zorns lemma, to show that if A is a ring for
which the regular representation A A is semisimple, then every A-module is semisimple.
12. Let U be a module for a ring A with a 1. Show that the following three statements
are equivalent.
(1) U is a direct sum of simple A-submodules.
(2) U is a sum of simple A-submodules.
(3) every submodule of U is a direct summand of U .
[Use Zorns lemma to prove a version of Lemma 1.5 that has no finiteness hypothesis and
then copy Proposition 1.6. This deals with all implications except (3) (2). For that, use
the fact that A has a 1 and hence every (left) ideal is contained in a maximal (left) ideal,
combined with condition (3), to show that every submodule of U has a simple submodule.
Consider the sum of all simple submodules of U and show that it equals U .]
13. Let RG be the group algebra of a finite group G over a commutative ring R with
1. Let S be a simple RG-module and let I be the anihilator in R of S, that is
I = {r R rx = 0 for all x S}.
(y) =
0
1
Peter Webb
13
14
(a 7 ax) x.
There are several things here that need to be checked: that the second assignment does
take values in EndA (A A), that the morphisms are ring homomorphisms, and that they are
mutually inverse. We leave most of this to the reader, observing only that under the first
homomorphism a composite is sent to ()(1) = ((1)) = ((1)1) = (1)(1), so that
this is indeed a homomorphism to Aop .
Observe that the proof of Lemma 2.2 establishes that every endomorphism of the
regular representation is of the form right multiplication by some element.
A ring A with 1 all of whose modules are semisimple is itself called semisimple. By
Exercise 11 of Chapter 1 it is equivalent to suppose that the regular representation A A is
semisimple. It is also equivalent if A is a finite dimensional algebra over a field to suppose
that the Jacobson radical of the ring is zero, but we will not deal with this point of view
until Chapter 6.
(2.3) THEOREM (ArtinWedderburn). Let A be a finite-dimensional algebra over
a field k with the property that every finite-dimensional module is semisimple. Then A is
a direct sum of matrix algebras over division rings. Specifically, if
AA
= S1n1 Srnr
Peter Webb
15
= S1n1 Srnr
and so dim Si = ni .
16
Let us now restate what we have proved specifically in the context of group representations.
(2.5) COROLLARY. Let G be a finite group and k a field in which |G| is invertible.
(1) As a ring, kG is a direct sum of matrix algebras over division rings.
(2) Suppose in addition that k is algebraically closed. Let S1 , . . . , Sr be pairwise nonisomorphic simple kG-modules and let di = dimk Si be the degree of Si . Then di
equals the multiplicity with which Si is a summand of the regular representation of
G, and |G| d21 + + d2r with equality if and only if S1 , . . . , Sr is a complete set of
representatives of the simple kG-modules.
Proof. This follows from Maschkes Theorem 1.4, the ArtinWedderburn Theorem
2.3 and Corollary 2.4.
Part (2) of this result provides a numerical criterion that enables us to say when we
have constructed all the simple modules of a group over an algebraically closed field k in
P
which |G| is invertible: we check that d2i = |G|. While this is an easy condition to verify,
it will be superseded later on by the even more straightforward criterion that the number
of simple kG-modules (with the same hypotheses on k) equals the number of conjugacy
P 2
classes of elements of G. Once we have proved this, the formula
di = |G| allows the
degree of the last simple representation to be determined once the others are known.
(2.6) Example. Over R we have constructed for S3 the trivial representation, the sign
representation that is also of dimension 1 but not the same as the trivial representation, and
a 2-dimensional representation that is simple because visibly no 1-dimensional subspace of
the plane is invariant under the group action. Since 12 +12 +22 = |S3 | we have constructed
all the simple representations.
At this point we make a deduction about representations of finite abelian groups.
Looking ahead to later results, we will obtain a partial converse of this in Theorem 4.4,
and in Theorem 5.10 we will obtain an extension to fields in which |G| is not invertible.
A more detailed description of representations of abelian groups when |G| is not invertible
follows from Example 8.6.
(2.7) COROLLARY. Let G be a finite abelian group. Over an algebraically closed
field k in which |G| is invertible, every simple representation of G has degree 1 and the
number of non-isomorphic simple representations equals |G|. In particular we may deduce
that every invertible matrix of finite order, with order relatively prime to the characteristic
of k, is diagonalizable.
Proof. We know that kG is semisimple, and because kG is a commutative ring the
matrix summands that appear in Theorem 2.3 must all have size 1, and the division rings
that appear must be commutative. In fact, since we have supposed that k is algebraically
closed, the division rings must all be k. This means that the degrees of the irreducible
Peter Webb
17
representations are all 1, and so the number of them must be |G| since this is the sum of
the degrees.
A matrix of finite order gives a representation of the cyclic group it generates on the
space of vectors on which the matrix acts and, by invertibility of the order, the representation is semisimple. It is a direct sum of 1-dimensional spaces by what we have just shown.
On choosing basis vectors to lie in these 1-dimensional spaces, the matrix is diagonal.
We do not have to exploit the theory we have developed to show that a matrix of finite,
invertible order is diagonalizable over an algebraically closed field. A different approach
is to consider its Jordan canonical form and observe that all Jordan blocks have size 1,
because blocks of size 2 or more have order that is either infinite or a multiple of the field
characteristic.
Summary of Chapter 2
1. Endomorphism algebras of simple modules are division rings.
2. Semisimple algebras are direct sums of matrix algebras over division rings.
3. For a semisimple algebra over an algebraically closed field, the sum of the squares
of the degrees of the simple modules equals the dimension of the algebra.
18
is semisimple.
(d) Show that Mn (D) is a simple ring, namely one in which the only 2-sided ideals
are the zero ideal and the whole ring.
4. Prove the following extension of Corollary 2.4:
THEOREM. Let A be a finite-dimensional semisimple algebra, S a simple A-module
and D = EndA (S). Then S may be regarded as a module over D and the multiplicity of
S as a summand of A A equals dimD S.
5. Using the fact that Mn (k) has a unique simple module up to isomorphism, prove
the Noether-Skolem theorem: every algebra automorphism of Mn (k) is inner, i.e. of the
form conjugation by some invertible matrix.
6. Show that for any field k we have Mn (k)
= Mn (k)op , and in general for any division
ring D that given any positive integer n, Mn (D)
= Dop .
= Mn (D)op if and only if D
7. Let A be a ring with a 1, and let V be an A-module. An element e in any ring is
called idempotent if and only if e2 = e.
(a) Show that an endomorphism e : V V is a projection onto a subspace W if and
only if e is idempotent as an element of EndA (V ). (The term projection was defined at
the start of the proof of 1.4. It is a linear mapping onto a subspace that is the identity on
restriction to that subspace.)
(b) Show that direct sum decompositions V = W1 W2 as A-modules are in bijection
with expressions 1 = e + f in EndA (V ), where e and f are idempotent elements with
ef = f e = 0. (In case ef = f e = 0, e and f are called orthogonal.)
(c) A non-zero idempotent element e is called primitive if it cannot be expressed as a
sum of orthogonal idempotent elements in a non-trivial way. Show that e EndA (V ) is
primitive if and only if e(V ) has no (non-trivial) direct sum decomposition. (In this case
e(V ) is said to be indecomposable.)
(d) Suppose that V is semisimple with finitely many simple summands and let e1 , e2
EndA (V ) be idempotent elements. Show that e1 (V )
= e2 (V ) as A-modules if and only if
e1 and e2 are conjugate by an invertible element of EndA (V ) (i.e. there exists an invertible
A-endomorphism : V V such that e2 = e1 1 ).
(e) Let k be a field. Show that all primitive idempotent elements in Mn (k) are
conjugate under the action of the unit group GLn (k). Write down explicitly any primitive
idempotent element in M3 (k). (It may help to use Exercise 3.)
8. Prove the following theorem of Burnside: let G be a finite group, k an algebraically
closed field in which |G| is invertible, and let : G GL(V ) be a representation over
k. By taking a basis of V write each endomorphism (g) as a matrix. Let dim V = n.
Show that the representation is simple if and only if there exist n2 elements g1 , . . . , gn2 of
G so that the matrices (g1 ), . . . , (gn2 ) are linearly independent, and that this happens if
and only if the algebra homomorphism kG Endk (V ) is surjective. (Note that itself is
generally not surjective.)
Peter Webb
19
9. (We exploit results from a basic algebra course in our suggested approach to this
question.) Let G be a cyclic group of order n and k a field.
(a) By considering a homomorphism k[X] kG or otherwise, where k[X] is a polynomial ring, show that kG
= k[X]/(X n 1) as rings.
(b) Suppose that the characteristic of k does not divide n. Use the Chinese Remainder
Theorem and separability of X n 1 to show that when kG is expressed as a direct sum of
irreducible representations, no two of the summands are isomorphic, and that their degrees
are the same as the degrees of the irreducible factors of X n 1 in k[X]. Deduce, as a special
case of Corollary 2.7, that when k is algebraically closed all irreducible representations of
G have degree 1.
(c) When n is prime and k = Q, use irreducibility of X n1 + X n2 + X + 1 to
show that G has a simple module S of degree n 1, and that EndkG (S)
= Q(e2i/n ).
n1
(d) When k = R and n is odd show that G has 2 simple representations of degree
2 as well as the trivial representation of degree 1. When k = R and n is even show that G
has n2
simple representations of degree 2 as well as two simple representations of degree
2
1. If S is one of the simple representations of degree 2 show that EndkG (S) = C.
10. Let U be a module for a semisimple finite dimensional algebra A. Show that if
EndA (U ) is a division ring then U is simple.
11. Let H be the algebra of quaternions, that has a basis over R consisting of elements
1, i, j, k and multiplication determined by the relations
i2 = j 2 = k 2 = 1, ij = k, jk = i, ki = j, ji = k, kj = i, ik = j.
You may assume that H is a division ring. The elements {1, i, j, k} under multiplication form the quaternion group Q8 of order 8, and it acts on H by left multiplication, so
that H is a 4-dimensional representation of Q8 over R.
(a) Show that EndRQ8 (H)
= H, and that H is simple as a representation of Q8 over
R. [Consider the image of 1 H under an endomorphism.]
Lt
(b) In the decomposition RQ8 = i=1 Mni (Di ) predicted by Corollary 2.5, compute
the number of summands t, the numbers ni and the divisions rings Di . Show that RQ8
has no simple representation of dimension 2. [Observe that there are four homomorphisms
Q8 {1} R that give four 1-dimensional representations. Show that, together with
the representation of dimension 4, we have a complete set of simple representations.]
(c) The span over R of the elements 1, i H is a copy of the field of complex numbers
C, so that H contains C as a subfield. We may regard H as a vector space over C by letting
elements of C act as scalars on H by multiplication from the right. Show that with the
action of Q8 from the left and of C from the right, H becomes a left CQ8 -module. With
respect to the basis {1, j} for H over C, write down matrices for the action of the elements
i, j Q8 on H. Show that this 2-dimensional CQ8 -module is simple, and compute its
endomorphism ring EndCQ8 (H).
(d) Show that C R H
= M2 (C).
12. Let k be a field of characteristic 0 and suppose the simple kG-modules are
Pr
2
S1 , . . . , Sr with degrees di = dimk Si . Show that
i=1 di |G| with equality if and
only if EndkG (Si ) = k for all i.
20
3. Characters
Characters are an extremely important tool for handling the simple representations
of a group. In this chapter we will see them in the form that applies to representations
over a field of characteristic zero, and these are called ordinary characters. Since representations of finite groups in characteristic zero are semisimple, knowing about the simple
representations in some sense tells us about all representations. Later, in Chapter 10, we
will study characters associated to representations in positive characteristic, the so-called
Brauer characters.
Characters are very useful when we have some specific representation and wish to
compute its decomposition as a direct sum of simple representations. The information we
need to do this is contained in the character table of the group, which we introduce in
this chapter. We also establish many important theoretical properties of characters that
enable us to calculate them more easily and to check that our calculations are correct.
The most spectacular of these properties is the orthogonality relations, which may serve
to convince the reader that something extraordinary and fundamental is being studied.
We establish numerical properties of the character degrees, and a description of the center
of the group algebra that aids in decomposition the group algebra as a sum of matrix
algebras. This would be of little significance unless we could use characters to prove
something outside their own area. Aside from their use as a computational tool, we use
them to prove Burnsides pa q b theorem: every group whose order is divisible by only two
primes is solvable.
Peter Webb
21
22
Peter Webb
23
(3.2) Example. We present the character table of S3 . We saw at the end of Chapter
2 that we already have a complete list of the simple modules for S3 , and the values of
their characters on representatives of the conjugacy classes of S3 are computed from the
matrices that give these representations.
S3
ordinary characters
g
|CG (g)|
() (12) (123)
6
2
3
1
sign
2
1
1
1 1
2
0
1
1
1
24
a b
c d
e
g
f
h
ae
ag
=
ce
cg
af
ah
cf
ch
be
bg
de
dg
bf
bh
.
df
dh
f v 7 (u 7 f (u) v),
this being the specification on basic tensors. We show that if either M or N is free as an
R-module of finite rank then is an isomorphism.
Peter Webb
25
= (v 7 g(f (g 1v)w))
= g(v 7 f (v)w)
= g(f w).
Finally the last formula for characters follows from parts (2) and (3)
Of course, if R is taken to be a field in part (4) of 3.3 then the argument can be
simplified. Since M and N are always free, only one of the two arguments given is needed,
and after showing either that is surjective or injective, the other follows by observing
that the dimensions on the two sides are equal.
26
The next result is an abstraction of the idea that was used in proving Maschkes
theorem, where the application was to the RG-module HomR (V, V ). We will use this
idea a second time in proving the orthogonality relations for characters. The projection
operator about to be described appears throughout representation theory.
(3.5) LEMMA. Let V be an RG-module where R is a ring in which |G| is invertible.
Then
1 X
g :V VG
|G|
gG
1 X
gh)v
|G|
gG
= (v)
1 X
=(
hg)v
|G|
gG
= h(v)
since as g ranges through the elements of G so do gh and hg. The same equations show
that every vector of the form (v) is fixed by G. Furthermore, if v V G then
(v) =
1 X
1 X
gv =
v=v
|G|
|G|
gG
gG
Peter Webb
27
to C, a set of functions that we may denote Ccc(G) where cc(G) is the set of conjugacy classes
of G. These functions become an algebra when we define addition, multiplication and scalar
multiplication pointwise on the values of the function. In other words, ()(g) = (g)(g),
( + )(g) = (g) + (g) and ()(g) = (g) where , are class functions and C. If
G has n conjugacy classes, this algebra is isomorphic to Cn , the direct sum of n copies of
C, and is semisimple. We have seen in Proposition 3.1 that characters of representations
of G are examples of class functions on G.
We define a Hermitian form on the complex vector space of class functions on G by
means of the formula
1 X
h, i =
(g)(g).
|G|
gG
Since the functions and are constant on conjugacy classes, this can also be written
h, i =
=
1
|G|
g[cc(G)]
g[cc(G)]
|G|
(g)(g)
|CG (g)|
1
(g)(g)
|CG (g)|
where [cc(G)] denotes a set of representatives of the conjugacy classes of G, since the
number of conjugates of g is |G : CG (g)|. As well as the usual identities that express
bilinearity and the fact that the form is evidently Hermitian, it satisfies
h, i = h, i
where (g) = (g) is the class function obtained by complex conjugation. If and
happen to be characters of a representation we have (g) = (g 1 ), is the character of
the contragredient representation, and we obtain further expressions for the bilinear form:
h, i =
1 X
(g 1 )(g)
|G|
gG
1 X
(g)(g 1)
=
|G|
gG
1 X
(g)(g),
|G|
gG
where the second equality is obtained by observing that as g ranges over the elements of
G, so does g 1 .
With all this preparation we now present the orthogonality relations for the rows of
the character table. The picture will be completed once we have shown that the character
table is square and deduced the orthogonality relations for columns in 3.16 and 3.17.
28
1 X
g) in its action on HomC (V, W )
|G|
gG
1 X
(g)(g)
=
|G|
gG
= h, i.
1,
Peter Webb
29
0 1.
Thus
1
(4 1 + 0 + 2 1 1) = 1
6
1
hV V , i = (4 1 + 0 + 2 1 1) = 1
6
1
hV V , V i = (4 2 + 0 2 1 1) = 1
6
hV V , C i =
Pr
Proof. We may write V
= S1n1 Srnr and then h, i = i=1 n2i is a positive
integer, which equals 1 precisely if one ni is 1 and the others are 0.
30
1
sign
2
3a
3b
1
1
1 1
2
0
3 1
3
1
1
1
2
1
1
1
1
0
1
1
1
1
1
0
0
cos sin 0
sin cos 0
0
0
1
Peter Webb
31
where is the angle of rotation, and conjugate matrices have the same trace. Thus, for
example, (12) and (123) must act as rotations through and 2
3 , respectively, so act via
matrices that are conjugates of
1
2 23 0
1 0 0
0 1 0
3
and
12 0
2
0 0 1
0
0
1
4 2
0 0 1
and we compute
2
1
4
+ + 0 + 0 + = 1.
24 4
3
Thus = 1 + where is the character of a 3-dimensional representation:
h, 1i =
1 0.
Again we have
9
1 1 1
+ + + +0 =1
24 4 8 4
so is simple by 3.10, and this is the bottom row of the character table.
There are other ways to complete the calculation of the character table. Having
computed four of the five rows, the fifth is determined by the facts that it is orthogonal to
the other four, and that the sum of the squares of the degrees of the characters equals 24.
Equally we could have constructed the bottom row as where is the other character
of degree 3 and is the sign character.
h, i =
Our next immediate goal is to prove that the character table of a finite group is square,
and to deduce the column orthogonality relations. Before doing this we show in the next
two results how part of the column orthogonality relations may be derived in a direct way.
Consider the regular representation of G on CG, and let CG denote the character of
this representation.
(3.12) LEMMA.
CG (g) =
|G|
0
if g = 1
otherwise.
32
We may deduce an alternative proof of 2.5 (in case k = C), and also a way to do the
computation of the final row of the character table once the others have been determined.
(3.13) COROLLARY. Let 1 , . . . , r be the simple complex characters of G, with
degrees d1 , . . . , dr . Then hCG , i i = di , and hence
Pr
(1)
d2i = |G|, and
Pi=1
r
(2)
i=1 di i (g) = 0 if g 6= 1.
Proof. Direct evaluation gives
hCG , i i =
1
|G|i (1) = di
|G|
R}
= R.
Proof. (1) Let Eij denote the matrix which is 1 in place i, j and 0 elsewhere. If
X = (xij ) is any matrix then
Eij X = the matrix with row j of X moved to row i, 0 elsewhere,
XEij = the matrix with column i of X moved to column j, 0 elsewhere.
If X Z(Mn (R)) these two are equal, and we deduce that xii = xjj and all other entries
in row j and column i are 0. Therefore X = x11 I.
(2) In 2.3 we constructed an isomorphism
CG
= Mn1 (C) Mnr (C)
where the matrix summands are in bijection with the isomorphism classes of simple modules. On taking centers, each matrix summand contributes 1 to dim Z(CG).
Peter Webb
33
yxi
since as y runs through the elements of G conjugate to xi , so does gyg 1, and from this it
follows that xi is central.
P
Next suppose gG ag g Z(RG). We show that if g1 g2 then ag1 = ag2 . Suppose
P
P
that g2 = hg1 h1 . The coefficient of g2 in h( gG ag g)h1 is ag1 and in ( gG ag g) is
ag2 . Since elements of G are independent in RG, these coefficients must be equal. From
this we see that every element of Z(RG) can be expressed as an R-linear combination of
the xi .
Finally we observe that the xi are independent over R, since each is a sum of group
elements with support disjoint from the supports of the other xj .
(3.16) THEOREM. The number of simple complex characters of G equals the number
of conjugacy classes of elements of G.
Proof. In 3.14 we showed that the number of simple characters equals the dimension
of the center, and 3.15 we showed that this is equal to the number of conjugacy classes.
We conclude that the character table of a finite group is always square. From this we
get orthogonality relations between the columns of the character table.
(3.17) COROLLARY (Column orthogonality relations). Let X be the character table
of G as a matrix, and let
|C (x )|
0
G
C=
0
..
.
|CG (x2 )|
..
.
.
|CG (xr )|
..
34
The special case of this in which g = 1 has already been seen in Corollary 3.13.
(1)
(2)
(3)
(4)
Peter Webb
35
Since b divides all terms in this equation except perhaps the term an , b must also be a
factor of an . Since a and b are coprime, this is only possible if b = 1, and we deduce that
a
b Z.
(4) (g) is the sum of the eigenvalues of g in its action on the representation which
affords . Since g n = 1 for some n these eigenvalues are all roots of X n 1 and so are
integers.
In the next result we identify Z with the subring Z 1, which is contained in the centre
Z(ZG) of ZG.
(3.19) PROPOSITION. The center Z(ZG) is integral over Z. Hence if x1 , . . . , xr are
representatives of the conjugacy classes of G, xi ZG is the sum of the elements conjugate
Pr
to xi , and 1 , . . . , r C are algebraic integers then the element i=1 i xi Z(CG) is
integral over Z.
Proof. It is the case that every commutative subring of ZG is integral over Z, using
condition 1(b) of 3.18, since such a subring is in particular a subgroup of the finitelygenerated free abelian group ZG, and hence is finitely generated as a Z-module.
We have seen in 3.15 that the elements x1 , . . . , xr lie in Z(ZG), so they are integral
Pr
over Z, and by part (2) of 3.18 the linear combination i=1 i xi is integral also. (We note
that the xi are in fact a finite set of generators for Z(ZG) as an abelian group, but we did
not need to know this for the proof.)
36
homomorphism
i : CG =
r
M
i=1
projecting onto the ith matrix summand. The fact that the group homomorphism i
extends to an algebra homomorphism in this way comes formally from the construction
of the group algebra. The fact that this algebra homomorphism is projection onto the ith
summand arises from the way we decomposed CG as a sum of matrix algebras, in which
each matrix summand acts on the corresponding simple module as matrices on the space
of column vectors.
(3.20) PROPOSITION. Fixing the suffix i, if x Z(CG) then i (x) = I for some
C. In fact
i (x) =
1
tr(i (x)) I
di
1 X
ag i (g) I.
di
gG
ag g we have
gG
Proof. Since x is central the matrix i (x) commutes with the matrices i (g) for all
g G. Therefore by Schurs lemma, since i is a simple complex representation, i (x) = I
1
di tr(I).
the right hand side and multiplying both sides by I gives the first expression for i (x).
Replacing x by the expression in the statement of the proposition we obtain
=
X
1
tr(i (
ag g))
di
gG
1 X
ag tr(i (g))
=
di
gG
1 X
ag i (g)
di
gG
Peter Webb
37
1 X
i (g 1 )i (g) I
di
gG
|G|
I
di
the second equality arising from the fact that hi , i i = 1. Now x is integral over Z1 by 3.19
since the coefficients i (g 1 ) are algebraic integers, so i (x) is integral over i (Z 1) = Z I.
|G|
|G|.
is
integral
over
Z
and
hence
Z.
We
deduce
that
d
Thus |G|
i
di
di
The matrix summands of the complex group algebra
Given a set of rings with identity A1 , . . . , Ar we may form their direct sum A = A1
Ar , and this itself becomes a ring with componentwise addition and multiplication.
In this situation each ring Ai may be identified as the subset of A consisting of elements
that are zero except in component i, but this subset is not a subring of A because it does
not contain the identity element of A. It is, however, a 2-sided ideal. Equally, in any
decomposition of a ring A as a direct sum of 2-sided ideals, these ideals have the structure
of rings with identity.
Decompositions of a ring as direct sums of other rings are closely related to idempotent
elements in the center of the ring. We have seen idempotent elements introduced in the
exercises to Chapter 2, and they will retain importance throughout this book. An element
e of a ring A is said to be idempotent if e2 = e. It is a central idempotent element if it
lies in the center Z(A). Two idempotent elements e and f are orthogonal if ef = f e = 0.
An idempotent element e is called primitive if whenever e = e1 + e2 where e1 and e2 are
orthogonal idempotent elements then either e1 = 0 or e2 = 0. We say that e is a primitive
central idempotent element if it is primitive as an idempotent element in Z(A), that is,
e is central and has no proper decomposition as a sum of orthogonal central idempotent
elements.
We comment that the term idempotent element is very often abbreviated to idempotent, thereby elevating the adjective to the status of a noun. This does have justification,
in that idempotent element is unwieldy, and we will usually conform to the shorter usage.
38
as a sum of orthogonal central idempotent elements, in such a way that ei is the identity
element of Ai and Ai = Aei . The Ai are indecomposable as rings if and only if the ei
are primitive central idempotent elements. If every Ai is indecomposable as a ring then
the subsets Ai and also the primitive central idempotents ei are uniquely determined;
furthermore, every central idempotent can be written as a sum of certain of the ei .
Proof. Given any ring decomposition A = A1 Ar we may write 1 = e1 + + er
where ei Ai and it is clear that the ei are orthogonal central idempotent elements.
Conversely, given an expression 1 = e1 + + er where the ei are orthogonal central
idempotent elements we have A = Ae1 Aer as rings.
To say that the ring Ai is indecomposable means that it cannot be expressed as a
direct sum of rings, except in the trivial way, and evidently this happens precisely if the
corresponding idempotent element cannot be decomposed as a sum of orthogonal central
idempotent elements.
We now demonstrate the (perhaps surprising) fact that there is at most one decomposition of A as a sum of indecomposable rings. Suppose we have two such decompositions,
and that the corresponding primitive central idempotent elements are labelled ei and fj ,
so that
1 = e1 + + er = f1 + + fs .
We have
ei = ei 1 =
s
X
ei fj ,
j=1
k=1
Peter Webb
39
Notice that it follows from Proposition 3.22 that distinct primitive central idempotents
must necessarily be orthogonal, a conclusion that is false without the word central. The
primitive central idempotents, and also the indecomposable ring summands to which they
correspond according to Proposition 3.22, are known as blocks. We will study blocks in
detail in Chapter 12, and also at the end of Chapter 9. In the case of the complex group
algebra CG, that is a direct sum of matrix rings, the block idempotents are the elements
that are the identity in one matrix summand and zero in the others. We shall now give a
formula for these block idempotents. It is the same (up to a scalar) as the formula used
in the proof of Theorem 3.21.
(3.23) THEOREM. Let 1 , . . . , r be the simple complex characters of G with degrees
d1 , . . . , dr . The primitive central idempotent elements in CG are the elements
di X
i (g 1 )g
|G|
gG
di X
di X
i (g 1 )g) =
i (g 1 )j (g) I
|G|
|G|dj
gG
gG
di
hi , j i I
dj
di
= i,j I
dj
=
= i,j I,
so that the elements specified in the statement of the theorem do indeed project correctly
onto the identity matrices, and are therefore the primitive central idempotent elements.
40
While the identity matrix is a primitive central idempotent element in the matrix
ring Mn (k), where k is a field, it is never a primitive idempotent element if n > 1 since
it is the sum of the orthogonal (non-central) primitive idempotent elements I = E1,1 +
+ En,n . Furthermore, removing the hypothesis of centrality we can no longer say that
decompositions of the identity as a sum of primitive idempotent elements are unique;
indeed, any conjugate expression by an invertible matrix will also be a sum of orthogonal
primitive idempotent elements. Applying these comments to a matrix summand of CG, the
primitive idempotent decompositions of 1 will never be unique if we have a non-abelian
matrix summand which, of course, happens precisely when G is non-abelian. It is
unfortunately the case that in terms of the group elements there is in general no known
formula for primitive idempotent elements of CG lying in a non-abelian matrix summand.
Burnsides pa q b theorem
We conclude this chapter with Burnsides remarkable pa q b theorem, which establishes
a group-theoretic result using the ideas of representation theory we have so far developed,
together with some admirable ingenuity. In the course of the proof we again make use of
the idea of integrality, but this time we also require Galois theory at one point. This is
needed to show that if is a field element that is expressible as a sum of roots of unity,
then every algebraic conjugate of is again expressible as a sum of roots of unity. We
present Burnsides theorem here because of its importance as a theorem in its own right,
not because anything later depends on it. In view of this the proof (which is fairly long)
can be omitted without subsequent loss of understanding.
Recall that a group G is solvable if it has a composition series in which all of the
composition factors are cyclic. Thus a group is not solvable precisely if it has a nonabelian composition factor.
(3.24) THEOREM (Burnsides pa q b theorem). Let G be a group of order pa q b where
p and q are primes. Then G is solvable.
Proof. We suppose the result is false, and consider a group G of minimal order subject
to being not solvable and of order pa q b .
Step 1. The group G is simple, not abelian and not of prime-power order; for if it were
abelian or of prime-power order it would be solvable, and if G had a normal subgroup N
then one of N and G/N would be a smaller group of order p q which was not solvable.
Step 2. We show that G contains an element g whose conjugacy class has size q d
for some d > 0. Let P be a Sylow p-subgroup, 1 6= g Z(P ). Then CG (g) P so
|G : CG (g)| = q d for some d > 0, and this is the number of conjugates of g.
Step 3. We show that there is a simple non-identity character of G such that q6 (1)
and (g) 6= 0. To prove this, suppose to the contrary that whenever 6= 1 and q 6 (1)
Peter Webb
41
then (g) = 0. Let R denote the ring of algebraic integers in C. Consider the orthogonality
relation between the column of 1 (consisting of character degrees) and the column of g:
1+
(1)(g) = 0.
6=1
Then q divides every term apart from 1 in the sum on the left, and so 1 qR. Thus
1
q 1 R. But q 1 Q and
so q Z by 3.18, a contradiction. We now fix a non-identity
character for which q6 (1) and (g) 6= 0.
Step 4. Recall that the number of conjugates of g is q d . We show that
q d (g)
(1)
is an algebraic integer. To do this we use results 3.15 and 3.20. These imply that if
P
g = hg h CG is the sum of the elements conjugate to g and is a representation
affording the character then g Z(CG) and
(g) =
1 X
(h) I
(1)
hg
q d (g)
I,
(1)
where I is the identity matrix. Now by 3.19 this is integral over (Z) = Z I, which proves
what we want.
Step 5. We deduce that (g)
(1) is an algebraic integer. This arises from the fact that
q6 (1). We can find , Z so that q d + (1) = 1. Now
(g)
q d (g)
=
+ (g)
(1)
(1)
42
where is the simple non-identity character introduced in Step 3. We argue first that H
is a normal subgroup. If the eigenvalues of (h) are 1 , . . . , n then, since these are roots
of unity, |1 + + n | = n if and only if 1 = = n . Thus |(h)| = (1) if and only if
(h) is multiplication by some scalar, and from this we see immediately that H is a normal
subgroup. It also implies that H/ Ker is abelian. From Step 6 we see that H contains
the non-identity element g, so simplicity of G forces H = G. Since is not the trivial
representation Ker 6= G, so simplicity of G again forces Ker = 1, so that G must be
abelian. However G was seen not to be abelian in Step 1, and this contradiction completes
the proof.
Summary of Chapter 3
1. There is a tensor product operation on RG-modules whose result is an RG-module.
2. Characters are class functions.
3. The character of a direct sum, tensor product or dual is the sum, product or
complex conjugate of the characters.
4. The characters of the simple representations form an orthonormal basis for class
functions with respect to a certain bilinear form.
5. The character table is square and satisfies row and column orthogonality relations.
The number of rows of the table equals the number of conjugacy classes in the group.
6. The conjugacy class sums form a basis for the center of RG.
7. The simple character degrees are divisors of |G|. The sum of their squares equals
|G|.
8. There is a formula for the primitive central idempotents in CG.
9. Every group of order pa q b is solvable.
Peter Webb
43
V C W , and (CG CG) = CG CG as CG-modules.
(b) If k is any field and V , W are kG-modules, show that (V k W )
= V k W ,
and (kG kG) = kG kG as kG-modules.
3. Consider a ring with identity that is the direct sum (as a ring) of non-zero subrings
A = A1 Ar . Suppose that A has exactly n isomorphism types of simple modules.
Show that r n.
4. Let g be any non-identity element of a group G. Show that G has a simple complex
character for which (g) has negative real part.
5. Show that if every element of a finite group G is conjugate to its inverse, then every
character of G is real-valued.
Conversely, show that if every character of G is real-valued, then every element of G
is conjugate to its inverse.
[The quaternion group of order 8 in its action on the algebra of quaternions provides
an example of a complex representation that is not equivalent to a real representation,
but whose character is real-valued (see Chapter 2 Exercise 11). In this example, the
representation has complex dimension 2, but there is no basis over C for the representation
space such that the group acts by matrices with real entries. A real-valued character does
not necessarily come from a real representation.]
6. (Jozsef Pelikan) While walking down the street you find a scrap of paper with the
following character table on it:
1
1
1
1
2 1
3
1
3
1
All except two of the columns are obscured, and while it is clear that there are five rows
you cannot read anything of the other columns, including their position. Prove that there
is an error in the table. Given that there is exactly one error, determine where it is, and
what the correct entry should be.
7. A finite group has seven conjugacy classes with representatives c1 , . . . , c7 (where
c1 = 1), and the values of five of its irreducible characters are given by the following table:
c1 c2 c3 c4 c5 c6 c7
1
1
1 1
1
1
1
1
1
1 1 1 1 1
4
1 1 0
2 1
0
4
1 1 0 2
1
0
5 1
0 1
1
1 1
44
Calculate the numbers of elements in the various conjugacy classes and the remaining
simple characters.
8. Let g G.
(a) Use Proposition 3.1 to prove that g lies in the center of G if and only if |(g)| =
|(1)| for every simple complex character of G.
(b) Show that if G has a faithful simple complex character (one whose kernel is 1)
then the center of G is cyclic.
9. Here is a column of a character table:
g
1
1
0
1
1
1+i 11
2
1i 11
2
0
1
0
Peter Webb
45
of V is the unique largest CS3 -submodule of V that is a direct sum of copies of the simple
2-dimensional CS3 -module.
12. Let G = hxi be cyclic of order n.
(a) Write down a complete set of primitive (central) idempotents in CG.
(b) Let Cn be an n-dimensional space with basis v1 , . . . , vn . Let g : Cn Cn be the
linear map of order n specified by gvi = vi+1 , 1 i n 1, gvn = v1 , so that g has matrix
0 0 1
.
0
1 ..
.
T =. .
..
..
..
.
0 1 0
Let = e2i/n be a primitive nth root of unity. Show that for each d the vector
ed =
n
X
ds vs
s=1
15. Let G permute a set and let R denote the permutation representation of G
over R determined by . This means R has a basis in bijection with and each element
g G acts on R by permuting the basis elements in the same way that g permutes .
(a) Show that when H is a subgroup of G and = G/H is the set of left cosets of H
in G, the kernel of G in its action on R is H if and only if H is normal in G.
(b) Show that the normal subgroups of G are precisely the subgroups of the form
Ker i1 Ker it where 1 , . . . , n are the simple characters of G. Use Proposition
3.1 to deduce that the normal subgroups of G are determined by the character table of G.
(c) Show that G is a simple group if and only if for every non-trivial simple character
and for every non-identity element g G we have (g) 6= (1).
46
Peter Webb
47
We next show how to obtain the simple characters of a product of groups in terms of
the characters of the groups in the product. Combining this with the last result we obtain
the character table of any finite abelian group. We describe a construction that works over
any ring R. Suppose that 1 : G1 GL(V1 ) and 2 : G2 GL(V2 ) are representations of
groups G1 and G2 . We may define an action of G1 G2 on V1 R V2 by the formula
(g1 , g2 )(v1 v2 ) = g1 v1 g2 v2
where gi Gi and vi Vi . When R is a field we may choose bases for V1 and V2 , and
now (g1 , g2 ) acts via the tensor product of the matrices by which g1 and g2 act. It follows
when R = C that
V1 V2 (g1 , g2 ) = V1 (g1 )V2 (g2 ).
(4.2) THEOREM. Let V1 , . . . , Vm and W1 , . . . , Wn be complete lists of the simple
complex representations of groups G1 and G2 . Then the representations Vi Wj with 1
i m and 1 j n form a complete list of the simple complex G1 G2 representations.
Remark. Theorem 4.2 is false in general when the field over which we are working is
not algebraically closed (see Exercise 15 to Section 4). The theorem is an instance of a
more general fact to do with representations of finite-dimensional algebras A and B over an
algebraically closed field k: the simple representations of A k B are precisely the modules
S k T , where S is a simple A-module and T is a simple B-module. This is proved in
[CRI, Theorem 10.38] and [Lemma 6.8 of R. Steinberg, Lectures on Chevalley Groups,
Yale University Notes, 1967.]. The connection between this result for abstract algebras
and group representations is that the group algebra R[G1 G2 ] over any commutative ring
R is isomorphic to RG1 R RG2 , which we may see by observing that this tensor product
has a basis consisting of elements g1 g2 with gi Gi that multiply in the same way as
the elements of the group G1 G2 .
Proof. We first verify that the representations Vi Wj are simple using the criterion
of 3.10:
hVi Wj , Vi Wj i =
=
=
1
|G1 G2 |
1
|G1 ||G2 |
1
|G1 |
= 1.
g1 G1
1 X
Vj (g2 )Vj (g2 )
|G2 |
g2 G2
48
The characters of these representations are distinct, since by a similar calculation if (i, j) 6=
(r, s) then hVi Wj , Vr Ws i = 0. To show that we have the complete list, we observe that
if dim Vi = di and dim Wj = ej then Vi Wj is a representation of degree di ej and
n
m
m X
n
X
X
X
2
2
e2j = |G1 ||G2 |.
di
(di ej ) =
i=1
i=1 j=1
j=1
Putting the last two results together enables us to compute the character table of any
finite abelian group. To give a very small example, let
G = hx, y x2 = y 2 = [x, y] = 1i
= C2 C2 .
The character tables of hxi and hyi are
hyi
hxi
g
|CG (g)|
1
2
x
2
1
2
1 1
1 1
g
|CG (g)|
1
2
y
2
1
2
1 1
1 1
1
4
x y
4 4
xy
4
1 1
2 1
1 2
2 2
1 1 1 1
1 1 1 1
1 1 1 1
1 1 1 1
We immediately notice that this construction gives the character table of C2 C2 as the
tensor product of the character tables of C2 and C2 , and evidently this is true in general.
(4.3) COROLLARY. The character table of a direct product G1 G2 is the tensor
product of the character tables of G1 and G2 .
We may see from this theory that all simple complex characters of an abelian group
have degree 1 (a result already shown in Corollary 2.10), and that in fact this property
characterizes abelian groups. We will give a different argument later on in Theorem 5.10
that shows that over any algebraically closed field the simple representations of abelian
groups all have degree 1.
Peter Webb
49
50
(4.6) Example. Neither implication of 4.4 holds if we do not assume that our representations are defined over an algebraically closed field of characteristic prime to |G|, such as
C. We have seen in examples before now that over R the 2-dimensional representation of
the cyclic group hx x3 = 1i, in which x acts as rotation through 2
3 , is simple since there
is no 1-dimensional subspace stable under the group action. We need to pass to C to split
it as a sum of two representations of degree 1. It is also possible to find a non-abelian group
all of whose simple representations do have degree 1: we shall see in Proposition 6.3 that
this happens whenever G is a p-group and we consider representations in characteristic p.
Induction and Restriction
We now consider how to construct representations of a group from representations of
its subgroups. Let H be a subgroup of G and V an RH-module where R is a commutative
ring with 1. We define an RG-module
V G
H = RG RH V
with the action of G coming from the left module action on RG:
X
X
x(
ag g v) = (x
ag g) v
gG
gG
as R-modules, where gi V = {gi v v V } RG RH V . Each gi V is isomorphic
to V as an R-module, and in case V is free as an R-module we have
rankR V G
H = |G : H| rankR V.
Peter Webb
51
|G:H|
M
i=1
|G:H|
M
i=1
|G:H|
M
i=1
gi RH) RH V
(gi RH RH V )
gi RH V
and as R-modules gi RH RH V
= RH RH V
=V.
We next show that with its left action on RG RH V coming from the left action on
RG, G permutes these R-submodules. If x G and xgi = gj h with h H then
x(gi v) = xgi v
= gj h v
= gj hv,
so that x(gi v) gj V . We argue that we have equality using the invertibility of
x. For, by a similar argument to the one above, we have x1 gj V gi V , and so
gj V = xx1 (gj V ) x(gi V ). This action of G on the subspaces is transitive since
given two subspaces gi V and gj V we have (gj gi1 )gi V = gj V .
Now to compute the stabilizer of g1 V where g1 H, if x H then x(g1 V ) =
g1 (g11 xg1 ) V = g1 V , and if x 6 H then x gi H for some i 6= 1 and so x(g1 V ) =
gi V . Thus StabG (g1 V ) = H.
The structure of induced modules described in the last result in fact characterizes
these modules, giving an extremely useful criterion for a module to be of this form that
we will use several times later on.
(4.8) PROPOSITION. Let M be an RG-module that has an
R-submodule V with
the property that M is the direct sum of the R-submodules {gV g G}. Let H = {g
G gV = V }. Then M
= V G
H.
We are using the notation {gV g G} to indicate the set of distinct possibilities for
gV , so that if gV = hV with g 6= h we do not count gV twice.
Proof. We define a map of R-modules
RG RH V M
g v 7 gv
52
1
|H|
(t1 gt)
tG
t1 gtH
t[G/H]
t1 gtH
(t1 gt).
Peter Webb
53
Proof. The two formulas on the right are in fact the same, since if t1 gt H and
h H then (th)1 gth H also, and so {t G t1 gt H} is a union of left cosets of
H. Since (t1 gt) = ((th)1 gth) the terms in the first sum are constant on the cosets
of H, and we obtain the second sum by choosing one representative from each coset and
multiplying by |H|.
Using the vector space decomposition of 4.7 we obtain that the trace of g on V G
H is
the sum of the traces of g on the spaces t V that are invariant under g, where t [G/H].
This is because if g does not leave t V invariant, we get a matrix of zeros on the diagonal
at that point in the block matrix decomposition for the matrix of g. Thus we only get a
non-zero contribution from subspaces t V with gt V = t V . This happens if and only
if t1 gt V = 1 V , that is t1 gt H. We have
X
G
(g)
=
trace of g on t V.
H
t[G/H]
t1 gtH
Now g acts on t V as
g(t v) = t(t1 gt) v = t (t1 gt)v
and so the trace of g on this space is (t1 gt). Combining this with the last expression
gives the result.
We see in the above proof that g leaves invariant t V if and only if t1 gt H,
or in other words g tHt1 . Thus StabG (t V ) = tHt1 . Furthermore, if we identify
t V with V by means of the bijection t v v, then g acts on t V via the composite
homomorphism
ct1
hgiH
GL(V )
where is the homomorphism associated to V and ca (x) = axa1 is the automorphism of
G that is conjugation by a G.
(4.11) Example. To make clearer what the terms in the expression for the induced
character are, consider G = S3 and H = h(123)i, the normal subgroup of order 3. To avoid
expressions such as (()) we will write the identity element of S3 as e. We may take the
coset representatives [G/H] to be {e, (12)}. If is the trivial character of H then
e
(12)
G
) = 2,
H (e) = (e ) + (e
G
H ((12)) = the empty sum = 0,
e
(12)
G
) = 2.
H ((123)) = ((123) ) + ((123)
54
first principles, so we now describe some formalism that will enable us to compute with
them more easily. The companion notion to induction is that of restriction of representations. If H is a subgroup of G and W is a representation of G we denote by W G
H the
representation of H obtained by letting the elements of H the way they do when regarded
as elements of G. In other words, we just forget about the elements of G not in H. Restriction and induction are a particular case of the following more general situation. Whenever
we have a (unital) homomorphism of rings A B, an A-module V and an B-module
W , we may form the B-module B A V and the A-module W B
A . On taking A = RH
and B = RG we obtain the induction and restriction we have been studying. There is, in
fact, a further operation we mention at this point, that of coinduction: we may form the
B-module HomA (B, V ), which acquires the structure of a left B-module because of the
right action of B on itself. We are going to prove that induction and coinduction are the
same in the case of group rings, so that we will not need to consider coinduction for long.
(4.12) LEMMA. Let A B be a homomorphism of rings, V an A-module and W a
B-module.
(1) (Left adjoint of restriction) HomB (B A V, W )
= HomA (V, W A ).
(2) (Right adjoint of restriction) HomA (W A , V ) = HomB (W, HomA (B, V )).
(3) (Transitivity of induction) If : B C is another ring homomorphism then
C B (B A V )
= C A V.
Proof. In the case of (1) the mutually inverse isomorphisms are
f 7 (v 7 f (1 v))
and
(b v 7 bg(v))g.
In the case of (2) the mutually inverse isomorphisms are
f 7 (w 7 (b 7 f (bw)))
and
(w 7 g(w)(1))g.
In the case of (3) the mutually inverse isomorphisms are
c b v 7 c(b) v
and
c 1 vc v.
There is checking to be done to show that morphisms are indeed well-defined homomorphisms of A-modules and B-modules, and that maps are mutually inverse, but it is all
routine.
Peter Webb
55
It is worth knowing that 4.12(1) and (2) are instances of a single formula to do with
bimodules. An (A, B)-bimodule T is defined to have the structure of both a left A-module
and a right B-module, and such that these two module actions commute: for all a A,
b B, t T we have (at)b = a(tb). The basic adjoint relationship between A-modules V
and B-modules W is an isomorphism
HomA (T B W, V )
= HomB (W, HomA (T, V ))
that is given by mutually inverse maps
f 7 (w 7 (t 7 f (t w)))
and
(t w 7 g(w)(t))g.
Supposing that we have a ring homomorphism A B, one such bimodule T is the set B
with left action of A given by left multiplication after applying the homomorphism to B,
and right action of B given by right multiplication. We denote this bimodule A BB . There
is a similarly defined (B, A)-bimodule B BA on which B acts by left multiplication and A
acts by right multiplication after applying the homomorphism to B. We have isomorphisms
of A-modules
HomB (B BA , V )
= V B
A = A BB B V
whereas B BA A W and HomA (A BB , W ) are by definition the induction and coinduction
of W . Applying the single adjoint isomorphism in the two cases of these bimodules yields
the relationships of 4.12(1) and (2)
(4.13) COROLLARY. Let H K G be subgroups of G, let V be an RH-module
and W an RG-module.
(1) (Frobenius reciprocity)
G
HomRG (V G
H , W ) = HomRH (V, W H )
and
G
HomRG (W, V G
H ) = HomRH (W H , V ).
(2)
(3)
(4)
(5)
G
G
(Transitivity of induction) (V K
H ) K = V H as RG-modules.
K
G
(Transitivity of restriction) (W G
K ) H = W H as RH-modules.
G
G
V G
H R W = (V R W H ) H as RG-modules.
Proof. The first isomorphism of (1) and part (2) follow from the relationships in 4.12,
in the case of the ring homomorphism RH RG. Part (3) also holds in this generality
56
and is immediate. The second isomorphism in (1) as well as (4) and (5) are special for
group representations.
Part (4) is the isomorphism
(RG RH V ) R W
= RG RH (V R W )
and it is not a corollary of 4.12. Here the mutually inverse isomorphisms are
(g v) w 7 g (v g 1 w)
and
(g v) gwg (v w).
We prove (5) by exhibiting mutually inverse isomorphisms
RG RH V
= HomRH (RG, V )
given by
g v 7 g,v
where g,v (x) = (xg)v if x Hg 1 and is 0 otherwise. Here g, x G and v V . In the
opposite direction the map is
X
g (g 1 )
g[G/H]
where the sum is taken over a set of representatives for the left cosets of H in G. We must
check here that g,v is a homomorphism of RH-modules, does specify a map on the tensor
product, that is well defined, and that the two morphisms are mutually inverse. We have
shown that induced and coinduced modules are the same.
Finally the second isomorphism of part (1) follows from 4.12(2), using (5).
In the case of representations in characteristic zero all of these results may be translated into the language of characters. In this setting the second Frobenius reciprocity
formula, which used the fact that coinduced modules and induced modules are the same,
becomes much easier. If our interest is only in character theory, we did not need to read
about left and right adjoints, coinduction and bimodules. By analogy with the notation
G
K
G
K for the character of an induced representation V K , let us write H for the
character of V K
H.
Peter Webb
57
=
(
.
H
H
H
Proof. In (1) we write h , iG and h , iH to denote the inner product of characters
of G and H, respectively. The four parts are translations of the first four parts of 4.13
into the language of characters. In part (1) we use the fact that the inner products are the
dimensions of the Hom groups in 4.13(1). We may deduce the second formula from the
first here because the inner product is symmetric.
Frobenius reciprocity for complex characters is equivalent to saying that if and
are simple characters of G and H respectively then the multiplicity of as a summand of
G
G
H equals the multiplicity of as a summand of H .
At a slightly more sophisticated level we may interpret induction, restriction and
Frobenius reciprocity in terms of the space Ccc(G) of class functions introduced in Chapter
3, that is, the vector space of functions cc(G) C where cc(G) is the set of conjugacy
classes of G. Since each conjugacy class of H is contained in a unique conjugacy class
of G we have a mapping cc(H) cc(G) and this gives rise by composition to a linear
cc(G)
map G
Ccc(H) that on characters is the restriction operation we have already
H: C
cc(H)
defined. We may also define a linear map G
Ccc(G) that on characters sends a
H: C
character of H to the character G
H . It would be possible to define this on arbitrary
class functions of H by means of the explicit formula given in 4.10, but the trouble with
this is that transitivity of induction is not entirely obvious. It is perhaps easier to observe
that the characters of simple representations of H form a basis of Ccc(H) . We have defined
G
G
H on these basis elements, and we may define H on arbitrary class functions so that
it is a linear map. With these definitions the formulas of 4.14 hold for arbitrary class
functions. We may also interpret Frobenius reciprocity within this framework. The inner
products h , iG and h , iH provide us with the notion of the transpose of a linear
map between the vector spaces Ccc(H) and Ccc(G) . Now Frobenius reciprocity states that
induction and restriction are the transpose of each other. We know that the characters
of simple modules form orthonormal bases of these vector spaces. Taking matrices with
respect to these bases, the matrix of induction is the transpose of the matrix of restriction.
(4.15) Examples. 1. Frobenius reciprocity is a most useful tool in calculating with
induced characters. In the special case that V and W are simple representations over C of
58
x
n
x2
n
1
1a
1
1
1
1
1
1
ns G
hxi
g
|CG (g)|
(1 s
n1
2 )
n1
2
y
2
n
1
1
n1
ns + ns n2s + n2s n 2
n1
+ n 2
1
1
0
if n6 s, using Frobenius reciprocity (or a direct calculation), and hence the characters s
must be simple when n 6 s, because otherwise they would be a sum of two characters of
degree 1, that must be 1 or 1, and evidently this would not give the correct character
values. For 1 s n1
they are distinct, and so we have constructed n1
+ 2 = n+3
2
2
2
simple characters. This equals the number of conjugacy classes of G, so we have the
complete character table.
Peter Webb
59
(extended by linearity to the whole of V n ). This map contains I in its kernel, so there is
induced a map
S n (V ) k[u1 , . . . , ur ]n .
1
2
r
This is now an isomorphism since, modulo I, the tensors ua
ua
ua
r
1
2
Pr
n
where
, and they map to the monomialsthat form a basis of
i=1 ai = n span V
dimk k[u1 , . . . , ur ]n . As is well-known, dimk k[u1 , . . . , ur ]n = n+r1
.
n
The nth exterior power of V is the vector space
n (V ) = V n /J
where J is the subspace spanned by tensors (( vi vj )+( vj vi ))
and ( vi vi ) where vi , vj V . We write the image of v1 vn in
n (V ) as v1 vn , so that interchanging vi and vj changes the sign of the symbol,
and if two of vi and vj are equal the symbol is zero. If the characteristic of k is not 2
the second of these properties follows from the first, but for the sake of characteristic 2 we
impose it anyway. By an argument similar to the one used for symmetric powers we see
n
that
(V ) has as anbasis {ui1 uin 1 i1 < < in r}, and its dimension is
r
n . In particular, (V ) = 0 if n > dim V .
Suppose now that a group G acts on V and consider the diagonal action of G on V n .
The subspaces of relations I and J are preserved by this action, and so there arise actions
of G on S n (V ) and n (V ):
g (v1 v2 vn ) = (gv1 )(gv2 ) (gvn )
60
Because we substitute the expressions for gvi into the monomials that form the bases of
S n (V ) and n (V ), we say that G acts on these spaces by linear substitutions. With these
actions we have described the symmetric and exterior powers of the representation V .
(4.16) Example. Consider the representation of G = hx x3 = 1i on the vector space
V with basis {u1 , u2 } given by
xu1 = u2
xu2 = u1 u2 .
Then S 2 (V ) has a basis {u21 , u1 u2 , u22 } and
x u21 = u22
(1, 3)( ) = .
From the above very convincing formulas and the fact that (1, 2, 3) = (1, 3)(1, 2) we deduce
that
(1, 2, 3)( ) =
which is evidence that if Sn then
(v1 vn ) = v1 (1) v1 (2) v1 (n) ,
a formula that is not quite so obvious. With this action it is evident that S n (V ) is the
largest quotient of V n on which Sn acts trivially, and when char(k) 6= 2, n (V ) is the
largest quotient of V n on which Sn acts as a sum of copies of the sign representation.
We define the symmetric tensors or divided powers to be the fixed points (V n )Sn , and
when char k 6= 2 we define the skew-symmetric tensors to be the largest kSn -submodule of
V n that is a sum of modules isomorphic to the sign representation. Thus
symmetric tensors = {w V n (w) = w for all Sn },
skew-symmetric tensors = {w V n (w) = sign()w for all Sn }.
Peter Webb
61
ij vi vj =
1X
1X
ij (vi vj + vj vi ) +
ij (vi vj vj vi ).
2
2
62
S (g) =
r
X
2i +
i=1
i j
1i<jr
1
((1 + + r )2 + (21 + + 2r ))
2
1
= ((g)2 + (g 2 )).
2
1
((1 + + r )2 (21 + + 2r ))
2
1
= ((g)2 (g 2 )).
2
Peter Webb
63
There is a formula due to Molien for the generating function of characters of the
symmetric powers of V . We present Moliens theorem in exercise 14.
64
Summary of Chapter 4
1. The characters of representations obtained by induction, tensor product, symmetric
and exterior powers are all useful in constructing character tables and there are formulas
for these characters.
2. The character table of G1 G2 is the tensor product of the character tables of G1
and G2 .
3. The degree 1 characters of G are precisely the characters of the simple representations of G/G .
4. The simple complex characters of a cyclic group of order n are the n homomorphisms to the group of nth roots of unity.
5. Induced characters may be decomposed using Frobenius reciprocity.
6. An induced module may be identified by the fact that it is a direct sum of subspaces
that are permuted by G.
Show that G has 5 conjugacy classes, and find its character table.
3. Find the character table of the following group of order 36:
G = ha, b, c a3 = b3 = c4 = 1, ab = ba, cac1 = b, cbc1 = a2 i.
[It follows from these relations that ha, bi is a normal subgroup of G of order 9.]
4. Compute the character table of the symmetric group S5 by the methods of this
chapter. To help in doing this, consider especially the decomposition of the permutation
representation on 5 symbols, the symmetric and exterior square of the summands, as well
as tensor product with the sign representation.
Peter Webb
65
that V G
H = (V H ) . Deduce from this that kG = (kG) and (more generally) that
permutation modules are self-dual (i.e. isomorphic to their dual).
6. Let k be any field, and V any representation of G over k. Prove that V kG is
isomorphic to a direct sum of copies of kG.
7. The tensor product V = R3 R R3 R R3 is a vector space of dimension 27 with basis
the tensors ei ej ek where e1 , e2 , e3 is a standerd basis for R3 . The symmetric group S3
acts on V by permuting the positions of the suffixes, so for instance (1, 2) (e3 e1 e2 ) =
e1 e3 e2 .
(a) Find the multiplicity of each simple representation of S3 in a decomposition of V as
a direct sum of simple representations. [Observe that V is a permutation representation.]
(b) Give also the decomposition of V as a direct sum of three subspaces consisting of
tensors with different symmetry properties under the action of S3 . What are the dimensions
of these subspaces? Find a basis for each subspace. [Use the result of Chapter 3 Exercise
11.]
(c) The Schur algebra SC (3, 3) may be defined to be the endomorphism ring HomCS3 (V, V ).
Show that SC (3, 3) is semisimple and find the dimensions of its simple representations.
8. Let V be a representation of G over a field k of characteristic zero. Prove that the
symmetric power S n (V ) is isomorphic as a kG-module to the space of symmetric tensors
in V n .
9. Let U, V be kG-modules where k is a field, and suppose we are given a nondegenerate bilinear pairing
h , i:U V k
that is G-invariant, that is, hu,vi = hgu, gvi for all u U , v V , g G. If U1 is a
subspace of U let U1 = {v V hu, vi = 0 for all u U1 } and if V1 is a subspace of V let
V1 = {u U hu, vi = 0 for all v V1 }.
(a) Show that V
= U as kG-modules, and that there is an identification of V with
U so that h , i identifies with the canonical pairing U U k.
(b) Show that if U1 and V1 are kG-submodules, then so are U1 and V1 .
(c) Show that if U1 U2 are kG-submodules of U then
U1 /U2
= (U2 /U1 )
as kG-modules.
(d) Show that the composition factors of U are the duals of the composition factors
of U .
10. Let be a finite G-set and k the corresponding permutation module, where
k is a field. Let h , i : k k k be the symmetric bilinear form specified on the
elements of as
n
1 if 1 = 2 ,
h1 , 2 i =
0 otherwise.
66
(a) Show that this bilinear form is G-invariant, i.e. h1 , 2 i = hg1 , g2 i for all g G.
(b) Deduce from this that k is self-dual, i.e. k
= (k) .
11. Let V be a kG-module where k is a field, and let h , i : V V k be the
canonical pairing between V and its dual, so hv, f i = f (v).
(a) Show that the specification hv1 vn , f1 fn i = f1 (v1 ) fn (vn ) determines a non-degenerate bilinear pairing h , i : V n (V )n k that is invariant both
for the diagonal action of G and the action of Sn given by permuting the positions of the
tensors.
(b) Let I and J be the subspaces of V n that appear in the definitions of the symmetric
and exterior powers, so S n (V ) = V n /I and n = V n /J. Show that I (defined in
question 9) equals the space of symmetric tensors in (V )n , and that J equals the space
of skew-symmetric tensors in (V )n (at least, when char k 6= 2).
(c) Show that (S n (V ))
= STn (V ), and that (n (V ))
= SSTn (V ), where STn
denotes the symmetric tensors, and in general we define the skew-symmetric tensors
SSTn (V ) to be J .
12. Let G = C2 C2 be the Klein four group with generators a and b, and k = F2
the field of two elements. Let V be a 3-dimensional space on which a and b act via the
matrices
1 0 0
1 0 0
1 1 0 and 0 1 0 .
0 0 1
1 0 1
HC
Ccc(H)
HC
Ccc(H) Ccc(G)
and induction
cc(H)
G
Ccc(G)
H: C
Peter Webb
67
(a) With respect to the usual inner product h , iG on Ccc(G) and the inner product
L
on HC Ccc(H) that is the orthogonal sum of the h , iH , show that resC and indC are
the transpose of each other.
(b) Show that resC is injective.
[Use the fact that Ccc(G) has a basis consisting of characters, that take their information
from cyclic subgroups.]
(c) Prove Artins induction theorem: In Ccc(G) every character can be written as a
rational linear combination
X
=
aH, G
H
where the sum is taken over cyclic subgroups H of G, ranges over characters of H and
aH, Q.
[Deduce this from surjectivity of indC and the fact that it is given by a matrix with integer
entries. A stronger version of Artins theorem is possible: there is a proof due to Brauer
which gives an explicit formula for the coefficients aH, ; from this we may deduce that
when is the character of a QG-module the that arise may all be taken to be the trivial
character.]
(d) Show that if U is any CG-module then there are CG-modules P and Q, each a
direct sum of modules of the form V G
H where H is cyclic, for various V and H, so that
n
Q
for
some
n,
where
U
is
the
direct sum of n copies of U .
Un P
=
14. (Moliens Theorem) (a) Let : G GL(V ) be a complex representation of G, so
that V is a CG-module, and for each n let S n (V ) be the character of the nth symmetric
power of V . Show that for each g G there is an equality of formal power series
S n (V ) (g)tn =
n=0
1
.
det(1 t(g))
X
i=0
Show that
MV (W ) =
hS n (V ) , W itn .
1 X W (g 1 )
.
|G|
det(1 t(g))
gG
68
(c) When G = S3 and V is the 2-dimensional simple CS3 -module show that
MV (C) =
1
(1
t2 )(1
t3 )
t3
MV () =
= t3 + t5 + t6 + t7 + t8 + 2t9 + t10 +
2
3
(1 t )(1 t )
t(1 + t)
MV (V ) =
= t + t2 + t3 + 2t4 + 2t5 + 2t6 + 3t7 + 3t8 + 3t9 + 4t10 +
(1 t2 )(1 t3 )
where C denotes the trivial module and the sign representation. Deduce, for example,
that the eighth symmetric power S 8 (V )
= C2 V 3 .
15. Show that every simple representation of C3 C3 over R has dimension 1 or 2.
Deduce that if V is a simple 2-dimensional representation of C3 over R then V V is not
a simple R[C3 C3 ]-module.
16. Let H be a subgroup of G.
P
(a) Write H = hH h for the sum of the elements of H, as an element of RG. Show
that RG H
= R G
H as left RG-modules. Show also that RG H equals the fixed points
of H in its action on RG from the right.
(b) More generally let : H R be a 1-dimensional representation of H (that is,
:= P
a group homomorphism to the units of R). Write H
hH (h)h RG. Show that
= as RG-modules.
RG H
H
17. Let H be a subgroup of G and V an RH-module. Show that if V can be generated
by d elements as an RH-module then V G
H can be generated by d elements as an RGmodule.
18. Compute the character tables of the alternating groups A4 and A5 using the
following procedure. You may assume that A5 is a simple group that is isomorphic to
the group of rotations of a regular icosahedron, and that A4 is isomorphic to the group of
rotations of a regular tetrahedron.
(a) Compute the conjugacy classes by observing that each conjugacy class of even
permutations in Sn is either a single class in An or the union of two classes of An , and
that this can be determined by computing centralizers of elements in An and comparing
them with the centralizers in Sn .
(b) Compute the abelianization of each group, and hence the 1-dimensional representations.
(c) Obtain further representations using the methods of this section. We have natural
3-dimensional representations in each case. It is also helpful to consider d induced representations from the Sylow 2-subgroup in the case of A4 , and from the subgroup A4 in the
case of A5 .
Peter Webb
69
If is a left G-set we use the notation G\ for the set of orbits of G on , and denote a
set of representatives for the orbits by [G\]. Similarly if is a right G-set we write /G
and [/G]. We will use all the time the fact that if is a transitive G-set and then
70
is evidently well-defined and surjective. If H(g1 K) = H(g2 K) then g2 K = hg1 K for some
h H, so g2 Hg1 K and Hg1 K = Hg2 K by (2). Hence the mapping is injective.
The proof that double cosets biject with (H\G)/K is similar.
In view of (3) we denote the set of (H, K)-double cosets in G by H\G/K. We denote
a set of representatives for these double cosets by [H\G/K].
(5.2) Example. Consider S2 = {1, (12)} as a subgroup of S3 . We have
S2 \S3 /S2 = {{1, (12)}, {(123), (132), (13), (23)}},
while, for example,
[S2 \S3 /S2 ] = {1, (123)}.
S3 acts transitively on {1, 2, 3} with StabS3 (3) = S2 , so as S3 -sets we have
S3 /S2
= {1, 2, 3}.
Thus the set of orbits on this set under the action of S2 is
S2 \(S3 /S2 ) {{1, 2}, {3}}.
We observe that these orbits are indeed in bijection with the double cosets S2 \S3 /S2 .
This example illustrates the point that when computing double cosets it may be
advantageous to identify G/K as some naturally occurring G-set, rather than as the set
of left cosets.
In the next result we distinguish between conjugation on the left and on the right:
g
x = gxg 1 and xg = g 1 xg. Later on we will write cg (x) = gx, so that cg : H gH is
the homomorphism that is left conjugation by g, and cg 1 (x) = xg .
(5.3) PROPOSITION. Let H, K be subgroups of G and g G an element. We have
isomorphisms
HgK/K
= H/(H gK) as left H-sets
and
H\HgK
= (H g K)\K
as right K-sets.
Thus the double coset HgK is a union of |H : H gK| left K-cosets and |K : H g K|
right H-cosets. We have
X
|G : K| =
|H : H gK|
g[H\G/K]
and
|G : H| =
g[H\G/K]
|K : H g K|.
Peter Webb
71
Proof. HgK is the union of a single H-orbit of left K-cosets. The stabilizer in H of
one of these is
StabH (gK) = {h H hgK = gK}
= {h H hg K = K}
= {h H hg K}
= H gK.
Thus HgK/K
= H/(H gK) as left H-sets and the number of left K-cosets in HgK
equals |H : H gK|. By summing these numbers over all double cosets we obtain the total
number of left K-cosets |G : K|.
The argument with right H-cosets is similar.
We next introduce conjugation of representations, a concept we have in fact already
met with induced representations. Suppose H is a subgroup of G, g G and V is a
representation of H. We define a representation gV of gH by specifying that gV = V as a
set, and if gh gH then ghv = hv. Thus if : H GL(V ) was the original representation,
cg 1
( g(V K
=
K
H
H g K )) H gK
g[H\G/K]
as RH-modules.
Proof. We have V G
K=
The terms
x[G/K] x
x[G/K]
xHgK
xV
form an R-submodule invariant under the action of H, since it is the direct sum of an orbit
of R-submodules permuted by H. Now
StabH (g V ) = {h H hg V = g V }
= {h H g 1 hg StabG (1 V ) = K}
= H gK.
72
Peter Webb
73
In the equivalence of statements that forms the third sentence, neither side is true if
G has more than one orbit on , so we may assume = G/H. The character of C is
1 G
H , and we compute
G
G
G
h1 G
H , 1 H iG = h(1 H ) H , 1iH
X
=h
( g1) H
H gH , 1iH
g[H\G/H]
h1 H
H gH , 1iH
g[H\G/H]
h1, 1iH gH
g[H\G/H]
g[H\G/H]
= |H\G/H|,
using Frobenius reciprocity twice and Mackeys formula. Now |H\G/H| is the number of
orbits of H (the stabilizer of a point) on G/H. By Lemma 5.5 this number is 2 if G acts
2-transitively on , and otherwise it is greater than 2 (since || 2 was a hypothesis).
Writing C[G/H] = S1 Sn as a direct sum of simple representations we have
G
h1 G
H , 1 H iG n,
and we get the value 2 for the inner product if and only if there are 2 simple representations
in this expression, and they are non-isomorphic. This is equivalent to requiring that V is
simple, because it could only be the trivial representation if G acts trivially on G/H, which
our hypotheses exclude. In any case we deduce that V is not the trivial representation.
(5.7) EXAMPLE. Let = {1, . . . , n} acted upon transitively by Sn and also by An .
Then C
= C V where V is a simple representation of Sn , which remains simple on
restriction to An provided n 4.
We now turn to Cliffords theorem, which we present in a weak and a strong form.
The weak form is used as a step in proving the strong form a little later, and as a result
in its own right it only has force in a situation where |G| is not invertible in the ground
ring. In these versions of Cliffords theorem we make the hypothesis that the ground ring
is a field, but this is no loss of generality in view of Exercise 13 from Chapter 1.
(5.8) THEOREM (Weak form of Cliffords theorem). Let k be any field, U a simple
kG-module and N a normal subgroup of G. Then U G
N is semisimple as a kN -module.
Proof. Let V be any simple kN -submodule of U G
N . For every g G, gV is also
1
a kN -submodule since if n N we have n(gv) = g(g ng)v gV , using the fact that
74
Peter Webb
75
(5.10) THEOREM. Let k be any algebraically closed field. If G is abelian then every
simple kG-module has dimension 1.
Proof. Consider a simple kG-module S and let g G. In its action on S, g has
an eigenvalue , with non-zero eigenspace S . Since all elements h G commute with g
we have hS = S (by the argument that if v S then gv = v, so g(hv) = h(gv) =
hv = hv, so hv S ; but also the action of h is invertible). Thus S is a kG-submodule
of S, so S = S by simplicity of S. It follows that every element g G acts by scalar
multiplication on S, and such a simple module S must have dimension 1.
As a consequence of the last result and Proposition 4.5, over an algebraically closed
field the degree 1 representations of any group G are the same as the representations of
G/G , lifted to G via the quotient homomorphism G G .
The next result about simple representations of p-groups is true as stated when k
has characteristic p, but it has no force in that situation because (as we will see in the
next section) the only simple representation of a p-group in characteristic p is the trivial
representation. We are thus only really interested in the following result over fields of
characteristic other than p, and in particular over fields of characteristic 0.
(5.11) COROLLARY. Let k be any algebraically closed field and G a p-group. Then
every simple module for G has the form U G
H where U is a 1-dimensional module for some
subgroup H.
Proof. We proceed by induction on |G|. Let : G GL(S) be a simple representation
of G over k and put N = Ker . Then S is really a representation of G/N . If N 6= 1 then
G/N is a group of smaller order than G, so by induction S has the claimed structure as
a representation of G/N , and hence also as a representation of G. Thus we may assume
N = 1 and G embeds in GL(S).
If G is abelian then all simple representations are 1-dimensional, so we are done.
Assume now that G is not abelian. Then G has a normal abelian subgroup A that is not
central. To construct this subgroup A, let Z2 (G) denote the second centre of G, that is,
the preimage in G of Z(G/Z(G)). If x is any element of Z2 (G) Z(G) then A = hZ(G), xi
is a normal abelian subgroup not contained in Z(G).
We apply Cliffords theorem:
a1
ar
S G
A = S1 Sr
a1
a1
and S = V G
K where V = S1 and K = StabG (S1 ). We argue that V must be a
simple kK-module, since if it had a proper submodule W then W G
K would be a proper
submodule of S, which is simple. If K 6= G then by induction V = U K
H where U is
K
G
G
1-dimensional, and so S = (U H ) K = U H has the required form.
76
We show finally that the case K = G cannot happen. For if it were to happen then
S
S1a1 , and since A is abelian dim S1 = 1. The elements of A must therefore act via
scalar multiplication on S. Since such an action would commute with the action of G,
which is faithfully represented on S, we deduce that A Z(G), a contradiction.
G
A=
The above result is evidently useful if we are constructing the character table of a
p-group, because it says that we need look no further than induced characters. We note
that the conclusion of Corollary 5.11 also applies to supersolvable groups, which again
have the property, if they are not abelian, that they have a non-central normal abelian
subgroup.
A representation of the form U G
H for some subgroup H and with U a 1-dimensional
representation of H is said to be monomial. A group G all of whose irreducible complex
representations are monomial is called an M-group. Thus p-groups (and also supersoluble
groups) are M-groups.
Summary of Chapter 5
1. The Mackey formula: induction followed by restriction is a sum over double cosets
of restriction followed by conjugation followed by induction.
2. A permutation representation is 2-transitive if and only if the complex permutation
module has two summands.
3. Cliffords theorem: the restriction of a simple module to a normal subgroup is
semisimple, and the module is induced from the stabilizer of a homogeneous component.
4. For a p-group over an algebraically closed field, every simple module is induced
from a 1-dimensional module.
Peter Webb
77
is a well-defined map that is a projection of the H-fixed points onto the G-fixed points.
In particular, this map is surjective.
(b) Show that if V G
H is semisimple as a kH-module then V is semisimple as a kG-module.
5. Let H be a normal subgroup of G and suppose that k is a field of characteristic p.
(a) Let p6 |G : H|. Show that if U is a semisimple kH-module then U G
H is a semisimple
kG-module.
(b) Let p |G : H|. Show by example that if U is a semisimple kH-module then it need
not be the case that U G
H is a semisimple kG-module.
6. Let H be a subgroup of G of index 2 (so that H is normal in G) and let k be a
field whose characteristic is not 2. The homomorphism G {1} k with kernel H is a
1-dimensional representation of G that we will call . Let S, T be simple kG-modules and
G
let U, V be simple kH-modules. You may assume that U G
H and V H are semisimple
(this is proved as Exercise 5(a)). Let g G H.
(a) Show that S G
H is the direct sum of either 1 or 2 simple kH-modules.
(b) Show that U G
H is the direct sum of either 1 or 2 simple kG-modules.
In the following questions, notice that
G
G
S G
H H = S (k H ) = S (k ) = S (S ).
and
G
HomkG (U G
H , V H ).
78
(iii) S = S .
(d) Show that the following are equivalent:
(i) U is the restriction to H of a kG-module,
(ii) U G
H is not simple,
(iii) U = g U .
G
U = V.
(g) We place an equivalence relation 1 on the simple kG-modules and an equivalence
relation 2 on the simple kH-modules:
S 1 T S
= T or S
=T
U 2 V U
= V or U
= g V.
G
Show that induction G
H and restriction H induce mutually inverse bijections between the
equivalence classes of simple kG-modules and of simple kH-modules in such a way that an
equivalence class of size 1 corresponds to an equivalence class of size 2, and vice-versa.
(h) Show that the simple kG-modules of odd degree restrict to simple kH-modules,
and the number of such modules is even.
(i) In the case where G = S4 , H = A4 and k = C, show that there are three equivalence
classes of simple characters under 1 and 2 . Verify that SA44 and SA44 give mutually inverse
bijections between the equivalence classes.
7. Let G = GL(3, 2) be the group of 3 3 invertible matrices over k = F2 and let
a b
H ={ c d
e f
0
0 a, b, c, d, e, f, 1 F2 , (ad bc) 6= 0}
1
You may assume from group theory that |G| = 168. Let V be the natural 3-dimensional
space of column vectors on which G-acts.
(a) Show that |H| = 24, so that |G : H| = 7.
(b) Show that V is simple as a kG-module.
(c) Show that as a kH-module V G
H has a simple socle with trivial H action, and such
that the quotient of V by the socle is a simple 2-dimensional module.
G
(d) Show that dim HomkG (k G
H , V ) = 1 and dim HomkG (V, k H ) = 0.
(e) Show that k G
H is not semisimple, thereby showing that even if p 6 |G : H| it need
not be the case that the induction of a simple module is semisimple when H is not
normal.
Peter Webb
79
G
(f) Show that dim HomkG (k G
H , V ) = 0 and dim HomkG (V , k H ) = 1. Show that
V 6
= V . Show that k G
H is the direct sum of the trivial module k and a 6-dimensional
module that has socle V and socle quotient V .
8. Find the complete list of subgroups H of the dihedral group D8 such that the
2-dimensional simple representation over C can be written U G
H for some 1-dimensional
representation U of H. Do the same thing for the quaternion group Q8 .
9. Compute the character tables of the generalized quaternion group of order 16
Q16 = hx, y x8 = 1, x4 = y 2 , yxy 1 = x1 i
SD16 = hx, y x8 = y 2 = 1, yxy 1 = x3 i.
80
Cyclic p-groups
We start by describing representations of cyclic p-groups because it is possible to give
the complete picture in an elementary fashion. In the first proposition we reduce their
study to that of modules for a principal ideal domain. When G is cyclic of order N we
have already made use of an isomorphism between kG and k[X]/(X N 1) in Exercise 9
from Section 2. When N is a power of p we can express this slightly differently.
n
(6.1) PROPOSITION. Let k be a field of characteristic p and let G = hg g p = 1i
n
be cyclic of order pn . Then there is a ring isomorphism kG
= k[X]/(X p ), where k[X] is
the polynomial ring in an indeterminate X.
Proof. We define a mapping
n
G k[X]/(X p )
g s 7 (X + 1)s .
Since
this mapping is a group homomorphism to the unit group of k[X]/(X p ), and hence it
extends to a linear map
n
kG k[X]/(X p )
that is an algebra homomorphism. Since g s is sent to X s plus terms of lower degree, the
n
n
images of 1, . . . , g p 1 form a basis of k[X]/(X p ). The mapping therefore gives a bijection
n
between a basis of kG and a basis of k[X]/(X p ), and so is an isomorphism.
Peter Webb
81
Direct sum decompositions of modules are the first consideration in describing their
structure. We say that a module U for a ring A is indecomposable if it cannot be expressed
as a direct sum of two modules except in a trivial way, that is, if U
= V W then either
V = 0 or W = 0. When A is an algebra over a field, by repeatedly expressing summands of
a module as further direct sums we can express any finite dimensional module as a direct
sum of indecomposable direct summands. It is useful to know, but we will not prove it until
Theorem 11.16, that for each module these summands are determined up to isomorphism,
independently of the choice of direct sum decomposition. This is the content of the KrullSchmidt Theorem. We point out that, since we need not be in characteristic zero when
group representations are semisimple, indecomposable modules need not be simple. This
is the point of introducing the new terminology! An example of an indecomposable module
that is not simple was given 1.3, and we will see many more examples.
A module over a ring is said to be cyclic if it can be generated by one element. We now
exploit the structure theorem for finitely-generated modules over a principal ideal domain,
which says that such modules are direct sums of cyclic modules.
(6.2) THEOREM. Let k be a field of characteristic p. Every finitely-generated
n
k[X]/(X p )-module is a direct sum of cyclic modules Ur = k[X]/(X r ) where 1 r pn .
The only simple module is the 1-dimensional module U1 . Each module Ur has a unique
composition series, and hence is indecomposable. From this it follows that if G is cyclic of
order pn then kG has exactly pn indecomposable modules, one of each dimension i with
1 i pn , each having a unique composition series.
n
Proof. The modules for k[X]/(X p ) may be identified with the modules for k[X] on
n
which X p acts as zero. Every finitely-generated k[X]-module is a direct sum of modules
n
k[X]/I where I is an ideal. Hence every k[X]/(X p )-module is a direct sum of modules
n
n
k[X]/I on which X p acts as zero, which is to say (X p ) I. The ideals I that satisfy this
n
last condition are the ideals (a) where a X p . This forces I = (X r ) where 1 r pn ,
and k[X]/I = Ur .
The submodules of Ur must have the form J/(X r ) where J is some ideal containing
r
(X ), and they are precisely the submodules in the chain
0 (X r1 )/(X r ) (X r2 )/(X r ) (X)/(X r ) Ur .
This is a composition series, since each successive quotient has dimension 1, and since it
is a complete list of submodules, it is the only one. If we could write Ur = V W as
a non-trivial direct sum, then Ur would have at least 2 composition series, obtained by
taking first a composition series for V , then one for W , or vice-versa. Hence each Ur is
indecomposable and we have a complete list of the indecomposable modules. The only Ur
that is simple is U1 , which is the trivial module.
The final identification of the indecomposable kG-modules comes from the isomorphism in Proposition 6.1.
82
1 0
.
.. ..
.
.
1
Translating now to modules for kG where G is a cyclic p-group, the generator g acts on
Ur as X + 1, which has matrix
1
1
1
..
..
1
.
1
Thus we see that the indecomposable kG-modules are exactly given by specifying that the
generator g acts via a matrix that is a single Jordan block, of size up to pn . It is helpful
to picture Ur using a diagram
X=g1
y
X=g1
y
Ur = .
..
X=g1
y
that may be interpreted by saying that the vertices are in bijection with a basis of Ur , and
the action of X or g 1 is given by the arrows. Where no arrow is shown starting from a
particular vertex (as happens in this case only with the bottom vertex), the interpretation
is that X and g 1 act as zero.
Peter Webb
83
84
element of order prime to p. For, suppose h H were to have order prime to p. Then
G
kG G
hhi would be a semisimple khhi-module that is the direct sum of modules S hhi with
S a simple kG-module. Since h acts trivially on all of these, it must act trivially on kG,
which is a contradiction. Therefore H is a p-group, and since it is normal, Op (G) H.
We therefore have equality. The last sentence is immediate.
(6.5) Examples. When G is nilpotent (and this applies, of course, if G is abelian) it
is a direct product G = H K where H is a Sylow p-subgroup of G, and so Op (G) = H.
Thus the simple kG-modules may be identified with the simple kK-modules when k has
characteristic p. For a different example, let k be a field of characteristic 2, and consider
the representations of A4 over k. Since O2 (A4 ) = C2 C2 , the simple kA4 modules are
the simple C3 = A4 /O2 (A4 )-representations, made into representations of A4 . Now kC3
is semisimple, and if k contains a primitive cube root of unity (i.e. if F4 k) there are
three 1-dimensional simple representations, on which the generator of C3 acts as 1, or
2.
M
iI
U/Mi
Peter Webb
85
(2) Suppose further that U has the descending chain condition on submodules. Then
U/ Rad U is a semisimple module, and Rad U is the unique smallest submodule of U
with this property.
Proof. (1) Let I be a subset of {1, . . . , n} maximal with the property that the
T
T
quotient homomorphisms U/( iI Mi ) U/Mi induce an isomorphism U/( iI Mi )
=
L
T
Mn and argue by contradiction. If it
iI U/Mi . We show that
iI Mi = M1
T
were not the case, there would exist Mj with iI Mi 6 Mj . Consider the homomorphism
f :U (
M
iI
U/Mi ) U/Mj
whose components are the quotient homomorphisms U U/Mk . This has kernel Mj
T
iI Mi , and it will suffice to show that f is surjective, because this will imply that the
larger set I {j} has the same property as I, thereby contradicting the maximality of I.
T
To show that f is surjective let g : U U/ iI Mi U/Mj and observe that
T
( iI Mi ) + Mj = U since the left-hand side is strictly larger than Mj , which is maxT
imal in U . Thus if x U we can write x = y + z where y iI Mi and z Mj . Now
T
T
g(y) = (0, x + Mj ) and g(z) = (x + iI Mi , 0) so that both summands U/ iI Mi and
U/Mj are contained in the image of g and g is surjective. Since f is obtained by composing
T
L
g with the isomorphism that identifies U/ iI Mi with iI U/Mi , we deduce that f is
surjective.
(2) By the assumption that U has the descending chain condition on submodules,
Rad U must be the intersection of finitely many maximal submodules. Therefore U/ Rad U
is semisimple by part (1). If V is a submodule such that U/V is semisimple, say U/V
=
proj.
S1 Sn where the Si are simple modules, let Mi be the kernel of U U/V Si .
Then Mi is maximal and V = M1 Mn . Thus V Rad U , and Rad U is contained
in every submodule V for which U/V is semisimple.
We define the radical of a ring A to be the radical of the regular representation
Rad A A and write simply Rad A. We present some identifications of the radical that are
very important theoretically, and also in determining what it is in particular cases.
(6.7) PROPOSITION. Let A be a ring. Then,
(1) Rad A = {a A a S = 0 for every simple A-module S}, and
(2) Rad A is a 2-sided ideal of A.
(3) Suppose further that A is a finite-dimensional algebra over a field. Then
(a) Rad A is the smallest left ideal of A such that A/ Rad A is a semisimple A-module,
(b) A is semisimple if and only if Rad A = 0,
(c) Rad A is nilpotent, and is the largest nilpotent ideal of A.
86
(d) Rad A is the unique ideal U of A with the property that U is nilpotent and A/U
is semisimple.
Proof. (1) Given a simple module S and 0 6= s S, the module homomorphism
A A S given by a 7 as is surjective and its kernel is a maximal left ideal Ms . Now if
a Rad A then a Ms for every S and s S, so as = 0 and a annihilates every simple
module. Conversely, if a S = 0 for every simple module S and M is a maximal left ideal
then A/M is a simple module. Therefore a (A/M ) = 0, which means a M . Hence
T
a maximalM M = Rad A.
(2) Being the intersection of left ideals, Rad A is also a left ideal of A. Suppose that
a Rad A and b A, so a S = 0 for every simple S. Now a bS a S = 0 so ab has the
same property that a does.
(3) (a) and (b) are immediate from 6.6. We prove (c). Choose any composition series
0 = An An1 A1 A0 = A A
of the regular representation. Since each Ai /Ai+1 is a simple A-module, Rad A Ai Ai+1
by part (1). Hence (Rad A)r A Ar and (Rad A)n = 0.
Suppose now that I is a nilpotent ideal of A, say I m = 0, and let S be any simple
A-module. Then
0 = I m S I m1 S IS S
is a chain of A-submodules of S that are either 0 or S since S is simple. There must be
some point where 0 = I r S 6= I r1 S = S. Then IS = I I r1 S = I r S = 0, so in fact
that point was the very first step. This shows that I Rad A by part (1). Hence Rad A
contains every nilpotent ideal of A, so is the unique largest such ideal.
Finally (d) follows from (a) and (c): these imply that Rad A has the properties stated
in (d); and, conversely, these conditions on an ideal U imply by (a) that U Rad A, and
by (c) that U Rad A.
Note that if I is a nilpotent ideal of A then it is always true that I Rad(A) without
the assumption that A is a finite dimensional algebra. The argument given to prove part
3c of Proposition 6.7 shows this.
For any group G and commutative ring R with a 1, the ring homomorphism
: RG R
g 7 1 for all g G
is called the augmentation map. As well as being a ring homomorphism it as a homomorphism of RG-modules, in which case it expresses the trivial representation as a homomorphic image of the regular representation. The kernel of is called the augmentation ideal,
P
and is denoted IG. Evidently IG consists of those elements gG ag g RG such that
P
gG ag = 0. We now show that when k is a field of characteristic p and G is a p-group
this construction gives the radical of kG.
Peter Webb
87
gG
ag g =
gG
ag g
gG
ag 1 =
16=gG
ag (g 1)
88
Let us show that the submodule Rad A U in (1)(c) satisfies condition (1)(b). Firstly
U/(Rad AU ) is a module for A/ Rad A, which is a semisimple algebra. Hence U/(Rad AU )
is a semisimple module and so Rad A U contains the submodule of (1)(b). On the other
hand if V U is a submodule for which U/V is semisimple then Rad A (U/V ) = 0 by
6.7, so V Rad A U . In particular, the submodule of (1)(b) contains Rad A U . This
shows that the descriptions in (1)(b) and (1)(c) are equivalent.
To show that they give the same submodule as (1)(a), observe that if V is any maximal
submodule of U , then as above (since U/V is simple) V Rad A U , so the intersection of
maximal submodules of U contains Rad A U . The intersection of maximal submodules of
the semisimple module U/(Rad A U ) is zero, so this gives a containment the other way,
since they all correspond to maximal submodules of U . We deduce that the intersection
of maximal submodules of U equals Rad A U .
For the conditions in (2), observe that {u U Rad A u = 0}, this is the largest submodule of U annihilated by Rad A. It is thus an A/ Rad A-module and hence is semisimple.
Since every semisimple submodule of U is annihilated by Rad A, it is the unique largest
such submodule.
(6.10) Example. Consider the situation of 6.1 and 6.2 in which G is a cyclic group of
order pn and k is a field of characteristic p. We see that Rad Ur
= Ur1 and Soc Ur
= U1
n
for 1 r p , taking U0 = 0.
We now iterate the notions of socle and radical: for each A-module U we define
inductively
Radn (U ) = Rad(Radn1 (U ))
Socn (U )/ Socn1 (U ) = Soc(U/ Socn1 U ).
It is immediate from 6.9 that
Radn (U ) = (Rad A)n U
Socn (U ) = {u U (Rad A)n u = 0}
Rad2 U Rad U U
0 Soc U Soc2 U
that are called, respectively, the radical series and socle series of U . The radical series of
U is also known as the Loewy series of U . The quotients Radn1 (U )/ Radn (U ) are called
the radical layers, or Loewy layers of U , and the quotients Socn (U )/ Socn1 (U ) are called
the socle layers of U .
The next corollary is a deduction from 6.9, and again it is true without the hypothesis
that the modules be finite-dimensional.
Peter Webb
89
90
The common length of the radical series and socle series of U is called the Loewy
length of the module U , and from the description of the terms of these series we see it is
the least integer n such that (Rad A)n U = 0.
We conclude this section by mentioning without proof the theorem of Jennings on the
Loewy series of kG when G is a p-group and k is a field of characteristic p and summarize
its implications. For proofs, see Bensons book [Ben]. Jennings constructs a decreasing
series of subgroups
r (G) = {g G g 1 modulo Radr (kG)},
that is sometimes called the Jennings series of G. This series of subgroups has the properties
1) [r , s ] r+s ,
2) g p ip for all g i ,
3) r /2r is elementary abelian.
Furthermore, we may generate r as
(p)
where 1 = G, r/p is the least integer greater than or equal to r/p, and r is the set
of pth powers of elements of r . Evidently the first term in this series is 1 (G) = G and
we may see that the second term 2 (G) is the Frattini subgroup of G (the smallest normal
subgroup of G for which the quotient is elementary abelian). After that the terms need to
be calculated on a case-by-case basis.
For each i 1 let di be the dimension of i /i+1
as a vector space over Fp , and choose
any elements xi,s of G such that the set {xi,s i+1 1 s di } forms a basis for i /i+1 .
Q i,s
Let x
i,s = xi,s 1 kG. There are |G| products of the form x
i,s , where the factors are
taken in some predetermined order, and 0 i,s p 1. The weight of such a product
P
is defined to be
ii,s . Jennings theorem states that the set of products of weight w is
contained in Radw (kG), and forms a basis of Radw (kG) modulo Radw+1 (kG). Thus the
set of all these products is a basis for kG compatible with the powers of the radical.
For example, if we take an element x of order 4 and an element y of order 2 that
generate the dihedral group of order 8, so D8 = hx, y x4 = y 2 = 1, yxy = x1 i, we have
1 = D8 , 2 = hx2 i, 3 = 1. We may choose x1,1 = x, x1,2 = y, x2,1 = x2 , and note that
x
2 = x2 . Now the products x
1,1 x
22,1 y1,2 = x
1,1 +22,1 y1,2 , where 0 i,s 1, form a
basis of kD8 that is compatible with the powers of the radical. In this special case, these
elements may be simplified as x
i yj with 0 i 3 and 0 j 1.
When doing calculations with group rings of p-groups in characteristic p the basis
given by Jennings is often to be preferred over the basis given by the group elements. This
basis gives a description of the powers of the radical in group-theoretic terms, and it allows
us to deduce a result about the socle series as well. Since each element x
i,s of weight i
Peter Webb
91
contributes factors of weights 0, i, 2i, 3i, . . . , (p 1)i in Jennings basis , the total number
of products of weight w is the coefficient of tw in
2
(1 + t + t + + t
p1 d1
) (1 + t + + t
2(p1) d2
Y (1 tip ) di
i1
(1 ti )
P
and this equals the polynomial w0 (dim Radw (kG)/ dim Radw+1 (kG))tw . We see from
this that the dimensions of the factors in the radical series of kG are symmetric, in that
they are the same if taken in reverse order. Jumping ahead of ourselves for a moment and
using the fact that kG
= kG (see Chapter 8) and also using Exercise 6 to this section,
we may see that when G is a p-group and k is a field of characteristic p the terms of the
radical series and the socle series of kG coincide, although these terms appear in the two
series in the opposite order.
Summary of Chapter 6
n
92
and
Radn (U ) = {f U f (Socn (U )) = 0}.
n
n+1
Deduce that Socn+1 (U )/ Socn (U )
(U )) as kG-modules. [Hint:
= (Rad (U )/ Rad
recall Exercise 9 to Chapter 4.]
7. Show that for each RG-module U , U/(IG U ) is the largest quotient of U on which
G acts trivially. Prove also that U/(IG U )
= R RG U .
[The first sentence means that G does act trivially on the given quotient; and if V is any
submodule of U such that G acts trivially on U/V , then V IG U . By analogy with the
notation for fixed points, this largest quotient on which G acts trivially is called the fixed
quotient of U and is denoted UG := U/(IG U ).]
The next five exercises give a direct proof of the result that is part of Proposition 6.8,
that for a p-group in characteristic p the augmentation ideal is nilpotent.
8. Show that if elements g1 , . . . , gn generate G as a group, then (g1 1), . . . , (gn 1)
generate the augmentation ideal IG as a left ideal of kG.
[Use the formula (gh 1) = g(h 1) + (g 1).]
9. Suppose that k is a field of characteristic p and G is a p-group. Prove that each
element (g 1) is nilpotent. (More generally, every element of IG is nilpotent.)
Peter Webb
93
modulo kG (IG )
Let k be a field of characteristic 2. Show that when n is a power of 2, each power (ID2n )r
of the augmentation ideal is spanned modulo (ID2n )r+1 by the two products (x 1)(y
1)(x 1)(y 1) and (y 1)(x 1)(y 1)(x 1) of length r. Hence calculate the
Loewy length of kD2n and show that Rad(kD2n )/ Soc(kD2n ) is the direct sum of two
kD2n -modules that are uniserial.
15. When n 3, the generalized quaternion group of order 2n has a presentation
n1
n2
= 1, y 2 = x2 , yxy 1 = x1 i.
Q2n = hx, y x2
Let k be a field of characteristic 2. Show that when r 1 each power (IQ2n )r of the
augmentation ideal is spanned modulo (IQ2n )r+1 by (x 1)r and (x 1)r1 (y 1). Hence
calculate the Loewy length of kQ2n .
16. Let H be a subgroup of G and let IH be the augmentation ideal of RH, which
we may regard as a subset of RG. Show that RG IH
= IH G
H as RG-modules, and
G
that RG/(RG IH) = R H as RG-modules. Show also that RG/(RG IH) is the largest
quotient of RG on which H acts trivially when acting from the right.
94
17. (a) Let G be any group and IG ZG the augmentation ideal over Z. Prove that
IG/(IG)2
= G/G as abelian groups.
[Consider the homomorphism of abelian groups IG G/G given by g 1 7 gG . Use
the formula ab 1 = (a 1) + (b 1) + (a 1)(b 1) to show that (IG)2 is contained in
the kernel, and that the homomorphism G/G IG/(IG)2 given by gG 7 g 1 + (IG)2
is well defined.]
(b) For any group G write d(G) for the smallest size of a set of generators of G as a
group, and if U is a ZG-module write d(U ) for the smallest size of a set of generators of
U as a ZG-module. Use Exercise 8 to show that d(G/G ) d(IG) d(G) with equality
when G is a p-group. [For the final equality use properties of the Frattini subgroup of G.]
(c) If now R is any commutative ring with 1 and IG RG is the augmentation ideal
of G over R, show that IG/(IG)2
= R Z G/G as R-modules. When G is a p-group and
R is a field of characteristic p, show again that d(G) = d(IG).
18. Let be a transitive G-set and k a field. Let k be the corresponding permutation
P
module. There is a homomorphism of kG-modules : k k defined as ( a ) =
P
P
a . Let =
k.
(a) Show that every kG-module homomorphism k k is a scalar multiple of .
G
(b) Show that the fixed points of G on k are k
= k .
(c) Show that () = 0 if and only if char k ||, and that if this happens then
Rad k and the trivial module k occurs as a composition factor of k with multiplicity
2.
(d) Show that if () 6= 0 then is a split epimorphism and 6 Rad k.
(e) Show that kG is semisimple if and only if the regular representation kG has the
trivial module k as a direct summand (i.e. k is a projective module).
19. Let be a transitive G-set for a possibly infinite group G and let R be the
corresponding permutation module. Show that is infinite if and only if (R)G = 0 and
deduce that G is infinite if and only if (RG)G = 0.
20. Let Ur be the indecomposable kCp -module of dimension r, 1 r p, where k is
a field of characteristic p. Prove that Ur
= S r1 (U2 ), the (r 1) symmetric power.
[One way to proceed is to show that if Cp = hgi then (g 1)r1 does not act as zero on
S r1 (U2 ) and use the classification of indecomposable kCp -modules.]
21. Let G = SL(2, p), the group of 2 2 matrices over Fp that have determinant 1,
where p is a prime. The subgroups
P1 =
1
0
Fp ,
P2 =
0
1
Fp
have order p. Let U2 be the natural 2-dimensional module on which G acts. When
0 r p 1 prove that S r (U2 ) is a uniserial Fp P1 -module, and also a uniserial Fp P2 module, but that the only subspaces of S r (U2 ) that are invariant under both P1 and P2
are 0 and S r (U2 ). Deduce that S r (U2 ) is a simple Fp G-module.
Peter Webb
95
1 0
Show further when p is odd that the matrix
acts as the identity on S r (U2 )
0 1
if and only if r is even, and hence that we have constructed (p+1)/2 simple representations
of
1 0
P SL(2, p) := SL(2, p)/
.
0 1
Printed Oct. 13, 2014
[Background to the question that is not needed to solve it: |G| = p(p2 1); both P1 and P2
are Sylow p-subgroups of G. In fact, the simple modules constructed here form a complete
list of the simple Fp SL(2, p)-modules.]
22. Let g be an endomorphism of a finite-dimensional vector space V over a field k
of characteristic p, and suppose that g has finite order pd for some d. Show that as a
khgi-module, V has an indecomposable direct summand of dimension at least pd1 + 1.
[You may assume the classification of indecomposable modules for cyclic p-groups in characteristic p.]
Deduce that if such an endomorphism g fixes pointwise a subspace of V of codimension 1 then g has order p or 1.
[An endomorphism (not necessarily of prime-power order) that fixes a subspace of codimension 1 is sometimes referred to as a reflection in a generalized sense.]
23. Let k be a field of characteristic p and suppose N is a normal subgroup of G that
is a Sylow p-subgroup of G. Show that Rad kG = kG Rad kN .
[Use questions 3 and 10, and show that k[G/N ] is the largest semisimple quotient of kG.]
24. Let A be a finite dimensional algebra over a field and let U be an arbitrary Amodule. As in the text, we define Rad U to be the intersection of the maximal submodules
of U .
(a) Use exercises 11 and 12 from Chapter 1 to show that Proposition 6.9 holds without
the hypothesis of finite generation. That is, show that Rad U = Rad A U , and that Rad U
is the smallest submodule of U with semisimple quotient. Show also that Soc U , defined as
the sum of the simple submodules of U , is the largest semisimple submodule of U , and it is
the set of elements of U annihilated by Rad A. [It is a question of copying the arguments
from 6.9. Note that the property of A being used is that A/ Rad A is semisimple.]
(b) Show that if U 6= 0 then U has a non-zero finite-dimensional homomorphic image.
[Use the fact that Rad A is nilpotent.]
(c) Show that each proper submodule of U is contained in a maximal submodule.
25. Let A be a finite dimensional algebra over a field and let U be an A-module. Write
(U ) for the Loewy length of U .
(a) Suppose V is a submodule of R. Show that (V ) (U ) and (U/V ) (U ).
Show by example that we can have equality here even when 0 < V < U .
(b) Suppose that U1 , . . . , Un are submodules of U for which U = U1 + + Un . Show
that (U ) = max{(Ui) 1 i n}.
26. (This exercise extends the theorem of Burnside presented in Chapter 2 Exercise
8 to non-semisimple algebras.) Let A be a finite dimensional algebra over a field k, let
96
b
p
Ur,G G
H = (Ua+1,H ) (Ua,H )
nt
b
nt
nt
to show
as kH-modules. [If G = hxi, H = hxp i use the fact that xp 1 = (x 1)p
pnt
1 that is non-zero on Ur,G is the ath power, and use this
that the largest power of x
to identify the summands in a decomposition of Ur,G G
H .]
Peter Webb
97
Proof. The proof is standard, and we merely remind the reader. If {xi i I} is
P
P
a basis, then given we may define ( iI ai xi ) = iI ai (xi ) and this is evidently
the unique module homomorphism extending
. Conversely if condition (b) holds we may
construct the free module F with {xi i I} as a basis and use the condition to construct
a homomorphism from M F that is the identity on {xi i I}. The fact just shown
that the free module also satisfies condition (b) allows us to construct a homomorphism
F M that is again the identity on {xi i I}, and the two homomorphisms have
composites in both directions
that are the identity, since these are the unique extensions
of the identity map on {xi i I}.
98
k
0 L
LN
where 1 and 2 are inclusion into the first summand and projection onto the second
summand,
(4) for every module U the sequence
0 HomA (U, L) HomA (U, M ) HomA (U, N ) 0
is exact,
(4 ) for every module U the sequence
Peter Webb
99
two conditions because the commutative diagram produces similar commutative diagrams
after applying HomA (U, ) and HomA (, U ).
Conversely if condition (1) is satisfied, so that = 1L for some homomorphism
: M L, we obtain a commutative diagram as in (3) on taking the components of
to be and . If condition (2) is satisfied we obtain a commutative diagram similar to
the one in (3) but with a homomorphism : L N M in the wrong direction, whose
components are and a splitting of . We obtain the diagram of (3) on showing that in
any such diagram the middle vertical homomorphism must be invertible.
The fact that the middle homomorphism in the diagram must be invertible is a consequence of both the five lemma and the snake lemma in homological algebra. We leave
it here as an exercise.
Finally if (4) holds then on taking U to be N we deduce that the identity map on N
is the image of a homomorphism : U M , so that 1N = and is split epi, so that
(2) holds. Equally if (4 ) holds then taking U to be L we see that the identity map on L
is the image of a homomorphism : M U , so that 1L = and (1) holds.
In the event that and are split, we say that the short exact sequence in Proposition
7.2 is split. Notice that whenever : M N is an epimorphism it is part of the short exact
y
V
100
We say that a module P satisfying any of the four conditions of 7.3 is projective.
Notice that direct sums and also direct summands of projective modules are projective.
An indecomposable module that is projective is an indecomposable projective module,
and these modules will be very important in our study. In other texts the indecomposable
projective modules are also known as PIMs, or Principal Indecomposable Modules, but we
will not use this terminology here.
We should also mention injective modules, which enjoy properties similar to those of
projective modules, but in a dual form. We say that a module I is injective if and only if
whenever there are morphisms
I
x
= A1 Ar
= Ae1 Aer ,
Peter Webb
101
for the Aei are evidently submodules of A, and their sum is A since if x A then x =
P
xe1 + + xer . The sum is direct since if x Aei j6=i Aej then x = xei and also
P
P
x = j6=i aj ej so x = xei = j6=i aj ej ei = 0.
Conversely, suppose that A A = A1 Ar is a direct sum of submodules. We
may write 1 = e1 + + er where ei Ai is a uniquely determined element. Now
ei = ei 1 = ei e1 + + ei er is an expression in which ei ej Aj , and since the only such
expression is ei itself we deduce that
n
ei if i = j,
ei ej =
0 otherwise.
102
but in fact many of the observations we are about to make also hold over the field F2 . By
Proposition 4.5 the 1-dimensional representations of S3 are the simple representations of
S3 /S3
= C2 , lifted to S3 . But F4 C2 has only one simple module, namely the trivial module,
by Proposition 6.3, so this is the only 1-dimensional F4 S3 -module. The 2-dimensional
representation of S3 constructed in Chapter 1 over any coefficient ring is now seen to be
simple here, since otherwise it would have a trivial submodule; but the eigenvalues of the
element (1, 2, 3) on this module are and 2 , where F4 is a primitive cube root of 1,
so there is no trivial submodule.
Let K = h(1, 2, 3)i be the subgroup of S3 of order 3. Now F4 K is semisimple with
three 1-dimensional representations on which (1, 2, 3) acts as 1, and 2 , respectively. In
fact
F4 K = F4 Ke1 F4 Ke2 F4 Ke3
where
e1 = () + (1, 2, 3) + (1, 3, 2)
e2 = () + (1, 2, 3) + 2 (1, 3, 2)
e3 = () + 2 (1, 2, 3) + (1, 3, 2)
are orthogonal idempotents in F4 K. We may see that these are orthogonal idempotents
by direct calculation, but it can also be seen by observing that the corresponding elements
2i
of CK with replaced by e 3 are orthogonal and square to 3 times themselves (Theorem
2i
2i
3.23), and lie in Z[e 3 ]K. Reduction modulo 2 gives a ring homomorphism Z[e 3 ] F4
that maps these elements to e1 , e2 and e3 , while retaining their properties. Thus
F4 S3 = F4 S3 e1 F4 S3 e2 F4 S3 e3
and we have constructed modules F4 S3 ei that are projective. We have not yet shown that
they are indecomposable.
We easily compute that
(1, 2, 3)e1 = e1 ,
(1, 2, 3)e2 = 2 e2 ,
(1, 2, 3)e3 = e3
and from this we see that K F4 ei = F4 ei for all i. Since S3 = K (1, 2)K we have
F4 S3 ei = F4 ei F4 (1, 2)ei , which has dimension 2 for all i. We have already seen that
when i = 2 or 3, ei is an eigenvector for (1, 2, 3) with eigenvalue or 2 , and a similar
calculation shows that the same is true for (1, 2)ei . Thus when i = 2 or 3, F4 S3 ei has
no trivial submodule and hence is simple by the observations made at the start of this
example. We have an isomorphism of F4 S3 -modules
F4 S3 e2 F4 S3 e3
e2 7 (1, 2)e3
(1, 2)e2 7 e3 .
Peter Webb
103
P
On the other hand F4 S3 e1 has fixed points F4 gS3 g of dimension 1 and so has two
composition factors, which are trivial. On restriction to F4 h(1, 2)i it is the regular representation, and it is a uniserial module.
We see from all this that F4 S3 = 11 2 2, in a diagrammatic notation. Thus the
2-dimensional simple F4 S3 -module is projective, and the trivial module appears as the
unique simple quotient of a projective module of dimension 2 whose socle is also the trivial
module. These summands of F4 S3 are indecomposable, and so e1 , e2 and e3 are primitive
P
idempotents in F4 S3 . We see also that the radical of F4 S3 is the span of gS3 g.
Projective covers, Nakayamas lemma and lifting of idempotents
We now develop the theory of projective covers. We first make the definition that an
essential epimorphism is an epimorphism of modules f : U V with the property that
no proper submodule of U is mapped surjectively onto V by f . An equivalent formulation
is that whenever g : W U is a map such that f g is an epimorphism, then g is an
epimorphism. One immediately asks for examples of essential epimorphisms, but it is
probably more instructive to consider epimorphisms that are not essential. If U V
is any epimorphism and X is a non-zero module then the epimorphism U X V
constructed as the given map on U and zero on X can never be essential. This is because
U is a submodule of U X mapped surjectively onto V . Thus if U V is essential then
U can have no direct summands that are mapped to zero. One may think of an essential
epimorphism as being minimal, in that no unnecessary parts of U are present.
The greatest source of essential epimorphisms is Nakayamas lemma, given here in
a version for modules over non-commutative rings. Over an arbitrary ring a finiteness
condition is required, and that is how we state the result here. We will see in the exercises
that when the ring is a finite-dimensional algebra over a field, the result is true for arbitrary
modules without any finiteness condition.
(7.6) THEOREM (Nakayamas Lemma). If U is any Noetherian module, the homomorphism U U/ Rad U is essential. Equivalently, if V is a submodule of U with the
property that V + Rad U = U , then V = U .
Proof. Suppose V is a submodule of U . If V 6= U then V M U where
M is a maximal submodule of U . Now V + Rad U M and so the composite V
U U/ Rad U has image contained in M/ Rad U , which is not equal to U/ Rad U since
(U/ Rad U )/(M/ Rad U )
= U/M 6= 0.
104
is an essential epimorphism.
Proof. (a) Suppose f and g are essential epimorphisms. Then gf is an epimorphism
also, and it is essential because if U0 is a proper submodule of U then f (U0 ) is a proper
submodule of V since f is essential, and hence g(f (U0)) is a proper submodule of S since
g is essential.
Next suppose f and gf are essential epimorphisms. Since W = Im(gf ) Im(g) it
follows that g is an epimorphism. If V0 is a proper submodule of V then f 1 (V0 ) is a
proper submodule of U since f is an epimorphism, and now g(V0 ) = gf (f 1 (V0 )) is a
proper submodule of S since gf is essential.
Suppose that g and gf are essential epimorphisms. If f were not an epimorphism
then f (U ) would be a proper submodule of V , so gf (U ) would be a proper submodule of
W since gf is essential. Since gf (U ) = W we conclude that f is an epimorphism. If U0 is
a proper submodule of U then gf (U0 ) is a proper submodule of W , since gf is essential,
so f (U0 ) is a proper submodule of V since g is an epimorphism. Hence f is essential.
(b) Consider the commutative square
U
U/ Rad U
V / Rad V
where the vertical homomorphisms are essential epimorphisms by Nakayamas lemma. Now
if either of the horizontal arrows is an essential epimorphism then so is the other, using
Peter Webb
105
part (a). The bottom arrow is an essential epimorphism if and only if it is an isomorphism;
for U/ Rad U is a semisimple module and so the kernel of the map to V / Rad V has a direct
complement in U/ Rad U , which maps onto V / Rad V . Thus if U/ Rad U V / Rad V is
an essential epimorphism its kernel must be zero and hence it must be an isomorphism.
(c) The map
(i Ui )/ Rad(i Ui ) (i Vi )/ Rad(i Vi )
induced by fi may be identified as a map
M
M
(Vi / Rad Vi ),
(Ui / Rad Ui )
i
and it is an isomorphism if and only if each map Ui / Rad Ui Vi / Rad Vi is an isomorphism. These conditions hold if and only if fi is an essential epimorphism, if and only if
each fi is an essential epimorphism by part (b).
We define a projective cover of a module U to be an essential epimorphism P U ,
where P is a projective module. Strictly speaking the projective cover is the homomorphism, but we may also refer to the module P as the projective cover of U . We are justified
in calling it the projective cover by the second part of the following result, which says that
projective covers (if they exist) are unique.
(7.8) PROPOSITION.
(1) Suppose that f : P U is a projective cover of a module U and g : Q U is an
epimorphism where Q is a projective module. Then we may write Q = Q1 Q2 so
that g has components g = (g1 , 0) with respect to this direct sum decomposition and
g1 : Q1 U appears in a commutative triangle
Q1
g1
y
where is an isomorphism.
(2) If any exist, the projective covers of a module U are all isomorphic, by isomorphisms
that commute with the essential epimorphisms.
Proof. (1) In the diagram
g
y
106
Peter Webb
107
= 0.
108
We now classify the indecomposable projective modules over a finite-dimensional algebra as the projective covers of the simple modules. We first describe how these projective
covers arise, and then show that they exhaust the possibilities for indecomposable projective modules. We postpone explicit examples until the next section, in which we consider
group algebras.
(7.13) THEOREM. Let A be a finite-dimensional algebra over a field and S a simple
A-module.
(a) There is an indecomposable projective module PS with PS / Rad PS
= S, of the form
PS = Af where f is a primitive idempotent in A.
(b) The idempotent f has the property that f S 6= 0 and if T is any simple module not
isomorphic to S then f T = 0.
(c) PS is the projective cover of S, it is uniquely determined up to isomorphism by this
property and has S as its unique simple quotient.
(d) It is also possible to find an idempotent fS A so that fS S = S and fS T = 0 for
every simple module T not isomorphic to S.
Proof. Let e A/ Rad A be any primitive idempotent such that eS 6= 0. It is possible
to find such e since we may write 1 as a sum of primitive idempotents and some term in
the sum must be non-zero on S. Let f be any lift of e to A, possible by 7.12. Then f is
primitive, f S = eS 6= 0 and f T = eT = 0 if T
6 S since a primitive idempotent e in the
=
semisimple ring A/ Rad A is non-zero on a unique isomorphism class of simple modules.
We define PS = Af , an indecomposable projective module. Now
PS / Rad PS = Af /(Rad A Af )
= (A/ Rad A) (f + Rad A) = S,
the isomorphism arising because the map Af (A/ Rad A) (f + Rad A) defined by
af 7 (af + Rad A) has kernel (Rad A) f . The fact that PS is the projective cover of
S is a consequence of Nakayamas lemma, and the uniqueness of the projective cover was
dealt with in 7.8. Any simple quotient of PS is a quotient of PS / Rad PS , so there is only
one of these. Finally we observe that if we had written 1 as a sum of primitive central
idempotents in A/ Rad A, the lift of the unique such idempotent that is non-zero on S is
the desired idempotent fS .
(7.14) THEOREM. Let A be a finite-dimensional algebra over a field k. Up to
isomorphism, the indecomposable projective A-modules are exactly the modules PS that
are the projective covers of the simple modules, and PS
= PT if and only if S
= T . Each
projective PS appears as a direct summand of the regular representation, with multiplicity
equal to the multiplicity of S as a summand of A/ Rad A. As a left A-module the regular
representation decomposes as
M
A
(PS )nS
=
simple S
Peter Webb
109
PS 1 PS n
.
yh
U/ Rad U
110
We should really learn more from 7.15 than simply that U has a projective cover: the
projective cover of U is the same as the projective cover of U/ Rad U .
(7.16) Example. The arguments that show the existence of projective covers have a
sense of inevitability about them and we may get the impression that projective covers
always exist in arbitrary situations. In fact they fail to exist in general for integral group
rings. If G = {e, g} is a cyclic group of order 2, consider the submodule 3Z e + Z (e + g)
of ZG generated as an abelian group by 3e and e + g. We rapidly check that this subgroup
is invariant under the action of G (so it is a ZG-submodule), and it is not the whole of ZG
since it does not contain e. Applying the augmentation map : ZG Z we have (3e) = 3
and (e + g) = 2 so (3Z e + Z (e + g)) = 3Z + 2Z = Z. This shows that the epimorphism
is not essential, and so it is not a projective cover of Z. If Z were to have a projective
cover it would be a proper summand of ZG by Proposition 7.8. On reducing modulo 2 we
would deduce that F2 G decomposes, which we know not to be the case by Corollary 6.12.
This shows that Z has no projective cover as a ZG-module.
a finite-dimensional algebra A
A-module.
if S
= T,
otherwise.
Peter Webb
111
Proof. (1) If PS T is any non-zero homomorphism, the kernel must contain Rad PS ,
being a maximal submodule of PS . Since PS / Rad PS
= S is simple, the kernel must be
112
Summary of Chapter 7
1. Direct sum decompositions of A A as an A-module (with indecomposable summands) correspond to expressions for 1A as a sum of orthogonal (primitive) idempotents.
2. U U/ Rad U is essential.
3. Projective covers are unique when they exist. For modules for a finite dimensional
algebra over a field they do exist.
4. Idempotents can be lifted through nilpotent ideals.
5. The indecomposable projective modules for a finite dimensional algebra over a field
are exactly the projective covers of the simple modules. Each has a unique simple quotient
and is a direct summand of the regular representation. Over an algebraically closed field
PS occurs as a summand of the regular representation with multiplicity dim S.
Peter Webb
113
114
Peter Webb
115
The next lemma is a consequence of the fact that induction is both the left and the
right adjoint of restriction, as shown in Lemma 4.12. Without using that language we give
a direct proof.
(8.2) LEMMA. Let H be a subgroup of G.
(1) If P is a projective RG-module then P G
H is a projective RH-module.
(2) If Q is a projective RH-module then Q G
H is a projective RG-module.
Proof. (1) As a RH-module,
RG H
=
RHg
= (RH)|G:H| ,
g[H\G]
(RH) G
H = (R 1 ) H = R 1 = RG
so that direct summands of RH n induce to direct summands of RGn .
We now put together 8.1 and 8.2 to obtain an important restriction on projective
modules.
(8.3) COROLLARY. Let k be a field of characteristic p and
let pa be the exact power
of p that divides |G|. If P is a projective kG-module then pa dim P .
We are about to describe in detail the structure of projective modules for some particular groups that are semidirect products, and the next two results will be used in our
proofs. The first is valid over any commutative ring R.
(8.4) PROPOSITION. Suppose that V is any RG-module that is free as an R-module
and P is a projective RG-module. Then V R P is projective as an RG-module.
Proof. If P P
= RGn then V RGn
= V P V P and it suffices to show that
V RGn is free. We offer two proofs of the fact that V RG
= RGrank V . The first is
G
G
that V RG
= V (R G
1 ) = (V R) 1 = V 1 , with the middle isomorphism coming
from Corollary 4.13. As a module for the identity group, V is just a free R-module and so
G rank V
V G
= RGrank V .
1 = (R 1 )
116
The second proof is really the same as the first, but we make the isomorphism explicit.
Let V triv be the same R-module as V , but with the trivial G-action, so V triv
= Rrank V as
RG-modules. We define a linear map
V RG V triv RG
v g 7 g 1 v g
which has inverse gw g w g. One checks that these mutually inverse linear maps
are RG-module homomorphisms. Finally V triv RG
= RGrank V .
In the calculations that follow we will need to use the fact that for representations
over a field, taking the tensor product with a fixed representation preserves exactness.
(8.5) LEMMA. Let 0 U V W 0 be a short exact sequence of kG-modules
and X another kG-module, where k is a field. Then the sequence
0 U k X V k X W k X 0
is exact. Thus if U is a submodule of V then (V /U ) k X
= (V k X)/(U k X).
Proof. Since the tensor products are taken over k, the question of exactness is independent of the action of G. As a short exact sequence of vector spaces 0 U V W 0
is split, so that V
= U W with the morphisms in the sequence as two of the component inclusions and projections. Applying k X to this we get V k X
= (U k X) (W k X)
with component morphisms given by the morphisms in the sequence 0 U k X
V k X W k X 0. This is enough to show that the sequence is (split) exact as a
sequence of vector spaces, and hence exact as a sequence of kG-modules.
Another approach to the same thing is to suppose that U is a submodule of V and
take a basis v1 , . . . , vn for V such that v1 , . . . , vd is a basis for U and let x1 , . . . , xm be a
basis for X. Now the vi k xj with 1 i n and 1 j m form a basis for V k X, and
the same elements with 1 i d and 1 j m form a basis for U k X. This shows
that U k X is a submodule of V k X, and the quotient has as a basis the images of the
vi k xj with d + 1 i n and 1 j m, which is in bijection with a basis of W k X.
Peter Webb
117
as k-algebras,
which arises because kG has as a basis the elements (h, k) where h H, k K, and
kH kK has as a basis the corresponding elements h k. These two bases multiply
together in the same fashion, and so we have an algebra isomorphism.
Let us write kK = S1n1 Srnr , where S1 , . . . , Sr are the non-isomorphic simple
kK-modules, bearing in mind that kK is semisimple since K has order relatively prime to
p. Since H = Op (G), these are also the non-isomorphic simple kG-modules, by Corollary
6.4. We have
kG = kH kK = (kH S1 )n1 (kH Sr )nr
as kG-modules, and so the kH Si are projective kG-modules. Each does occur with
multiplicity equal to the multiplicity of Si as a summand of kG/ Rad(kG), and so must
be indecomposable, using 7.14. We have therefore constructed all the indecomposable
projective kG-modules, and they are the modules PSi = kH Si .
Suppose that 0 P1 Pn = kH is a composition series of the regular representation of H. Since H is a p-group, all the composition factors are the trivial representation,
k. Because k Si preserves exact sequences, the series 0 P1 Si Pn Si = PSi
has quotients k Si = Si , which are simple, and so this is a composition series of PSi .
There is only one isomorphism type of composition factor.
We can also see this from the ring-theoretic structure of kG. Assuming that k is
algebraically closed (to make the notation easier) we have EndkG (Si ) = k for each i and
Lr
kK
= i=1 Mni (k) as rings where ni = dim Si . Now
kG = kH kK
=
r
M
i=1
Mni (kH)
118
r
M
Mni (Rad(kH)).
i=1
r
M
i=1
for each n.
As a very specific example, suppose that H is cyclic of order ps . Then kH has a
unique composition series by Theorem 6.2, with terms Pj = Rad(kH)j kH. We see from
the above discussion and the fact that Rad kG = Rad(kH) kK that the terms in the
composition series of PSi are Pj Si = Rad(kG)j PSi . Thus the radical series of PSi is in
fact a composition series, and it follows (as in Exercise 5 of Chapter 6) that PSi also has
a unique composition series, there being no more submodules of PSi other than the ones
listed.
We move on now to describe the projective kG-modules where k is a field of characteristic p and where G is a semidirect product with Sylow p-subgroup H, doing this first
when G has the form G = H K and afterwards doing it when G = K H. Before this
we give a module decomposition of the group ring of a semidirect product that holds in
all cases.
Peter Webb
119
G
where the structure of R G
K has just been described, and R H = RK with H acting
trivially.
120
(b)
(c)
(d)
(e)
where kH is taken to have the kG-module action where H acts by multiplication and
K acts by conjugation, as in Lemma 8.7.
Rad(kG) is the kernel of the ring homomorphism kG kK given by the quotient
homomorphism G K, so that kG/ Rad(kG)
= kK as rings, and also as kG-modules.
Furthermore, Rad(kG) is generated both as a left ideal and a right ideal by Rad(kH) =
IH, the augmentation ideal of kH.
The simple kG-modules are precisely the simple kK-modules, regarded as kG-modules
via the quotient homomorphism G K. If S is any simple kG-module with projective
cover PS then PS
= Pk S
= S G
K as kG-modules.
For each simple kG-module S with projective cover PS , the radical series of PS has
terms Radn (PS ) = (IH)n S with radical layers ((IH)n1 /(IH)n ) S, where IH is
taken to have the restriction of the action on RH given in Lemma 8.7. Consequently
all of the indecomposable projective modules have the same Loewy length.
There is an isomorphism of kG-modules Pk kK
= kG.
Proof. (a) Since |K| is invertible in k we have kGeK = kG K (with the notation
of Lemma 8.7), and since eK is an idempotent this is a projective kG-module, and it has
the stated identifications by Lemma 8.7. This module is indecomposable as a kH-module
since H is a p-group, by Theorem 8.1, so it is indecomposable as a kG-module. Since its
unique simple quotient is represented by 1 H, and K conjugates this trivially, kGeK is
the projective cover of the trivial module.
(b) Since H = Op (G) the simple kG-modules are precisely the simple kK-modules,
by 6.4. Now G acts on these via the ring surjection kG kK, so the kernel of this map
acts as zero on all simple modules and hence is contained in Rad kG. But also kK is a
semisimple ring, so the kernel equals Rad kG.
The fact that the kernel of the ring homomorphism kG kK (and hence Rad kG)
has the description kG IH is shown in Exercise 10 of Chapter 6, and we also give the
argument here. We know that Rad kH = IH from 6.8. Taking coset representatives
F
G = g[G/H] gH and writing k[gH] for the span in kG of the elements gh, h H, we
have that
X
X
M
kG IH =
k[gH] IH =
gIH =
gIH,
g[G/H]
g[G/H]
g[G/H]
the sum being direct since each term gIH has basis {g(h 1) h H} with support in the
coset gH. The span of the elements of K inside kG is a space complementary to kG IH
Peter Webb
121
that is mapped isomorphically to kK. From this we see that kG IH is the kernel of the
homomorphism kG kK. From the fact that H is a normal subgroup we may calculate
directly that kG Rad(kH) = kG Rad(kH) kG = Rad(kH) kG. We may also show this
by observing that kG Rad(kH) is the kernel of a ring homomorphism, hence is a 2-sided
ideal and so kG Rad(kH) = kG Rad(kH) kG. We could have argued with Rad(kH) kG
just as well, and so this also is equal to kG Rad(kH) kG.
(c) The fact that the simple kG-modules are the same as the simple kK-modules was
observed at the start of the proof of (b). If S is a simple module, by 8.4 Pk S is projective.
Tensoring the epimorphism Pk k with S gives an epimorphism Pk S S, so Pk S
contains PS as a summand. We show that Pk S equals PS . Now
Rad(Pk S) = Rad(kG) (Pk S) = Rad(kH) kG (Pk S) = IH (Pk S).
We show that this equals (IH Pk ) S. The reason for this is that IH is spanned by
elements h 1 where h H, and if x s is a basic tensor in Pk S then
(h 1) x s = h (x s) 1 (x s)
= hx hs x s
= hx (h 1)s + (h 1)x s
= (h 1)x s
since H is a normal p-subgroup which thus acts trivially on S. Note that this argument
does not depend on the first module in the tensor product being Pk . To continue the
argument, we have
(IH Pk ) S = (IH kG Pk ) S = (Rad(kG) Pk ) S = Rad(Pk ) S.
Now Rad(Pk ) has codimension 1 in Pk , so Rad(Pk ) S has codimension dim S in Pk S.
From all this it follows that Pk S is the projective cover of S, and so is isomorphic to
G
PS . We have k G
K S = S K by 4.13.
(d) We have seen in the proof of part (c) that Rad(PS ) = (IH Pk )S. It was observed
there that the validity of the equation Rad(Pk S) = (IH Pk )S did not depend on upon
the particular structure of Pk . Thus we see by induction that Radn (PS ) = ((IH)n Pk )S.
Using the identification of Pk as kH that we established in (a) gives the result.
(e) This is immediate from 8.7 and part (a).
Note in 8.8(a) that when Pk is identified with kH using the module action from 8.7,
it is not being identifed with the subring kH of kG, which does not immediately have the
structure of a kG-module.
Proposition 8.8 allows us to give very specific information in the case of groups with
a normal Sylow p-subgroup that is cyclic. This extends the description of the projectives
that we already gave in 8.6 for the case when the Sylow p-subgroup is a direct factor (as
happens, for instance, when G is cyclic). We recall from Chapter 6 that a uniserial module
is one with a unique composition series, and that some equivalent conditions to this were
explored in the exercises to Chapter 6.
122
= (x 1)(xr1 + + x + 1 r) + r(x 1)
r(x 1) (mod IH 2 ).
More generally for some IH 2 ,
y(x 1)s = (xr 1)s
= (r(x 1) + )s
r s (x 1)s
(mod IH s+1 ),
and so y acts on the quotient IH s /IH s+1 as multiplication by r s . One way to describe
this is that IH s /IH s+1 = W s , the s-fold tensor power. Thus by 8.8(d) the radical layers
of PS are of the form W s S. These are simple because W has dimension 1, so the
radical series of PS is its unique composition series, by Exercise 5 from Chapter 6.
An algebra for which all indecomposable projective and injective modules are uniserial
is called a Nakayama algebra, so that we have just shown that k[H K] is a Nakayama
algebra when k has characteristic p, H is a cyclic p-group and K has order prime to p.
At least, we have shown that the projectives are uniserial, and the injectives are uniserial
because they are the duals of the projectives. We will see in Corollary 8.13 that for group
algebras over a field, injective modules and projective modules are, in fact, the same thing.
In Proposition 11.8 we will give a complete description of the indecomposable modules
for a Nakayama algebra. It turns out they are all uniserial and there are finitely many of
them.
Peter Webb
123
Let us continue further with our analysis of the composition factors of the indecomposable projective module PS in the situation of Proposition 8.9. Observe that the
isomorphism types of these composition factors occur in a cycle that repeats itself: for
each element y K the map x 7 y x = xr is an automorphism of Cpn and there is a least
positive integer fy such that r fy 1 (mod pn ). This fy divides pn 1, and letting f be
the l.c.m. of all fy with y K we put pn 1 = ef . Then the modules k, W, W 2 , W 3 , . . .
give rise to f different representations. They repeat e times in Pk , except for k which
appears e + 1 times. A similar repetition occurs with the composition factors W i S of
PS .
As a specific example of this, consider the non-abelian group
G = hx, y x7 = y 3 = 1, yxy 1 = x2 i = C7 C3
of order 21 over the field F7 . For the element y we have r = 2 and fy = 3, so that f = 3
and e = 2. There are three simple F7 G-modules by Proposition 8.8, all 1-dimensional,
which we will label k1 , k2 , k4 . The element x acts trivially on all of these modules, and y
acts trivially on k1 , as multiplication by 2 on k2 and as multiplication by 4 on k4 . In the
previous notation, W = k2 , W 2 = k4 and W 3 = k1 . The three projective covers are
uniserial with composition factors as shown.
Pk1
k1
|
k2
|
k4
|
= k1
|
k2
|
k4
|
k1
Pk2
k2
|
k4
|
k1
|
= k2
|
k4
|
k1
|
k2
Pk4
k4
|
k1
|
k2
|
= k4
|
k1
|
k2
|
k4
124
In Example 7.5 we have already seen an instance of the situation described in 8.10.
In that example we took G = S3 = K H where H = h(1, 2)i and K = h(1, 2, 3)i and we
worked with a field k of characteristic 2, which for a technical reason was F4 . Note that
if V is the 2-dimensional simple kS3 -module then V kH
= V V since V is projective.
We see from this that the module S kH that appeared in the proof of 8.10 need not be
indecomposable.
Peter Webb
125
(1)
(2)
(3)
(4)
The first part of the next proposition has already been proved in two different ways in
the exercises to Chapter 4. The proof given here is really the same as one of those earlier
proofs, but it is presented a little differently.
(8.12) PROPOSITION. Let k be a field. Then
(1) kG
= kG as kG-modules, and
(2) a finitely generated kG-module P is projective if and only if P is projective as a
kG-module.
Proof. (1) We denote the elements of kG dual to the basis elements {g g G} by
g, so that g(h) = g,h k, the Kronecker . We define an isomorphism of vector spaces
X
gG
kG kG
X
ag g 7
ag g .
gG
for g, h G, so that x
g=x
cg.
(2) Since P
= P as kG-modules it suffices to prove one implication. If P is a
n
summand of kG then P is a summand of (kGn )
= kGn , and so is also projective.
126
Peter Webb
127
Since dim End(S ) = dim End(S) this implies that P has a unique simple submodule and
Soc(P ) is simple.
An algebra for which injective modules and projective modules coincide is called selfinjective or quasi-Frobenius, so we have just shown that group rings of finite groups over
a field are self-injective. An equivalent condition on a finite dimensional algebra A that
it should be self-injective is that the regular representation A A should be an injective
A-module.
(8.14) COROLLARY. Suppose U is a kG-module, where k is a field, for which there
are submodules U0 U1 U with U1 /U0 = P a projective module. Then U
= P U for
some submodule U of U .
P U0 . Thus
Proof. The exact sequence 0 U0 U1 P 0 splits, and so U1 =
P is isomorphic to a submodule of U , and since P is injective the monomorphism P U
must split.
We will now sharpen part (2) of 8.13 by showing that Soc PS
= S for group algebras,
and we will also show that the Cartan matrix for group algebras is symmetric. These
are properties that hold for a class of algebras called symmetric algebras, of which group
algebras are examples. We say that a finite dimensional algebra A over a field k is a
symmetric algebra if there is a non-degenerate bilinear form ( , ) : A A k such that
(1) (symmetry) (a, b) = (b, a) for all a, b A,
(2) (associativity) (ab, c) = (a, bc) for all a, b, c A.
The group algebra kG is a symmetric algebra with the bilinear form defined on the
basis elements by
1 if gh = 1,
(g, h) =
0 otherwise,
as is readily verified. Notice that this bilinear form may be described on general elements
a, b kG by (a, b) = coefficient of 1 in ab. Having learned that group algebras are symmetric it will be no surprise to learn that matrix algebras are symmetric. When A = Mn (k)
is the algebra of n n matrices over a field k, the trace bilinear form (A, B) = tr(AB)
gives the structure of a symmetric algebra.
We will use the bilinear form on kG in the proof of the next result. Although we only
state it for group algebras, it is valid for symmetric algebras in general.
(8.15) THEOREM. Let P be an indecomposable projective module for a group algebra kG. Then P/ Rad P
= Soc P .
Proof. We may choose a primitive idempotent e kG so that P
= kGe as kG-modules.
We claim that Soc(kGe) = Soc(kG) e, since Soc(kG) e Soc(kG) and Soc(kG) e kGe
so Soc(kG) e kGe Soc(kG) = Soc(kGe), since the last intersection is the largest
128
semisimple submodule of kGe. On the other hand Soc(kGe) Soc(kG) since Soc(kGe) is
semisimple so Soc(kGe) = Soc(kGe) e Soc(kG) e.
Next, Hom(kGe, Soc(kG)e) and e Soc(kG)e have the same dimension by 7.17(3), and
since Soc(kG)e is simple, by 8.13, this is non-zero if and only if Soc(kG)e
= kGe/ Rad(kGe)
by Theorem 7.13. We show that e Soc(kG)e 6= 0.
If e Soc(kG)e = 0 then
0 = (1, e Soc(kG)e)
= (e, Soc(kG)e)
= (Soc(kG)e, e)
= (kG Soc(kG)e, e)
= (kG, Soc(kG)e e)
= (kG, Soc(kGe)).
Since the bilinear form is non-degenerate this implies that Soc(kGe) = 0, a contradiction.
Recall that for any RG-module U we have defined the fixed points of G on U to be
U := {u U gu = u for all g G}. We also define the fixed quotient of G on U to be
UG := U/{(g 1)u 1 6= g G}. Then U G is the largest submodule of U on which G acts
trivially and UG is the largest quotient of U on which G acts trivially.
G
Peter Webb
129
(8.17) THEOREM. Let k be a field and let S, T be simple kG-modules. The Cartan
invariants satisfy
cST dim EndkG (T ) = cT S dim EndkG (S).
If dim EndkG (S) = 1 for all simple modules S (for example, if k is algebraically closed)
then the Cartan matrix C = (cST ) is symmetric.
Proof. We recall from 7.18 that
cST = dim HomkG (PS , PT )/ dim EndkG (S)
and in view of this we must show that dim HomkG (PS , PT ) = dim HomkG (PT , PS ). Now
HomkG (PS , PT ) = Homk (PS , PT )G
= (PS k PT )G
by 3.3 and 3.4. Since PS k PT is projective by 8.4, this has the same dimension as
(PS k PT )G
= HomkG (PT , PS ),
= (PS k PT )G
using 8.16.
We conclude this section by summarizing some further aspects of injective modules.
We define an essential monomorphism to be a monomorphism of modules f : V U
with the property that whenever g : U W is a map such that gf is a monomorphism
then g is a monomorphism. An injective hull (or injective envelope) of U is an essential
monomorphism U I where I is an injective module. By direct arguments, or by taking
the corresponding results for essential epimorphisms and projective covers and applying
the duality U 7 U , we may establish the following properties for finitely-generated kGmodules.
The inclusion Soc U U is an essential monomorphism.
g
130
Summary of Chapter 8
1. When G is a p-group and k is a field of characteristic p, the regular representation
kG is indecomposable.
2. The property of projectivity is preserved under induction and restriction.
3. Tensor product with a projective modules gives a projective module.
4. When k is a field of characteristic p we have an explicit description of the projective
kG-modules when G has a normal Sylow p-subgroup, and also when G has a normal pcomplement.
5. G has a normal p-complement if and only if for every simple module S the composition factors of PS are all isomorphic to S.
6. Projective kG-modules are the same as injective kG-modules.
7. kG is a symmetric algebra. Soc PS
= S always. The Cartan matrix is symmetric.
Exercises for Chapter 8
1. Prove that if G is any finite group then the only idempotents in the integral group
ring ZG are 0 and 1.
[If e is idempotent consider the rank of the free abelian group ZGe and also its image
under the homomorphism ZG Fp G for each prime p dividing |G|, which is a projective
Fp G-module. Show that rankZ ZGe is divisible by |G|. Deduce from this that if e 6= 0 then
e = 1.]
2. (a) Let H = C2 C2 and let k be a field of characteristic 2. Show that (IH)2 is a
P
one-dimensional space spanned by hH h.
(b) Let G = A4 = (C2 C2 ) C3 and let F4 be the field with four elements. Compute the
radical series of each of the three indecomposable projectives for F4 A4 and identify each
of the quotients
Radn PS / Radn+1 PS .
Now do the same for the socle series. Hence determine the Cartan matrix of F4 A4 .
[Start by observing that F4 A4 has 3 simple modules, all of dimension 1, which one might
denote by 1, and 2 . This exercise may be done by applying the kind of calculation that
led to Proposition 8.9.]
(c) Now consider F2 A4 where F2 is the field with two elements. Prove that the 20 1
dimensional F2 -vector space on which a generator of C3 acts via
is a simple
1 1
F2 C3 -module. Calculate the radical and socle series for each of the two indecomposable
projective modules for F2 A4 and hence determine the Cartan matrix of F2 A4 .
3. Let G = H K where H is a p-group, K is a p -group, and let k be a field of
characteristic p. Regard kH as a kG-module via its isomorphism with Pk , so H acts as
usual and K acts by conjugation.
Peter Webb
131
(a) Show that for each n, (IH)n is a kG-submodule of kH, and that (IH)n /(IH)n+1 is a
kG-module on which H acts trivially.
(b) Show that
Pk = kH IH (IH)2 (IH)3
is the radical series of Pk as a kG-module.
(c) Show that there is a map
IH/(IH)2 k (IH)n /(IH)n+1 (IH)n+1 /(IH)n+2
x + (IH)2 y + (IH)n+1 7 xy + (IH)n+2
132
gG
g) U ) =
1 if U
= Pk ,
0 otherwise.
gG
g) V ) is the multi-
Peter Webb
133
P
[Observe that kGG = PkG = k gG g. Remember that Pk is injective and has socle
isomorphic to k.]
10. Let U be a finite-dimensional kG-module, where k is a field, and let PS be an
indecomposable projective kG-module with simple quotient S. Show that in any decomposition of U as a direct sum of indecomposable modules, the multiplicity with which PS
occurs is equal to
dim HomkG (PS , U ) dim HomkG (PS / Soc PS , U )
dim EndkG (S)
and also to
Pk1
k1
|
k
|
= k1
|
k
|
k1
Pk
k
|
k1
|
= k
|
k1
|
k
U
|
U
|
PU = U
|
U
|
U
Finite Group
3 2
Representations
134
0
0 and that F5 D30
= F5 D10 M2 (F5 C5 ) as
5
Peter Webb
135
9. Changing the ground ring: splitting fields and the decomposition map
We examine the relationship between the representations of a fixed group over different
rings. Often we have have assumed that representations are defined over a field that is
algebraically closed. What if the field is not algebraically closed? Such a question is
significant because representations arise naturally over different fields, which might not be
algebraically closed, and it is important to know how they change on moving to an extension
field such as the algebraic closure. It is also important to know whether a representation
may be defined over some smaller field. We introduce the notion of a splitting field, showing
that such a field may always be chosen to be a finite extension of the prime field.
After proving Brauers theorem that over a splitting field of characteristic p the number of non-isomorphic simple representations equals the number of conjugacy classes of
elements of order prime to p, we turn to the question of reducing representations from
characteristic 0 to characteristic p. The process involves first writing a representation in
the valuation ring of a p-local field and then factoring out the maximal ideal of the valuation ring. This gives rise to the decomposition map between the Grothendieck groups
of representations in characteristic 0 and characteristic p. We show that this map is welldefined, and then construct the so-called cde triangle. This provides a very effective way
to compute the Cartan matrix in characteristic p from the decomposition map.
In the last part of this section we describe in detail the properties of blocks of defect
zero. These are representations in characteristic p that are both simple and projective.
They always arise as the reduction modulo p of a simple representations in characteristic
zero, and these are also known as blocks of defect 0. The blocks of defect zero have
importance in character theory, accounting for many zeroes in character tables, and they
are also the subject of some of the deepest investigations in representation theory.
136
Some definitions
Suppose that A is an algebra A over a commutative ring R and that U is an A-module.
If R R is a homomorphism to another commutative ring R we may form the R -algebra
R R A, and now R R U becomes an R R A-module in an evident way. In this section
we study the relationship between U and R R U . When we specialize to a group algebra
A = RG we will identify R R RG with R G.
We will pay special attention to two particular cases of this construction, the first being
when R is a subring of R . If U is an A-module, we say that the module V = R R U
is obtained from U by extending the scalars from R to R ; and if an R R A-module V
has the form R R U we say it can be written in R. In this situation, when U is free as
an R-module we may identify U with the subset 1R R U of R R U , and an R-basis of
U becomes an R -basis of R R U under this identification. In case A = RG is a group
ring, with respect to such a basis of V = R R U the matrices that represent the action
of elements g G on V have entries in R, and are the same as the matrices representing
the action of these elements on U (with respect to the basis of U ). Equally, if we can find
a basis for an R G-module V so that each g G acts by a matrix with elements in R then
RG preserves the R-linear span of this basis, and this R-linear span is an RG-module U
for which V = R R U . Thus an R -free module V can be written in R if and only if V
has an R -basis with respect to which G acts via matrices with entries in R.
The second situation to which we will pay particular attention arises when R = R/I
for some ideal I in R. In this case applying R R to a module U is the same as reducing
U modulo I. If V is an R R A-module of the form R R U for some A-module U we
say that V can be lifted to U , and that U is a lift of V . Most often we will perform this
construction when R is a local ring and I is the maximal ideal of R.
Splitting fields
We start by considering the behaviour of representations over a field. It is often a
help to know that a representation can be written in a small field.
(9.1) PROPOSITION. Let F E be fields where E is algebraic over F and let A be
a finite-dimensional F -algebra. Let V be a finite-dimensional E F A-module. Then there
exists a field K with F K E, of finite degree over F , so that V can be written in K.
Proof. Let a1 , . . . , an be a basis for A and let at act on V with matrix (atij ) with
respect to some basis of V . Let K = F [atij , 1 t n, 1 i, j d]. Then [K : F ] is finite
since K is an extension of F by finitely many algebraic elements, and A acts by matrices
with entries in K.
Peter Webb
137
138
(1) (2): We prove this implication here only in the situation where F is a perfect
field, so that all irreducible polynomials with coefficients in F are separable. The result
is true in general and is not difficult but requires some technicality that we wish to avoid
(see [CR]). This implication will not be needed for our application of the result.
Suppose that EndA (U ) is larger than F , so there exists an endomorphism : U U
that is not scalar multiplication by an element of F . Let be a root of the characteristic
polynomial of in some field extension E F : in other words, is an eigenvalue of .
Then 1E : E F U E F U is not scalar multiplication by , because if it were the
minimal polynomial of over F would be a factor of (X )n where n = dimF U and by
separability of the minimal polynomial we would deduce F . Now 1E 1U
EndEA (E F U ) is a non-zero endomorphism with non-zero kernel, and since E F U is
simple this cannot happen, by Schurs lemma.
The next theorem is the main result about splitting fields that we will need for the
process of reduction modulo p to be described later in this section.
(9.3) THEOREM. Let A be a finite-dimensional algebra over a field F . Then A has
a splitting field of finite degree over F . If G is a finite group, it has splitting fields that
are finite degree extensions of Q (in characteristic zero) or of Fp (in characteristic p).
Proof. The algebraic closure F of F is a splitting field for A, since by Schurs lemma
condition (2) of Proposition 9.2 is satisfied for each simple F F A-module. By Proposition
9.1 there is a finite extension E F so that every simple F F A-module can be written
in E. The simple E F A-modules U that arise like this are absolutely simple, because if
K E is an extension field for which K E U is not simple then K E U is not simple,
where K is an algebraic closure of K, and since K contains a copy of F , F F U cannot
be simple since F is a splitting field for A, a contradiction.
It is also true that every simple E F A-module is isomorphic to one of the simple
modules U that arise in this way from the algebraic closure F . For, if V is a simple E F Amodule let e2 = e E F A be an idempotent with the property that eV 6= 0 but eV = 0
for all simple modules V not isomorphic to V . Let W be a simple F F A-module that
is not annihilated by e. We must have W
= F E V since V is the only possible simple
module that would give a result not annihilated by e.
Group algebras are defined over the prime field Q or Fp (depending on the characteristic), and by what we have just proved QG and Fp G have splitting fields that are finite
degree extensions of the prime field.
Peter Webb
139
We see from the above that every simple representation of a finite group may be
written over a field that is a finite degree extension of the prime field. In characteristic
zero this means that every representation can be written in such a field, by semisimplicity.
It is not always true that every representation of a finite group in positive characteristic
can be written in a finite field, and an example of this is given as exercise 8 to this section.
Some other basic facts about splitting fields are left to the exercises at the end of this
section. Thus, if A is a finite dimensional algebra over a field F that is a splitting field
for A and E F is a field extension, it is the case that every simple E F A-module
can be written in F (exercises 3 and 4). It is also true for a finite dimensional algebra
that no matter which splitting field we take, after extending scalars we always have the
same number of isomorphism classes of simple modules (exercise 4). Thus in defining the
character table of a finite group, instead of working with complex representations we could
have used representations over any splitting field and obtained the same table.
In positive characteristic the situation is not so straightforward. It is usually not the
case that indecomposable modules always remain indecomposable under all field extensions
(those that do are termed absolutely indecomposable), even when all fields concerned are
splitting fields. We can, however, show that once we are working over a splitting field, the
indecomposable projective modules remain indecomposable under field extension (exercise
7). The consequence of this is that the Cartan matrix does not change once we have a
splitting field, being independent of the choice of the splitting field. Just as when we speak
of the character table of group we mean the character table of representations over some
splitting field, so in speaking of the Cartan matrix of a group algebra we usually mean the
Cartan matrix over some splitting field.
In the case of group algebras there is a finer result about splitting fields than 9.3. It
was first conjectured by Schur and later proved by Brauer as a deduction from Brauers
induction theorem. We state the result, but will not use it and do not prove it. The
exponent of a group G is the least common multiple of the orders of its elements.
(9.4) THEOREM (Brauer). Let G be a finite group, F a field, and suppose that F
contains a primitive mth root of unity, where m is the exponent of G. Then F is a splitting
field for G.
2i
This theorem tells us that Q(e m ) and Fp () are splitting fields for G, where is a
primitive mth root of unity in an extension of Fp . Often smaller splitting fields than these
can be found, and the determination of minimal splitting fields must be done on a caseby-case basis. For example, we may see as a result of the calculations we have performed
earlier in this text that in every characteristic the prime field is a splitting field for S3
the same is in fact true for all the symmetric groups. However, if we require that a field
2i
be a splitting field not only for G but also for all of its subgroups, then Q(e m ) and Fp ()
are the smallest possibilities, since as we have seen earlier that a cyclic group of order n
requires the presence of a primitive nth root of 1 in a splitting field.
Again we will not use it, but it is important to know the following theorem about field
extensions. For a proof see [CRI, p. 139 Section 6 Exercise 6].
140
Peter Webb
141
mn
ap = (ap )p
zero. On the other hand when i 6= j every matrix Eij (zero everywhere except for a 1
in position (i, j)) can be written as a commutator: Eij = Eik Ekj Ekj Eik , and also
Eii Ejj = Eij Eji Eji Eij . Since these matrices span the matrices of trace zero we
deduce that S consists exactly of the matrices of trace 0. Now S T A and S has
codimension 1 so either T = S or T = A. The matrix E11 is idempotent and does not lie
in T , so T = S.
(9.8) PROPOSITION. Let A be a finite-dimensional algebra over a field of characteristic p that is a splitting field for A. The number of non-isomorphic simple representations
of A equals the codimension of T in A.
Proof. Let us write T (A), S(A), T (A/ Rad(A)), S(A/ Rad(A)) for the constructions
S, T applied to A and A/ Rad(A). Since Rad(A) is nilpotent it is contained in T (A). Also
(S(A) + Rad(A))/ Rad(A) = S(A/ Rad(A))
n
is easily verified. We claim that T (A)/ Rad(A) = T (A/ Rad(A)). For, if ap S(A) then
n
(a + Rad(A))p (S(A) + Rad(A))/ Rad(A) = S(A/ Rad(A)) and this shows that the
n
Now A/ Rad(A) is a direct sum of matrix algebras. It is apparent that both S and
T preserve direct sums, so the codimension of T (A/ Rad(A)) in A/ Rad(A) equals the
number of simple A-modules, and this equals the codimension of T (A) in A.
142
Peter Webb
143
P pn
p-regular. Now
i xi S. But x1 , . . . , xr are independent modulo S by Lemma 9.10
pn
so i = 0 for all i, and hence i = 0 for all i. This shows that x1 + T, . . . , xr + T are
linearly independent.
(9.12) COROLLARY. Let F be a splitting field of characteristic p for a finite groups
G1 and G2 . The simple F [G1 G2 ]-modules are precisely the tensor products S1 S2 where
Si is a simple F Gi -module, i = 1, 2, and the action of G1 G2 is given by (g1 , g2 )(s1 s2 ) =
g1 s1 g2 s2 . Two such tensor products S1 S2 and S1 S2 are isomorphic as F [G1 G2 ]modules if and only if Si
= Si as F Gi -modules, i = 1, 2.
We have commented before, after Theorem 4.2, that this kind of result is a special case
of a more general statement about the simple modules for a tensor product of algebras,
and this can be found in [CRI, Theorem 10.38]. The general argument is not hard, but we
can use our expression for the number of simple representations of a group to eliminate
half of it.
Proof. We verify that the modules S1 S2 are simple. This is so since the image of
each F Gi in EndF (Si ) given by the module action is the full matrix algebra EndF (Si ),
by Burnsides theorem that was presented as Exercise 8 in Section 2, and which follows
from the ArtinWedderburn theorem. The image of F [G1 G2 ] in EndF (S1 S2 ) contains
EndF (S1 ) EndF (S2 ) and so by counting dimensions it is the whole of EndF (S1 S2 ).
This implies that S1 S2 is simple.
On restriction to G1 , S1 S2 is a direct sum of copies of S1 , and similarly for G2 ,
so S1 S2
= S1 as F G1 -modules and
= S1 S2 as F [G1 G2 ]-modules if and only if S1
S2
= S2 as F G2 -modules.
We conclude by checking that this gives the right number of simple modules for
G1 G2 . By Brauers theorem this is the number of p-regular conjugacy classes of G1 G2 .
Since (g1 , g2 ) is p-regular if and only if both g1 and g2 are p-regular, and this element is
conjugate in G1 G2 to (g1 , g2 ) if and only if g1 G1 g1 and g2 G2 g2 , the number of
p-regular classes in G1 G2 is the product of the numbers for G1 and G2 .
144
Peter Webb
145
146
Peter Webb
147
y
U
U/ Rad U
148
we have seen that such a module lifts to PS , which is a lattice and is the projective cover of
PS . Suppose that L is any RG-lattice for which L/L
= PS . Since L Rad L the radical
quotient of L is S. The projective cover morphism PS S factors as PS L S, giving
an isomorphism on radical quotients. It follows that PS L is surjective by Nakayamas
lemma, and since the ranks of PS and L are the same, this map is an isomorphism.
We next examine the relationship between RG-modules and F G-modules where R is
a principal ideal domain and F is its field of fractions. Given an RG-lattice L we may
regard L as a subset of F R L. In this situation the F G-module F R L may be written
in R, according to the terminology introduced at the start of this chapter. Conversely, if
U is an F G-module, a full RG-lattice U0 in U is defined to be an RG-lattice U0 U that
has an R basis which is also an F -basis of U . In this situation U
= F R U0 and we say
that U0 is an R-form of U . Thus an F G-module that has an R-form can be written in R.
We now show that every finitely-generated F G-module has an R-form.
(9.17) LEMMA. Let R be a principal ideal domain with field of fractions F , and let
U be a finite-dimensional F -vector space. Any finitely-generated R-submodule of U that
contains an F -basis of U is a full lattice in U .
Proof. Let U0 be a finitely-generated R-submodule of U that contains an F -basis of U .
Since U0 is a finitely-generated torsion-free R-module, U0
= Rn for some n, and it has an
R-basis x1 , . . . , xn . Since U0 contains an F -basis of U it follows that x1 , . . . , xn span U over
F . We show that x1 , . . . , xn are independent over F . Suppose that 1 x1 + n xn = 0
for certain i F . We may write i = abii where ai , bi R, since F is the field of fractions
Q
of R. Now clearing denominators we have ( bi )(1 x1 + n xn ) = 0 which implies that
Q
( bi )i = 0 for each i since x1 , . . . , xn is an R-basis. This implies that i = 0 for all i
and hence that n = dim U and x1 , . . . , xn is an F -basis of U .
The kind of phenomenon that the last result is designed to exclude is exemplified by
considering subgroups
of R generated by elements that are independent over Q, such as
Peter Webb
149
We should expect that much of the time when p |G| an F G-module U will contain
various non-isomorphic full sublattices. To show how such non-isomorphic sublattices may
come about, consider an indecomposable projective RG-module P . It often happens that
F R P is not simple as an F G-module. Writing F R P = S1 St as a direct
sum of simple F G-modules and taking a full RG-lattice in each Si , the direct sum of these
lattices will not be isomorphic to P , because P is indecomposable. In this manner we may
construct non-isomorphic R-forms of F R P .
For another specific example, consider a cyclic group G = hgi of order 2 and let
(F, R, k) be a 2-modular system. The regular representation F G contains the full lattice
R 1 + R g that is indecomposable since its reduction kG is indecomposable. It also
contains the full lattice R(1 + g) + R(1 g), which is a direct sum of RG-modules and
hence is decomposable.
The following result is crucial to the definition of the decomposition map, which will
be given afterwards.
(9.19) THEOREM (Brauer-Nesbitt). Let (F, R, k) be a p-modular system, G a finite
group, and U a finitely-generated F G-module. Let L1 , L2 be full RG-lattices in U . Then
L1 /L1 and L2 /L2 have the same composition factors with the same multiplicities, as
kG-modules.
Proof. We observe first that L1 + L2 is also a full RG-lattice in U , by Lemma 9.17,
so by proving the result first for the pair of lattices L1 and L1 + L2 and then for L2 and
L1 + L2 we see that it suffices to consider the case of a pair of lattices, one contained in
the other. We now assume that L1 L2 .
As R-modules, L1 and L2 are free of the same rank, and so L2 /L1 is a torsion module.
Hence L2 /L1 has a composition series as an R-module, and hence also as an RG-module,
because every series of RG-modules can be extended to give a composition series of Rmodules. By working down the terms in a composition series, we see that it suffices to
assume that L1 is a maximal RG-submodule of L2 , and we now make this assumption.
150
1 0
1 0
1 0 and 0 1 .
0 1
1 1
At the end of Chapter 2 we observed that S3 has three simple representations over a field
of characteristic 0: the trivial representation, the sign representation and a 2-dimensional
representation that we have exhibited as a ZS3 -lattice. These three representations index
the rows of the decomposition matrices. In characteristic 2 the simple representations are
the trivial module and the 2-dimensional irreducible with the same matrices as in characteristic 0, but with the entries interpreted as lying in F2 . The latter representation is
simple because S3 permutes the three non-zero vectors in the representation space transitively. In characteristic 3 the simple representations are the trivial representation and
the sign representation. We may show that these lists are complete in various ways. For
example, we could use Theorem 9.11. Another approach is to say that all 1-dimensional
representations are representations of the abelianization C2 of S3 , giving one such representation in characteristic 2 and two in characteristic 3. Now use a count of dimensions of
the simple modules together with Theorem 7.14 to show that the lists are correct.
It is now a question of calculating composition factors of the reductions from characteristic 0. The 1-dimenional representations remain simple on reduction. The 2-dimensional
representation has been given as a Z-form, and it suffices to compute composition factors
when the matrices are interpreted as have entries in F2 and F3 . Over F2 this representation
is simple, as has been observed.
Peter Webb
151
152
Coming out of the proof of the last result we see that in the situation where |G| is
relatively prime to p, an0 RG-module L is projective if and only if it is projective as an
R-module, and furthermore that for each F G-module U , all R-forms of U are isomorphic
as RG-modules. We leave the details of this as an exercise.
When |G| is divisible by p the decomposition matrix cannot be the identity, because
as a consequence of Theorem 9.11 it is not even square. We state without proof a theorem
which says that when G is p-solvable the decomposition matrix does at least contain the
identity matrix as a submatrix of the maximum possible size. A group G is said to be
p-solvable if it has a chain of subgroups
1 = Gn G1 G0 = G
so that each factor Gi /Gi+1 is either a p-group or a group of order prime to p. A proof
can be found in [CRI, 22.1]
(9.21) THEOREM (Fong, Swan, Rukolaine). Let (F, R, k) be a splitting p-modular
system for a p-solvable group G. Then every simple kG-module is the reduction modulo
() of an RG-lattice.
Peter Webb
153
simple F G-module we write [T ] for the corresponding basis element of G0 (F G). Similarly
if S is a simple kG-module we write [S] for the corrsponding basis element of G0 (kG), and
if P is an indecomposable projective kG-module we write [P ] for the corresponding basis
element of K0 (kG). Extending this notation, if U is any kG-module with composition
factors S1 , . . . , Sr occurring with multiplicities n1 , . . . , nr in some composition series of U ,
we write
[U ] = n1 [S1 ] + + nr [Sr ] G0 (kG).
The fact that this is well defined depends on the JordanHolder theorem. There is a similar
interpretation of [V ] G0 (F G) if V happens to be an F G-module. This time because V
is semisimple it is the direct sum of its composition factors, and so if V = T1n1 Trnr
we put [V ] = n1 [T1 ] + + nr [Tr ] G0 (F G). In the same way if P = P1n1 Prnr
where the Pi are indecomposable projective kG-modules we put
[P ] = n1 [P1 ] + + nr [Pr ] K0 (kG).
Since simple F G-modules biject with their characters, we may identify G0 (F G) with the
subset of the space of class functions Ccc(G) consisting of the Z-linear combinations of the
characters of the simple modules as considered in Chapter 3. Such Z-linear combinations of
characters are termed virtual characters of G, so G0 (F G) is the group of virtual characters
of F G.
We now define the homomorphisms of the cde triangle, which is as follows:
G0 (F G)
e
K0 (kG)
G0 (kG)
154
Peter Webb
155
and in characteristic 3
156
P1 = 1 + 2
P = + 2 .
We now have a second proof of the symmetry of the Cartan matrix, but perhaps more
importantly an extremely good way to calculate it. The effectiveness of this approach will
be increased once we know about Brauer characters, which are treated in the next section.
(9.27) COROLLARY. Let (F, R, k) be a splitting p-modular system for G. Then the
Cartan matrix C = DT D. Thus C is symmetric.
Examples. 1. When G = S3 the Cartan matrices in characteristic 2 and in characteristic 3, expressed as a product DT D, are
1 0
1 0
1 1 0
2 0
1 0 1
2 1
1 0 =
and
0 1 =
.
0 0 1
0 1
0 1 1
1 2
0 1
1 1
The decomposition matrices were calculated earlier in this chapter. In characteristic 2 the
Cartan matrix follows from Example 7.5, and in characteristic 3 it follows from Proposition
8.9. The decomposition matrix factorization provides a new way to compute the Cartan
matrices.
2. Results 9.25 and 9.27 fail without the hypothesis that the p-modular system is
splitting. An example of this is provided by the alternating group A4 with the 2-modular
system (Q2 , Z2 , F2 ). The Cartan matrix of F2 A4 was computed in Exercise 2 of Chapter
8, and we leave it as a further exercise to compute the matrices of d and e here.
The equality of dimensions that played the key role in the proof of Theorem 9.25 can
be nicely expressed in terms of certain bilinear pairings between the various Grothendieck
groups. On the vector space of class functions on G we already have defined a Hermitian
form and on the subgroup G0 (F G) it restricts to give a bilinear form
h ,
i : G0 (F G) G0 (F G) Z
specified by h[U ], [V ]i = dim HomF G (U, V ) when U and V are F G-modules. We also have
a pairing
h , i : K0 (kG) G0 (kG) Z
Peter Webb
157
158
where T is the character of T . Observe that n/|G| R because pd n, and also T (g 1 )
lies in R since it is a sum of roots of unity, and roots of unity (in some extension ring if
necessary) have valuation 1, so lie in R. Thus e RG. It follows that RG = eRG (1
e)RG as a direct sum of rings.
The homomorphism : F G EndF (T ) that expresses the action of G on T identifies
eF G with the matrix algebra EndF (T ), and has kernel (1 e)F G. The restriction of to
RG takes values in EndR (T0 ) because T0 is a full RG-sublattice of T . The kernel of this
restriction is (1 e)F G RG = (1 e)RG which is a direct summand of RG. We will
show that this restricted homomorphism is surjective to EndR (T0 ) and from this it will
follow that eRG
= EndR (T0 )
= Mn (R), a direct summand of RG.
As an extension of the formula in Theorem 3.23 for the primitive central idempotent
corresponding to T , we claim that if EndF (T ) then
=
n X
tr((g 1 ))(g).
|G|
gG
To demonstrate this it suffices to consider the case = (h) where h G, since these
elements span EndF (T ). In this case
n X
n X
tr((g 1 )(h))(g) = (h)
tr((g 1 h))(h1 g)
|G|
|G|
gG
gG
= (h)(e)
= (h)
using the previously obtained formula for e. This shows that the claimed formula holds
when = (h), and hence holds in general.
Finally we may see that the restriction of is a surjective homomorphism RG
P
n
1
EndR (T0 ), since any EndR (T0 ) is the image under of |G|
))g RG.
gG tr((g
This completes the proof of this implication.
(2) (3) Certainly T0 is projective as a module for EndR (T0 ) since it identifies with
the module of column vectors for this matrix algebra. Assuming (2), we have that T0 is
a projective RG-module, since RG acts via its summand eRG which identifies with the
matrix algebra. Furthermore, F G acts on T
= F R T0 as column vectors for a matrix
algebra over F , so T is a simple F G-module.
(2) (4) The decomposition RG = eRG(1e)RG with eRG
= Mn (R) is preserved
on reducing modulo (), and we obtain kG = ekG (1 e)kG where e is the image of e
in kG. Furthermore ekG
= Mn (k) because it is the reduction module () of Mn (R), and
the action of kG on T0 /T0 is via projection onto ekG.
(3) (5) Since T0 is a direct summand of a free RG-module it follows that T0 /T0
is a direct summand of a free kG-module, and hence is projective. Furthermore, we claim
that T0 /T0 is indecomposable. To see this, write T0 /T0 = PS1 PSt where the PSi
are indecomposable projectives, so that T0 /T0 is the projective cover of S1 St as
Peter Webb
159
160
hgi
hgi
so by uniqueness of projective covers P
= S hsi . We deduce that = S hsi . It follows
that (g) = 0 from the formula for an induced character, since no conjugate of g lies in
hsi.
Peter Webb
161
Summary of Chapter 9
1. Every finite dimensional algebra over a field has a splitting field of finite degree.
2. If two modules are isomorphic after extending the ground field, they were originally
isomorphic.
3. If k is a splitting field of characteristic p, the number of simple kG-modules equals
the number of p-regular conjugacy classes in G.
4. Idempotents lift from the group ring over the residue field to the group ring over a
complete discrete valuation ring.
5. Projective covers exist over RG when R is a complete discrete valuation ring. The
indecomposable projective RG-modules are the projective covers PS , where S is simple.
6. When R is a principal ideal domain with field of fractions F , every F G-module
can be written in R.
7. The decomposition map is well defined and C = DT D. The decomposition matrix
also computes the ordinary characters of the projectives PS .
8. If k is a splitting field of characteristic p where p 6 |G|, the representations of G
over C and over k are the same in a certain sense.
9. Blocks of defect zero are identified by the fact that the degree of their ordinary
character is divisible by the p-part of |G|. Such characters vanish on elements of order
divisible by p. They remain irreducible on reduction mod p, where they give a projective
module and a matrix ring summand of kG.
162
Peter Webb
163
1
0
1
1
(b) =
1 t
0 1
Show that this representation is absolutely indecomposable, meaning that it remains indecomposable under all field extensions. Show also that this representation cannot be written
in any proper subfield of F2 (t).
9. Let (F, R, k) be a p-modular system and suppose that R is complete. Let L and
M be RG-lattices, where G is a finite group.
(a) Show that L is a projective RG-module if and only if L/L is a projective kGmodule.
[Consider the projective cover of L.]
(b) Deduce that if L/L
= M/M as kG-modules and that L/L is a projective kGM
as
RG-modules.
In other words, projective kG-modules lift uniquely
module then L
=
to RG-lattices.
10. Let (F, R, k) be a p-modular system and G a finite group. Show that if U = U1 U2
is a finite-dimensional F G-module and L is a full RG-lattice in U then L U1 , L U2 are
full RG-lattices in U1 and U2 , but that it need not be true that L = (L U1 ) (L U2 ).
[Consider the regular representation when G = C2 .]
11. Let U be the 2-dimensional representation of S3 over Q that is defined by requiring
that with respect to a basis u1 , u2 the elements (1, 2, 3) and (1, 2) act by matrices
0 1
1 1
and
1
0
1
1
164
13. Let (F, R, k) be a splitting p-modular system for the group G and suppose that
|G| is relatively prime to p. Let L be a finitely generated RG-module. Show that L
is projective if and only if it is projective as an R-module. Show further that for each
finite-dimensional F G-module U , all R-forms of U are isomorphic as RG-modules.
14. Consider the cde triangle for A4 with the 2-modular system (Q2 , Z2 , F2 ). Compute
the matrices DT and E of the maps d and e with respect to the bases for the Grothendieck
groups described in this section. Verify that E 6= D, but that the Cartan matrix does
satisfy C = DT E and is not symmetric. [Compare Chapter 8 Exercise 2. You may assume
that Q2 contains no primitive third root of 1. This fact follows from the discussion of roots
of unity at the start of the next section, together with the fact that F2 does not contain a
primitive third root of 1.]
15. Give a proof of the following result by following the suggested steps.
THEOREM. Let E F be a field extension of finite degree and let A be an F algebra. Let U and V be A-modules. Then
E F HomA (U, V )
= HomEF A (E F U, E F V )
via an isomorphism F f 7 ( F u 7 F f (u)).
(a) Verify that there is indeed a homomorphism as indicated.
(b) Let x1 , . . . , xn be a basis for E as an F -vector space. Show that for any F -vector
Pn
space M , each element of E F M can be written uniquely in the form i=1 xi F mi
with mi M .
Pn
Pn
(c) Show that if an element i=1 xi fi E F HomA (U, V ) maps to 0 then i=1 xi
fi (u) = 0 for all u U . Deduce that the homomorphism is injective.
(d) Show that the homomorphism is surjective as follows: given an E F A-module
Pn
homomorphism g : E F U E F V , write g(1F u) = i=1 xi fi (u) for some elements
fi (u) V . Show that this defines A-module homomorphisms fi : U V . Show that g is
Pn
the image of i=1 xi fi .
16. Let E F be a field extension of finite degree and let A be an F -algebra.
(a) Let 0 U V W 0 be a short exact sequence of A-modules. Show that
the sequence is split if and only if the short exact sequence 0 E F U E F V
E F W 0 is split.
(b) Let U be an A-module. Show that U is projective if and only if E F U is projective
as an E F A-module.
(c) Let U be an A-module, S an absolutely simple A-module and let LS (U ) denote the
largest submodule of U that is a direct sum of copies of S, as in Corollary 1.8. Show that
LEF S (E F U )
= E F (U G ).
= E F LS (U ). Deduce for fixed points that (E F U )G
Prove a similar result for the largest quotient of U that is a direct sum of copies of S and
hence a similar result for the fixed quotient of U .
Peter Webb
165
(d) Show that if F is assumed to be a splitting field for A and U is an A-module then
Rad(E F U )
= E F Rad(U ) and Soc(E F U )
= E F Soc(U ). Explain why this does
not contradict the conclusion of question 1.
[Hint: use the most promising of the equivalent conditions of Propositions 7.2 and 7.3
in combination with the result of question 15.]
17. Let R be a complete discrete valuation ring with residue field k of characteristic
p. Let g GL(n, k) be an n n-matrix with entries in k. Suppose g has finite order s
and suppose that both R and k contain primitive sth roots of unity. Show that there is
an n n-matrix g GL(n, R) of order s whose reduction to k is g.
18. Show that, over a splitting 2-modular system (F, R, k), the dihedral group D30
has seven 2-blocks of defect zero and two further 1-dimensional ordinary characters. Hence
find the degrees of the simple kD30 -modules. Find the decomposition matrix and Cartan
matrix in characteristic 2. Show that kD30
= kC2 M2 (k)7 as rings.
166
The polynomial X a 1 is separable both in F [X] and k[X] since its formal derivative
d
a
a1
is not zero and has no factors in common with X a 1, so both
dX (X 1) = aX
F and k are cyclic groups of order a. Also F R since roots of unity have value 1
under the valuation. We claim that the quotient homomorphism R R/() = k gives
. This is because X a 1 reduces to the
an isomorphism F k , which we write
polynomial that is written the same way, and so its linear factors over F must map to the
over F and
complete set of linear factors over k. These linear factors have the form X
X over k, so we obtain a bijection between the two groups of roots of unity.
Let g G be a p-regular element, and let : G GL(U ) be a representation over k.
Then (g) is diagonalizable, since khgi is semisimple and all eigenvalues of (g) lie in k,
being ath roots of unity. If the eigenvalues of (g) are 1 , . . . , n we put
1 + +
n,
U (g) =
Peter Webb
167
and this is the Brauer character of U . It is a function that is only defined on the p-regular
elements of G, and takes values in a field of characteristic zero, which we may always take
to be C.
0 1
Example. Working over F2 , the specification (g) =
provides a 2-dimensional
1 1
representation U of the cyclic group hgi of order 3. The characteristic polynomial of this
matrix is t2 + t + 1 and its eigenvalues are the primitive cube roots of unity in F4 . These
lift to primitive cube roots of unity in C, and so U (g) = e2i/3 + e4i/3 = 1. It is very
tempting in this situation to observe that the trace of (g) is 1, which can be lifted to
1 C, and thus to suppose that U (g) = 1; however, this supposition would be incorrect.
We list the immediate properties of Brauer characters.
and
(1)
(2)
(3)
(4)
(5)
(6)
Proof. (1) In its action on U the identity has dimk U eigenvalues all equal to 1. They
all lift to 1 and the sum of the lifts is dimk U .
(2) This follows because g and xgx1 have the same eigenvalues.
(3) The eigenvalues of g 1 on U are the inverses of the eigenvalues of g on U , as are
the eigenvalues of g on U (since here g acts by the inverse transpose matrix). The lifting
lifts then
1
of roots of unity is a group homomorphism, so the result follows since if
lifts 1 .
(4) If g is a p-regular element then U and V have bases u1 , . . . , ur and v1 , . . . , vs
consisting of eigenvectors of g with eigenvalues 1 , . . . , r and r , . . . , s , respectively. Now
the tensors ui uj form a basis of eigenvectors of U V with eigenvalues i j . Their lifts
j since lifting is a group homomorphism, and P
j = (P
P
are d
i j = i
i,j i
i i )(
j j )
so that UV (g) = U (g)V (g).
(5) If g is a p-regular element then khgi is semisimple so that V
= U W as khgimodules. It follows that the eigenvalues of g on V are the union of the eigenvalues on V
and on W (taken with multiplicity), and from this V (g) = U (g) + W (g) follows. If
U
= U1 we may consider the sequence 0 U1 U 0 0 to see that U = U1 . The
final sentence follows by an inductive argument.
168
then g acts on U = U
/ U
S3
ordinary characters
S3
Brauer simple p = 2
S3
Brauer projective p = 2
g
|CG (g)|
() (12) (123)
6 2
3
g
|CG (g)|
() (123)
6
3
g
|CG (g)|
() (123)
6
3
1
sign
2
1 1
1 1
2 0
1
2
1
1
2 1
1
2
2
2
2 1
1
1
1
S3
Brauer projective p = 3
g
|CG (g)|
() (12)
6 2
g
|CG (g)|
() (12)
6 2
1
sign
1 1
1 1
1
sign
3 1
3 1
Peter Webb
169
1
sign
2
3a
3b
1
1
1 1
2
0
3 1
3
1
1
1
2
1
1
1
1
0
1
1
1
1
1
0
0
as seen in Example 3.11. In characterstic 2, S4 has two simple modules, namely the trivial
module and the 2-dimensional module of S3 , made into a module for S4 via the quotient
homomorphism S4 S3 . Both of these lift to characteristic zero, and so the Brauer
characters of the simple 2-modular simple representations are
S4
Brauer simple p = 2
g
|CG (g)|
() (123)
24
3
1
2
1
2
1
1
which is the same as for S3 . The Brauer characters of the reductions modulo 2 of the
ordinary characters of S4 are
S4
Reductions from characteristic 0 to 2
g
|CG (g)|
() (123)
24
3
1
sign
2
3a
3b
1
1
2
3
3
1
1
1
0
0
We see from this that the sign representation reduces modulo 2 to the trivial representation, and reductions of the two 3-dimensional representations each have the 2-dimensional
representation and the trivial representation as composition factors with multiplicity 1.
This is because the corresponding Brauer characters are expressible as sums of the simple
170
Brauer characters in this way, and since the simple Brauer characters are visibly independent there is a unique such expression. This expression has to be the expression given by
the composition factor multiplicities in the manner of Proposition 10.1 part (5). It follows
that the decomposition and Cartan matrices for S4 at the prime 2 are
1 0
1 0
1 0
1 0
4 2
1 1 0 1 1
.
D = 0 1 and C =
0 1 =
2 3
0 0 1 1 1
1 1
1 1
1 1
1 1
Knowing the Cartan matrix and the simple Brauer characters we may now compute the
Brauer characters of the indecomposable projective representations.
In characteristic 3 the trivial representation and the sign representation are distinct
1-dimensional representations, and we also have two non-isomorphic 3-dimensional representations that are the reductions modulo 3 of the two 3-dimensional ordinary representations. This is because these 3-dimensional representations are blocks of defect zero, and
by Theorem 9.28 they remain simple on reduction modulo 3. This constructs four simple
representations in characteristic 3, and this is the complete list by Theorem 9.11 because
S4 has four 3-regular conjugacy classes. Thus the table of Brauer characters of simple
modules in characteristic 3 is
S4
Brauer simple p = 3
g
|CG (g)|
1
sign
3a
3b
1
1
1 1
3 1
3
1
1
1
1
1
1
1
1
1
and each is the reduction of a simple module from characteristic zero. The remaining 2dimensional ordinary representation has Brauer character values 2, 0, 2, 0 and since this
Brauer character is uniquely expressible as a linear combination of simple characters,
namely the trivial Brauer character plus the sign Brauer character, these two 1-dimensional
modules are the composition factors of any reduction modulo 3 of the 2-dimensional representation. We see that the decomposition and Cartan matrices for S4 in characteristic 3
are
1 0 0 0
2 1 0 0
0 1 0 0
1 2 0 0
D = 1 1 0 0 and C = DT D =
.
0 0 1 0
0 0 1 0
0 0 0 1
0 0 0 1
This information now allows us to compute the Brauer characters of the projective representations.
Peter Webb
171
Proof. We may assume without loss of generality that R is complete. For, if necessary,
replace R by its completion at (), and let F be the field of fractions of R. Making this
change does not alter the residue field k or the Brauer characters of representations, so the
equation we have to establish is unaltered. Assuming that R is complete we may now lift
projective modules from kG to RG.
We make use of the isomorphism HomkG (U, V )
= HomkG (U k V , k) whenever U
and V are finite-dimensional kG-modules, which holds since both sides are isomorphic to
(U k V )G , using results 3.3 and 3.4. Now HomkG (P, U )
= HomkG (P U , k) and P U
d
is a projective kG-module by 8.4. Thus it lifts to a projective RG-lattice P
k U and we
have
dim HomkG (P, U ) = dim HomkG (P U , k)
d
= rank HomRG (P
k U , R)
d
= dim HomF G (F R (P
k U ), F )
1 X
(g 1 )k (g).
=
d
F R (P
kU )
|G|
gG
1 X
(g 1 ).
)
d
F
(P
U
R
k
|G|
gG
We claim that
d
F R (P
kU )
(g
)=
P (g 1 )U (g) if g is p-regular,
0
otherwise,
and from this the result follows. If g is not p-regular it has order divisible by p and the
d
character value is zero by Proposition 9.29, since P
k U is projective. When g is p-regular
172
we calculate the character value by using the fact that it depends only on the structure
of P and U as khgi-modules. Since khgi is a semisimple algebra both P and U are now
c is isomorphic to
projective, and they lift to Rhgi-lattices whose tensor product P R U
d
P
k U as Rhgi-modules, since both of these are projective covers as Rhgi-modules of
P k U . From this we see that
F
d (g
R P k U
1
) = (F P )(F U
c ) (g )
1
= F P (g 1 )F U
c (g )
= P (g 1 )U (g)
as required.
It is convenient to interpret the formula of the last proposition in terms of an inner
product on a space of functions, in a similar way to what we did with ordinary characters.
Let preg(G) denote the set of conjugacy classes of p-regular elements of G, so that
preg(G) cc(G) where the latter denotes the set of all conjugacy classes of G. Since
Brauer characters are constant on the conjugacy classes of p-regular elements we may
regard them as elements of the vector space Cpreg(G) of functions
preg(G) C.
We define a Hermitian form on this vector space by
h, i =
1
|G|
(g)(g)
pregular gG
and just as with the similarly-defined bilinear form on Ccc(G) we note that
h, i = h, i
where (g) = (g) is the complex conjugate of (g). If and are the Brauer characters
of representations we have (g) = (g 1 ) so that h, i = h, i = h , i = h , i.
With the notation of this bilinear form the last result now says that if P and U are
finite-dimensional kG-modules with P projective then
dim HomkG (P, U ) = hP , U i.
(10.3) THEOREM (Row orthogonality relations for Brauer characters). Let G be a
finite group and k a splitting field for G of characteristic p. Let S1 , . . . , Sn be a complete
list of non-isomorphic simple kG-modules, with projective covers PS1 , . . . , PSn . Then the
Brauer characters S1 , . . . , Sn of the simple modules form a basis for Cpreg(G) , as do also
Peter Webb
173
the Brauer characters PS1 , . . . , PSn of the indecomposable projective modules. These
two bases are dual to each other with respect to the bilinear form, in that
hPSi , Sj i = i,j .
The bilinear form on Cpreg(G) is non-degenerate.
Proof. Since PSi has Si as its unique simple quotient we have HomkG (PSi , Sj ) = 0
unless i = j, in which case HomkG (PSi , Sj )
= EndkG (Si ). Because k is a splitting field,
this endomorphism ring is k and so hPSi , Sj i = i,j . Everything follows from this and
the fact that the number of non-isomorphic simple modules equals the number of p-regular
Pn
Pn
conjugacy classes of G. Thus if i=1 j Sj = 0 we have j = hPSi , i=1 j Sj i = 0,
which shows that the Sj are independent, and hence form a basis. By a similar argument
the PSi also form a basis. The matrix of the bilinear form with respect to these bases is
the identity matrix and it is non-degenerate.
This result says that the rows of the matrices of Brauer characters of simples and projectives are orthogonal to each other, provided that entries from each column are weighted
by the reciprocal of the centralizer order of the element that parametrizes the column, this
being the number of conjugates of the element divided by |G|. We will give examples of this
later. The result implies, of course, that the Brauer characters of the simple kG-modules
are linearly independent as functions on the set of p-regular conjugacy classes of G, a fact
we observed and used in the earlier examples. It has the consequence that the information
contained in a Brauer character is exactly that of composition factor multiplicities and we
see this as part (c) of the next result. Since there is the tacit assumption with Brauer
characters that we might have extended the field of definition to include roots of unity, we
describe the behaviour of composition factors under field extension.
(10.4) COROLLARY. Let E k be a field extension and let U and V be finitedimensional kG-modules.
(a) If S and T are non-isomorphic simple kG-modules then E k S and E k T have no
composition factors in common.
(b) U and V have the same composition factors as kG-modules if and only if E k U and
E k V have the same composition factors as EG-modules.
(c) U and V have the same composition factors if and only if their Brauer characters U
and V are equal.
Proof. (a) As in Theorem 7.13, let fS kG be an idempotent such that fS S = S and
fS T = 0. Now fS (E k S) = E k S and fS (E k T ) = 0, so fS acts as the identity on
all composition factors of E k S, but as zero on all composition factors of E k T . Thus
E k S and E k T can have no composition factors in common.
174
(b) The implication from left to right is immediate since extending scalars is exact.
Conversely, the composition factors of E k U are precisely the composition factors of the
E k S where S is a composition factor of U . Since non-isomorphic S and T give disjoint
sets of composition factors in E k S and E k T , if the composition factors of E k U
and E k V are the same, the composition factors of U and V that give rise to them must
be the same.
(c) By part (b) we may extend k if necessary to assume that it is a splitting field. We
know from part (5) of 10.1 that U is the sum of the Brauer characters of the composition
factors of U , and similarly for V , so if U and V have the same composition factors then
U = V . Conversely, the composition factors of U are determined by U since by Theorem
10.3 the Brauer characters of simple modules form a basis for Cpreg(G) . Hence if U = V
then U and V must have the same composition factors.
Computing the Brauer character of a module and expressing it as a linear combination
of simple characters is a very good way of finding the composition factors of the module,
and some examples of this appear in the exercises. As another application we may deduce
that if S is a simple kG-module then S
= S if and only if the Brauer character S takes
real values, since this is the condition that S = S = S .
It also follows from Theorem 10.3 that the Brauer characters of the indecomposable
projective modules are independent functions on preg(G), and so Brauer characters enable us to distinguish between projective modules. Since the Brauer characters of a module
are determined by its composition factors this has the following consequence.
(10.5) COROLLARY. Let P and Q be finitely generated projective kG-modules,
where k is a splitting field of characteristic p. Then P and Q are isomorphic if and
only if they have the same composition factors. The Cartan matrix is invertible. The
decomposition matrix has maximum rank.
Pn
Proof. Write P = PSa11 PSann and Q = PSb11 PSbnn so that P = i=1 ai PSi
Pn
and Q =
b . Then P
= Q if and only if ai = bi for all i, if and only if
i=1
Pn
Pni PSi
i=1 ai PSi =
i=1 bi PSi since the PSi are linearly independent, if and only if P = Q .
By the last corollary this happens if and only if P and Q have the same composition factors.
We claim that the kernel of the Cartan homomorphism c : K0 (kG) G0 (kG) is zero.
Any element of K0 (kG) can be written [P ] [Q] where P and Q are projective modules,
and such an element lies in the kernel if and only if P and Q have the same composition
factors. This forces P
= Q, so that the kernel is zero. It now follows that the Cartan
homomorphism is an isomorphism.
For the final assertion about the decomposition matrix we use the fact that C = DT D.
Thus rank C rank D. The number of columns of D equals the number of columns of C,
and this number must be the rank of D.
Peter Webb
175
A finer result than this is true. The determinant of the Cartan matrix over a splitting field of characteristic p is known to be a power of p, and the decomposition map
d : G0 (F G) G0 (kG) is, in fact, a surjective homomorphism of abelian groups. The
assertion in 10.5 that its matrix has maximum rank implies only that the cokernel of the
decomposition map is finite. These stronger results may be proved by a line of argument that originates with the induction theorem of Brauer, a result that we have omitted.
The surjectivity of of the decomposition map is also implied by the Fong-Swan-Rukolaine
theorem 9.21 in the case of p-solvable groups.
We may write the row orthogonality relations in various ways. The most rudimentary
way, for simple kG-modules S and T , is the equation
n
X
1
1 if S
= T,
PT (g 1 )S (g) =
|G|
0 otherwise.
pregular gG
We can also express this as a matrix product. Let be the table of Brauer character
values of simple kG-modules, the table of Brauer character values of indecomposable
projective modules, and let B be the diagonal matrix whose entries are |CG1(g)| as g ranges
through the p-regular classes. The row orthogonality relations may now be written as
BT = I.
Note that the independence of the Brauer characters of simple modules and of projective
modules is equivalent to the property that and are invertible matrices, and also that
= C where C is the Cartan matrix. From this we now deduce the column orthogonality
relations for Brauer characters, which say that each column of is orthogonal to the
remaining columns of , and has product with the corresponding column of equal to
the order of the centralizer of the group element that indexes the column.
(10.6) PROPOSITION (Column orthogonality relations for Brauer characters). With
the above notation,
|C (x )|
0
G
T
=
0
..
.
|CG (x2 )|
..
.
.
|CG (xn )|
..
Proof. We take the equation BT = I and multiply on the left by and on the
1
right by (T )1 to get B = (T )1 . Inverting both sides gives B 1 = T . We
finally take the complex conjugate of both sides, observing that B is a real matrix.
176
1
sign
3a
3b
1
1
1 1
3 1
3
1
1
1
1
1
1
1
1
1
2
1
0
0
1
2
0
0
0
0
1
0
0
0
0
1
Peter Webb
177
1
sign
3a
3b
3
1
3 1
3 1
3
1
3
3
1
1
1
1
1
1
3a3a = 3a 3a
Taking first the inner products hPS , 3a3a i with the Brauer characters of projectives we
obtain the numbers
27 1 3 1
+ + + =2
24 4 8 4
27 1 3 1
+ =1
24 4 8 4
27 1 1 1
+ =1
24 4 8 4
27 1 1 1
+ =1
24 4 8 4
which means that 3a 3a has composition factors 1 (with multiplicity 2), , 3a, 3b where
is the sign representation. On the other hand 3a 3a is itself projective, and we get more
information if we take the inner products h3a3a , S i with the simple Brauer characters,
giving numbers
1 1 1
9
+ + + =1
24 4 8 4
1 1 1
9
+ =0
24 4 8 4
27 1 1 1
+ =1
24 4 8 4
27 1 1 1
+ = 1.
24 4 8 4
178
G0 (F G) Ccc(G)
[S] 7 S ,
and
[T ] 7 T
and in fact these same formulas hold whenever P is an arbitrary finitely-generated projective module and S, T are arbitrary finitely-generated modules. Extending scalars from Z
to C they define isomorphisms of vector spaces
C Z K0 (kG)
= Cpreg(G)
C Z G0 (kG)
= Cpreg(G)
C Z G0 (F G)
= Ccc(G) .
There is further structure that we have not mentioned yet, which is that there is a
multiplication defined on the three Grothendieck groups by [U ] [V ] := [U V ], the same
formula working for all three groups but with the modules U and V interpreted suitably
over F G or kG and projective or not as appropriate. Since tensor product over the ground
field preserves exact sequences, this definition makes sense for arbitrary (finitely generated)
modules. This multiplication makes G0 (F G) and G0 (kG) into rings with identity, the
identity element being the class of the trivial module. We see that the isomorphisms just
given preserve the product structure.
(10.7) PROPOSITION. After extending scalars to C, the Grothendieck groups C Z
G0 (F G) and C Z G0 (kG) are semisimple rings.
Proof. The isomorphisms to Ccc(G) and Cpreg(G) are ring isomorphisms, and the
latter algebras are direct sums of copies of C, which are semisimple.
Peter Webb
179
This means that after extending scalars to C we may identify the cde triangle with
the following diagram of vector spaces
Ccc(G)
e
Cpreg(G)
Cpreg(G)
Summary of Chapter 10
1. Brauer characters take values in C. They satisfy similar properties to ordinary
characters, except that they are only defined on p-regular elements of G.
2. Two Brauer characters U and V are equal if and only if U and V have the same
composition factors.
3. The tables of Brauer characters of simple modules and of projective modules satisfy
orthogonality relations.
180
1
|H|
(t1 gt)
tG
t1 gtH
(t1 gt).
t[G/H]
t1 gtH
This is the same as the formula in Proposition 4.10 that was used to define induction of
ordinary class functions.
(a) Let U be a finite-dimensional kH-module with Brauer character U , where k is
G
G
a field of characteristic p. Prove that UG
= U G
H , and that e( H ) = e() H and
H
G
d( G
H ) = d() H if is a class function.
preg(G)
(b) Similarly define the restriction G
and show that similar
H where C
formulas hold.
(c) Using the Hermitian forms defined on these spaces of functions, show that h G
H
, i = h, G
i
always
holds.
H
2. Let (F, R, k) be a splitting p-modular system for G. Let P and Q be finitelygenerated projective RG-modules such that F R P
= F R Q as F G-modules. Show that
P = Q.
3. The simple group GL(3, 2) has order 168 = 8 3 7. The following is part of its
ordinary character table (the numbers that label the conjugacy classes of elements in the
top row indicate the order of the elements):
GL(3, 2)
ordinary characters
g
|CG (g)|
1
3
?
6
7
8
1 2 4
168 8 4
1 1 1
3 1 1
3 7a 7b
3 7 7
1
6 2 0
1 1
7 1 1 1
8
1 1 1
Peter Webb
181
(c) Find the decomposition matrix and Cartan matrix of GL(3, 2) at the prime 2.
(d) Write down the table of Brauer characters of projective F2 [GL(3, 2)]-modules.
(e) Determine the direct sum decomposition of the module 83 (where 8 and 3 denote
the simple F2 [GL(3, 2)]-modules of those dimensions shown in the table), as a direct sum
of indecomposable modules.
(f) Determine the composition factors of 3 3 and 3 3 .
[Note that 3 can be taken to be the natural 3-dimensional F2 [GL(3, 2)]-module. One
approach is to use the orthogonality relations.]
4. It so happens that GL(3, 2)
= P SL(2, 7). In this exercise we regard this group as
P SL(2, 7) and assume the construction of the simple modules over F7 given in exercise 21
of Chapter 6.
(a) Construct the table of Brauer characters of simple F7 [P SL(2, 7)]-modules.
[It will help to observe that all elements of order prime to 7 in SL(2, 7) act semisimply
on F27 by Maschkes theorem. From this it follows that there is a unique element of order
2 in SL(2, 7), namely I, since its eigenvalues must both be 1, and it is semisimple.
We may deduce that an element of order 8 in SL(2, 7) represents an element of order 4
in P SL(2, 7), and its square represents an element of order 2, since I represents 1 in
P SL(2, 7). We may now determine the eigenvalues of 7-regular elements in their action
on the symmetric powers of the 2-dimensional module.]
(b) Compute the decomposition and Cartan matrices for P SL(2, 7) in characteristic
7. Show that the projective cover of the trivial module, P1 , has just four submodules,
namely 0, P1 and two others.
5. Let (F, R, k) be a p-modular system for some prime p. For the following two
statements, show by example that the first is false in general, and that the second is true.
(a) Suppose that A is an invertible matrix with entries in R and let A be the matrix
with entries in k obtained by reducing the entries of A modulo (). Then the eigenvalues
of A are the lifts of the eigenvalues of A.
(b) Suppose that A is an invertible matrix with entries in R and let A be the matrix
with entries in k obtained by reducing the entries of A module (). Suppose further that
A has finite order, and that this order is prime to p. Then the eigenvalues of A are the
lifts of the eigenvalues of A. Furthermore, the order of A is the same as the order of A.
6. (Modular version of Moliens Theorem. Copy the approach of Chapter 4 Exercise 14
and assume Exercise 5.) Let G be a finite group and let (F, R, k) be a splitting p-modular
system for G, for some prime p.
(a) Let : G GL(V ) be a representation of G over k, and for each n let S n (V ) be
d
the Brauer character of the nth symmetric power of V . For each p-regular g G let (g)
be a matrix with entries in R of the same order as (g) that reduces modulo () to (g).
Show that for each p-regular g G there is an equality of formal power series
n=0
S n (V ) (g)tn =
1
d
det(1 t(g))
182
X
i=0
Show that
MV (W ) =
1
|G|
hPW , S n (V ) itn .
gpreg(G)
PW (g 1 )
.
d
det(1 t(g))
Show that S 8 (V ) is the direct sum as a F2 S3 -module of three copies of V and a 3dimensional module whose composition factors are all trivial. [Note that the ring of
1
invariants (S (V ))G has series (1t2 )(1t
3 ) in this case, that is not the same as MV (k).]
(d) When G = GL(3, 2) and V is the natural 3-dimensional F2 GL(3, 2)-module, use
the information from Exercise 3 to show that
1 t + t4 t7 + t8
(1 t)(1 t3 )(1 t7 )
t(1 t3 + t4 + t5 )
MV (V ) =
(1 t)(1 t3 )(1 t7 )
t2 (1 + t t2 + t5 )
MV (V ) =
(1 t)(1 t3 )(1 t7 )
t4
MV (8) =
(1 t)(1 t3 )(1 t7 )
MV (k) =
7. Show that when k is an arbitrary field of characteristic p, the number of isomorphism classes of simple kG-modules is at most the number of p-regular conjugacy classes
in G.
8. Let k be a splitting field for D30 in characteristic 3. Using the fact that D30
=
C3 D10 has a normal Sylow 3-subgroup, show that kD30 has four simple modules of
Peter Webb
183
dimensions 1, 1, 2 and 2 and that the values of the Brauer characters on the simple
modules are the same as the ordinary character table of D10 . We will label the simple
modules k1 , k , U a and U b respectively. Using the method of Proposition 8.9, show that
the indecomposable projectives have the form
Pk1
k1
|
= k
|
k1
Pk
k
|
= k1
|
k
PUa
U a
U b
|
|
= U a PUb = U b
|
|
U a
U b
2
1
C=
0
0
1
2
0
0
0
0
3
0
0
0
0
3
184
Peter Webb
185
(1)
(2)
(3)
(4)
(5)
(6)
186
Proof. (1) (3) Let I be the unique maximal left ideal of B. Since Rad(B) is the
intersection of the maximal left ideals, it follows that I = Rad(B). If a B I then
Ba is a left ideal not contained in I, so Ba = B. Thus there exists x B with xa = 1.
Furthermore x 6 I, so Bx = B also and there exists y B with yx = 1. Now yxa = a = y
so a and x are 2-sided inverses of one another. This implies that B/I is a division ring.
(1) (6) The argument just presented shows that the unique maximal left ideal I is
in fact a 2-sided ideal, and every element not in I is invertible. This implies that every
non-invertible element is contained in I. Equally, no element of I can be invertible, so I
consists of the non-invertible elements, and they form a 2-sided ideal.
(3) (1) If I is a maximal left ideal of B then I Rad(B) and so corresponds to
a left ideal of B/ Rad(B), which is a division ring. It follows that either I = Rad(B) or
I = B, and so Rad(B) is the unique maximal left ideal of B.
(4) (1) Let J be the set of non-invertible elements of B and I a maximal left ideal.
Then no element of I is invertible, so I J. Since J is an ideal, we have equality, and I
is unique.
(6) (4) This implication is immediate, and so we have established the equivalence
of conditions (1), (3), (4) and (6).
Since conditions (3) and (6) are left-right symmetric, it follows that they are also
equivalent to conditions (2) and (5), by analogy with the equivalence with (1) and (4).
We will call a ring B satisfying any of the equivalent conditions of the last proposition
a local ring. Any commutative ring that is local in the usual sense (i.e. it has a unique
maximal ideal) is evidently local in this non-commutative sense. As for non-commutative
examples of local rings, we see from Proposition 6.8 part (3) that if G is a p-group and k
is a field of characteristic p then the group algebra kG is a local ring. This is because its
radical is the augmentation ideal and the quotient by the radical is k, which is a division
ring, thus verifying condition (3) of Proposition 11.3.
We have seen in Corollary 11.2 a characterization of indecomposable modules as modules whose endomorphism ring only has idempotents 0 and 1. We now make the connection
with local rings.
Peter Webb
187
(11.4) PROPOSITION.
(1) In a local ring the only idempotents are 0 and 1.
(2) Suppose that B is an R-algebra that is finitely generated as an R-module, where R is
a complete discrete valuation ring or a field. If the only idempotents in B are 0 and
1 then B is a local ring.
Proof. (1) In a local ring B, any idempotent e other than 0 and 1 would give a nontrivial direct sum decomposition of B = Be B(1 e) as left B-modules, and so B would
have more than one maximal left ideal, a contradiction.
(2) Suppose that 0 and 1 are the only idempotents in B, and let () be the maximal
ideal of R. Just as in the proof of part (1) of Proposition 9.13 we see that annihilates
every simple B-module, and so B Rad(B). This implies that B/ Rad(B) is a finitedimensional R/()-algebra. If e B/ Rad(B) is idempotent then by the argument of
Proposition 9.15 it lifts to an idempotent of B, which must be 0 or 1. Since e is the image
of this lifting, it must also be 0 or 1. Now B/ Rad(B)
= Mn1 (1 ) Mnt (t ) for
certain division rings i , since this is a semisimple algebra, and the only way this algebra
would have just one non-zero idempotent is if t = 1 and n1 = 1. This shows that condition
(3) of the last proposition is satisfied.
We put these pieces together:
(11.5) COROLLARY. Let U be a module for a ring A.
(1) If EndA (U ) is a local ring then U is indecomposable.
(2) Suppose that R is a complete discrete valuation ring or a field, A is an R-algebra,
and U is finitely-generated as an R-module. Then U is indecomposable if and only if
EndA (U ) is a local ring. In particular this holds if A = RG where G is a finite group.
Proof. (1) This follows from 11.2 and 11.4.
(2) From 11.2 and 11.4 again all we need to do is to show that EndA (U ) is finitelygenerated as an R-module. Let Rm U be a surjection of R-modules. Composition with
this surjection gives a homomorphism EndA (U ) HomR (Rm , U ), and it is an injection
since Rm U is surjective (using the property of Hom from homological algebra that
it is left exact and the fact that A-module homomorphisms are a subset of R-module
homomorphisms). Thus EndA (U ) is realized as an R-submodule of HomR (Rm , U )
= U m,
which is a finitely generated R-module. Since R is Noetherian, the submodule is also
finitely-generated.
188
s
X
j=1
j j ) =
s
X
j j
j=1
and since EndA (U1 ) is a local ring it follows that at least one term j j must be
invertible. By renumbering the Vj if necessary we may suppose that j = 1, and we
write = 1 1 . Now (1 1 )(1 ) = 1U1 and so 1 : U1 V1 is split mono and
1 1 : V1 U1 is split epi. It follows that 1 (U1 ) is a direct summand of V1 . Since V1
is indecomposable we have 1 (U1 ) = V1 and 1 : U1 V1 must be an isomorphism.
that r = s and the summands are isomorphic in pairs, which completes the proof.
Peter Webb
189
Note that the proof of 11.6 shows that an exchange lemma property holds for the
indecomposable summands in the situation of the theorem. After the abstraction of general
rings, we state the Krull-Schmidt theorem in the context of finite group representations,
to make things clear.
(11.7) COROLLARY. Let R be a complete discrete valuation ring or a field and G
a finite group. Suppose that U is a finitely-generated RG-module that has two decompositions
U = U1 Ur = V1 Vs
where the Ui and Vj are indecomposable RG-modules. Then r = s and the summands Ui
and Vj are isomorphic in pairs when taken in a suitable order.
Proof. We have seen in Corollary 11.5 that the rings EndRG (Ui ) are local, so that
Theorem 11.6 applies.
190
Peter Webb
191
Relative projectivity
The relationship between the representations of a group and those of its subgroups
are one of the most important tools in representation theory. In the context of modular
representations of groups this relationship shows itself in a refinement of the notion of
projectivity, namely relative projectivity. We will use it to give a criterion for actual
projectivity and to determine the group rings of finite representation type. Finally we will
describe the theory of vertices and sources of indecomposable modules, which appear as
part of Green correspondence.
Let H be a subgroup of G and R a commutative ring with 1. An RG-module is said to
be H-free if it has the form V G
H for some RH-module V . It is H-projective, or projective
relative to H, if it is a direct summand of a module of the form V G
H for some RH-module
V.
For example, the regular representation RG
= R G
1 is 1-free, and projective modules
are 1-projective. If R is a field then every 1-projective module is projective, but if R is not
a field and V is an R-module that is not projective as an R-module, then V G
1 is 1-free
but not projective as an RG-module. Every RG-module is G-projective.
In order to investigate relative projectivity we first deal with some technicalities. We
P
have seen the pervasive importance of the group ring element gG g at every stage of the
development of representation theory. As an operator on any representation of G it has
image contained in the G-fixed points. We now consider something more general, and for
H
a subgroup H of G and an RG-module U we define the relative trace map trG
U G.
H :U
To define this we choose a set of representatives g1 , . . . , gn of the left cosets of H in G,
Pn
so G = g1 H gn H. If u U H we define trG
H (u) =
i=1 gi u. To complete the
picture we mention that when H G there is an inclusion of fixed points that we denote
g
G
resG
U H . When H G and g G there is also a map cg : U H U ( H) given by
H :U
u 7 gu. These operations behave like induction, restriction and conjugation of modules.
(11.10) LEMMA. Let U be an RG-module and let K H G and L be subgroups
of G.
H
(1) The homomorphism trG
U G is well-defined.
H :U
H
G
(2) trG
H trK = trK ,
G
G
resH
K resH = resK ,
g
H
cg trH
K = trg K cg ,
g
H
cg resH
K = resg K cg and
H
trH
H = 1 = resH .
H
(3) trH
: K|.
K resK is multiplication by |H
P
g
G G
H
(4) (Mackey formula) resL trH = g[L\G/H] trL
Lg H resLg H cg .
192
Pt
elements h1 , . . . , ht in H. Now the definition of trG
would be trG
(u) = i=1 gi hi u, but
H
H
Pt
this equals i=1 gi u as before, since u is fixed by all of the hi .
(2) Many of these formulas are quite straightforward and we only prove the first and
the third. For the first, if we take left coset representatives h1 , . . . , hm of K in H and left
coset representatives g1 , . . . , gn of H in G then the set of all elements gi hj is a set of left
coset representatives for K in G and their sum is (g1 + gn )(h1 + + hm ). Now
H
G H
trG
K (u) = (g1 + gn )(h1 + + hm )u = (g1 + gn )trK (u) = trH trK (u).
For the third formula, the elements g h1 , . . . , g hm are a set of left coset representatives
Pm g
Pm
g
H
for K in g H and so trg H
K cg (u) =
i=1 hi gu =
i=1 ghi u = cg trK (u).
(3) Continuing with the notation in use, if u is fixed by H then
g
H
trH
K resK (u)
m
X
hi u =
i=1
m
X
u = mu
i=1
where m = |H : K|.
(4) Considering the action from the left of L on the cosets G/H, for any g G the
L-orbit containing the coset gH consists of the cosets agH where a lies in L and ranges
through representatives of the cosets L/(L g H). This is because the cosets in the L-orbit
g
are the agH with a L, and a1 gH = a2 gH if and only if a1
2 a1 StabL (gH) = L H,
which happens if and only if a1 and a2 lie in the same coset of L g H in L. Thus on
partitioning the cosets G/H into L-orbits we see that g[L\G/H] {ag a [L/(L g H)]}
is a set of left coset representatives for H in G. Hence if u U H ,
X
X
X
trG
agu =
trL
Hu =
Lg H cg u.
gL\G/H] a[L/(Lg H)]
g[L\G/H]
The most important situation where we will use the relative trace map is when the
RG-module U is a space of homomorphisms HomR (X, Y ) between RG-modules X and
Y . In this situation HomR (X, Y )H = HomRH (X, Y ) for any subgroup H, so that the
relative trace map from H to G is a homomorphism of R-modules trG
H : HomRH (X, Y )
HomRG (X, Y ).
In the following result it would have been technically correct to insert resG
H in several places, but since this operation is simply the inclusion of fixed points it seems more
transparent to leave it out.
(11.11) LEMMA. Suppose that : U V and : W X are homomorphisms
G
of RG-modules and that : V G
H W H is an RH-module homomorphism. Then
G
G
G
G
(trH ) = trH ( ) and (trH ) = trH ( ).
Proof. Let g1 , . . . , gn be a set of left coset representatives for H in G and u U .
Pn
Pn
1
1
G
Then (trG
=
u) =
Similarly
H )(u)
i=1 gi (gi P
i=1 gi (gi u) = trH ()(u).
P
n
n
1
1
G
G
(trH )(u) = i=1 gi (gi u) = i=1 gi (gi u) = trH ()(u).
Peter Webb
193
trG
H
It will help us to consider the adjoint properties of induction and restriction of modules
in detail. We have seen in Corollary 4.13 that when H G, U is an RH-module and V is
G
HomRG (U1 G
H , V )
x
HomRG (U2 G
H , V )
HomRH (U1 , V G
H)
x
HomRH (U2 , V G
H)
G
HomRG (U G
H , V1 ) HomRH (U, V1 H )
y
y
G
HomRG (U G
H , V2 ) HomRH (U, V2 H )
194
In the language of category theory, and are the unit and counit of the adjuntion that
G
shows that G
H is left adjoint to H , and , are the unit and counit of the adjuntion that
G
shows that G
H is right adjoint to H .
To define these homomorphisms, choose a set of left coset representatives {g1 , . . . , gn }
Ln
G
of H in G with g1 = 1. For each RH-module U let : U U G
H H =
i=1 gi U be the
G G
inclusion into the summand 1 U , so (u) = 1 u, and let : U H H U be projection
onto this summand. We see that is a monomorphism, is an epimorphim and their
composite is the identity. If V is an RG-module we define the RG-module homomorphisms
Pn
P
: V V G
G
and : V G
G
V by (v) = i=1 gi gi1 v and ( x x u) =
H
H
H
H
P
G G
G G
x xu. In fact, regarding trG
H : HomRH (V, V H H ) HomRG (V, V H H ) we have
G
= trG
H where has domain V regarded as an RH-module. Similarly = trH , and this
shows that and are defined independently of the choice of coset representatives. We
see that is a monomorphism, is an epimorphism and their composite is multiplication
by |G : H|.
We now construct the natural isomorphism
G
HomRG (U G
H , V ) = HomRH (U, V H ).
H
G
G
U U G
H H V H
H
G G
U G
H V H H V.
We may check that these two constructions are mutually inverse, and are natural in U and
V.
We next construct a natural isomorphism
G
HomRG (V, U G
H ) = HomRH (V H , U )
G G
V G
H U H H U
H
G
G
V V G
H H U H .
Again we check that these operations are natural and mutually inverse.
Peter Webb
195
x u 7 xu
is split.
G
(5) U is a direct summand of U G
H H .
(6) (Higmans criterion) 1U lies in the image of trG
H : EndRH (U ) EndRG (U ).
Proof. (1) (2) We first prove this implication in the special case when U is an
induced module T G
H . Suppose we have a diagram of RG-modules
T G
H
y
W
G
G
and a homomorphism of RH-modules : T G
H H V H so that = . Under the
G
G
adjoint correspondence corresponds to the composite T T G
H H W H and we
have a commutative triangle of RH-module homomorphisms
V G
H
y .
W G
H
196
()G
H
G
Now consider a module U that is a summand of T G
H , and let U T H U be
G
inclusion and projection. We suppose there is a homomorphism : U G
H V H so that
G
= . The homomorphism : T H W has the property that = () and so by
what we proved there is a homomorphism of RG-modules : T G
H V so that = .
Now = = so that : U V is an RG-module homomorphism that makes the
triangle commute.
(2) (3) This follows immediately on applying (2) to the diagram
1
y .U
U
G
(3) (4) We know that : U G
H H U is split as an RH-module homomorphism
G
by : U U G
H H . Applying condition (3) it splits as an RG-module homomorphism.
(4) (5) and (5) (1) are immediate.
G
(5) (6) We may prove that 1UG
G = trH () by direct computation. Writing
H H
G
U G
H H = V1 V2 where V1 = U , we can represent as a matrix
f11 f21
=
f12 f22
trG
H f11
trG
H f12
trG
H f21
trG
H f22
1 0
0 1
G
and from this we see that for every summand of U G
H H (and in particular for U ) the
G
identity map on that summand is in the image of trH .
G
G
(6) (5) Write 1U = trG
H for some morphism : U H U H . Now corresponds
by the adjoint correspondence to the composite homomorphism
H
G G
G
U U G
H H U H H .
We claim that G
H splits : for
G
G
G
G
G
G
H = trH () H = trH ( H ) = trH () = 1U .
Peter Webb
197
1
trG 1 ,
|G:H| H U
RH n G
H = RG . Therefore U is projective.
EXAMPLES. 1. This criterion for projectivity would have simplified matters when we
were considering the projective modules for groups of the form G = H K in Chapter 8. In
this situation we saw that RH becomes an RG-module where H acts by left multiplication
and K acts by conjugation. If K has order prime to p and R is a field of characteristic p (or
a discrete valuation ring with residue field of characteristic p) it follows from the corollary
that RH is projective as an RG-module, because on restriction to H it is projective and
the index of H in G is invertible. On the other hand, we may also regard RK as an
RG-module via the homomorphism G K, and if now H has order prime to p then RK
is a projective RG-module, because it is projective on restriction to RK and the index of
K in G is prime to p.
2. In the exercises to Chapter 6 the simple Fp SL(2, p)-modules were considered.
The goal of Exercise 21 of Chapter 6 was to show that the symmetric powers S r (U2 )
are all simple Fp SL(2, p)-modules when 0 r p 1, where U2 is the 2-dimensional
space on which SL(2, p) acts as invertible transformations of determinant 1. The order
of SL(2, p) is p(p2 1) and so a Sylow p-subgroup of this group is cyclic of order p. In
Exercise 20 of Chapter 6 one shows that on restriction to a certain Sylow p-subgroup,
S r (U2 ) is indecomposable of dimension r + 1 when 0 r p 1. From the classification of
indecomposable modules for a cyclic group of order p we deduce that S p1 (U2 ) is projective
198
as a module for the Sylow p-subgroup. It follows from the last corollary that S p1 (U2 ) is
projective as an Fp SL(2, p)-module. This module is thus a simple projective Fp SL(2, p)module, or in other words a block of defect zero, which therefore lifts to characteristic
zero. Start from information in characteristic p we have thus deduced the existence of an
ordinary simple character of SL(2, p) of degree p.
Peter Webb
199
indecomposable. Now if RG has finite representation type there are only finitely many
isomorphism types of summands of modules V G
P , by the Krull-Schmidt theorem, and
hence RP has finite representation type.
We have already seen in Theorem 6.2 that cyclic p-groups have finite representation
type over a field of characteristic p, so by Proposition 11.16 groups with cyclic Sylow psubgroups have finite representation type. We will show the converse in Theorem 11.18,
and as preparation for this we now show that k[Cp Cp ] has infinitely many non-isomorphic
indecomposable modules, where k is a field of characteristic p. Apart from the use of this
in establishing infinite representation type, it is useful to see how indecomposable modules
may be constructed, and in the case of C2 C2 it will lead to a classification of the
indecomposable modules.
We will first describe infinitely many modules of different dimensions for k[Cp Cp ],
and after that we will prove that they are indecomposable. Let G = Cp Cp = haihbi. For
each n 1 we define a module M2n+1 of dimension 2n + 1 with basis u1 , . . . , un , v0 , . . . , vn
and an action of G given as follows:
a(ui ) = ui + vi1 , b(ui ) = ui + vi
a(vi ) = vi ,
b(vi ) = vi
where 1 i n
where 0 i n.
where 1 i n
where 0 i n
a1
v0
u1
b1
a1
v1
u2
b1
a1
un
b1
vn
200
(11.17) PROPOSITION. The quotient EndkG (M2n+1 )/ Rad EndkG (M2n+1 ) has dimension 1. Thus EndkG (M2n+1 ) is a local ring and M2n+1 is indecomposable.
Proof. We will show that EndkG (M2n+1 )/I has dimension 1 for a certain nilpotent
ideal I. Such an ideal I must be contained in the radical, being nilpotent, and the fact
that it has codimension 1 will then force it to equal the radical. This will prove the result.
Observe that
Soc(M2n+1 ) = Rad(M2n+1 ) = kv0 + kvn .
The ideal I in question is HomkG (M2n+1 , Soc(M2n+1 )), and this squares to zero since if
: M2n+1 Soc(M2n+1 ) then Rad(M2n+1 ) Ker and so Soc(M2n+1 ) = 0.
We now show that I has codimension 1. It is easy to get an intuitive idea of why this
is so, but not so easy to write it down in technically correct language. The intuitive idea is
that the basis elements of M2n+1 lie in a string as shown diagrammatically, each element
related to those on either side by the action of a 1 and b 1. To within elements of A,
any endomorphism must send this string to another string of elements that are similarly
related. If the endomorphism shifts the string either to the left or the right, then part
of the string must fall off the end of the module, or in other words be sent to zero. The
connection between adjacent basis elements then forces the whole shift to be zero, so that
no shift to the left or right is possible. From this we deduce that, modulo I, endomorphisms
are scalar multiples of the identity.
We now attempt to write down this intuitive idea in formal terms. If is any endomorphism of M2n+1 then (Soc(M2n+1 )) Soc M2n+1 so induces an endomorphism
of M2n+1 / Soc(M2n+1 ). We show that is necessarily a scalar multiple of the identity. To
establish this we will exploit the equations
(a 1)ui = (b 1)ui1
when 2 i n
and also the fact that (a 1) and (b 1) both map ku1 + + kun injectively into
Soc(M2n+1 ). Applying to the last equations we have for 2 i n,
((a 1)ui ) = (a 1)(ui ) = ((b 1)ui1 ) = (b 1)(ui1 ).
It follows from this that is completely determined once we know (u1 ), since then
2 ) by injectivity of a 1 on the span of the ui ,
(a 1)(u2 ) = (b 1)(u1 ) determines (u
3 ) similarly, and so on.
(a 1)(u3 ) = (b 1)(u2 ) determines (u
Suppose that
(u1 ) 1 u1 + + r ur (mod Soc(M2n+1 ))
where the i are scalars and r 6= 0. Multiplying both sides by b1 and using the equations
(a 1)(ui ) = (b 1)(ui1 ) as before, as well as injectivity of multiplication by a 1 on
the span of the ui , we see that
(u2 ) 1 u2 + + r ur+1 (mod Soc(M2n+1 ))
Peter Webb
201
and inductively
(unr+1 ) 1 unr+1 + + r un (mod Soc(M2n+1 )).
If it were the case that r > 1 then the equation
(b 1)(unr+1 ) = (a 1)(unr+2 ) = 1 vnr+1 + + r vn
would have no solution, since no such vector where the coefficient of vn is non-zero lies in
the image of a 1.
We conclude that r = 1 and (ui ) 1 ui (mod Soc(M2n+1 )) for some scalar 1 , for
all i with 1 i n. Thus
1 1M2n+1 : M2n+1 Soc(M2n+1 )
and so EndkG (M2n+1 )/ HomkG (M2n+1 , Soc(M2n+1 )) has dimension 1.
We can now establish the following:
(11.18) THEOREM (D.G. Higman). Let k be a field of characteristic p. Then kG
has finite representation type if and only if Sylow p-subgroups of G are cyclic.
Proof. By Proposition 11.16 it suffices to show that if P is a p-group then kP has
finite representation type if and only if P is cyclic. We have seen in Theorem 6.2 that kP
has finite representation type when P is cyclic. If P is not cyclic then P has the group
Cp Cp as a homomorphic image. (This may be proved using the fact that if (P ) is
the Frattini subgroup of P then P/(P )
= (Cp )d for some d and that P can be generated
by d elements. Since P cannot be generated by a single element, d 2 and so (Cp )2 is
an image of P .) The infinitely-many non-isomorphic indecomposable k[Cp Cp ]-modules
become non-isomorphic indecomposable kP -modules via the quotient homomorphism, and
this establishes the result.
Even when the representation type is infinite, the arguments that we have been using
still yield the following result.
(11.19) THEOREM. Let k be a field of characteristic p. For any finite group G, the
number of isomorphism classes of indecomposable modules that are projective relative to
a cyclic subgroup is finite.
We now turn to group rings of infinite representation type, namely the group rings
in characteristic p for which the Sylow p-subgroups of the group are not cyclic. We might
expect that, even though the technical difficulties may be severe, a classification of indecomposable modules is about to be revealed. Perhaps the surprising thing about infinite
202
representation type is that in some sense, most of the time, a classification of indecomposable modules is not merely something that is beyond our technical capabilities, it is rather
something that may never be possible in any meaningful sense, because of inherent aspects
of the problem.
Discussion of such matters raises the question of what we mean by a classification.
There are many instances of classification in mathematics, but whichever one we are considering, we might reasonably understand that it is a description of some objects that is
simpler than the objects themselves and that allows us to identify in some reasonable way
some significant aspects. Thus it would be inadequate to say that we had a parametrized
indecomposable modules by using the same set of indecomposable modules to achieve the
classification, because this would provide no simplification. It would also be inadequate
to put the set of isomorphism classes of indecomposable modules in bijection with some
abstract set of the same cardinality. Although we could certainly do this, and for some
people it might be a classification, it would provide no new insight. Those who maintain
that it is always possible to classify perhaps have such possibilities in mind. In reality it
might be very difficult or impossible to classify in any meaningful sense because there are
simply too many objects to classify: if we were to put the objects in a list in a book, the list
would simply be too long and structureless to have any meaning. Of course, just because
we cannot see how to make sense of a set of objects does not mean it cannot be done.
However, it remains the case that for most group algebras in positive characteristic no
one has been able to provide any reasonable classification of the indecomposable modules.
Furthermore, we will present reasons why it would be hard to do so.
Infinite representation type divides up into two possibilities: tame and wild. For
the tame group algebras we can (in principle) classify the indecomposable modules. For
the wild algebras no one can see how to do it. Before we address these general questions we will describe the indecomposable representations of C2 C2 in over a field k
of characteristic 2, since this exemplifies tame representation type. In constructing infinitely many indecomposable modules for Cp Cp we already constructed some of the
indecomposable k[C2 C2 ]-modules, but now we complete the picture. As before we let
Peter Webb
kG =
W2n+1 =
W1 = M1 = ()
M2n+1 =
Ef,n =
E0,n =
E,n
203
and for n 1
In these diagrams each node represents a basis element of a vector space, a southwest
arrow emanating from a node indicates that a 1 sends that basis element to the basis
element at its tip, and similarly a southeast arrow indicates the action of b 1 on a
basis element. Where no arrow in some direction emanates from a node, the corresponding
element a 1 or b 1 acts as zero.
The even-dimensional indecomposable representations Ef,n require some further explanation. They are parametrized by pairs (f, n) where f k[X] is an irreducible monic
polynomial and n 1 is an integer. Let the top row of nodes in the diagram correspond
to basis elements u1 , . . . , un , and the bottom row to basis elements v1 , . . . , vn . Let
(f (X))n = X mn + amn1 X mn1 + + a0 .
The right-most arrow starting at umn that has no terminal node is supposed to indicate
that (b 1)umn = vmn+1 where
vmn+1 = amn1 vmn1 a1 v2 a0 v1 ,
so that with respect to the given bases b 1 has matrix
0
1
0
..
.. ..
...
.
.
.
.
0
0
1
a0
amn1
204
Peter Webb
205
an integer n 1 we may construct the k[C2 C2 ]-module M k[X] k[X]/(f n), which is a
module isomorphic to (k[X]/(f n))2 as a k[X]-module, and which is acted on by k[C2 C2 ]
as a module isomorphic to Ef, n. This construction accounts for all but finitely many of
the indecomposable k[C2 C2 ]-modules in each dimension.
With our understanding improved by the last example, we now divide infinite representation type into two kinds: tame and wild. Let A be a finite-dimensional algebra over
an infinite field k. We say A has tame representation type if it has infinite type and for each
dimension d there are finitely many (A, k[X])-bimodules Mi that are free as k[X]-modules
so that all but finitely many of the indecomposable A-modules of dimension d have the
form Mi k[X] k[X]/(f n) for some irreducible polynomial f and integer n. If the bimodules
Mi can be chosen independently of d (as happens with representations of C2 C2 ) we say
that A has domestic representation type, and otherwise it is non-domestic.
We say that the finite-dimensional algebra A has wild representation type if there
is a finitely-generated (A, khX, Y i)-bimodule M that is free as a right khX, Y i)-module,
such that the functor M khX,Y i from finite-dimensional khX, Y i-modules to finitedimensional A-modules preserves indecomposability and isomorphism type. Here khX, Y i
is the free algebra on two non-commuting variables, having as basis the non-commutative
monomials X a1 Y b1 X a2 Y b2 where ai , bi 0.
In view of the following theorem it would have been possible, over an algebraically
closed field, to define wild to be everything that is not finite or tame. We state the next
three results without proof, since they take us outside the scope of this book.
(11.21) THEOREM (Drozd [Dro]; Crawley-Boevey [CB]). Let A be a finite-dimensional
algebra over an algebraically-closed field. Then A has either finite, tame or wild representation type.
When A has wild representation type, the idea is that phenomena that occur with
representations of khX, Y i also appear with A-modules. Thus A has at least as many isomorphism types indecomposable modules as khX, Y i does. The indecomposable khX, Y imodules are quite diverse, and in some sense it is a hopeless task to try to classify them.
It is known that the theory of finite-dimensional khX, Y i-modules is undecidable, meaning that there exists a sentence in the language of finite-dimensional khX, Y i-modules
that cannot be decided by any Turing machine (an account of this result can be found in
[Prest]). It is also the case in a heuristic sense that the khX, Y i-modules are as badly behaved as those of any algebra: it is possible to embed the category of A-modules and their
homomorphisms, for any algebra A, into the category of khX, Y i-modules (see [Brenner,
Gabriel]).
In the context of group algebras the division into finite, tame and wild representation
type is given by the following theorem, in which we include the result of D.G. Higman
already proven.
206
Peter Webb
207
G
( g(U G
K g H )) K gH
g[K\G/H]
using transitivity of restriction and induction, and now U must be a summand of some
module induced from one of the groups K gH. If both H and K happen to be minimal
subject to the condition that U is projective relative to these groups, we deduce that
trG
K gH (cg ).
g[K\G/H]
Since U is indecomposable its endomorphism ring is local and so some term trG
K gH (cg )
must lie outside the unique maximal ideal of EndRG (U ) and must be an automorphism.
This implies that trG
K gH : EndR[K gH] (U ) EndRG (U ) is surjective, since the image of
G
trK gH is an ideal, and so U is K gH-projective.
208
We now deduce as in the first proof that if K and H are minimal subgroups relative
to which U is projective, then H and K are conjugate.
(2) Let Q be a minimal subgroup relative to which U is projective. We know that U
G
G
is a summand of U G
Q Q and hence it is a summand of T Q for some indecomposable
summand T of U G
Q . Suppose that T is another indecomposable module for which U is a
L
Q
G G
g
g
summand of T G
Q . Now T is a summand of T Q Q =
g[Q\G/Q] ( (T Q Q )) Q gQ
g 1
deduce from the fact that T is a summand of U G
T must be a summand
Q that T =
g 1
G
G
g 1
A minimal subgroup Q of G relative to which the indecomposable module U is projective is called a vertex of U , and it is defined up to conjugacy in G. We write vtx(U ) to
denote a subgroup Q that is a vertex of U . An RQ-module T for which U is a summand
of T G
Q is called a source of U , and given the vertex Q it is defined up to conjugacy by
elements of NG (Q).
We record some immediate properties of the vertex of a module.
(11.25) PROPOSITION. Let R be a field of characteristic p or a complete discrete
valuation ring with residue field of characteristic p.
(1) The vertex of every indecomposable RG-module is a p-group.
(2) An indecomposable RG-module is projective if and only if it is free as an R-module
and its vertex is 1.
(3) A vertex of the trivial RG-module R is a Sylow p-subgroup of G.
Proof. (1) We know from Proposition 11.14 that every module is projective relative
to a Sylow p-subgroup, and so vertices must be p-groups.
(2) If an indecomposable module is projective it is a summand of RG, which is induced
from 1, so it must be free as an R-module and have vertex 1. Conversely, if U has vertex 1
G
G
it is a summand of U G
1 1 , so if U is free as an R-module it is a summand of R 1 = RG
and hence is projective.
(3) Let Q be a vertex of R and P a Sylow p-subgroup of G containing Q. Then R is
L
G
G G
P
a summand of R G
Q , so R P is a summand of R Q P =
g[P \G/Q] R P gQ and hence
is a summand of R P
P gQ for some g G. We claim that for every subgroup H P ,
P
R H is an indecomposable RP -module. From this it will follow that R = R P
P gQ and
that Q = P . The only simple RP -module is the residue field k with the trivial action
Peter Webb
209
210
V G
L L = V V ,
G
Z G
L L = Z Z
= V V Z Z
M
L
=
( x(T Q
Lx Q )) L xQ .
x[L\G/Q]
There is one summand in the last direct sum with x L and it is isomorphic to T L
Q=
x
V Z. The remaining summands are all induced from subgroups L Q with x 6 L, and
it follows that all indecomposable summands of V and Z are projective relative to these
subgroups. This in particular implies the assertion we have to prove in this step, since
such subgroups cannot be conjugate to Q by elements of L.
Step 2. We show that in any decomposition as a direct sum of indecomposable modules, V G
L has a unique indecomposable summand with vertex Q and that the remaining
summands are projective relative to subgroups of the form Q xQ where x 6 L. To show
this, write V G
L as a direct sum of indecomposable modules and pick an indecomposable
summand U for which U G
L has V as a summand. This summand U must have vertex Q;
for it is projective relative to Q, since V is, and if U were projective relative to a smaller
group then V would be also, contradicting the fact that Q is a vertex of V . This shows
that the direct sum decomposition of V G
L has at least one summand with vertex Q.
G
Let U be another summand of V L . Then U G
L must be a summand of V , in the
notation of Step 1, and every indecomposable summand of U G
L is projective relative to a
y
G
subgroup L Q with y 6 L. Since U is a summand of T Q it is projective relative to Q,
and hence has a vertex Q that is a subgroup of Q. Since L Q it follows that U G
L has
=
(T
.
This
is
because
U
is
a
summand
of
T
Peter Webb
211
G
summand of V G
L L , and by Step 1 this has just one direct summand with vertex Q,
namely V . In fact U G
L must have an indecomposable summand that on further restriction
to Q has T as a summand, and this summand has vertex Q. It follows that this summand
must be isomorphic to V , and in any expression for U G
L as a direct sum of indecomposable
modules, one summand is isomorphic to V and the rest are projective relative to subgroups
of the form L xQ with x 6 L. This completes the proof of assertion (1) of the theorem.
Step 4. The final assertion of the theorem follows from the first two and the fact that
G
G G
U is isomorphic to a summand of U G
L L and V is isomorphic to a summand of V L L .
posable kG-modules are projective, and there are lk (G) of them, giving (p1)lk (NG (Q)) +
lk (G) indecomposable modules in total.
212
Proof. (1) If 1 6 Ij for all j then every Ij is contained in the unique maximal ideal of
P
B and so 1 6 jJ Ij since 1 does not lie in the maximal ideal.
(2) The only idempotents in eBe are 0 and e since e is primitive. By Proposition 11.4
since eBe is finitely generated as an R-module, it is a local ring. It contains the family of
ideals {eIj e j J} and e lies in their sum. Therefore by part (1), e lies in one of the
ideals.
Peter Webb
213
trH
Hg Q Eg Q (V ),
g6H
Q
trH
Hg Q resHg Q cg ()
g[H\G/Q]
and all of the terms in the sum lie in J except for the one represented by 1 which gives
trH
Q ().
Let e EH (V ) be the idempotent that is projection onto M . We claim that e
H
trQ EQ (V ) and also that e 6 J. These come about because Q is a vertex of M and e is
the identity in the ring eEH (V )e, which may be identified with the local ring EH (M ).
H
Thus e trH
Q EQ (M ) trQ EQ (V ) since M is projective relative to Q. If e were to lie in
J, which is a sum of ideals trH
Hg Q Eg Q (V ) with g 6 H, it would lie in one of them by
Rosenbergs Lemma, and hence M would be projective relative to H g Q for some g 6 H.
The vertex Q of M must be conjugate in H to a subgroup of H g Q, so that g Q = h Q for
some h H. This implies that g 1 h NG (Q) H, so g H, a contradiction. Therefore
e 6 J.
resG
H
Let be the composite ring homomorphism EG (V )E
H (V ) EH (V )/J where the
second morphism is the quotient map. Since it is the image of a local ring, (EG (V )) is
also a local ring. It contains the ideal (trG
Q EQ (V )). We claim that this ideal contains the
214
(mod J)
by an earlier calculation.
Any ideal of a local ring containing a non-zero idempotent must be the whole ring, so
G
that (trG
Q EQ (V )) = (EG (V )). Again since EG (V ) is local, trQ EQ (V ) = EG (V ).
We have just shown that V is projective relative to Q, so Q contains a vertex of V .
The vertex cannot be smaller than Q, because otherwise by a calculation using the Mackey
formula M would have vertex smaller than Q. We conclude that Q is a vertex of V .
Peter Webb
215
(5) Let M be an RG-lattice with no non-zero projective summands. Then M is indecomposable if and only if M is indecomposable, if and only if 1 M is indecomposable.
Proof. (1) The short exact sequence splits as a sequence of R modules, because W is
projective as an R-module, and so the dual sequence 0 W V U 0 is exact.
Suppose that V W is a projective cover and W has no non-zero projective summand.
Then V is projective, and if V U is not a projective cover then by Proposition 7.8
we may write V = X Y where X U is a projective cover and Y maps to zero. This
would mean that V
= (V ) Y and where Y is projective, and on dualizing we deduce
that V has a non-zero projective summand, which is not the case. Hence V U is a
projective cover, so that U V is a (relative) injective hull. The second statement follows
from the first on applying it to the dual sequence 0 W V U 0.
(2) The dual of the sequence 0 (M ) PM M 0 that computes (M )
is 0 M (PM ) (M ) 0 and it computes 1 (M ). Thus 1 (M )
= (M )
and dualizing again gives the result.
(3) These isomorphisms follow immediately from (1) since the same sequence that
constructs M also constructs 1 (M ) and the sequence that constructs 1 M also
constructs 1 (M ).
(4) This comes from the fact that the projective cover of M1 M2 is the direct sum
of the projective covers of M1 and M2 , and similarly with (relative) injective hulls.
(5) If (M ) were to decompose then so would M
= 1 (M ) by (4), so the indecomposability of M implies the indecomposability of M . The reverse implication and
the equivalence with the indecomposability of 1 M follow similarly.
216
We see from this that permutes the isomorphism types of indecomposable RGlattices, with inverse permutation 1 . Let us write i for the ith power of when i > 0
and the ith power of 1 when i < 0. When i = 0 we put 0 M = M .
Many further methods in representation theory that go beyond the scope of this book
make implicit reference to , especially the ones that have to do with homological algebra,
and the connection is the following. A key notion in homological algebra is that of a
projective resolution of a module M . This is a sequence of projective modules
d
d1
3
2
1
0
P
2 P1 P0 00
that is exact everywhere except at P0 , where its homology (in this case the cokernel
of the map d1 ) is M . When R is a field or a complete discrete valuation ring we can
construct always a minimal projective resolution, which has the property that each map
Pi Ker di1 is a projective cover. We see by induction that for a minimal projective
resolution, i M
= Ker di1 when i 1.
We will use properties of in the exercises as part of a proof that the list of indecomposable modules for k[C2 C2 ] is complete, where k is a field of characteristic 2. Before
we leave we mention its properties under the Kronecker product.
(11.31) PROPOSITION. Let R be a field or a complete discrete valuation ring and
G a finite group. For any RG-lattice M we have M R i R
= i M Qi for each i, where
Qi is a projective RG-module.
Proof. Let P2 P1 P0 0 be a projective resolution of k. We claim that
the sequence M R P2 M R P1 M R P0 0 is exact except in position
0. This is because every map in the resolution must split as a map of R-modules since
M is projective as an R-module. The homology in position 0 is M R k
= M for the
same reason. Furthermore, all of the modules M R Pi are projective RG-modules by
Proposition 8.4. Thus we have a projective resolution of M , but there is no reason why
it should be minimal. Assuming inductively that the kernel of the map in position i 1
has the form i M Qi for some projective module Qi (and this is true when i = 0), we
deduce from Proposition 7.8 that the kernel in position i has the form i+1 M Qi+1 ,
which completes the proof by induction on i when i 0. When i < 0 we can deduce the
result by duality using part (2) of Proposition 11.30.
We see from this result that up to isomorphism the modules i R where i Z, together
with the projective modules, are closed under the operations of taking indecomposable
summands of tensor products.
Peter Webb
217
218
looking at representations. The reader can find more about this in the work of [Alperin],
[Benson-Carlson], [Benson-Conway].
Summary of Chapter 11
1. When R is a complete discrete valuation ring or a field the indecomposable summands in any direct sum decomposition of a finitely generated RG-module have isomorphism type and multiplicity determined independently of the particular decomposition.
2. When R is a field of characteristic p or a discrete valuation ring with residue field
of characteristic p, every RG-module is projective relative to a Sylow p-subgroup. An
RG-module is projective if and only if it is projective on restriction to a Sylow p-subgroup.
3. The representation type of kG is characterized in terms of the Sylow p-subgroups
of G.
4. Every indecomposable RG-module has a vertex and a source, which behave well
with respect to the Heller operator .
5. Green correspondence gives a bijection between isomorphism types of indecomposable RG-modules with vertex D and indecomposable R[NG (D)]-modules with vertex
D.
Peter Webb
219
220
G
indecomposable summand with vertex Q . Deduce that if U H is indecomposable then
subgroups of H that are conjugate in G to Q are all conjugate in H.
(b) Show that there is an indecomposable RH module V with vertex Q so that U is
a direct summand of V G
H.
11. Suppose that H is a subgroup of G and that V is an indecomposable RH-module
with vertex Q, where Q H. Show that V G
H has an indecomposable direct summand
with vertex Q. Show that for every p-subgroup Q of G there is an indecomposable RGmodule with vertex Q.
12. Given an indecomposable RG-module U , let X be the set of pairs (Q, T ) such
that Q is a vertex of U and T is a source of U with respect to Q. For each g G define
g
(A, T ) := (g Q, g T ). Show that this defines a permutation action of G on X, and that it
is transitive.
13. Let U be an indecomposable RG-module. Show that U and U have the same
vertex.
14. Let U be an indecomposable kG-module where k is a field. Assuming that U is
not projective, show that U/ Rad U
= Soc U .
15. Let U be an indecomposable RG-module with vertex Q, let H NG (Q) and
let f (U ) be the RH-module that is the Green correspondent of U . Show that f (U ) =
(f (U )).
16. The group G := P SL(2, 5) (which is isomorphic to A5 ) has three simple modules
over k = F5 , of dimensions 1, 3 and 5, and has Cartan matrix (with the simples taken in
the order just given)
2 1 0
1 3 0
0 0 1
Peter Webb
221
You may assume that a Sylow 5-subgroup is cyclic of order 5 and that its normalizer H is
dihedral of order 10.
(a) Show that has two orbits, each of length 4, on the non-projective kD10 -modules.
Deduce that for kG there are 8 indecomposable non-projective modules in two orbits of
length 4.
(b) Show that apart from a block of defect zero, each indecomposable projective
module for kG has Loewy length 3, and that its radical series equals its socle series.
(c) Show that each non-projective indecomposable module for kG has Loewy length
2.
(d) Identify the socle and the radical quotient of each of the 8 indecomposable nonprojective kG-modules.
17. The group G := P SL(2, 7) (which is isomorphic to GL(3, 2)) has four simple
modules over k = F7 , of dimensions 1, 3, 5 and 7, and has Cartan matrix (with the simples
taken in the order just given)
2 0 1 0
0 3 1 0
1 1 2 0
0 0 0 1
222
19. Let H be a subgroup of G and let U , V be RG-modules. We say that an RGmodule homomorphism f : U V factors through an H-projective module if there is an
G
(3) f factors as U U G
H H V for some RG-module homomorphism , where is the
map constructed after 11.12,
G
(4) f factors as U V G
H H V for some RG-module homomorphism , where is the
map constructed after 11.12,
(5) f lies in the image of trG
H : HomRH (U, V ) HomRG (U, V ).
(b) When R is a field or a complete discrete valuation ring, use question 9 from Chapter
8 to show that the space of homomorphisms U V that factor through a projective has
dimension equal to the multiplicity of PR as an RG-module summand of HomR (U, V ).
Peter Webb
223
12. Blocks
Block theory is one of the deepest parts of the representation theory of finite groups,
and in this chapter we can only scratch the surface of the sophisticated constructions
and techniques that have been developed. We present the basic parts of the theory, in
line with one of the goals of this book, which is to have relevance for the needs of the
non-specialist in dealing with representations that naturally arise. For such purposes it
is useful to have an idea of the defect group of a block and its relation to vertices of
indecomposable modules in the block. It is useful to see the different methods used in
block theory - sometimes ring-theoretic, sometimes module-theoretic, and to be able to
relate them. After a discussion of different definitions of a block involving idempotents,
ring direct summands and equivalence classes of representations, and having presented
some examples, we move on to define the defect group in different ways, showing that
these ways are equivalent. We are able to see that the blocks of defect zero introduced in
Chapter 9 do indeed deserve the name. We conclude with Brauers First Main Theorem,
which establishes a correspondence between blocks with a certain defect group D and
blocks of NG (D) with defect group D.
224
Peter Webb
225
(2) We have seen the lifting argument before. Since n1 Z(RG)/ n Z(RG) is a nilpotent ideal in Z(RG)/ n Z(RG) we may lift any idempotent en1 + n1 Z(RG) to an
idempotent en + n Z(RG), thereby obtaining from any idempotent e1 + Z(RG) Z(kG)
a Cauchy sequence e1 , e2 , . . . of elements of Z(RG) whose limit is the required lift. We have
also seen before that primitive idempotents correspond to primitive idempotents. In the
present situation the lift of each primitive idempotent is unique since the primitive central
idempotents of RG themselves are unique, so that there is only one primitive idempotent
of Z(RG) that reduces to each primitive idempotent of Z(kG).
Because of this, it is the same thing to study the blocks of RG and of kG since they
correspond to each other under reduction modulo . If U is a kG-module we may regard
it also as an RG-module via the surjection RG kG and if e is a block of RG with image
the block e Z(kG) there is no difference between saying that U belongs to e or that U
belongs e. We may also partition the simple F G-modules into blocks in a way consistent
with the blocks for RG and kG. Regarding RG as a subset of F G, a primitive central
idempotent e of RG is also a central idempotent of F G. We say that an F G module U
belongs to e if eU = U . Evidently each simple F G-module belongs to a unique block. We
see that if U is an RG-module and U0 U is any R-form of U (i.e. a full RG-lattice in
U ) then U0 belongs to e if and only if U belongs to e.
Examples. 1. When A is a finite-dimensional semisimple algebra over a field, the
blocks correspond to the matrix summands of A, each block being the idempotent that
is the identity element of a matrix summand. Each simple module lies in its own block,
and so there is only one indecomposable module in each block. We saw in Theorem 3.23 a
formula in terms of characters for the primitive central idempotent in CG corresponding
to each simple complex representation.
2. When G is a p-group and k is a field of characteristic p the regular representation
kG is indecomposable and the identity element is a block. There is only one block in this
situation and block theory does nothing for us if we are interested only in representations
of p-groups in characteristic p.
3. We have seen at the end of Chapter 9 that, when we have a block of defect zero
for G over a splitting p-modular system (F, R, k), there is a 2-sided direct summand of
RG that is isomorphic to a matrix algebra Mn (R), and also a matrix summand Mn (k)
of kG that is the reduction modulo of the summand of RG. There is a unique simple
kG-module in this block, and it is projective. It lifts to a unique RG-lattice, and all RGsublattices of it are isomorphic to it. A key fact that we used in identifying the features of
this situation is that the primitive central idempotent of CG corresponding to the matrix
summand in fact lies in RG. All this explains why we made reference to a block in that
situation, but not why we used the term defect zero. This too will soon be explained.
4. When G = S3 in characteristic 2 there are two blocks, since the simple module of
degree 2 is a block of defect zero, and the only other simple module is the trivial module,
which lies in the principal block. This has been seen in Example 7.5, the examples after
9.27 and Theorem 9.28. The projective cover of the trivial module as an RG-module has
226
character equal to the sum of the characters of the trivial representation and the sign
representation, and so the principal block idempotent in RG acts as the identity on this,
meaning that they are the ordinary characters in the principal block. The other ordinary
character, of degree 2, lies in the other block.
In characteristic 3, S3 has only one block since there are two simple modules (trivial
and sign) and the sign representation appears as a composition factor of the projective
cover of the trivial module. This means that it belongs to the principal block.
(12.4) LEMMA. Let e be a block of a ring A with identity. If 0 U V W 0
is a short exact sequence of A-modules then V belongs to e if and only if U and W belong
to e.
In other words, every submodule and factor module of a module that belongs to e
also belong to e, and an extension of two modules that belong to e also belongs to e.
Proof. A module belongs to e if and only if multiplication by e is an isomorphism of
that module, and this property holds for V if and only if it holds for U and W .
In what follows we characterize blocks in terms of a relation on the simple modules.
We could work with modules for an algebra over a complete discrete valuation ring R, but
since the simple modules are naturally defined over the residue field k we will assume that
A is a finite-dimensional k-algebra. The next result is an elaboration of the observation
from the last lemma that each indecomposable projective module PS lies in the same block
as the simple module S, and in fact all of the composition factors of PS lie in this block.
(12.5) PROPOSITION. Let A be a finite-dimensional algebra over a field k. The
following are equivalent for simple A-modules S and T .
(1) S and T lie in the same block.
(2) There is a list of simple A-modules S = S1 , S2 , . . . , Sn = T so that Si and Si+1
are both composition factors of the same indecomposable projective module, for each
i = 1, . . . , n 1.
(3) There is a list of simple A-modules S = S1 , S2 , . . . , Sn = T so that for each i =
1, . . . , n 1, Si and Si+1 appear in a non-split short exact sequence of A-modules
0 U V W 0 with {U, W } = {Si , Si+1 }.
Proof. The implications (3) (2) (1) are straightforward in that all the composition factors of any indecomposable module necessarily belong to the same block, and
whenever we have a non-split extension of the kind that appears in condition (3), the
middle module is uniserial and hence indecomposable.
To show that (1) (2) we will write S T to mean that the simple modules S and
T satisfy the condition of (2), which is an equivalence relation. All the composition factors
of any particular indecomposable projective module are equivalent in this sense. Suppose
that S and T lie in the same block. We need to show that S T . We may write the
Peter Webb
227
228
Example. Suppose that G = K H where K has order prime to p and H is a pgroup. In other words, G has a normal p-complement and is termed p-nilpotent. We saw in
Theorem 8.10 that this is precisely the situation in which each indecomposable projective
module has only one isomorphism type of composition factor. Another way of expressing
this is to say that the Cartan matrix is diagonal. In view of the last results we see that
we have characterized the groups for which each p-block contains just one simple module
over a field of characteristic p as being the p-nilpotent groups.
We can describe the primitive central idempotents explicitly in this situation.
(12.7) PROPOSITION. Let G = K H be a p-nilpotent group, where K has order
prime to p and H is a p-group. Let k be a field of characteristic p. Each block of kG lies
in kK, and is the sum of a G-conjugacy class of blocks of kK.
Proof. Observe that if 1 = e1 + + en is the sum of blocks of kK then for each
i and g G the conjugate gei g 1 is also a block of kK. For this we verify that this
element is idempotent, and also that it is central in kK, which is so since if x K then
e=gei g 1
e=gei g 1
the elements in the G-orbit of ei . We now show that f is primitive in Z(kG). Suppose
instead that f = f1 + f2 is a sum of orthogonal idempotents in Z(kG). Then there are
non-isomorphic simple kG-modules U1 and U2 with f1 U1 = U1 and f2 U2 = U2 . Since
f f1 = f1 and f f2 = f2 we have f U1 = U1 and f U2 = U2 .
a
a
By Cliffords theorem U1 G
K = S1 St for some integer a, where S1 , . . . , St are G-
conjugate simple kK-modules. Since f kK we have f Si = Si for all i, and this identifies
Since K consists of the p-regular elements of G it follows that the Brauer characters of
U1 and U2 are scalar multiples of one another. Since Brauer characters of non-isomorphic
simple modules are linearly independent we deduce that U1
= U2 , a contradiction. This
shows that f is primitive in Z(kG).
Peter Webb
229
We may see the phenomenon described in the last result in many examples, of which
the smallest non-trivial one is G = S3 = K H where K = h(1, 2, 3)i and H = h(1, 2)i,
taking p to be 2 and k = F4 . By Theorem 3.23 the blocks of kK are
e1 = () + (1, 2, 3) + (1, 3, 2),
e2 = () + (1, 2, 3) + 2 (1, 3, 2), and
e3 = () + 2 (1, 2, 3) + (1, 3, 2)
where is a primitive cube root of 1 in F4 . (In case the reader expects some factors 31 ,
note that 3 = 1 in characteristic 2. In the next expression 1 has been written as 1 and
2 as 0.) In the action of G on these idempotents there are two orbits, namely {e1 } and
{e2 , e3 }. The blocks of kG are thus e1 and e2 + e3 = (1, 2, 3) + (1, 3, 2). These idempotents
have already appeared in Example 7.5, where it was calculated that kGe1 = 11 = P1 and
kG(e2 + e3 ) = 2 2 is the direct sum of two simple projective kG-modules. This structure
has also been considered in Exercise 5 of Chapter 8. From Theorem 9.28 we know the
second summand is a block of defect zero, and kG(e2 + e3 )
= M2 (k) as rings. We also see
2
that kGe1 = kC2 = k[X]/(X ) as rings. An approach to this can be found in Exercise 11
of Chapter 8.
where g1 , g2 G, x kG.
230
(12.8) PROPOSITION.
(1) The submodules of RG, regarded as a module for R[G G], are precisely the 2-sided
ideals of RG.
(2) Any decomposition of RG as a direct sum of indecomposable R[G G]-modules is a
decomposition as a direct sum of blocks.
(3) Regarded as a representation of G G, RG is a transitive permutation module in
GG
which the stabilizer of 1 is (G). Thus RG
as R[G G]-modules.
= R (G)
Proof. (1) The 2-sided ideals of RG are precisely the R-submodules of RG that are
closed under multiplication from the left and from the right by G. These is equivalent to
being an R[G G]-submodule
(2) This follows from (1) and 12.1.
Proof. By 12.8 eRG is indecomposable as an R[G G]-module, and since it is (G)projective it has a vertex contained in (G). Such a subgroup is necessarily a p-group, so
has the form (D) for some p-subgroup D of G where (G) is determined up to conjugacy
in G G. If D1 is another subgroup of G for which (D1 ) is a vertex of eRG then
(D1 ) = (g1 ,g2 ) (D) for some g1 , g2 G, and so for all x D, (g1 x, g2 x) (D1 ). Thus
g1
x D1 for all x D and since D and D1 have the same order it follows that D1 = g1 D.
We define a subgroup D of G to be a defect group of the block e if (D) is a vertex of
eRG, and according to Corollary 12.9 it is defined up to conjugacy in G. If |D| = pd we
say that d is the defect of e, but often we abuse this terminology and say simply that the
defect of the block is D. According to these definitions, a block of defect 0 is one whose
defect group is the identity subgroup. It is not immediately apparent that this definition
of a block of defect zero coincides with the previous one, and this will have to be proved.
We now start to investigate the kinds of groups that can be defect groups.
Peter Webb
231
(12.10) THEOREM (Green). Let e be a p-block of G with defect group D, and let
P be a Sylow p-subgroup of G that contains D. Then D = P gP for some element
g CG (D).
Proof. Let P be a Sylow p-subgroup of G containing D. We consider
GG GG
GG
RG P
P = R (G) P P
M
=
(G)
P P
( y(R (P P )y (G) )) (P
P ) y(G)
y[P P \GG/(G)]
y[P P \GG/(G)]
P P
R (P
P ) y(G) .
P P
Now each R (P
P ) y(G) is indecomposable since P P is a p-group, as shown in the
proof of part (3) of Proposition 11.24. The summand eRG of RG on restriction to P P
has a summand that on further restriction to (D) has a source of eRG as a summand.
GG
This summand of eRG P
P also has vertex (D) (an argument that is familiar from the
P P
proof of the Green correspondence 11.26), and must have the form R (P
P ) y(G) for some
y
y GG. Thus (D) is conjugate in P P to (P P ) (G), so (D) = z((P P ) y(G))
for some z P P .
The elements 1 G = {(1, t) t G} form a set of coset representatives for (G) in
G G, and so we may assume y = (1, t) for some t G. Write z = (r, s), where r, s P .
Now
(P P ) (1,t)(G) = {(x, tx) x P and tx P }
= {(x, tx) x P P t }
(1,t)
(P P t )
and (D) = (r,s)(1,t)(P P t ). The projection onto the first coordinate here equals D =
1
1
r
(P P t ) = rP rt P = P rt P since r P . At this point though the proof is complete,
1
1 1
apart from the fact that rt1 might not centralize D. Now (1,t )(r ,s )(D) (G),
1 1
1
so that r x = t s x for all x D, and rt1 s1 CG (D). Since s P we have
1 1
D = P rt s P and this completes the proof.
(12.11) COROLLARY. If e is a p-block of G with defect group D then
D = Op (NG (D)) Op (G),
where Op (G) denotes the largest normal p-subgroup of G.
Proof. We start by observing that OP (G) is the intersection of the Sylow p-subgroups
of G; for the intersection of the Sylow p-subgroups is a normal p-subgroup since the Sylow
p-subgroups are closed under conjugation, and on the other hand Op (G) P for some
232
Sylow p-subgroup P , and hence g(Op (G)) = Op (G) gP for every element g G. Since
g
P accounts for all Sylow p-subgroups by Sylows theorem it follows that Op (G) is contained
in their intersection.
We see immediately that D Op (G) since D is the intersection of two Sylow psubgroups and hence contains the intersection of all Sylow p-subgroups.
To prove that D = Op (NG (D)), let P be a Sylow p-subgroup of G that contains a
Sylow p-subgroup of NG (D). Such a P necessarily has the property that P NG (D) is a
Sylow p-subgroup of NG (D). Now D = P gP for some g CG (D) and so in particular
g NG (D). Thus gP NG (D) = g(P NG (D)) is also a Sylow p-subgroup of NG (D), and
D = (P NG (D)) g(P NG (D)) is the intersection of two Sylow p-subgroups of NG (D).
Thus D Op (NG (D)). But on the other hand D is a normal p-subgroup of NG (D), and
so is contained in Op (NG (D)). Thus we have equality.
The condition on a subgroup D of G that appeared in Corollary 12.11, namely that
D = Op NG (D), is quite restrictive. Such subgroups have been called p-radical subgroups in
recent years, and they have also been called p-stubborn subgroups. They play an important
role in many questions about groups to do with topology, for example in studying the
properties of classifying spaces and in studying the topological properties of the partiallyordered set of p-subgroups of G. The terminology p-radical comes from the fact that when
G is a finite group of Lie type in characteristic p the p-radical subgroups are precisely
the unipotent radicals of parabolic subgroups. Thus the definition of p-radical subgroup
extends this notion to all finite groups. The name p-stubborn refers to the fact that in
certain calculations with the partially-ordered set of p-subgroups of G, many subgroups
can be ignored, but the p-stubborn ones cannot. The reader should be warned that the
term p-radical has also been used in some different senses, one of which is described in
the book by Feit [Fei].
It is immediate that Sylow p-subgroups of a group are p-radical, as is Op (G). An
exercise in group theory (presented at the end of this section) shows that other p-radical
subgroups must lie between these two extremes. It is always the case that G has a p-block
with defect group a Sylow p-subgroup (the principal block, for example, as will be seen in
Corollary 12.18) but it need not happen that Op (G) is the defect group of any block. An
example of this is S4 in characteristic 2: from the Cartan matrix computed in Chapter 10
and Corollary 12.6 we see that the only block is the principal block, and it also follows
from Corollary 12.26.
Peter Webb
233
234
G
(3) If : A B is a homomorphism of G-algebras then (trG
H (b)) = trH ((b)).
Proof. The calculations are the same as for 11.11 and 11.12, and part (3) is immediate.
For most of the arguments we will present it will be sufficient to consider G-algebras,
but sometimes a stronger property is needed which block algebras possess. We define an
interior G-algebra over R (a notion due to L. Puig) to be an R-algebra A together with
a group homomorphism u : G A where A denotes the group of units of A. As an
example, the group algebra RG is itself an interior G-algebra where u is the inclusion of
G as a subset of RG. Whenever A is an interior G-algebra and : A B is an algebra
homomorphism (sending 1A to 1B ) we find that B becomes an interior G-algebra via the
homomorphism u : G B . Thus if B is a block of G the algebra homomorphism
RG B makes B into an interior G-algebra. As a further example, if U is any RGmodule its structure is determined by an algebra homomorphism RG EndR (U ) that
expresses the action of RG. In fact, to specify a module action of G on U is the same as
specifying the structure of an interior G-algebra on EndR (U ).
Given an interior G-algebra A we obtain the structure of a G-algebra on A by letting
each g G act as a 7 ga := u(g)au(g 1). We see immediately that our three examples of
the G-algebra structure on RG, on a block, and on EndR (U ) are obtained in this way.
Why should we consider interior G-algebras? The reason here is that if A is an interior
G-algebra and U is an A-module then we can recover the structure of U as an RG-module
u
from the composite homomorphism GA EndR (U ), and without the extra property
of an interior G-algebra we cannot do this. This property is used in the next lemma, which
will be used in Corollary 12.17.
(12.13) PROPOSITION. Let A be an interior G-algebra over R, let U be an AH
module and let H be a subgroup of G. Suppose that 1A = trG
H a for some element a A .
Then regarded as an RG-module, U is H-projective.
Proof. The representation of A on U is given by a G-algebra homomorphism : A
EndR (U ). This homomorphism is a homomorphism of G-algebras and by Proposition
G
12.12 we have 1U = (1A ) = (trG
H a) = trH (a). Thus by Higmans criterion (part (6) of
11.13) U is H-projective.
Our goal now is Theorem 12.16, which makes a connection between the moduletheoretic approach to the defect group as we have defined it, and the ring-theoretic ap
proach. The proof will follow from the next two results. As before, (G) = {(g, g) g G}.
Peter Webb
235
GG
(G)
GG
GG
Proof. The hypothesis says that U is a summand of some module V (H)
, so U (G)
is a summand of
GG GG
(G) =
V (H)
(G)
(H)
x[(G)\GG/(H)]
In this formula we may write each x G G as x = (a, b) = (a, a)(1, a1b) and now
(G) x(H) =
(a,a)
((G)
(a,a)
(a,a)
{(H, a
(1,a1 b)
(H))
1
b
h) h = a bh}
(a,a)
(H).
GG GG
(G) is projective relative
It follows that every summand of the decomposition of V (H)
GG
is
to a (G)-conjugate of (H), which is the same as being (H)-projective. Thus U (G)
(H)-projective.
The point about these lemmas is that we are relating the structure of a block algebra
eRG as a representation of G G to its structure as a representation of G via the isomor-
then trG
H and tr(H) are exactly the same thing, from the definitions. This is because G
is taken to act on EndR (G) via .
element a (eRG)H .
GG
is (H)-projective. By Higmans criterion there is an endomorLemma 12.14 (eRG) (G)
(G)
236
trG
H . Now
e = 1eRG (e)
(G)
= (tr(H) )(e)
X
g
=
(e)
g[G/H]
(( g e))
g[G/H]
g((g 1 eg))g 1
g[G/H]
g(e)g 1
g[G/H]
= trG
H ((e)),
using the fact in the middle that e is central. We take a = (e) and the proof in this
direction is complete.
H
1
Conversely, suppose e = trG
= a for all h H
H a for some a (eRG) . Thus hah
P
1
and e = g[G/H] gag . Now : eRG eRG specified by (x) = ax is an endomorphism of R[(H)]-modules (which is the same thing as an endomorphism of RH-modules),
since
((h, h)x) = ahxh1 = h(h1 ah)xh1 = haxh1 = (h, h)(x).
(G)
(G)
(tr(H) )(x) =
g[G/H]
g[G/H]
g[G/H]
gag 1xgg 1
g[G/H]
gag 1x
g[G/H]
= ex
= x.
GG
This shows that eRG (G)
is (H)-projective, and hence since eRG is (G)-projective by
Proposition 12.8 it implies that eRG is in fact (H)-projective.
Peter Webb
237
Putting the last results together we obtain a proof of the following characterization of
the defect group in terms of the effect of the relative trace map on the interior G algebra
eRG. This characterization could have been used as the definition of the defect group.
(12.16) THEOREM. Let e be a block of RG and D a p-subgroup of G. Then D
is a defect group of e if and only if D is a minimal subgroup with the property that
D
trG
(eRG)G is surjective. Equivalently, D is a minimal subgroup with the
D : (eRG)
H
property that e = trG
D a for some a (eRG) .
Proof. From the definition, D is a defect group of e if and only if D is a minimal
subgroup of G such that eRG is (D)-projective. By Lemma 12.15 it is equivalent to say
(G)
that D is minimal such that e can be written e = tr(H) a for some a (eRG)H . With
the understanding that G acts on eRG via (G), we may write this as e = trG
H a. Since
G
H
G
D
tr(eRG) is an ideal of (eRG) , it is equivalent to say that trD : (eRG) (eRG)G is
surjective.
(12.17) COROLLARY. Let e be a block of RG with defect group D. Then every
eRG-module is projective relative to D and hence has a vertex that is a subgroup of D.
Proof. We see from Theorem 12.16 that it is possible to write e = trG
D a for some
D
a (eRG) . Thus by Proposition 12.13 every eRG-module is projective relative to D.
Note that it follows by the remarks about Greens indecomposability theorem at the
end of Chapter 11 that if |G| = |D|pa q where q is relatively prime to p, then pa divides
the dimension of every F G-module and kG-module in the block.
It is a fact (which we will not prove in general) that every block has an indecomposable module with vertex exactly the defect group D, so that the defect group may be
characterized as the unique maximal vertex of modules in the block. In the case of the
principal block the trivial module provides an example of such a module and it allows us
to deduce the next corollary.
(12.18) COROLLARY. The defect groups of the principal block are the Sylow psubgroups of G.
Proof. A defect group is a p-subgroup of G, and it must contain a vertex of the trivial
module, which is a Sylow p-subgroup by 11.24 part (3).
238
We are now in a position to show that our previous use of the term block of defect
zero in Chapter 9 is consistent with the definitions of this section. A block of defect zero
was taken to be a block with a representation satisfying any of the equivalent conditions
of Theorem 9.28. These were seen to be equivalent to the condition that over the field k
the block has a simple projective representation, and also equivalent to the condition that
over k the block is a matrix algebra.
(12.19) COROLLARY. Assume that k is a splitting field for G. A block of kG has
defect zero in the sense of this chapter if and only if it has defect zero in the sense of
Chapter 9, which can be characterized as saying that the block is a matrix algebra over k.
Proof. If the block has defect zero in the sense of this chapter its defect group is 1 and
by Corollary 12.17 every module in the block is 1-projective, or in other words projective.
Thus the block has a simple projective module and by Theorem 9.28 and the comment
immediately after its proof the block has defect zero in the sense considered there.
Conversely, suppose the block has defect zero in the sense of Chapter 9, so that
ekG is a matrix algebra over k. We will show that ekG is projective as a k[G G]module, and will make use of the isomorphism k[G G]
= kG k kG. Let : kG kG
be the algebra anti-isomorphism that sends each group element g to its inverse, so that
P
P
( g g) =
g g 1 . An element x in the second kG factor in the tensor product acts on
ekG as right multiplication by x , and it follows that e e acts as the identity on ekG
(Since e = e). Now e e is a central idempotent in kG k kG that generates the 2-sided
ideal ekG k e kG = ekG k (ekG) . This is the tensor product of two matrix algebras,
since the image of a matrix algebra under an anti-isomorphism is a matrix algebra. Such
a tensor product is again a matrix algebra, for if Eij and Fkl are two matrix algebra bases
(consisting of the matrices that are non-zero in only one place, where the entry is 1), then
the tensors Eij k Fkl are a basis for the tensor product that multiply together in the
manner of a matrix algebra basis. We see that the block ekG k e kG of kG k kG is
semisimple, and so ekG is a projective k[G G]-module. This shows that the defect group
of ekG is 1, and so ekG has defect zero as defined in this chapter.
Peter Webb
239
K
trH
K (U ).
K<H
G
We also define the Brauer morphism BrG
U (H) to be the composite
H :U
resG
H
H
U G U
U (H)
where the latter map is the quotient homomorphism. Here U G is simply a vector space
but U (H) has further structure: it is a k[NG (H)]-module with the action determined
g
by the maps cg : U H U H . If g NG (H) then g H = H so cg preserves H-fixed
g
K
K
K
points, and since cg trG
= trH
= trH
we see that cg permutes the terms
g K cg U
gK U
KU
H
in the sum being factored from U to produce U (H), so cg has a well-defined action on
U (H). We see also that the image of BrG
points under the action of
H lies in the fixed
P
K
NG (H). If U happens furthermore to be a G-algebra then K<H trH
K (U ) is an ideal of
U H by Proposition 12.12, the Brauer quotient U (H) is an NG (H)-algebra, and the Brauer
morphism is a ring homomorphism.
We will eventually apply these constructions to kG regarded as a G-algebra, in which
case we regard kG as a representation of G via the conjugation action and (kG)G is the
centre of kG. The Brauer quotient provides a means to express properties of the relative
trace map. We already have a characterization in Theorem 12.16 of the defect group of a
block in terms of this map. We now provide a corresponding characterization in terms of
the Brauer quotient. This is preceded by a more technical lemma.
240
The next result provides two more characterizations of the defect group of a block.
Compare parts (2) and (3) with the statement from Theorem 12.16 that D is equivalently
H
a minimal subgroup such that e = trG
D a for some a (ekG) . (Such D is therefore
necessarily unique to up to conjugacy.)
(12.21) THEOREM. Let e be a block of kG and D a subgroup of G. The following
are equivalent.
(1) e has defect group D.
D
(2) e = trG
and BrG
D (e) 6= 0.
D (a) for some element a (kG)
(3) D is up to conjugacy the unique maximal subgroup of G such that BrG
D (e) 6= 0.
Proof. (1) (2) If e has defect group D then D is minimal among groups for which
D
e = trG
by Theorem 12.16. Thus certainly e = trG
D (a) for some a (kG)
D (a). Suppose
P
P
G
D
K
that BrD (e) = 0. then e K<D trK (kG) and we may write e = K<D trD
K (uK ) where
K
uK (kG) . Note that although this expression only suggests that e (kG)D , in fact
e (kG)G . Now
e = ee
X
= (trG
trD
D (a))(
K (uK ))
K<D
= trG
D (a
trD
K (uK ))
K<D
trG
D
trD
K (auK )
K<D
trG
K (auK ).
K<D
P
K
K
Thus e K<D trG
and since e is a primitive idempotent, e trG
for some
K (kG)
K (kG)
K < D, by Rosenbergs Lemma 11.28. This contradicts the minimal property of D and so
the supposition BrG
D (e) = 0 was false.
Peter Webb
241
D
(2) (3) Suppose that e = trG
and BrG
D (e) 6= 0. By Lemma
D (a) for some a (kG)
12.20, if K is any subgroup for which BrG
(e)
=
6
0
then
K
is
a
subgroup
of a conjugate of D,
K
and this shows that D is up to conjugacy the unique maximal subgroup with BrG
D (e) 6= 0.
(3) (1) Suppose that condition (3) holds and let D1 be a defect group of e. By the
implication (1) (3) we know that D1 is up to conjugacy the unique maximal subgroup
for which BrD1 (e) 6= 0. Since D has this property, D and D1 are conjugate, and D is a
defect group.
The relative trace map and the Brauer morphism have the convenient theoretical
properties we have just described, but so far it does not appear to be easy to calculate
with them in specific cases. We remedy this situation by showing that the Brauer morphism
for the module kG acted upon by conjugation has an interpretation in terms of subgroups
of G and group elements. We start with a general lemma about permutation modules.
Since we will apply this in the situation of a subgroup H of G we denote our group by
:= P
H. If is an H-set we write
w as an element of the permutation module k.
(Elsewhere we might have denoted this element , but the bar notation is already in use
here.)
(12.22) LEMMA. Let be an H-set and k the corresponding permutation module.
1, . . . ,
n where =
(1) The fixed point space (k)H has as a basis the H-orbit sums
1 n is a disjoint union of H-orbits.
P
K
(2) If H is a p-group then K<H trH
of (k)H with basis the
K ((k) ) is the subspace
P
H
K
i where |i | > 1. Thus (k)H = k[H ]
orbit sums
K<H trK ((k) ) in this
case. The Brauer quotient k(H) may be identified with k[H ], the span of the fixed
points of the H-set .
Proof. (1) For each transitive H-set i we know that (ki )H is the 1-dimensional
i . From this it follows that if = 1 n where the i are the
space spanned by
orbits of H on then (k)H = (k1 )H (kn )H has as a basis the orbit sums.
(2) Suppose H is a p-group. Since the relative trace map preserves direct sums it
suffices to assume that H acts transitively on , so
H/J for some subgroup J. We
P
P=
H
K
have = trJ (0 ) for any fixed 0 . Thus K<H trH
contains the orbit
K (k)
sums for orbits of size larger than 1. On the other hand, if = {} has size 1 and K < H
then trH
K = |H : K| = 0 since H is a p-group so that |H : K| = 0 in k.
(12.23) COROLLARY. Let G act on kG via conjugation and let H be a p-subgroup
of G. Then kG(H)
= k[CG (H)]. With this identification the Brauer morphism is a
NG (H)
ring homomorphism BrG
that truncates a group ring element
H : Z(kG) k[CG (H)]
P
P
gG g g to
gCG (H) g g.
Proof. In the conjugation action kG is a permutation module and the set of fixed
points of H on G is CG (H). The Brauer morphism may be identified as inclusion of
242
fixed points followed by projection onto the first factor in the decomposition (kG)H =
P
K
k[CG (H)] K<H trH
K ((kG) ) given in Lemma 12.22. Under this identification the Brauer
map is truncation to of support to CG (H).
In older treatments the Brauer map may be defined as the map Z(kG) k[CG (H)]
that truncates a group ring element to have support on CG (H). A disadvantage of this
direct approach is that it is less obvious that the Brauer morphism is a ring homomorphism
and has the other properties we have described. It also ignores the larger context in
which the Brauer quotient and morphism may be defined for any kG-module, and this has
significance beyond the scope of this text. When the module is kG with the conjugation
action, the interpretation of BrG
H as truncation of the support of a group ring element to
CG (H) does provide a concrete understanding of this homomorphism in the context of
blocks. We immediately see the following, for example.
(12.24) COROLLARY. Let e be a block of kG. A defect group of e is a maximal
p-subgroup D of G such that e has some part of its support in CG (D).
Proof. This follows from part (3) of Theorem 12.21 since the condition given is exactly
the requirement that BrG
D (e) 6= 0.
We illustrate with G = S3 and k = F2 where we have blocks e1 = ()+(1, 2, 3)+(1, 3, 2)
and e2 = (1, 2, 3) + (1, 3, 2). When H = h(1, 2)i we have CG (H) = H and BrG
H (e1 ) = (),
G
G
BrG
(e
)
=
0.
We
also
have
Br
(e
)
=
e
and
Br
(e
)
=
e
showing
that
h(1,
2)i and 1
1
1
2
2
H 2
1
1
G
are (respectively) the largest subgroups H (up to conjugacy) for which BrH (e1 ) 6= 0 and
BrG
H (e2 ) 6= 0. These subgroups are the defect groups of the blocks, as asserted by Corollary
12.18, Example 7.5 and the discussion after 12.7.
Notice that, as with Galois correspondence, for each containment of subgroups K H
we have the reverse containment of fixed points (kG)K (kG)H and also a containment
of centralizers CG (K) CG (H). This
means that we obtain a commutative diagram
resG
H
G
G
(kG)H kG(H) = kCG (H)
BrH : (kG)
k
yresK
y
G
BrG
K : (kG)
resG
K
(kG)K
G
where is the inclusion of the centralizer group rings, so that BrG
K is the composite BrH .
G
G
We see from this that if BrH (x) 6= 0 for some x (kG)G then BrK (x) 6= 0 for every K
G
H, since BrG
H (x) equals BrK (x) with its support truncated to CG (H). This observation
provides a strengthening of part (3) of Theorem 12.21: if K is any subgroup of a defect
group D of a block e of kG, then BrG
K (e) 6= 0.
The next result is important and interesting in its own right. It will be used when
we come to describe the Brauer correspondence of blocks. The statement of this result is
given in terms of Op (G), the largest normal p-subgroup of G. Although Op (G) could, in
principle, be 1, the result gives no information in that case, so it is really a result about
groups with a non-trivial normal p-subgroup.
Peter Webb
243
244
and elements that are fixed under NG (H) are also fixed under J. If 1 = e1 + + en is
G
the sum of blocks of kG then 1 = BrG
H (e1 ) + + BrH (en ) is a sum of orthogonal central
G
idempotents of kJ. Thus if b is a block of kJ then b = bBrG
H (e1 ) + + bBrH (en ) is a
decomposition of b as a sum of orthogonal idempotents in Z(kJ), and since b is primitive
G
in Z(kJ) there is a unique block ei of G so that bBrG
H (ei ) = b. We write b for this block
ei , and call it the Brauer correspondent of b. There may be several blocks b of kJ with the
same Brauer corresponding block of kG: bG = bG . We see that the blocks of kG partition
the blocks of kJ by this means.
There was a choice of p-subgroup H in the definition of bG , since it would be possible
to have another p-subgroup H1 with H1 CG (H1 ) J NG (H1 ), and perhaps the definition
of bG would be different using H1 . In fact the choice of H does not matter, as we now
show.
(12.27) PROPOSITION. If b is a block of kJ, the definition of bG is independent of
the choice of the p-group H satisfying HCG (H) J NG (H).
Proof. Since H is a normal p-subgroup of J we have H Op (J) and CG (H)
CJ (Op (J)). By Proposition 12.25 every central idempotent of kJ lies in k[CG (Op (J))].
For each block ei of kG the idempotent BrG
H (ei ) thus lies in k[CG (Op (J))] and could have
been computed by truncating the support of ei to lie in CJ (Op (J)) instead of CG (H). This
G
identification of BrG
H (ei ) is independent of the choice of H and hence so is b .
(12.28) PROPOSITION. Let H be a p-subgroup of G, J a subgroup of G with
HCG (H) J NG (H) and b a block of kJ. Then bG has a defect group that contains a
defect group of b.
Proof. If D J is a defect group of b then D H by Corollary 12.11 since H is a
normal p-subgroup of J, and so CG (H) CG (D) = CJ (D) since J CG (H). Let us write
G
BrG
H (b ) = b + b1 where b and b1 are orthogonal central idempotents of kJ. We can now
apply BrJD to this, and since this map truncates to CJ (D) we get the same thing as if we
J
J
G G
J
G G
had originally applied BrG
D . Thus BrD (b ) = BrD (BrH (b )) = BrD (b) + BrD (b1 ) is a sum
G
of orthogonal idempotents in k[CJ (D)]. Since BrJD (b) 6= 0 it follows that BrG
D (b ) 6= 0,
and hence D is contained in a defect group of bG by Theorem 12.21.
(12.29) LEMMA. Let H be a subgroup of G and U a kG-module. Then
N (H)
G
G
BrG
H trH = trH
H
BrH
U (H)NG (H) .
H :U
P
H
G
G G
Proof. BrG
H trH (x) is the image in U (H) of resH trH (x) =
g[H\G/H] trH gH (gx).
g
The only terms that contribute have H H = H, which happens if and only if g NG (H),
P
N (H)
so the image equals the image of g[NG (H)/H] gx = trHG x.
Peter Webb
245
Since trN
D preserves both of the two summands on the right and b k[CG (D)] by PropoD
N
sition 12.25, we may assume a k[CG (D)] and so BrD
D (a) = a. Thus b = trD BrD (a)
trN
D (k[CG (D)]).
Let e = bG Z(kG) be the Brauer correspondent of b, so that b = bBrG
D (e) is a
G
summand of BrD (e) and e has a defect group D1 D by Proposition 12.28. We will show
that D1 = D. Now using Lemma 12.29
D
G
D
N
D
N
BrG
D (trD ((kG) )) = trD BrD ((kG) ) = trD (k[CG (D)])
246
P = C2 C2 has NG (P )
= A4 , and we have seen in exercise 2 from Chapter 8 that kA4 has
only one block. Thus there is only one block of kG with defect group P : it is the principal
block. The subgroups C2 have Sylow p-subgroups P as their normalizers and there are no
blocks of kP with C2 as defect group, since P is a 2-group. Hence there are no blocks of
kG with this defect group. This duplicates information we know from a different source,
to the effect that C2 cannot be a defect group by Corollary 12.11 because it is not O2 of
its normalizer. Finally there remain the blocks of defect zero of kA5 , which have defect
group 1. There is in fact just one of these, as may be seen by inspecting the character
table of A5 for characters of degree divisible by 4.
Summary of Chapter 12
1. Blocks correspond to blocks of the Cartan matrix, indecomposable ring summands
of kG or RG, primitive central idempotents in kG or RG, and certain equivalence classes
of representations in characteristic 0 or in characteristic p.
2. The defect group of a block may also be characterized in several ways, using the
bimodule structure of the block as a ring summand, the relative trace map, the Brauer
morphism, and the vertices of modules in the block.
3. A defect group is always the intersection of two Sylow p-subgroups and is the
largest normal p-subgroup of its normalizer.
4. The principal block has Sylow p-subgroups as defect groups.
5. The Brauer correspondent bG provides a bijection between blocks of G and of
NG (D) with defect group D.
Peter Webb
247
248
that
(a) = 0 if and only if a = 0,
(ab) = (a)(b) for all a, b F , and
(a + b) (a) + (b) for all a, b F .
In many texts the theory is developed in terms of additive valuations. A suitable reference
which uses the multiplicative language is [Hua].
(A.1) Examples. 1. No matter what the field F is, we always have the valuation
(a) =
0 if a = 0
1 otherwise.
so that
a = pp (a)
if a = 0,
if p r,
otherwise,
r
s
where (r , p) = 1 = (s , p). Now let be any real number with 0 < < 1 and put
(a) = p (a) . Often is taken to be p1 , but the precise choice of does not affect the
properties of the valuation. This valuation is called the p-adic valuation on Q.
This last is an example of a valuation which satisfies the so-called ultrametric inequality
(a + b) max{(a), (b)},
which in the case of this example comes down to the fact that if pn a and pn b
where a, b Z, then pn (a + b). We say that is non-archimedean if it satisfies the
ultrametric
inequality. The valuations in the third example are also discrete, meaning that
{(a) a K, a 6= 0} is an infinite cyclic group under multiplication. It is the case that
discrete valuations are necessarily non-archimedean.
We deduce from the axioms for a valuation that (1) = (1) = 1. Using this we
see that every valuation gives rise to a metric d(a, b) = (a b) on the field F . We
say that two valuations are equivalent if and only if the metric spaces they determine are
equivalent, i.e. they give rise to the same topologies. In Example 3 above, changing the
value of between 0 and 1 gives an equivalent valuation.
Peter Webb
249
250
As for (6), given any element a F , either a R or (a) = ()n for some n > 0.
= {a F (a)
1} with unique maximal ideal
valuation we have a valuation ring R
P = {a F (a) < 1}, and P = R () since () = () generates the value group.
Peter Webb
251
the powers Pn .
is denoted Qp and is called
When is the p-adic valuation on Q, the completion Q
with respect to is denoted Zp and is
the field of p-adic rationals. The valuation ring Q
called the ring of p-adic integers.
(A.4) LEMMA. Let be a discrete valuation on a field F with valuation ring R .
induces an isomorphism R /P n
n
The inclusion R R
= R /P for all n.
R
/P n . Its kernel is
Proof. Consider the composite homomorphism R R
Pn and the desired isomorphism will follow if we can show that this homomorphism is
we know from the construction of the completion
surjective. To show this, given a R
that there exists b R with (b a) < ()n , that is, b a Pn . Now b maps to
a + Pn .
is by definition the set of equivalence classes of Cauchy sequences
The completion R
in R . We comment that a sequence (ai ) of elements of R is a Cauchy sequence if and
only if for every n there exists a number N so that whenever i, j > N we have ai aj Pn ,
that is, ai aj (mod Pn ).
(A.5) LEMMA. Let be a discrete valuation on a field F with valuation ring R ,
. Any element of R
is uniquely expressible as a series
maximal ideal P and completion R
a = a0 + a1 + a2 2 +
where the ai lie in a set of representatives S for R /P .
. Since R
/P
Proof. Let a R
= R /P , we have a+ P = a0 + P for some uniquely
. Repeating this
determined a0 S. Now a a0 P so a = a0 + b1 for some b1 R
construction we write b1 = a1 + b2 with a1 S uniquely determined, and in general
bn = an + bn+1 with an S uniquely determined. Now a0 , a0 + a1 , a0 + a1 + a2 2 , . . .
is a Cauchy sequence in R whose limit is a, and we write this limit as the infinite series.
The last result, combined with Proposition A.3, provides a very good way to realize
. For example, in the case of the p-adic valuation on Q we may take
the completion R
= Zp may be realized as the set of infinite
S = {0, 1, . . . , p 1}. The completion Z
sequences a3 a2 a1 a0 . of elements from S presented in positions to the left of a point,
analogous to the decimal point (which we write on the line following American convention).
Thus a0 is in the 1s position, a1 is in the ps position, a2 is in the p2 s position, and so
on. Unlike decimal numbers these strings are potentially infinite to the left of the point,
whereas decimal numbers are potentially infinite to the right of the point. Addition and
multiplication of these strings is performed by means of the same algorithms (carrying
252
values from one position to the next when p is exceeded, etc.) that are used with infinite
decimals. Note that p-adic integers have the advantage over decimals that certain real
numbers have more than one decimal representation, whereas distinct p-adic expansions
always represent distinct elements of Zp .
Exercises.
1. With the description of the p-adic integers as the set of infinite sequences
a3 a2 a1 a0 .
in positions to the left of a point, where ai {0, . . . , p 1}, show that when p = 2 we
have
1 = 1111. and
1
= 10101011.
3
Find the representation of the fraction 1/5 in the 2-adic integers. What fraction does
1100110011. represent?
2. Show that the field of p-adic rationals Qp may be constructed as the set of sequences
a3 a2 a1 a0 .a1 a2 an which may be infinite to the left of the point, but must be finite
to the right of the point, where ai {0, . . . , p 1} for all i. Show that the field of rational
numbers Q is the subset of these sequences which eventually recur.
Peter Webb
253
We let Cn = hx xn = 1i.
Characteristic 0:
Cn
ordinary characters
xn1
n
g
|CG (g)|
1 x
n n
ns
(0 s n 1)
1 ns n
s(n1)
This table was described in Proposition 4.1. The representations in positive characteristic
were described for abelian groups in Example 8.6 and the results which followed.
The symmetric group S3 .
This group is isomorphic to the dihedral group D6 and also to GL(2, 2) = SL(2, 2)
Characteristic 0:
S3
ordinary characters
g
|CG (g)|
() (12) (123)
6
2
3
1
sign
2
1
1
1 1
2
0
1
1
1
254
Characteristic 2:
S3
= SL(2, 2)
Brauer simple p = 2
S3
= SL(2, 2)
Brauer projective p = 2
g
|CG (g)|
() (123)
6
3
g
|CG (g)|
() (123)
6
3
1
2
1
1
2 1
1
2
2
2
2 1
S3
= SL(2, 2)
Decomposition matrix p = 2
S3
= SL(2, 2)
Cartan matrix p = 2
1 2
1 2
1
sign
2
1
1
0
0
0
1
1
2
2
0
0
1
Characteristic 3:
S3
Brauer simple p = 3
S3
Brauer projective p = 3
g
|CG (g)|
() (12)
6
2
g
|CG (g)|
() (12)
6
2
1
sign
1
1
1 1
1
sign
3
1
3 1
S3
Decomposition matrix p = 3
S3
Cartan matrix p = 3
1 sign
1 sign
1
sign
2
1
0
1
0
1
1
1
sign
2
1
1
2
Peter Webb
255
1 x2 x y xy
8 8 4 4 4
1
1a
1b
1c
2
1 1 1 1 1
1 1 11 1
1 1 1 1 1
1 1 11 1
2 2 0 0 0
1
1a
1b
3
1
1
1
3
1
1
1
1
1
3
32
0
1
32
3
0
Characteristic 2:
A4
Brauer simple p = 2
A4
Brauer projective p = 2
g
() (123) (132)
|CG (g)| 12 3
3
g
() (123) (132)
|CG (g)| 12 3
3
1
1a
1b
1
1a
1b
1
1
1
1
3
32
1
32
3
4
4
4
1
3
32
1
32
3
A4
Decomposition matrix p = 2
A4
Cartan matrix p = 2
1 1a 1b
1 1a 1b
1
1a
1b
3
1
0
0
1
0
1
0
1
0
0
1
1
256
1
1a
1b
2
1
1
1
2
1
1
1
2
Characteristic 3:
A4
Brauer simple p = 3
A4
Brauer projective p = 3
g
|CG (g)|
() (12)(34)
12
4
g
|CG (g)|
() (12)(34)
12
4
1
3
1
3
1
3
3
3
1
1
A4
Decomposition matrix p = 3
A4
Cartan matrix p = 3
1 3
1 3
1
1a
1b
3
1
1
1
0
0
0
0
1
3
1
1
3
3
0
0
1
Construction notes for A4 : the element (123) conjugates transitively the three nonidentity elements of order two in the unique Sylow 2-subgroup h(12)(34), (13)(24)i. The
remaining group elements have order three, conjugated in two orbits by the Sylow 2subgroup, which equals the derived subgroup. There are three degree 1 characters. The
remaining character can be found from the orthogonality relations; it can be constructed
by inducing a non-trivial character from h(12)(34), (13)(24)i; it is also the character of the
realization of A4 as the group of rotations of a regular tetrahedron.
The dihedral and quaternion groups of order 16.
Put
D16 = hx, y x8 = y 2 = 1, yxy 1 = x1 i,
Q16 = hx, y x8 = 1, x4 = y 2 , yxy 1 = x1 i.
Peter Webb
257
1 x4 x2
16 16 8
1
1a
1b
1c
2a
2b
2c
1 1 1 1
1 1 1
1 1 1 1
1 1 1
1 1 1 1 1 1 1
1 1 1 1 1 1 1
2 2 2 0 0 0 0
2 2 0 22 0 0
2 2 0 2
2 0 0
x
8
x5 y xy
8 4 4
Construction notes: as with D8 and Q8 , find the conjugacy classes by listing the elements.
The quotient by hx4 i is a copy of either D8 or Q8 (depending on the case), and we obtain
the top 5 rows of the character table by lifting (or inflating) the characters from the
quotient group. The final two characters are obtained by inducing the characters 8 and
83 from the cyclic subgroup hxi.
The character table of the group SD16 = hx, y x8 = y 2 = 1, yxy 1 = x3 i is
SD16
ordinary characters
g
|CG (g)|
1 x4 x2
16 16 8
1
1a
1b
1c
2a
2b
2c
1 1 1
1
1 1 1
1 1 1
1
1 1 1
1 1 1 1 1 1 1
1 1 1 1 1 1 1
2 2 2
0
0 0 0
2 2 0 i2i2 0 0
2 2 0 i 2 i 2 0 0
x
8
x5
8
y xy
4 4
Construction notes: the same as for D16 and Q16 . The quotient by hx4 i is a copy of D8 ,
and we obtain the top 5 rows of the character table by inflating (or lifting) the characters
from the quotient group. The final two characters are obtained by inducing the characters
8 and 1 from the cyclic subgroup hxi.
8
258
The representations of C7 C3 = hx, y x7 = y 3 = 1, yxy 1 = x2 i in characteristic 7
were explicitly described after Proposition 8.9.
C7 C3
ordinary characters
g
|CG (g)|
1
21
1
1a
1b
3a
3b
1
1
1
3
3
x1
7
x
7
y y 1
3
3
1
1
1
1
1
3
1
1
32
2
4
3
5
6
7 + 7 + 7 7 + 7 + 7 0
73 + 75 + 76 7 + 72 + 74 0
1
32
3
0
0
Characteristic 3:
C7 C3
Brauer simple p = 3
g
1
|CG (g)| 21
1
3a
3b
C7 C3
Brauer projective p = 3
x1
7
x
7
1
1
1
2
4 3
3 7 + 7 + 7 7 + 75 + 76
3 73 + 75 + 76 7 + 72 + 74
C7 C3
Decomposition matrix p = 3
g
1
|CG (g)| 21
1
3a
3b
3
3
3
2
4 3
3 7 + 7 + 7 7 + 75 + 76
3 73 + 75 + 76 7 + 72 + 74
C7 C3
Cartan matrix p = 3
1 3a 3b
1 3a 3b
1
1a
1b
3a
3b
1
1
1
0
0
0
0
0
1
0
0
0
0
0
1
x1
7
x
7
1
3a
3b
3
0
0
0
1
0
0
0
1
Characteristic 7:
C7 C3
Brauer simple p = 7
C7 C3
Brauer projective p = 7
g
1 y y 1
|CG (g)| 21 3 3
g
1 y y 1
|CG (g)| 21 3 3
1
1a
1b
1
1a
1b
1 1 1
1 3 32
1 32 3
7 1 1
7 3 32
7 32 3
Peter Webb
259
C7 C3
Cartan matrix p = 7
C7 C3
Decomposition matrix p = 7
1 1a 1b
1 1a 1b
1
1a
1b
3a
3b
1
0
0
1
1
0
1
0
1
1
1
1a
1b
0
0
1
1
1
3
2
2
2
3
2
2
2
3
Construction notes: the two ordinary characters of degree 3 are induced from the characters
7 and 73 of the cyclic subgroup of order 7. In characteristic 3 they are blocks of defect
zero, so remain simple on reduction. In characteristic 7 the indecomposable projectives
were already constructed after 8.9, giving the Cartan matrix by a different method. In
the context of characters we can simply say that the three 1-dimensional characters are all
distinct on the 7-regular classes so give the three simple Brauer characters.
The symmetric group S4 .
S4
ordinary characters
g
|CG (g)|
1
sign
2
3a
3b
1
1
1 1
2
0
3 1
3
1
1
1
2
1
1
1
1
0
1
1
1
1
1
0
0
Construction notes: this table was constructed in Chapter 3. It (and tables for the other
symmetric groups) can also be constructed using the combinatorial methods available for
representations of the symmetric groups, not described in this text.
Characteristic 2:
S4
Brauer simple p = 2
S4
Brauer projective p = 2
g
|CG (g)|
() (123)
24
3
g
|CG (g)|
() (123)
24
3
1
2
1
2
1
2
8
8
1
1
2
1
260
S4
Cartan matrix p = 2
S4
Decomposition matrix p = 2
1 2
1 2
1
sign
2
3a
3b
1
1
0
1
1
1
2
0
0
1
1
1
4
2
2
3
Characteristic 3:
S4
Brauer simple p = 3
S4
Brauer projective p = 3
g
|CG (g)|
g
|CG (g)|
1
sign
3a
3b
1
1
1 1
3 1
3
1
1
sign
3a
3b
3
1
3 1
3 1
3
1
1
1
1
1
1
1
1
1
S4
Decomposition matrix p = 3
1
0
1
0
0
0
1
1
0
0
0
0
0
1
0
1 sign 3a 3b
1
sign
3a
3b
0
0
0
0
1
2
1
0
0
1
2
0
0
0
1
1
1
1
1
S4
Cartan matrix p = 3
1 sign 3a 3b
1
sign
2
3a
3b
3
3
1
1
1
0
and
y=
1
0
1
1
0
0
1
0
0
0
0
1
Peter Webb
261
1 a2 a
24 24 4
1
1a
1b
2a
2b
2c
3
1 1 1 1
1
1 1 1 3
32
1 1 1 32
3
2 2 0 1 1
2 2 0 3 32
2 2 0 32 3
3 3 1 0
0
y
6
y2
6
ya2 y 2 a2
6
6
1
3
32
1
3
32
0
1
32
3
1
32
3
0
Characteristic 2:
SL(2, 3)
Brauer simple p = 2
SL(2, 3)
Brauer projective p = 2
g
1 y y 1
|CG (g)| 24 6 6
g
1 y y 1
|CG (g)| 24 6 6
1
1a
1b
1
1a
1b
1 1 1
1 3 32
1 32 3
SL(2, 3)
Decomposition matrix p = 2
SL(2,3)
Cartan matrix p = 2
1 1a 1b
1 1a 1b
1
1a
1b
2a
2b
2c
3
Characteristic 3:
1
0
0
0
1
1
1
0
1
0
1
0
1
1
0
0
1
1
1
0
1
8 2 2
8 23 232
8 232 23
1
1a
1b
4
2
2
2
4
2
2
2
4
262
SL(2, 3)
Brauer simple p = 3
SL(2, 3)
Brauer projective p = 3
g
1 a2 a
|CG (g)| 24 24 4
g
1 a2 a
|CG (g)| 24 24 4
1
2
3
1
2
3
1 1 1
2 2 0
3 31
SL(2, 3)
Decomposition matrix p = 3
SL(2,3)
Cartan matrix p = 3
1 2 3
1 3 5
1
1a
1b
2a
2b
2c
3
1
1
1
0
0
0
0
0
0
0
1
1
1
0
0
0
0
0
0
0
1
3 3 3
6 6 0
3 3 1
1
2
3
3
0
0
0
3
0
0
0
1
Peter Webb
263
Since SL(2, 3) has A4 as a quotient, the character table of A4 lifts to SL(2, 3) giving
the characters of degree 1 and 3. Calculating the sum of the squares of the degrees shows
that the remaining three characters have degree 2. Since a2 is central it must act as a
scalar on each simple representation, so it acts as 2 in the degree 2 representations, and
column orthogonal between columns 1 and 2 shows that the entries must be 2 for the
characters of degree 2. Column orthogonality now gives the values of these characters as 0
on a. If any remaining entry in the degree 2 characters is non-zero, it gives non-zero values
in the same column for the degree two characters, by multiplying by the three degree 1
characters. Thus any zero entry here gives three zero entries on the degree 2 characters.
This is not possible because the product of that column with its complex conjugate must be
the centralizer order, which is 6. Hence the remaining entries are all non-zero, and once one
of the degree 2 characters is determined the other two are obtained by multiplying by the
degree 1 characters. We deduce from column orthogonality that each degree 2 character
must have absolute value 1 on y, y 2, ya2 and y 2 a2 . The matrices by which the element y
acts has eigenvalues taken from 1, 3 , 32 and the only possible sums of two of these with
absolute value 1 are 1 = 3 + 32 , 3 = 1 + 32 and 32 = 1 + 3 , so these must be the
three values of the degree 2 characters on y. The values on y 2 are their conjugates. The
three degree 2 characters must arise as one complex conjugate pair and one real character.
A similar argument with the possible sums of the eigenvalues of the matrix of ya2 , which
must come from 1, 3 , 32 now determines the remaining character values.
When p = 2 it is immediate that the three degree 1 characters reduce to give three
distinct simple Brauer characters. When p = 3 we have the trivial Brauer character, and
there are no more degree 1 characters because the abelianization is C3 . The degree 3
character is a block of defect zero so also gives a simple Brauer character. Finally the
degree 2 character gives a simple Brauer character, because it cannot be written as a sum
of degree 1 characters.
The tables for characteristics 2 and 3 illustrate the theory described in Proposition 8.8
and Theorem 8.10 for semidirect products of a p-group and a p -group. In characteristic
3 they also illustrate the form of the simple representations, as described in Chapter 6
Exercise 21. In characteristic 2 more detailed information about the projectives is obtained
in Chapter 8 Exercise 4. However, the calculation of these tables is straightforward and
does not require this theory.
The alternating group A5 .
264
() (12)(34) (123)
60
4
3
1
3a
3b
4
5
1
3
3
4
5
1
1
1
0
1
1
0
0
1
1
(12345)
5
(13524)
5
1
1
3
+ 5 ) (5 + 54 )
(5 + 54 ) (52 + 53 )
1
1
0
0
(52
Characteristic 2:
A5
= SL(2, 4)
Brauer simple p = 2
A5
= SL(2, 4)
Brauer projective p = 2
g
|CG (g)|
g
|CG (g)|
() (123)
60
3
1
2a
2b
4
1
2
2
4
1
2a
2b
4
12
0
2
2
8 1 (52 + 53 )(5 + 54 )
8 1 (5 + 54 )(52 + 53 )
4
1
1
1
1
1
1
1
1
1
5 + 54 52 + 53
52 + 53 5 + 54
1
1
A5
= SL(2, 4)
Decomposition matrix p = 2
1
1
1
0
1
0
1
0
0
1
0
0
1
0
1
(13524)
5
A5
= SL(2, 4)
Cartan matrix p = 2
1 2a 2b 4
1 2a 2b 4
1
3a
3b
4
5
(12345)
5
1
2a
2b
4
0
0
0
1
0
4
2
2
0
2
2
1
0
2
1
2
0
0
0
0
1
Characteristic 3:
A5
Brauer projective p = 3
A5
Brauer simple p = 3
g
() (12)(34) (12345)
|CG (g)| 60
4
5
1
3a
3b
4
1
3
3
4
1
1
1
0
(13524)
5
1
1
3
+ 5 )(5 + 54 )
(5 + 54 )(52 + 53 )
1
1
(52
g
() (12)(34) (12345)
|CG (g)| 60
4
5
1
3a
3b
4
6
3
3
9
2
1
1
1
(13524)
5
1
1
3
+ 5 )(5 + 54 )
(5 + 54 )(52 + 53 )
1
1
(52
Peter Webb
A5
Decomposition matrix p = 3
265
A5
Cartan matrix p = 3
1 3a 3b 4
1 3a 3b 4
1
3a
3b
4
5
1
0
0
0
1
0
1
0
0
0
0
0
1
0
0
0
0
0
1
1
1
3a
3b
4
2
0
0
1
0
1
0
0
0
0
1
0
1
0
0
2
Characteristic 5:
A5
= P SL(2, 5)
Brauer simple p = 5
A5
= P SL(2, 5)
Brauer projective p = 5
g
() (12)(34) (123)
|CG (g)| 60
4
3
g
() (12)(34) (123)
|CG (g)| 60
4
3
1
3
5
1
3
5
1
3
5
1
1
1
1
0
1
A5
= P SL(2, 5)
Decomposition matrix p = 5
5
10
5
1
0
0
1
0
0
1
1
1
0
0
0
0
0
1
2
1
1
A5
= P SL(2, 5)
Cartan matrix p = 5
1 3 5
1 3 5
1
3a
3b
4
5
1
2
1
1
3
5
2
1
0
1
3
0
0
0
1
Construction notes for A5 : to compute the conjugacy classes, compute the centralizer
of each element first in S5 , then intersect that centralizer with A5 . The index in A5 is the
number of conjugates of that element. This enables us to see that the class of 5-cycles
splits into two in A5 , but the other classes of S5 do not. One of the three-dimensional
representations can be obtained via the realization of A5 as the group of rotations of the
icosahedron. Note that when computing the trace of a rotation matrix it simplifies the
calculation to choose the most convenient basis. The tensor square of this representation
contains two copies of itself, determined by taking an inner product. The remaining
three-dimensional summand is simple and new. It can also be obtained as an algebraic
266
1
1
2
2
0
1
1
1
1
1
1
0
1
1
1
1
0
0
2
1
1
1
1
0
0
0
1
1
1
1
1
1
0
1
1
1
1
1
1
1
0
0
Construction notes for S5 : the conjugacy classes are determined by cycle type, the derived subgroup is A5 and there are two 1-dimensional representations. The permutation
representation on the five symbols is the direct sum of the trivial representation and a 4dimensional simple representation 4a (by Lemma 5.5, but it can be checked using the orthogonality relations). Its tensor with the sign representation gives another 4-dimensional
simple. Its exterior square is simple of dimension 6. The symmetric square decomposes
with a new summand of dimension 5. The final character is obtained by multiplying by
the sign representation.
The general linear group GL(3, 2).
The group GL(3, 2) = SL(3, 2) is simple and is isomorphic to P SL(2, 7).
GL(3, 2)
ordinary characters
g
|CG (g)|
1
3a
3b
6
7
8
1 2 4
168 8 4
3 7a 7b
3 7 7
1 1 1 1 1 1
3 1 1 0
3 1 1 0
6 2 0 0 1 1
7 1 1 1 0 0
8 0 0 1 1 1
Peter Webb
267
GL(3, 2)
Brauer projective p = 2
g
|CG (g)|
g
|CG (g)|
1 3
168 3
1
3a
3b
8
8 2
1
1
16 1 1 1
16 1 1 1
8 1
1
1
1
3a
3b
8
1 3 7a 7b
168 3 7 7
1 1 1
3 0
3 0
8 1 1
GL(3, 2)
Decomposition matrix p = 2
1
1
0
0
1
0
0
0
0
1
1
0
0
0
1
1
1
0
0
0
0
0
0
1
7b
7
GL(3, 2)
Cartan matrix p = 2
1 3a 3b 8
1 3a 3b 8
1
3a
3b
6
7
8
7a
7
1
3a
3b
8
2
1
1
0
1
3
2
0
1
2
3
0
Characteristic 7:
GL(3, 2)
= P SL(2, 7)
Brauer simple p = 7
GL(3, 2)
= P SL(2, 7)
Brauer projective p = 7
g
|CG (g)|
g
|CG (g)|
1 2 4 3
168 8 4 3
1
3
5
7
7 3 1 1
14 2 21
14 2 01
7 11 1
1 2 4 3
168 8 4 3
1
3
5
7
1 1 1 1
3 1 1 0
5 111
7 11 1
GL(3, 2)
= P SL(2, 7)
Decomposition matrix p = 7
GL(3, 2)
= P SL(2, 7)
Cartan matrix p = 7
1 3 5 7
1 3 5 7
1
3a
3b
6
7
8
1
0
0
1
0
0
0
1
1
0
0
1
0
0
0
1
0
1
0
0
0
0
1
0
1
3
5
7
2
0
1
0
0
3
1
0
1
1
2
0
0
0
0
1
0
0
0
1
268
Construction notes for GL(3, 2): By counting ordered bases for F32 the order of
GL(3, 2) is 168 = (23 1)(23 2)(23 22 ). It is helpful to identify certain subgroups
of GL(3, 2). We describe these by indicating the form of the matrices in the subgroups.
These matrices must be invertible and, subject to that condition, can have any field element
from F2 (i.e. 0 or 1) where is positioned. Let
1 0
B = 1
and
Then
0 , P1 =
1
1 0
P2 =
0
1 0 0
0 , U1 = 0 1 0 , L 1 =
1
1
0 0
0
1
, U2 =
0 0
1
1 0 , L2 = 0
0 1
0
0
0
1
0 0
.
B
= D8 ,
L1
= L2
= GL(2, 2)
= S3 ,
U1
= U2
= C2 C2 ,
P1 = U 1 L 1
= S4 ,
P2 = U 2 L 2
= S4
because S4 also has this semidirect product structure. The subgroup B is a Sylow 2subgroup of GL(3, 2), since it has order 8.
We show that GL(3, 2) has a single class of elements of order 2 (involutions) by
showing that all involutions in B are conjugate in GL(3, 2). Since all Sylow 2-subgroups
are conjugate in GL(3, 2), this will be sufficient. Conjugacy of involutions within B follows
because in P1 they fall into two conjugacy classes, and in P2 they fall into two different
classes, so that combining this information we see they are all conjugate in GL(3, 2). Direct
calculation shows that
1 0 0
B = CGL(3,2) ( 0 1 0 )
1 0 1
Peter Webb
269
order 6, since this group appears as a subgroup of P1 . All elements of order 3 are thus
conjugate, and there are 56 = 168/3 of them.
Sylows theorem shows that there are 1 or 8 Sylow 7-subgroups K, and since GL(3, 2)
does not have a normal 7-cycle, there are 8. Thus NGL(3,2) (K)
= K H of order 21, where
H has order 3. The action of H on K is non-trivial because the centralizer of H contains
no element of order 7, so if K = hgi the action of a generator of H may be taken to send
g 7 g 2 7 g 4 7 g. This means the elements of order 7 fall into two conjugacy classes,
represented by g and g 1 , each of size 24.
Since 168 = 1 + 21 + 42 + 56 + 24 + 24 we have accounted for all the elements of
GL(3, 2).
We construct the three largest degree characters of GL(3, 2). The induced character
from the subgroup K H of order 21 is
g
|CG (g)|
GL(3,2)
1 KH
1 2 4 3 7a 7b
168 8 4 3 7 7
8
0 0 2
0 1
1 0
0 0
0
1 0
0, 1 1
1
0 1
0
0
0, 0
1
1
1 0
0 1
0 0
and find the number of fixed points. For elements of order 7 it is clear that they act
regularly on the 7 non-zero elements of F32 , so there are no fixed points. This shows that
270
the character is
g
|CG (g)|
1 2 4 3 7a 7b
168 8 4 3 7 7
GL(3,2)
1 P1
3 1 1
GL(3,2)
0 C B
GL(3,2)
C P 1
GL(3,2)
C P2
0.
The simplicial complex is, in fact, connected so that we have an exact sequence
GL(3,2)
0 H1 C B
GL(3,2)
C P 1
GL(3,2)
C P2
C0
GL(3,2)
and 1 P2
The computation of
is made easier by the fact that B is a 2-group, so that the
character is only non-zero on 2-power elements. It is
g
|CG (g)|
GL(3,2)
1 B
1 2 4 3 7a 7b
168 8 4 3 7 7
21
5 1 0
The value on an involution arises because there are 5 involutions in B, which is the centralizer one of them, and so 5 coset representatives of B conjugate this involution to an
Peter Webb
271
element of B. There are two elements of order 4 in B, and they determine B as the normalizer of the subgroup they generate. Since B is self-normalizing, there is only one coset
of B whose representative conjugates an element of order 4 into B. We compute that H1
is a simple character using the orthogonality relations, and it has degree 8.
It remains to compute the two 3-dimensional characters of GL(3, 2). To do this we
can say that the two columns indexed by elements of order 7 must be complex conjugates
of each other because the elements are mutually inverse. All the characters constructed so
far are real-valued, so the two remaining characters must not be real-valued in order that
those two columns should be distinct. This means that the two remaining characters must
be complex conjugates of each other. On elements of orders 1, 2, 3 and 4 these characters
are real, so must be equal. With this information we can now determine those characters
on these columns using orthogonality relations and the fact that the sum of the squares
of the degrees of the characters is 168. Orthogonality gives an equation for the missing
complex number which is also determined in this way.
Dihedral groups.
We let D2n = hx, y xn = y 2 = 1, yxy 1 = x1 i.
D2n , n odd
ordinary characters
1
2n
x
n
x2
n
1
1a
1
1
1
1
1
1
ns G
hxi
g
|CG (g)|
(1 s
y
2
n
1
1
n1
ns + ns n2s + n2s n 2
n1
2 )
n1
2
n1
+ n 2
1
1
0
D2n , n even
ordinary characters
g
|CG (g)|
1
2n
1
1a
1b
1c
ns G
hxi
(1 s
1
1
1
1
2
n
2
1)
x
n
x2
n
1
1
1
1
1
1
1
1
s
2s
n + ns n + n2s
x2
2n
y xy
4 4
1
1 1
1 1 1
n
(1) 2 1 1
n
(1) 2 1 1
2(1)s 0 0
Construction notes for D2n : these tables were constructed in Example 4.15 and Section 4
Exercise 1.