Representation Theory Old
Representation Theory Old
Contents
1. Group representations 2
1.1. Introduction 3
1.2. Group representations 5
1.3. Equivalence of representations 9
1.4. Subrepresentations 10
1.5. Direct sums of representations 11
1.6. Workshop Questions 12
1.7. Homework Questions 13
Summary 14
2. CG-modules 14
2.1. Definitions and equivalences of notions 14
2.2. Conventions 17
2.3. Faithful modules 18
2.4. New modules from old ones 19
2.5. Inner products 23
2.6. Maschke’s theorem 26
2.7. Workshop Questions 28
2.8. Homework Questions 29
Summary 30
3. CG-homomorphisms 30
3.1. Definitions 30
3.2. Schur’s lemma 39
3.3. Representations of finite abelian groups 42
3.4. Spaces of CG-homomorphisms 45
3.5. Workshop Questions 52
3.6. Homework Questions 54
Summary 55
4. Irreducible CG-modules and the Group Algebra 55
4.1. Composition factors 55
4.2. Group algebra as CG-module 58
Representation Theory of Finite Groups page 2
The course aims to introduce the basic concepts and results of the classical theory
of complex representations of finite groups. The study of the representations and
characters of a finite group combines two of the major strands of algebra covered
in the second year: linear algebra and groups. A group is an abstract algebraic
structure. A representation of a group gives a way of working with it in terms of
matrices, which are rather more concrete, and have the advantage that they can be
used in computational algebra. Characters condense the information encoded in the
matrices of a group representation.
Our main objectives in this course :
• Representations of finite groups.
• The group algebra of a finite group and its modules.
• Correspondence between modules and representations.
• Module homomorphisms and their spaces.
• Maschke’s theorem, Schur’s lemma and other key results.
• Characters and main results.
For convenience, there is an appendix comprising the essential notions from linear
algebra and group theory that we need.
1. Group representations
Representation theory concerns the study of groups in terms of matrices. The moti-
vation for this approach is that groups are abstract objects, while matrices are rather
more concrete. In other words, it is simpler to work with matrices. Also, matrices
enable us to use computational algebra to find answers to specific questions.
We first set up the notation that we will use throughout, unless otherwise speci-
fied, together with results assumed to be known from MATH103, MATH220 and
MATH225.
Representation Theory of Finite Groups page 3
1.1. Introduction.
Notation 1.1 (Notation and Assumed background).
That is, the j-th column of [ϕ]BB0 is the image of the j-th basis vector of the domain
V in the coordinates of the basis vectors of the range W . Explicitly, if
X X
v= λj ej then ϕ(v) = aij λj fi
1≤j≤n 1≤j≤n
1≤i≤m
Therefore,
This is a bijective correspondence which we will use henceforth, and which is essential
to the study of group representations.
As noted above, the matrix expression of a linear transformation depends on the
choice of a basis for the vector spaces. In particular, if ϕ = IdV and B = B 0 , then
[IdV ]BB = In , where n = dim V . But if B 6= B 0 , then [IdV ]BB0 6= In .
Example 1.2. Let V = C2 and take B = {e1 , e2 } be the standard basis of V and
another basis B 0 = {f1 , f2 } of V with
3 −1
f1 = 3e1 + e2 = and f2 = −e1 =
1 0
Conversely,
0
How do we get the matrix [ϕ]BB0 of ϕ expressed in the basis B 0 ? The answer is
That is,
ϕ(f1 ) =
=
ϕ(f2 ) = giving the matrix
0
[ϕ]BB0 = (cf. Figure 1)
From now on G denotes a finite group and we write 1 for the multiplicative identity
element in G.
In other words, a function ρ : G → GLn (C) is a representation of G if and only if
the following two conditions hold:
(i) for all g ∈ G, the matrix ρ(g) is an invertible linear transformation of the
same vector space of dimension n for some positive integer n, and
(ii) ρ(g)ρ(h) = ρ(gh) for all g, h ∈ G.
From elementary properties of group homomorphisms, if ρ : G → GLn (C) is a repre-
sentation, then
ρ(1) = In and ρ(g −1 ) = ρ(g)−1 for all g ∈ G.
are words in the gj ’s, i.e. the Ri ’s are monomials in the gj ’s. For instance a cyclic
group G of order n is presented as
G = hg | g n = 1i
while a dihedral group D2m of order 2m admits the presentations
D2m = ha, b | am = b2 = 1 , ba = a−1 bi = hs, t | s2 = t2 = 1 , (st)m = 1i
In particular a presentation of a group is not unique. However a function ϕ : G →
GLn (C) is a group homomorphism if and only if
Ri (ϕ(g1 ), . . . , ϕ(gd )) = 1
whatever the presentation.
Example 1.4.
(i) For any group G and any n ∈ N, the function ρ : G → GLn (C) defined by
ρ(g) = In for all g ∈ G is a representation of G, since for all g, h ∈ G we have
In particular, ρ(1) = In , while if e1 is the first basis vector of V , then for any g ∈ G,
the first column of ρ(g) has a 1 in the g-th row and 0 elsewhere.
Representation Theory of Finite Groups page 8
These two extreme examples of permutation representations show the crucial facts
about representation theory of groups: on the one hand the regular representation
gives an exhaustive list of a group in terms of matrices, but this becomes cumbersome
for large “complicated” groups (e.g. a large nonabelian finite simple group). On the
other hand, the trivial representation of a group is very easy to handle for computing,
but does not allow us to gather much information about the group structure, as all
finite groups “look” alike. This leads us to the following concept.
Definition 1.9. A representation ρ : G → GLn (C) of G is faithful if ker(ρ) = {1}.
In particular, the regular representation of any group is faithful, while the trivial
representation is faithful if and only if the group is trivial.
The fundamental isomorphism theorem for groups asserts that if φ : G → H is a
group homomorphism, then G/ ker(φ) ∼= im(φ).
Proposition 1.10. A representation ρ of G is faithful if and only if im(ρ) ∼
= G.
Proof.
Representation Theory of Finite Groups page 9
and we calculate
0
[Id]BB0 [ρ(a)]B [Id]BB =
T =
we conclude that ρ ∼ τ .
1.4. Subrepresentations.
Definition 1.14. Let ρ : G → GLn (C) be a representation of G and write V = Cn
for the underlying vector space. A subspace W ⊆ V is G-invariant if ρ(g)w ∈ W for
all w ∈ W and all g ∈ G.
If W is a G-invariant subspace of V , the representation ρ defines a subrepresentation
of G as follows. Let C = {e1 , . . . , em } be a basis for W , with m ≤ n. Put
ρW : G −→ GLm (C) , ρW (g) = [ρW (g)]C = (agij ) for all g ∈ G
where
X
ρW (g)ej = ρ(g)ej = agij ei for coefficients agij ∈ C, ∀ g ∈ G.
1≤i≤m
and similarly for the powers of a, which is immediate from the matrix [τ (a)]B0 .
Definition 1.15. A representation ρ : G → GLn (C) of G of degree n is irreducible
if n > 0 and ρ does not possess any proper nonzero subrepresentation. If ρ is not
irreducible, then we say that ρ is reducible.
where the matrix is expressed in the basis {e1 , e2 , e3 } of G. Then ρ is reducible since
the subspace W = hf1 = e1 + e2 + e3 i is G-invariant. We have
that is,
ρW : G → GL1 (C) ∼
= C× , ρW (a) = 1 is the trivial representation of G.
2πi
Similarly, let ζ = e 3 ∈ C be a primitive cube root of 1, and put f2 = e1 + ζ 2 e2 + ζe3
and f3 = e1 + ζe2 + ζ 2 e3 . Then
ρ(a)(f2 ) =
ρ(a)(f3 ) =
So hf2 i and hf3 i are G-invariant subspaces too. In the basis B 0 = {f1 , f2 , f3 } of C3 ,
the above calculation shows that
[ρ(a)]B0 =
Definition 1.17. Let ρj : G → GLnj (C) for j = 1, 2. The direct sum of ρ1 and ρ2 is
the representation
ρ1 (g) 0
ρ1 ⊕ ρ2 : G → GLn1 +n2 (C) , (ρ1 ⊕ ρ2 )(g) =
0 ρ2 (g)
Representation Theory of Finite Groups page 12
For instance, in Example 1.16, ρ is the direct sum of three irreducible representations
of degree 1.
Summary.
The new notions studied in this section are:
• representations of a finite group G
• special representations: regular, trivial, permutation, faithful
• equivalence of representations
• subrepresentations
• direct sums of representations
2. CG-modules
In this section we introduce the concept of a CG-module, and we show that CG-
modules and representations of G are equivalent notions.
We assume throughout that modules are finite dimensional C-vector spaces, and
that G acts on the left. In some textbooks, the modules are allowed to be infinite
dimensional vector spaces over any field, or the group G acts on the right (i.e. vg
Representation Theory of Finite Groups page 15
The above gives the method to obtain a representation of G from a CG-module, using
matrices. Conversely, suppose that ρ : G → GLn (C) is a representation of G of
degree n. Set V = Cn for the space of column vectors and define
gv = ρ(g)v for all v ∈ V and all g ∈ G.
The left-hand side of the equality denotes the group action, whereas the right-hand
side is a matrix multiplication. It is straightforward that this action of G makes the
vector space V into a CG-module. So
given a representation of G of degree n = dim V , we get a CG-module V .
The statements in the two boxes above say that CG-modules and representations of G
are equivalent notions, and make the equivalence explicit, i.e. provide the dictionary
CG-modules ↔ Representations of G.
Definition 2.2. If a representation ρ and a CG-module V are in such a correspon-
dence, we say that ρ affords (or corresponds to) V .
Notation 2.3. For convenience of notation, if we consider a CG-module V spanned
as vector space by nonzero vectors, v1 , . . . , vn ∈ V , for some n ∈ N and some finite
group G, we simply write V = hv1 , . . . , vn i.
Example 2.4. Let G = ha | a4 = 1i be a cyclic group of order 4.
Representation Theory of Finite Groups page 16
2.2. Conventions.
Let V be a vector space and G a finite group with a function G × V → V . We want
to find a minimum set of conditions to check in order to determine whether V is a
CG-module or not.
Suppose that G has a presentation
G = hg1 , . . . , gr | R1 = · · · = Rt = 1i where Ri = Ri (g1 , . . . , gr ) are the relations,
and that B = {v1 , . . . , vn } is a basis of V . Then V has a structure of CG-module,
not unique in general, if there exists a well-defined G-action on V . That is, if we can
define a product
gi vj for all 1 ≤ i ≤ r and all 1 ≤ j ≤ n
subject to the following conditions for all the indices:
(i) gi vj ∈ V
(ii) Ri (g1 , . . . , gr )vj = vj for all 1 ≤ i ≤ t. In particular, if gi has order ni , i.e.
gini = 1, then
gi · (gi · · · (gi ·vj )·) = vj = 1vj
| {z }
ni terms
We now translate in terms of modules some of the concepts defined for representations.
From a historical perspective, group representations using group homomorphisms
appeared earlier than modules, but the modern approach of representation theory of
finite groups is done via CG-modules instead. Also, for computational reasons, these
latter are much easier to handle, especially when looking for specific answers about a
given finite group.
Representation Theory of Finite Groups page 18
For completeness, the above definition can be somewhat misleading, because “the”
permutation CG-module is not unique. It depends on how we identify G with a sub-
group of the symmetric group. In particular there may be several possible symmetric
groups. The following examples show this fact. There is no canonical way to define
“the” permutation CG-module in general, unless for cyclic groups or full symmetric
groups. A cyclic group of order n is identified with the subgroup generated by the
n-cycle (1 2 . . . n) ∈ Sn unless otherwise specified.
Example 2.8.
(i) Let G = hg | g 4 = 1i ∼
= C4 . The permutation CG-module is the 4-dimensional
vector space V with natural basis B = {v1 , v2 , v3 , v4 } on which the action of
g is given by the matrix
[g]B =
Proof. We need to show that for any h ∈ ker(V ) and g ∈ G we have ghg −1 ∈ ker(V ).
That is, for each v ∈ V , we need to check that
We calculate
where (1) and (3) hold by associativity of the G-action on V and (2) holds by as-
sumption on h ∈ ker(V ).
Example 2.11.
(i) Let G = hg | g 2 = 1i ∼
= C2 and let V = hv1 , v2 i the CG-module defined by
gv1 = v2 and gv2 = v1
Then ker(V ) = and so V is
(ii) Let G = hg, h | g 2 = h2 = 1 , gh = hgi ∼
= C2 × C2 and let V = hv1 , v2 i the
CG-module defined by
gv1 = hv1 = v2 , gv2 = hv2 = v1
Then ker(V ) = and so V is
Note that the condition for a CG-module to be faithful concerns all v ∈ V . It is quite
possible for a nonidentity element of G to fix some vectors in V . We will see such
instances in the exercises.
2.4.1. CG-submodules.
Definition 2.12. Let V be a CG-module. A subset U of V is a CG-submodule of V ,
or simply a submodule of V , if U is a G-invariant subspace of V , that is, U is a
subspace of V such that gu ∈ U for all g ∈ G and all u ∈ U .
where the coefficients in the columns of A ∈ Mm×(n−m) (C) and B ∈ GL(n−m) (C),
stacked on top of each other, are the coordinates of gvm+1 , . . . , gvn in the basis B 0 .
Conversely, if a given CG-module V has a basis B such that each matrix [g]B has the
above form, then the subspace W spanned by the first m vectors of the basis B is
mapped to itself by multiplication by any [g]B . That is W is a CG-submodule of V ,
saying also that V is reducible.
In other words, V is reducible if and only if there is a basis B such that each matrix
[g]B has the above form, with a zero block in the bottom left corner.
Recall from linear algebra that a vector space V is a direct sum V1 ⊕ · · · ⊕ Vn if and
only if each element v ∈ V can be written in a unique way as a linear combination
v = v1 + · · · + vn with vj ∈ Vj for all 1 ≤ j ≤ n.
In other words, the following two conditions hold:
By definition
au2 =
au4 =
Thus, V1 is a CG-submodule of V , since a generates the entire group G. Similarly,
we have
implying that V1 + V2 = V .
Note that the inner product of two elements of V is a complex number. Thus in
general it is meaningless to write (v, w) ≥ 0, since inequalities of complex numbers
do not make sense. However, taking w = v in axiom (i) shows that (v, v) = (v, v). So
(v, v) is a real number for any v ∈ V , and axiom (iii) makes sense.
It is immediate that if some vector space V is equipped with the standard inner
product, then V ⊥ = {0}. Moreover, if U is a subspace of V , the equality V = U ⊕ U ⊥
shows that
dim V = dim U + dim U ⊥ (cf. MATH220)
Example 2.20. Let V = C3 with basis {v1 , v2 , v3 }. The standard inner product on
V is the function
!
X X X
V ×V →C , λj vj , µk vk = λj µj
1≤j≤3 1≤k≤3 1≤j≤3
We now prove that Definition 2.22 makes sense, provided G is finite (because of the
sum!).
Proof. We need to check the axioms of Definition 2.19. For (i), we note that for all
v, w ∈ V and all g ∈ G, we have
X
hv, wi = (gv, gw) =
g∈G
hu + v, wi =
=
and hλv, wi =
Thus (ii) holds. Finally, for axiom (iii), we note that for any v ∈ V , we have
with equality if and only if each summand (gv, gv) is zero, i.e. if and only if gv = 0
for all g ∈ G. Since 1v = v we must have v = 0. Therefore h , i is an inner product
as claimed.
For the last claim, let g ∈ G and v, w ∈ V . We have
hgv, gwi =
where (1) holds because we suppose that h , i is G-invariant and (2) holds because
U is a CG-module and so for any g ∈ G and any v ∈ U .
Example 2.24. Let G = S3 and V the permutation module for G with natural basis
{v1 , v2 , v3 }. Take the standard inner product ( , ) on V and use it to define the inner
product h , i of Definition 2.22. Then hvi , vj i = for all 1 ≤ i, j ≤ 3 since
|G| = 6 and (gvi , gvj ) = 1 if i = j and 0 otherwise.
Let u = v1 + v2 + v3 . Then U = hui is a trivial CG-submodule of V , since
for all g ∈ G. We have
hu, ui =
So v ∈ U ⊥ if and only if
Therefore W = U ⊥ = .
Representation Theory of Finite Groups page 27
Proof. By Proposition 2.23, the CG-module U ⊥ , with respect to the inner product
h , i defined using the standard inner product ( , ) on V , is the required direct sum
complement of U in V .
(iii) Suppose that G is not abelian and that V (i.e. ρ) is faithful. Prove the
following
(a) There exists no basis B of V for which all the matrices [ρ(g)]B are
simultaneously diagonal.
(b) Suppose that [ρ(g)]B is scalar for some g ∈ G. Prove that g ∈ Z(G),
where Z(G) = {z ∈ G | xz = zx , ∀ x ∈ G} is the centre of G.
Exercise 2.7.6. Let G = Sn and V the permutation CG-module with natural basis
{v1 , . . . , vn }. Find ker(V ).
Summary.
The notions studied in this section are:
• CG-modules
• CG-submodules
• quotients of CGmodules
• direct sums of CG-modules
• correspondence between modules and representations
• special and modules: regular, trivial, permutation, faithful
• irreducible and completely reducible CG-modules
• Maschke’s theorem
3. CG-homomorphisms
3.1. Definitions.
In this section we introduce and study the notion of homomorphism as it applies to
CG-modules.
Definition 3.1. Let G be a finite group and V, W two CG-modules. A function
ϕ : V → W is a CG-homomorphism if ϕ satisfies the following two conditions:
Representation Theory of Finite Groups page 31
Using linear algebra techniques, it turns out that ϕ is a linear transformation if and
only if ϕ can be written as the n × m matrix [ϕ] with the j-th column being the
column vector of the P coordinates of ϕ(vj ) in the basis {w1 , . . . , wn }. Then, the image
of a vector v = j aj vj ∈ V is the vector of W whose coordinates in the basis
{w1 , . . . , wn } are the coefficients of the column vector resulting from the product
a1
[ϕ] =
[ϕ] · v =
It is often easier to work with matrices. For this, choose bases B = {e1 , . . . , en } of V
and B 0 = {f1 , . . . , fm } of W . Then ϕ is given by the matrix
A = [ϕ]BB0 = (aij ) ∈ Mm×n (C) (recall the notation [ϕ]domain
range )
in which the j-th column is ϕ(ej ) written as linear combination of the fi ’s, for all
1 ≤ j ≤ n. That is, the coefficients
m
X
aij are defined by ϕ(ej ) = aij fi for all 1 ≤ j ≤ n.
i=1
Then if the set {g1 , . . . , gr } generates G, the linear transformation ϕ given by the
matrix A is a CG-homomorphism if and only if
A[gi ]B = [gi ]B0 A for all 1 ≤ i ≤ r.
Example 3.3. Let us review Example 3.2 in terms of matrices. Using the basis
B = {v1 , v2 } of V as basis for domain and range, the linear transformation ϕ is given
by the matrix [ϕ]B = The matrix of the action of g on V , using the same
[ϕ]B [g]B =
[g]B [ϕ]B =
Non-example.
Representation Theory of Finite Groups page 33
[ϕ][g] = whereas
[g][ϕ] =
For convenience, if the bases of domain and range are “obvious” and no confusion is
possible, we simply write [g], respectively [ϕ] for the matrix of the action of g on V ,
respectively a linear transformation ϕ, as in the Non-example above.
The well-known notions of kernel and image of a linear transformation generalise to
CG-modules in the natural way as follows.
Definition 3.4. Let ϕ : V → W be a CG-homomorphism.
(i) The kernel of ϕ is the subspace
ker(ϕ) = {v ∈ V | ϕ(v) = 0} of V .
(ii) The image of ϕ is the subspace
im(ϕ) = {ϕ(v) | v ∈ V } of W .
Proposition 3.5. Let V and W be CG-modules and let ϕ : V → W be a CG-
homomorphism. The following hold.
(i) ker(ϕ) is a CG-submodule of V .
(ii) im(ϕ) is a CG-submodule of W .
(iii) ϕ is injective if and only if ker(ϕ) = {0}.
(iv) ϕ is surjective if and only if im(ϕ) = W .
Proof. We know from MATH220 that ker(ϕ) is a subspace of V and im(ϕ) a subspace
of W . So to show the first two parts we only need to prove that gv ∈ ker(ϕ) for all
g ∈ G and v ∈ ker(ϕ), and gw ∈ im(ϕ) for all g ∈ G and w ∈ im(ϕ). Pick any
v ∈ ker(ϕ) and g ∈ G. We have
(∗)
ϕ(gv) =
where (∗) holds because ϕ is a CG-homomorphism. Therefore gv ∈ ker(ϕ) as required.
Now pick any w ∈ im(ϕ) and g ∈ G. So by definition of im(ϕ), there exists v ∈ V
with ϕ(v) = w. The equality
gw =
Representation Theory of Finite Groups page 34
giving ker(ψ) =
[πU ]B = since
im(πU ) =
Finally, ker(πU ) is the eigenspace of [πU ] for the eigenvalue 0, that is,
ker(πU ) =
πj : V → V by πj (v) = uj .
Then each πj is a CG-homomorphism.
Definition 3.9. Let V be a CG-module with CG-submodules U1 , . . . , Ur such that
V = ⊕ri=1 Ui . For each 1 ≤ j ≤ r, the CG-homomorphism πj : V → V mapping
πj (v) = uj is called the projection of V onto Uj .
Remark 3.10. These maps πj are called projections for obvious reasons. In linear
algebra, given a vector space V and any subspace U of V , we may choose a basis BV
of V which contains a basis BU of U . Thus, any vector v ∈ V can be written as a sum
v = u + w, where u ∈ U and w lies in a complement of U in V . The projection of V
onto U is the linear transformation πU : V → V which sends each v = u + w ∈ V to
u. (E.g. in the complex plane C2 , we have two obvious projection maps π1 , π2 , where
π1 (z1 , z2 ) = (z1 , 0) and π2 (z1 , z2 ) = (0, z2 ).) If moreover U and W are CG-modules,
Proposition 3.8 asserts that the projection maps are CG-homomorphisms.
Proof. Since each πj is a linear map, the only thing left to prove is that the equality
πj (gv)P= gπj (v) holds for all g ∈P G, all v ∈ V , and all j. Given g ∈ G and
r r
v = u
i=1 i ∈ V , we have gv = i=1 gui . Moreover gui ∈ Ui for all i because
each Ui is a CG-submodule of V . So πj (gv) = proving that πj is a
CG-homomorphism.
v1 = and v2 =
The projections onto U and W can be expressed in terms of matrices, using the basis
B = {u, w} of V . Then
Observe that [πU ]2 = [πU ]. This means that πU is the identity map on U and the zero
map on W . (Similarly for W and πW .)
We now bring together the notions of isomorphic CG-modules and equivalent repre-
sentations of G for a finite group G.
Theorem 3.13. Let G be a finite group and ρ, σ : G → GLn (C) two representations
of G of degree n. Let V and W be the CG-modules of dimension n afforded by ρ and
σ respectively, say in the bases B = {v1 , . . . , vn } of V and B 0 = {w1 , . . . , wn } of W .
Then V and W are isomorphic if and only if ρ and σ are equivalent, i.e.
V ∼
= W ⇐⇒ ρ ∼ σ
and these give the coordinates of the j-th columns of the matrices of ρ(g) and σ(g),
expressed in the bases B and B 0 ,
[ρ(g)]B = (aij ) and [σ(g)]B0 = (bij ) for all g ∈ G.
By definition ρ ∼ σ if and only if there exists T ∈ GLn (C) such that, for all g ∈ G,
T −1 [ρ(g)]B T = [σ(g)]B0 ⇐⇒
(1)
⇐⇒
and
We end this section with the version of the isomorphism theorem as it applies to
CG-modules.
Theorem 3.15. Let V and W be CG-modules and ϕ : V → W a CG-homomor-
phism. Then there exists a CG-submodule U of V such that
V = ker(ϕ) ⊕ U and U ∼ = im(ϕ)
∼
showing that w ∈ im(ϕ). Therefore ϕ : U
= / im(ϕ) is a CG-isomorphism.
For instance, the identity map and the zero map, respectively
IdV : V → V , IdV (v) = v and 0 : V → V , 0(v) = 0 for all v ∈ V
are linear tranformations of V for any vector space V . Moreover IdV has a unique
eigenvalue, namely 1, and every nonzero v ∈ V is an eigenvector for IdV . Similarly 0
is the unique eigenvalue of 0 and every nonzero vector is an eigenvector.
By the fundamental theorem of algebra, any non constant polynomial with coefficients
in C has a root in C and so any linear transformation of V has an eigenvalue, because
its characteristic polynomial splits, i.e. factorises into a product of linear factors:
f (x) = an xn + an−1 xn−1 + · · · + a1 x + a0 = an (x − α1 ) · · · (x − αn ) ∈ C[x]
where α1 , . . . , αn ∈ C are the roots of f (x), not necessarily all distinct.
Representation Theory of Finite Groups page 40
Proof. (i) Suppose that ϕ 6= 0, that is, there exists v ∈ V with ϕ(v) 6= 0. Then {0} =
6
im(ϕ) is a CG-submodule of W . Since W is irreducible we must have im(ϕ) = W ,
i.e. ϕ is surjective.
On the other hand, ker(ϕ) 6= V because ϕ 6= 0. Since ker(ϕ) is a CG-submodule
of V and V is irreducible, we must have ker(ϕ) = {0}, i.e. ϕ is injective. So ϕ is
invertible, and therefore ϕ is a CG-isomorphism.
(ii) As recalled above, the linear transformation ϕ has an eigenvalue, say λ ∈ C, and
so For any g ∈ G and v ∈ V , the equality
Drawing upon Lemma 3.17 (ii), we obtain the following very useful criteria for testing
the irreducibility of CG-modules.
Proposition 3.18. If V is a non-zero CG-module such that every CG-homomorphism
from V to V is a scalar multiple of IdV , then V is irreducible.
Proof. We prove the contrapositive, that is, we suppose that V is reducible, and we
need to show that there exists a CG-endomorphism of V which is not a scalar multiple
of IdV . Since V is reducible, we can choose a proper nonzero CG-submodule U of V .
By Maschke’s theorem 2.25, there is a CG-submodule W of V satisfying V = .
So we can write v = for some unique elements u ∈ U and w ∈ W .
By Proposition 3.8, the projection map πU : V → V onto U , defined by πU (u+w) = u
for all u ∈ U and w ∈ W , is a CG-endomorphism of V , and πU is not a scalar multiple
of IdV . Indeed, πU (u) = u = Id(u) for all u ∈ U , whereas πU (w) = 0 = 0 Id(w) for all
w ∈ W.
Thus we have proved that if V is reducible, then there are CG-endomorphisms of V
which are not a scalar multiple of IdV .
Representation Theory of Finite Groups page 41
We now give the matrix version of Schur’s lemma 3.17 and Proposition 3.18.
Recall that a square matrix is scalar if it is a multiple of the identity matrix, i.e. of
the form λIn for some λ ∈ C and n ∈ N. For instance,
2 0 2 0
is scalar, but is not scalar.
0 2 0 −1
Scalar matrices have the property that they commute with any other square matrix
of same size, that is
(λIn )A = A(λIn ) for all A ∈ Mn (C), all λ ∈ C and all n ∈ N.
Conversely a square matrix which commutes with any matrix of same size must be
a scalar matrix. (Scalar matrices form the centre of the ring Mn (C).) Now, a linear
map is a scalar multiple of the identity if and only if its matrix form is scalar and
this does not depend on the choice of a basis.
Corollary 3.19. Let ρ : G → GLn (C) be a representation of a finite group G. Then
ρ is irreducible if and only if every matrix A ∈ Mn (C) which satisfies
Aρ(g) = ρ(g)A for all g ∈ G is scalar.
Since the matrix commutes with ρ(g), Corollary 3.19 proves that
ρ is (See also Proposition 3.21 below.)
(ii) Let G = ha, b | a4 = b2 = 1 , ba = a−1 bi be a dihedral group of order 8 and
σ : G → GL2 (C) the representation of G defined by
0 1 1 0
σ(a) = and σ(b) =
−1 0 0 −1
Assume that the matrix
α β
A= for some α, β, γ, δ ∈ C
γ δ
Representation Theory of Finite Groups page 42
commutes with both σ(a) and σ(b). From the equalities Aσ(b) = σ(b)A and
Aσ(a) = σ(a)A, we obtain and , respectively. So
A=
c(λ) =
We then calculate the eigenspaces for both eigenvalues i and −i. These are the
irreducible submodules of V .
Representation Theory of Finite Groups page 43
Then V = Ei ⊕ E−i as vector space, by MATH220. But we also have guj ∈ Ej for
any uj ∈ Ej for j = ±i since E±i are the eigenspaces for the linear transformation
ρ(g) which defines the G-action on V . Therefore E±i are CG-submodules of V .
Since dim E±i = 1, the above is a direct sum decomposition of V into irreducible
submodules.
Proof. The above argument shows that any irreducible representation of G has the
form ρλ1 ,...,λr .
Conversely, if λj is an nj -th root of unity for all 1 ≤ j ≤ r, then the map ρ : G →
GL1 (C) given by
In terms of modules, Theorem 3.23 asserts that for any finite abelian group G as
above, there are exactly |G| distinct isomorphism classes of irreducible CG-modules,
all of dimension 1. Accordingly to the notation in the theorem, we call each of these
Vλ1 ,...,λr . The action of any element g1d1 · · · grdr ∈ G on Vλ1 ,...,λr is as follows:
Representation Theory of Finite Groups page 44
2πi
Example 3.24. Let G = hg | g n = 1i be a cyclic group of order n, and put ζ = e n
for a primitive n-th root of 1 in C. Here primitive means that n is the smallest
positive integer for which ζ n = 1. Theorem 3.23 asserts that there are n irreducible
representations of G, and they are all the ρζ j for 0 ≤ j ≤ n − 1, where
V0 = hv0 i, . . . , V5 = hv5 i
G × Vj → Vj ,
2 3
identification and the equalities λ = ζ and λ = −1, we get the correspondence table
(i.e. the isomorphisms, in technical terms):
Representation Theory of Finite Groups page 45
By assumption, dim Vi = 1 for all i. Suppose that {vi } is a basis of Vi for all i, so
that B = {v1 , . . . , vr } is a basis of V . Since each Vi is a CG-submodule of V , we
have gvi ∈ Vi for all g ∈ G. Thus each matrix [g]B is As matrices
(i.e. AB = BA for any diagonal matrices A and B), we have
Proof. We know that (ϕ + ϕ0 ) and (λϕ) are linear transformations since L(V, W ) is a
vector space. Also, for all v ∈ V ,
(1)
(ϕ + ϕ0 )(v) = ϕ(v) + ϕ0 (v) = ϕ0 (v) + ϕ(v) = (ϕ0 + ϕ)(v)
where (1) holds because W is a vector space, hence an abelian group. So addition
is a commutative binary operation on the set of CG-homomorphisms V → W . The
additive identity element is the zero map 0 : V → W and the additive inverse of ϕ is
the CG-homomorphism
−ϕ : V → W where (−ϕ)(v) = −ϕ(v).
Now, for all g ∈ G and all v ∈ V , we have by definition
(ϕ + ϕ0 )(gv) =
(λϕ)(gv) =
Definition 3.29. Let V and W be CG-modules. The vector space of all CG-
homomorphisms ϕ : V → W is written HomCG (V, W ).
Our focus in this section is on the vector space HomCG (V, W ), which of course depends
on V and W . Thanks to Maschke’s theorem 2.25, the strategy consists in “gluing
together” CG-homomorphisms between irreducible submodules of V and W . Then,
the key argument is given by Schur’s lemma 3.17. In particular, as we shall see below,
the dimension of the Hom-space “counts”, in some sense, the number of isomorphic
irreducible submodules of V and W .
We begin with a consequence of Schur’s lemma (Theorem 3.17), which gives the
dimension of this space if V and W are both irreducible.
1 if V ∼
=W
dim(HomCG (V, W )) =
0 if V W
Proof. The proposition is immediate from Schur’s lemma 3.17. Indeed, since V and W
are irreducible, any CG-homomorphism V → W is: zero if and only if or a
CG-isomorphism if and only if So suppose that V ∼= W and let ϕ : V → W
−1
be a CG-isomorphism. For any φ ∈ HomCG (V, W ) we have ϕ φ ∈ HomCG (V, V ). By
Theorem 3.17 (ii) there exists λ ∈ C such that ϕ−1 φ = Then φ = and
so we have
HomCG (V, W ) =
From irreducible modules to arbitrary ones, the result about the dimension of the
vector space HomCG (V, W ) uses the complete reducibility of CG-modules and Propo-
sition 3.30. We first prove a useful result.
(i) The vector spaces HomCG (V, W1 ⊕ W2 ) and HomCG (V, W1 ) ⊕ HomCG (V, W2 )
are isomorphic. In particular,
(ii) The vector spaces HomCG (V1 ⊕ V2 , W ) and HomCG (V1 , W ) ⊕ HomCG (V2 , W )
are isomorphic. In particular,
Proof. We only prove the first part, as the second is very similar, and so left as an
exercise. Consider the diagram:
π1 , π2 are the projection maps and
: W2T where
ϕ2
σ1 , σ2 the natural inclusions.
σ2 π2
ϕ
V / W1 ⊕ W2
J
σ1 π1
ϕ1
$
W1
that is, σ1 : W1 → W1 ⊕ W2 , σ1 (w) = for all w ∈ W1 , and similarly
σ2 : W2 → W1 ⊕ W2 , σ2 (w) = for all w ∈ W2 . Note that π1 , π2 , σ1 , σ2 are
CG-homomorphisms. Moreover,
πi σ j = , i, j = 1, 2 and σ1 π1 + σ2 π2 =
where δij = 0 whenever i 6= j and δii = 1. We prove that the following function f is
an invertible linear transformation:
f : HomCG (V, W1 ⊕ W2) −→ HomCG (V, W1 ) ⊕ HomCG (V, W2 )
ϕ : V → W1 ⊕ W2 7−→
For short, put H = HomCG (V, W1 ⊕ W2 ) and Hj = HomCG (V, Wj ) for j = 1, 2. Note
that f is well defined, as composition of CG-homomorphisms is a CG-homomorphism
and domains and ranges match up. Also, it is immediate from the axioms defining
linear transformations that f is a linear transformation, that is,
f (ϕ + ϕ0 ) = f (ϕ) + f (ϕ0 ) and f (λϕ) = λf (ϕ) for all ϕ, ϕ0 ∈ H and all λ ∈ C.
To show that f is bijective, we find its inverse. Put
g : H1 ⊕ H2 −→ H
, where for each v ∈ V ,
(ϕ1 , ϕ2 ) 7−→ σ1 ϕ1 + σ2 ϕ2
(σ1 ϕ1 + σ2 ϕ2 )(v) = . (It may be useful to track the
=
and similarly,
f g(ϕ1 , ϕ2 )(v) =
=
=
Therefore f is an invertible linear transformation (with inverse g), proving that H ∼
=
H1 ⊕ H2 as vector spaces.
Representation Theory of Finite Groups page 48
In the case when all Vi and Wj are irreducible, Proposition 3.30 and Corollary 3.32
enable us to calculate dim(HomCG (V, W )) for any CG-modules V and W .
Corollary 3.33. Let V be a CG-module which can be written as a direct sum V =
⊕ri=1 Ui of irreducible CG-submodules Ui , and let W be any irreducible CG-module.
Then
dim(HomCG (V, W )) = dim(HomCG (W, V ))
and this number is equal to the number of terms Ui with Ui ∼
= W.
dim HomCG (W, V ) =
In the next section, we will return to this result and its applications. We end this
section with a collection of thoroughly worked out examples.
2πi
Example 3.34. Let G = hg | g 3 = 1i be a group of order 3 and put ζ = e 3 ∈ C
for a primitive cube root of 1. (Recall that 1 + ζ + ζ 2 = 0 ans that ζ 2 = ζ −1 .) Let
V = hv1 , v2 , v3 i and W = hw1 , w2 i be the CG-modules for the G-action given by the
matrices
1 0 0
ζ 0
[g]V = 0 ζ 0 and [g]W =
0 1
0 0 ζ2
with respect to the given bases {v1 , v2 , v3 } of V and {w1 , w2 } of W . Let Vj = hvj i for
1 ≤ j ≤ 3 and Wk = hwk i for k = 1, 2. Then
V = V1 ⊕ V2 ⊕ V3 and W = W1 ⊕ W2
with V1 and W2 CG-modules and V2 ∼ = W1 the irreducible CG-module on
which g acts by multiplication by . In particular, V1 ∼
= W2 and V2 ∼
= W1 are the only
isomorphic pairs of irreducible submodules of V and W . So Corollary 3.33 implies
that
dim(HomCG (V, W )) = dim(HomCG (W, V )) = dim(HomCG (W, W )) =
and dim(HomCG (V, V )) = .
Representation Theory of Finite Groups page 49
More precisely, we can find a basis of the above vector spaces by “matching up”
isomorphic pairs. For example, we work out a basis of HomCG (V, W ) as follows: we
only have isomorphisms
θ1 θ2
∼ ∼
V1
= / W2 , v1 7→ w2 , and V2
= / W1 , v2 7→ w1
A CG-homomorphism V → W is given by a matrix, where the (k, l) coeffi-
cient represents a CG-homomorphism Vl → Wk . Thus Corollary 3.32 says that to
find a basis of HomCG (V, W ) we “glue” together the bases of the nontrivial spaces
HomCG (Vk , Wl ). In our case, we get the basis
and finally,
HomCG (W, W ) =
Next example shows the two different approaches possible when looking for bases of
Hom-spaces.
Example 3.35. Let G = hg | g 2 = 1i, and V, W, X the CG-modules of dimension
2, 1 and 4 respectively, say
V = hv1 , v2 i , W = hwi and X = hx1 , x2 , x3 , x4 i
with G-action
1 0 0 0
0 1 0 −1 0 0
[g]V = , [g]W = (−1) and [g]X =
1 0 0 0 0 1
0 0 1 0
Find a bases of HomCG (V, W ), of HomCG (V, V ) and of HomCG (X, V ).
Solution. We have two methods. We find bases of HomCG (V, W ) and HomCG (V, V )
using both methods, and use the second method to find a basis of HomCG (X, V ).
method 1: direct calculation
a b
Consider first HomCG (V, V ). We want to find all A = such that
c d
which shows that any such A must be of the form , i.e. a basis of HomCG (V, W )
is the set
Similarly for HomCG (V, W ), we note that W ∼ = E−1 via w 7→ u2 . So a basis for
HomCG (V, W ) with respect to the bases {u1 , u2 } of V and {w} of W is the set
Finally, we want to find a basis of HomCG (X, V ). We know that V = E1 ⊕ E−1 .
We find the eigenspaces of [g]X which gives a direct sum decomposition of X into
irreducible submodules. The characteristic polynomial of [g]X is
We calculate the eigenspaces F±1 . These are immediate from the previous computa-
tions:
F1 = and F−1 =
So, the “matching picture” of all CG-homomorphisms X → V is as follows,
indicating that dim HomCG (X, V ) = . In terms of matrices with respect to the bases
of eigenvectors
We see from this last example that if one of the spaces is “large”, here dim X = 4,
then it is more convenient to decompose it into a direct sum of irreducible submodules
and then match up the isomorphic irreducible modules.
We end with another example, slightly more complicated because the group is not
abelian and because the modules are not given as direct sum of irreducible submod-
ules. This last example therefore encompasses various difficulties.
Example 3.36. Let G = S3 be the symmetric group of degree 3, and let V and W
be the permutation CG-module and the trivial CG-module respectively. Find bases
of HomCG (V, W ) and of HomCG (W, V ).
since
Because G = h(1 2), (1 3)i these two matrices determine the G-action on any CG-
module. The block diagonal form of both matrices shows that V = U0 ⊕ U1 . As
dim U0 = 1, we know that U0 is irreducible. We prove that U1 is also irreducible using
Corollary 3.19. Consider the ordered basis {u1 , u2 } of U1 . The matrices of a and b
are respectively
A= ,B =
x y
We want to prove that a matrix M = ∈ M2 (C) such that AM = M A and
z t
BM = M B is necessarily scalar. The equations AM = M A and BM = M B give
If we take the natural basis of V instead then θ = (x, y z) ∈ HomCG (V, W ) if and
only
(i) Check that the G-action on V is well-defined, that is, the function g 7→ [g]B
is a group homomorphism G → GL3 (C).
(ii) Decompose V into a direct sum of irreducible submodules.
(iii) Find a basis for EndCG (V ).
Summary. The notions studied in this section summarise the basic properties of
CG-homomorphisms and some applications.
• CG-homomorphisms and basic properties;
• isomorphism theorem;
• Schur’s lemma and its corollaries;
• classification of the irreducible representations of finite abelian groups;
• vector spaces of CG-homomorphisms.
In particular, X X
0CG = 0g and 1CG = δ1,g g
g∈G g∈G
where δ1,g = 1 for g = 1 and δ1,g = 0 otherwise.
Moreover CG is also a C-vector space of dimension |G|. The natural basis of CG is
the set G = {g | g ∈ G}.
Example 4.2. Let G = hg | g 3 = 1i. Then CG is the vector space with basis
G = {1, g, g 2 } and consists of all the elements of the form
a1 1 + ag g + ag 2 g 2 for a1 , ag , ag2 ∈ C.
In the ring CG, we calculate for instance for x = 1 + 2g and y = −1 + g + g 2 ,
Representation Theory of Finite Groups page 56
Proof. (i) Let u be a nonzero vector in U . By definition of the direct sum, we can
write for unique vectors ui ∈ Ui for all i. Since u 6= 0, we know
that there is at least one i for which Consider the projection πi : U → U
defined in Definition 3.9, i.e. πi (v1 + · · · + vr ) = vi for all vj ∈ Uj , for 1 ≤ j ≤ r.
In particular, im(πi ) is a submodule of the irreducible CG-module U and
isomorphic to a submodule of the irreducible CG-module The only
possibility is therefore im(πi ) =
(ii) We proceed by induction on the number r of summands U1 , . . . , Ur of V . If r = 1,
then V is so that V = and the assertion holds. Suppose
that r ≥ 2. The induction hypothesis states that if a CG-module X can be written
p q
∼
M M
as direct sums X = Xi = Yj of irreducible CG-submodules Xi and Yj with
i=1 j=1
p < r, then p = q and Xi ∼= Yi for all i up to a permutation of the indices. Let
r s
Ui and suppose that there is an isomorphism V ∼
M M
V = = Wj for some positive
i=1 j=1
integer s and irreducible CG-modules Wj . Note that s ≥ 2 because we assume that
V is By part (i), we have an isomorphism for some 1 ≤ k ≤ r.
Without loss, we re-label the indices of the Ui ’s so that Let
By induction we have
Representation Theory of Finite Groups page 57
Proposition 4.4 shows that every CG-module is characterised by its composition fac-
M r
tors, that is, if V is a CG-module, then we can write V = Ui with each Ui an
i=1
irreducible CG-module, and each Ui is unique up to isomorphism. Moreover the mul-
tiplicity of an irreducible summand isomorphic to Ui is independent of the direct sum
decomposition of V . In other words, if
V = U1 ⊕ · · · ⊕ Ur = W1 ⊕ · · · ⊕ Wr
then for any 1 ≤ j ≤ r
{i | 1 ≤ i ≤ r , Ui ∼
= Uj } = {i | 1 ≤ i ≤ r , Wi ∼
= Uj }
Example 4.6. Let G = hg | g 3 = 1i be a group of order 3, and let V = hv1 , v2 i and
W = hw1 , w2 , w3 i be the CG-modules with G-action
V : gv1 = v2 gv2 = −v1 − v2
W : gw1 = w2 gw2 = w3 gw3 = w1
That is, as matrices expressed in the bases B = {v1 , v2 } of V and B 0 = {w1 , w2 , w3 }
of W we have
2πi
Let ζ = e 3 be a complex primitive cube root of unity. Recall that ζ 3 = 1 6= ζ so that
ζ satisfies the equality ζ 2 + ζ + 1 = 0. This is because ζ is a root of the polynomial
t3 − 1 ∈ C[t] , which factorises as (t − 1)(t2 + t + 1)
Since ζ 6= 1 we must have that ζ is a root of t2 + t + 1.
Let u = v1 − ζv2 ∈ V and x = w1 + ζ 2 w2 + ζw3 ∈ W . Put U = hui and X = hxi. We
calculate
gu =
gx =
saying that u and x are of [g]B and [g]B0 respectively and both for
the eigenvalue ζ. So U and X are isomorphic CG-submodules of V and W . More
precisely, the linear transformation
ϕ : U →X defined by ϕ(u) = x is a CG-isomorphism.
Since U and X have dimension they are both Therefore V and W
have a common composition factor.
Representation Theory of Finite Groups page 58
Using Schur’s lemma 3.17, we can detect common composition factors in two given
modules using homomorphisms.
Proposition 4.7. Let V and W be CG-modules. Then, V and W have a common
composition factor if and only if HomCG (V, W ) 6= {0}.
r
M s
M
Proof. Write V = Vi and W = Wj with each Vi and each Wj irreducible. By
i=1 j=1
Corollary 3.32,
In particular, X X
0CG = 0g and 1CG = δ1,g g
g∈G g∈G
where δ1,g = 1 for g = 1 and δ1,g = 0 otherwise.
Representation Theory of Finite Groups page 59
Given any finite group G, the regular CG-module is the vector space CG with basis
{G} and G-action induced by left multiplication in G, that is, for each g ∈ G left
multiplication by g is a linear transformation of CG, and so gives a well-defined
G-action:
X X X
µg : CG → CG , µg ( ah h) = ah gh = ag− 1h h
h∈G h∈G h∈G
g(λ1 1 + λ2 g + λ3 g 2 ) =
g 2 (λ1 1 + λ2 g + λ3 g 2 ) =
Taking matrices with respect to the natural basis G of CG gives the regular repre-
sentation of G. Note that all the matrices we get are permutation matrices
1 7→ g 7→ and g 2 7→
It follows that
ϕ(g) =
=
In other words,
ϕ is entirely determined by
More generally:
Definition 4.14. Let G be a finite group and V a CG-module. Then V is cyclic if
there exists v ∈ V such that
V = hvi = {αv | α ∈ CG} = CG · v
Any such v is called a generator of V .
So the above shows that CG is a cyclic CG-module, with generator 1CG . But also the
zero CG-module {0} is cyclic, “generated” by the zero vector 0.
For any cyclic CG-module, say V = hvi, then any CG-homomorphism ϕ : V → W
with domain V is determined by ϕ(v) because
V = hvi = CGhvi = {αv | α ∈ CG}
so that for any CG-homomorphism ϕ : V → W and every u ∈ V , there exists
αu ∈ CG such that u = αu v and by definition of CG-homomorphism we calculate
ϕ(u) =
that is, the image ϕ(v) of v determines the image ϕ(u) of u for all u ∈ V .
The property of a CG-homomorphism being determined by a single value is valid only
for cyclic modules hvi. In particular, as vector space a cyclic module has dimension
at most |G| and any irreducible module is cyclic (cf. exercises) but the converse is
false in general because CG = h1CG i is cyclic but not irreducible in general.
Representation Theory of Finite Groups page 61
The conclusion we draw from the above discussion is the following result.
Proposition 4.15. Let V be a CG-module. Then HomCG (CG, V ) is isomorphic to
V as vector space.
Proof. Define
f : HomCG (CG, V ) −→ V , f (ϕ) = ϕ(1) (where 1 = 1CG )
We have seen above that ϕ(1) ∈ V determines ϕ, in the sense that ϕ(α) =
for all α ∈ CG. By definition, for any ϕ, ψ ∈ HomCG (CG, V ) and any λ ∈ C, we have
f (ϕ + ψ) =
f (λϕ) =
Therefore f is a linear transformation. To prove that f is bijective, define
CG → V
s : V −→ HomCG (CG, V ) , s(v) =
α 7→ αv
We first need to check that s is well defined, i.e. that s(v) is indeed a CG-homomor-
phism. This is immediate from the module axioms, indicated as (∗) over the equalities
below: for any α, β ∈ CG, any λ and any g ∈ G, we have
s(v) (α + β) =
s(v) (λα) =
(saying that s(v) is a linear transformation)
g s(v) (α) =
Similarly, we check that s is a linear transformation, i.e. s(λv + µw) = λs(v) + µs(w),
for all λ, µ ∈ C and all v, w ∈ V , and this too is immediate from the module axioms
(left as exercise). To show that s = f −1 we calculate
f s(v) =
sf (ϕ) =
Therefore f and s are inverse of each other and since they are linear transformations,
we conclude that f : HomCG (CG, V ) → V is an isomorphism of vector spaces.
The main property of the regular CG-module of any finite group is the following.
Theorem 4.16. Let G be a finite group. Any irreducible CG-module is a composition
factor of the regular CG-module.
where the right hand side is a nonzero vector space. So Corollary 3.33 says that there
is an index i such that W ∼= Ui . Therefore W is a composition factor of CG.
The overarching conclusion of our findings is that to get a complete list of all the
representations of a finite group G, it suffices to find all the irreducible CG-modules.
Moreover, all the irreducible CG-modules are composition factors of the regular CG-
module. In practice this is hardly doable for an arbitrary group since we want to find
a basis of CG such that each square |G| × |G| matrix giving the action of a group
element has a block diagonal form. The case of finite abelian groups is manageable,
because we know that square matrices can be put in diagonal form simultaneously.
But difficulties arise already for the simplest nonabelian group, S3 ∼ = D6 .
Example 4.18.
(i) Let G = hg | g 3 = 1i. Then dim CG = 3, with basis {1, g, g 2 }. The matrix of
g ∈ G is
cg (λ) =
We calculate the eigenspaces (calculs abbreviated, gaps to fill in exercise):
−1 0 1
E1 : 1 −1 0 E1 =
0 0 0
−ζ 0 1
Eζ : 1 −ζ 0 Eζ =
0 0 0
2
−ζ 0 1
Eζ 2 : 1 −ζ 2 0 Eζ 2 =
0 0 0
Then CG = E1 ⊕ Eζ ⊕ Eζ 2 and each irreducible CG-module must be isomor-
phic to one of the three composition factors E∗ for ∗ ∈ {1, ζ, ζ 2 }. Then any
CG-module is isomorphic to a direct sum of composition factors.
Representation Theory of Finite Groups page 63
au1 =
=
Therefore the 1-dimensional spaces hvi i and hui i are CA-modules. Moreover
we calculate
bv0 =
Therefore the 2-dimensional spaces
U0 = hv0 , u0 i U3 = hv1 , u2 i and U4 = hv2 , u1 i
are still CA-modules and also CB-modules. In other words U0 , U3 and U4 are
all three CG-submodules of CG because they are simultaneously CA- and
CB-modules and G is generated by a and b.
Next we prove that U3 = hv1 , u2 i is irreducible using Corollary 3.19, i.e. by
finding all the matrices which commute with those of a and b. Let
x y
C= ∈ GL2 (C) and suppose that
z t
ζ 0 ζ 0 0 1 0 1
C = C and that C = C
0 ζ2 0 ζ2 1 0 1 0
ζ 0 0 1
where [a]{v1 ,u2 } = and [b]{v1 ,u2 } =
0 ζ2 1 0
Solving for x, y, z, t these equations gives respectively
Representation Theory of Finite Groups page 64
So far, we have proved that every irreducible CG-module occurs as composition factor
of the regular CG-module. The above examples show that some irreducible modules
appear more than once in CG. Our next objective is to determine how often each
irreducible CG-module occurs as composition factor of CG.
r
M
Theorem 4.19. Suppose that CG = Ui is a decomposition of CG into a direct
i=1
sum of irreducible CG-submodules. Let U be any irreducible CG-module. Then the
number of terms Ui isomorphic to U is equal to dim U .
Proof. Proposition 4.15 states that dim U = dim(HomCG (CG, U )). Corollary 3.33
shows that dim(HomCG (CG, U )) is equal to the number of Ui ’s with Ui ∼
= U.
Example 4.20. In each of the three regular modules in Example 4.18, the theorem
holds.
Representation Theory of Finite Groups page 65
Our last main result establishes a relation between the dimensions of the irreducible
CG-modules of a finite group G and the order of G. Corollary 4.17 asserts that
any irreducible CG-module is a composition factor of the regular CG-module CG,
so that we obtain an exhaustive finite list of all the isomorphism types of irreducible
CG-modules by decomposing the regular CG-module.
Definition 4.21. Let V1 , . . . , Vk be irreducible CG-modules. Assume that no two are
isomorphic and that any irreducible CG-module is isomorphic to some Vi . Then, we
say that {V1 , . . . , Vk } form a complete set of nonisomorphic irreducible CG-modules.
Example 4.22. In example 4.18, a complete set of irreducible CG-modules for the
dihedral group G ∼
= D6 is {U1 , U2 , U3 }. Indeed, since U3 ∼
= U4 , we can write
CG ∼ = U1 ⊕ U2 ⊕ U3 ⊕ U3
Theorem 4.23. Suppose that V1 , . . . , Vk form a complete set of nonisomorphic irre-
ducible CG-modules. Then
X k
dim(Vi )2 = |G|
i=1
Moreover we may assume that V1 is the trivial CG-module of dimension 1.
|G| =
Finally, any group has a trivial representation ρ : G → C× where ρ(g) = 1 for all
g ∈ G. That is, each group has a trivial CG-module. Since G is finite, we can let
* +
X
V1 = v1 = g ⊆ CG be the trivial CG-submodule of CG of dimension 1.
g∈G
Example 4.24.
(i) Let G be a finite abelian group. We have seen in Theorem 3.23 that G has
non-equivalent representations, all of degree That is, in terms of CG-
modules, a complete set of nonisomorphic CG-modules has irreducible
CG-modules, all of dimension
|G| =
(ii) Let G be a dihedral group of order 6 as in Example 4.18. We have seen that
CG ∼= U1 ⊕ U2 ⊕ U3 ⊕ U3
is the direct sum of four irreducible CG-modules: the trivial U1 of dimension
, a nontrivial -dimensional CG-module U2 and two isomorphic copies of a
CG-module U2 of dimension Thus
6 = |G| =
Representation Theory of Finite Groups page 66
Nonisomorphic groups can have the same list of dimensions of a complete set of their
nonisomorphic irreducible CG-modules. In particular, any two abelian groups of same
order will have the same list of dimensions of irreducible modules, namely one.
Now given a nonabelian finite group G, we know that there must be at least one
irreducible module of dimension one. But how many 1-dimensional CG-modules are
there in total? Can we find this number, or bound it, using the group structure of
G? The answer is positive and provided by the next result.
Proposition 4.25. Let G be a finite group and
[G, G] = h[g, h] = ghg −1 h−1 | g, h ∈ Gi
the commutator, or derived, subgroup of G. There are exactly |G/[G, G]| irreducible
CG-modules of dimension 1. Equivalently there are exactly |G/[G, G]| irreducible
representations G → C× of G of degree 1.
Recall that [G, G] is the smallest normal subgroup N of G such that the factor group
G/N is abelian.
Example 4.26. Let G = Sn be the symmetric group of degree n (and order n!),
for some n ≥ 3. Then [G, G] = An is the alternating group of degree n, i.e. the
subgroup of Sn formed by all the even permutations on n letters. Since Sn /An ∼ = C2 ,
any symmetric group of degree at least 3 has exactly two 1-dimensional CG-modules.
One of these is the trivial module, and the other is the signature, i.e.
× 1 if σ ∈ An is even
ε : Sn → C , ε(σ) =
−1 if σ ∈
/ An is odd
(Compare with the dihedral group in Example 4.18 and recall that D6 ∼ = S3 .)
Summary. The notions studied in this section are concerned with the structure of
the group algebra CG of a finite group G. We record in particular the following
concepts and results:
• CG as CG-module and its decomposition;
• composition
Pk factors;
• i=1 dim(Vi )2 = |G|, for any finite group G, where {V1 , . . . , Vn } is a complete
set of nonisomorphic CG-modules.
Representation Theory of Finite Groups page 68
Summary. The notions studied in this section summarise the basic properties of the
group algebra CG of a finite group G.
• composition factors and detection using CG-homomorphisms;
• CG-module structure of CG;
• decomposition of CG and complete set of nonisomorphic CG-modules.
5. Characters
Given an abstract finite group G, we have investigated CG-modules, which roughly
allow us to associate to G a set of matrices in a consistent way reflecting the group
structure and providing us with a way to extract some properties of G. For instance, a
finite group all of whose irreducible modules have dimension 1 must be abelian. In this
section we go a step further in extracting from each matrix a single number: its trace.
This seemingly ingenuous procedure has very useful and enlightening consequences.
Recall the following properties of the trace learned in MATH103 and MATH220: for
any A, B ∈ Mn (C) and λ ∈ C,
(i) Tr(In ) = n,
(ii) Tr(A + B) = Tr(A) + Tr(B),
(iii) Tr(AB) = Tr(BA),
(iv) if A is invertible, then Tr(ABA−1 ) = Tr(B). In particular, if λ1 , . . . , λd are
the distinct eigenvalues of B, with λi of multiplicity mi for each i, then
X
Tr(B) = mi λi
1≤i≤d
The proofs are left in exercise. Let us point out a subtlety of part (iii). Namely,
the trace is not multiplicative, i.e. Tr(AB) 6= Tr(A) Tr(B) for A, B ∈ Mn (C) in
general. Take for instance A = B = I2 . Then Tr(AB) = Tr(I2 ) = 2, whereas
Tr(A) Tr(B) = Tr(I2 )2 = 22 = 4.
Part (iv) says that the trace is an invariant of similitude. That is, if B and B 0 are
similar, then Tr(B) = Tr(B 0 ). Recall that similar precisely means that there exists
an invertible matrix A such that B 0 = ABA−1 .However matrices having the same
2 0
trace need not be similar. For instance I2 and have the same trace but are
0 0
not similar.
Definition 5.1. Let ρ : G → GLn (C) be a representation of degree n ≥ 1 of a finite
group G and V the corresponding CG-module.
Representation Theory of Finite Groups page 69
Using the elementary properties of the trace, we note that characters are invariants
of groups in the following sense.
Proposition 5.4. Let χ : G → C be the character of a finite group G afforded by
some representation ρ : G → GLn (C), and V the corresponding CG-module.
(i) Suppose that g, g 0 ∈ G are conjugate. Then χ(g) = χ(g 0 ). In other words, the
character of a representation takes the same value on all the elements in a
given conjugacy class of G.
(ii) Suppose that ρ0 ∼ ρ. Then χρ0 = χ. That is, equivalent representations have
the same character.
(iii) Suppose that V 0 ∼
= V . Then χV 0 = χ. That is, isomorphic CG-modules have
the same character.
Proof. For part (i) let h ∈ G such that g 0 = hgh−1 . Choosing a basis of Cn , this
equality translates in terms of matrices as:
ρ(g 0 ) = since ρ is a group homomorphism.
Now, by property (iv) of the trace, we have
Tr(ρ(g 0 )) =
For part (ii), suppose that ρ0 ∼ ρ. By definition, this means that there exists T ∈
GLn (C) such that ρ0 (g) = T ρ(g)T −1 for all g ∈ G. As above we have
χρ0 (g) =
The last part is the translation of part (ii) in terms of modules.
Remark 5.5.
(i) Proposition 5.4 (i) says that a character is a class function. The definition of
a class function is precisely a function f : G → C which takes the same value
on all the elements in a conjugacy class of G (and therefore f is determined
by the values it takes on a set of representatives of the conjugacy classes).
Recall that the conjugacy classes of G are all the subset of G of the form
g G = {hgh−1 | h ∈ G} for g ∈ G
In particular, two conjugacy classes are either equal or disjoint. Hence if
{g1 , . .[
. , gc } is a set of representatives of the conjugacy classes of G, then
G= giG is the disjoint union of its distinct conjugacy classes. Thus
1≤i≤c
a character is entirely determined by the images of g1 , . . . , gc .
A well-known theorem asserts that the irreducible characters of a finite group
form a basis of the C-vector space of class functions on G.
(ii) Proposition 5.4 (iii) implies that
there can be at most as many distinct irreducible characters as
there are nonisomorphic irreducible CG-modules.
We will prove below that there are in fact exactly as many irreducible char-
acters as there are nonisomorphic irreducible CG-modules.
Example 5.6. Let G = hg | g 3 = 1i and V the permutation CG-module with natural
basis B = {v1 , v2 , v3 }. Thus
[g]B = , [g 2 ]B =
[g]B0 = , [g 2 ]B0 =
We will come back to such tables shortly. An immediate observation from this exam-
ple is that if G is abelian, then each conjugacy class is reduced to a single element,
and each irreducible character has degree 1. Hence let us take the example of a
non-abelian group.
Example 5.7. Let G = S3 be the symmetric group of degree 3. Put g = (1 2 3) and
h = (1 2). Recall that two elements of the symmetric group Sn are conjugate if and
only if they have the same cycle type. So there are 3 conjugacy classes, represented by
1, g, h, and it follows that any character χ of G is determined by the values χ(1), χ(g)
and χ(h). Let V be the permutation CG-module with natural basis B = {v1 , v2 , v3 }.
We calculate
W1 = , W2 = and W3 =
• the top left vertex of our “triangle” is formed by a “nice set” of matrices, one
for each element of G,
• the top right vertex is the corresponding vector space, essentially some set
Cn on which each g ∈ G acts by a linear transformation, and
• the bottom vertex is a complex function which takes constant values on con-
jugacy classes of elements of G.
These three vertices allow us to use computational algebra in order to study the
structure of an abstract group given by a set of generators and relations.
Representation Theory of Finite Groups page 72
ρ : G → GLn (C) o / V
(3)
g 7→ ρ(g) g · v = ρ(g)v
h 7
( w
χ:G → C
g 7→ Tr(ρ(g))
We now present some results on characters of finite groups, and see what “informa-
tion” about G we can extract from them.
Let ρ : G → GLn (C) be a representation of degree n ≥ 1 and affording the character
χ : G → C. Suppose that ρ is not irreducible, say ρ = ρ1 ⊕ ρ2 with ρj of degree nj ,
and let χj be the character afforded by ρj for j = 1, 2. So n = n1 + n2 and we can
find a basis of Cn such that
ρ1 (g) 0
ρ(g) = for all g ∈ G.
0 ρ2 (g)
In particular,
χ(g) =
Extending from two to a finite number of subrepresentations and using the theorem
of complete reducibility, we have shown the following fact.
Proposition 5.8. Let ρ : G → GLn (C) be a representation of a finite group G. Sup-
pose that ρ = ρ1 ⊕ · · · ⊕ ρt is a decomposition of ρ into irreducible subrepresentations.
Denote χj the character of ρj for each 1 ≤ j ≤ t and χ the character of ρ. Then
X
χ(g) = χj (g) for all g ∈ G.
1≤j≤t
In other words, the character of a representation is the sum of the characters of its
irreducible subrepresentations, counted with their multiplicities.
Therefore, continuing with the above notation for the eigenvalues of ρ(g), all of which
we know to have modulus 1, we calculate
χ(g) =
since ρ(g) has eigenvalues λ1 , . . . , λr if and only if the eigenvalues of ρ(g −1 ) = ρ(g)−1
are
Finally, if g and g −1 are conjugate, then
χ(g) = which implies that
Theorem 5.10. Let G be a finite group and χ : G → C the character of G
afforded by some representation ρ : G → GLn (C) of G. Let V be the corresponding
CG-module.
(i) χ(1) = dim(V ).
(ii) |χ(g)| ≤ χ(1) for all g ∈ G, with equality if and only if ρ(g) = λIn for some
λ ∈ C with |λ| = 1.
(iii) ker(ρ) = {g ∈ G | χ(g) = χ(1)}.
|χ(g)| = as required.
when we regard C as the 2-dimensional real vector space. Since all the λj have
modulus 1, they must be and so ρ(g) = for some complex number λ with
modulus
For part (iii), we have by definition ker(ρ) = {g ∈ G | ρ(g) = In }, and since Tr(In ) =
χ(1), we have an inclusion
ker(ρ) ⊆
Representation Theory of Finite Groups page 74
Conversely, the assumption χ(g) = χ(1) implies that χ(g) is a positive real number
(integer in fact), and by (ii) we must have
the nine irreducible representations has degree 1 and ρ0,0 is the trivial representation.
Take for instance ρ1,2 : G → C× , defined by ρ1,2 (g) = ζ and ρ1,2 (h) = ζ 2 . Let us
calculate χ1,2 :
We observe that |χ1,2 (x)| = 1 and that χ1,2 (x) = χ1,2 (x−1 ) for all x ∈ G, as expected.
In addition, we calculate
ker(χ1,2 ) =
In particular (as we already knew), χ1,2 is not faithful. We also observe that characters
of linear representations (i.e. of degree 1) are essentially “the same”, as 1 × 1 matrices
with complex coefficients are generally identified with their unique coefficient.
Let us pick any other irreducible representation, say ρ2,0 . We calculate:
Here ker(χ2,0 ) =
Because G is abelian, each element is its own conjugacy class. Let us compare χ1,2
and χ2,0 using the inner product of Definition 5.13 below. We define
1 X 1 X
hχ1,2 , χ2,0 i = χ1,2 (g j hk )χ2,0 (g j hk ) = χ1,2 (g j hk )χ2,0 (h−k g −j )
|G| 0≤j,k≤2 |G| 0≤j,k≤2
Representation Theory of Finite Groups page 75
by Corollary 5.9. Using that each element has order 3, i.e ζ 2 = ζ −1 , g 2 = g −1 and
h2 = h−1 we compute (following left to right, top to bottom of χ1,2 ):
1
hχ1,2 , χ2,0 i = χ1,2 (1)χ2,0 (1) + χ1,2 (g)χ2,0 (g 2 ) + χ1,2 (g 2 )χ2,0 (g)+
9
+ χ1,2 (h)χ2,0 (h2 ) + χ1,2 (gh)χ2,0 (g 2 h2 ) + χ1,2 (g 2 h)χ2,0 (gh2 )+
+ χ1,2 (h2 )χ2,0 (h) + χ1,2 (gh2 )χ2,0 (g 2 h) + χ1,2 (g 2 h2 )χ2,0 (gh)+
=
Similarly
hχ1,2 , χ1,2 i =
In the following sections we will explain the equalities obtained in Example 5.3, which
are not coincidences. As a motivation for this scope, let us mention that these equa-
tions are cornerstones in group theory, and they have been key in the early stages
of the classification of finite simple groups in particular. But also characters provide
very efficient tools in demonstrating important properties of finite groups, such as
Burnside’s pa q b theorem, which asserts that any group whose order is the product of
two distinct prime powers is soluble (or solvable).
hχ, χi =
Representation Theory of Finite Groups page 76
(ii) Let G = hg | g 3 = 1i ∼
= C3 and χ the regular character of G. That is,
0 0 1 0 1 0
χ(1) = Tr(I3 ) = 3 , χ(g) = Tr 1 0 0 = 0 , χ(g 2 ) = Tr 0 0 1 = 0
0 1 0 1 0 0
which give
hχ, χi =
Let χ1 be the trivial character of G and χ2 : G → C the linear character
2πi
afforded by ρ2 : G → GL1 (C) with ρ2 (g) = ζ = e 3 . We calculate
hχ1 , χ1 i = as noted above
hχ2 , χ2 i =
hχ1 , χ2 i =
because 1 + ζ + ζ 2 = 0
For the next result, let us recall a very important and useful equation.
Theorem 5.16 (Class equation). Let G be a finite group and g ∈ G. Let
g G = {hgh−1 | h ∈ G} denote the conjugacy class containing g ∈ G and
\
Z(G) = {g ∈ G| gh = hg , ∀ h ∈ G} = CG (g) the centre of G,
g∈G
where CG (g) = {h ∈ G | hg = gh} is the centraliser of an element g in G. Suppose
that g1 , . . . , gl form a set of representatives of the conjugacy classes of G of size greater
than 1. Note that Z(G) is the set of elements of G whose conjugacy class has size 1
and we have
G = Z(G) ∪ g1G ∪ · · · ∪ glG (disjoint union).
Then X
|G| = |Z(G)| + |giG | where |giG | = |G : CG (gi )|
1≤i≤l
|G|
and |G : CG (g)| denotes the index of CG (g) in G, i.e. the number of distinct
|CG (g)|
cosets G/CG (g), for any g ∈ G.
Example 5.17. Let G = S3 . Then G has 3 conjugacy classes represented by ,
s= and t = Their respective sizes are
, |G : CG (s)| = , |G : CG (t)| =
Now Z(G) = {1} by direct computation. We check
where for short we indicate in the indices representatives of the corresponding classes.
Example 5.18. Let G = ha, b | a6 = b2 = 1 , ba = a−1 bi be a dihedral group of
order 12. We calculate that G has 6 conjugacy classes, represented by 1, a, a2 , a3 , b
and ab, of size 1, 2, 2, 1, 3, 3 respectively, and that Z(G) = {1, a3 }. The centralisers
are as follows:
CG (a) = CG (a2 ) = CG (a4 ) = CG (a5 ) = hai ∼
= C6 of index 2,
CG (aj b) = ha3 , aj bi ∼
= C2 × C2 of index 3 for 0 ≤ j ≤ 5.
Representation Theory of Finite Groups page 77
For the second part, let g G = {hgh−1 | h ∈ G} as in Theorem 5.16. So χ(k) = χ(g)
for any k ∈ g G , and k ∈ g G if and only if k −1 ∈ (g −1 )G . It follows that
1 X
hχ1 , χ2 i = χ1 (g)χ2 (g −1 ) =
|G| g∈G
The main result about inner products shows that irreducible characters are orthonor-
mal vectors in the vector space of class functions, i.e. any two distinct irreducible
characters are orthogonal to each other and each one has norm 1. By “norm”, we
mean the norm defined by taking the square root of the inner product of a vector
with itself (as in the euclidean space).
Theorem 5.22 (Orthonormality of irreducible characters). Let χ1 , χ2 be irreducible
characters of a finite group G. Then
1 if χ1 = χ2
hχ1 , χ2 i =
0 otherwise
= as required.
In particular, if χ is the regular character of G, then we may assume that the dis-
tinct composition factors of the regular CG-module are V1 , . . . , Vr with dimensions
a1 , . . . , ar equal to the multiplicity of each Vi as composition factor of CG, by Theo-
rem 4.19. Now Theorem 4.23 asserts that
|G| =
Corollary 5.24. Let χ be a character of a finite group G. Then
χ is irreducible ⇐⇒ hχ, χi = 1
Let us end this section with the answer to the question raised above: How many
distinct irreducible characters does a finite group have?
Indeed, we have seen that if G is a finite group and V, W two irreducible CG-modules,
V ∼= W =⇒ χV = χW in Proposition 5.4.
The difficult assertion is the converse.
Theorem 5.26. Let G be a finite group. Suppose that V, W are irreducible CG-
modules. Then
V ∼= W ⇐⇒ χV = χW
Representation Theory of Finite Groups page 80
χV (1) = χW (1) =
χV (g) = and χW (g) =
χV (g 2 ) = and χW (g 2 ) =
Therefore χV = χW and V ∼ = W (which we already knew).
Much information about the group structure can be recovered from its character table,
and a few exercises below point in that direction. We end the material of the course
with a few examples and observations.
Example 5.30. Let G = hg | g n = 1i ∼ = Cn for some positive integer n. Then G
has n irreducible characters, χ1 , . . . , χn , all of degree 1, and where χk (g j ) = ζ (k−1)j
2πi
and ζ = e n ∈ C. Also, G has n distinct conjugacy classes, each of size 1. So the
character table of G has the form:
1 g ... gj . . . g n−1
χ1 1 1 ... 1 ... 1
j n−1
χ2 1 ζ ... ζ ... ζ
.. .. .. .. ..
. . . ... . ... .
χk 1 ζ k−1 . . . ζ (k−1)j . . . ζ n−k+1
.. .. .. .. ..
. . . ... . ... .
n−1 n−j
χn 1 ζ ... ζ ... ζ
The following result is useful in finding the character table of a finite group G with
a normal subgroup N such that we know the characters of the quotient group G/N ,
e.g. when G/N is abelian.
Proposition 5.32. Let G be a finite group and N a normal subgroup of G. Write
ĝ = gN for the elements of G/N , i.e. the cosets of N in G. Suppose that χ : G/N →
C is a character of G/N . The function
e:G→C
χ , e(g) = χ(ĝ) for all g ∈ G
χ
is a character of G of same degree as χ. Moreover, χ is irreducible if and only if χ
e
is irreducible. Here ĝ denotes the coset gN of G/N which contains g, for any group
element g ∈ G.
A= and that of b is B =
χ3 (b) =
2π
To summarise, the character table of G is, with α = ,
5
Representation Theory of Finite Groups page 84
Summary. The notions studied in this section summarise the basic properties of
characters of finite groups.
• definition and elementary results;
• inner products of characters and orthogonality relations;
• character table of a finite group.
References
[1] G. James, M. Liebeck, Representations and characters of groups, Second edition, Cam-
bridge University Press, New York, 2001.
[2] S. Lang, Algebra, revised third edition, Vol. 211, Springer-Verlag, New York, 2002.
[3] J. Rotman, Advanced modern algebra Prentice Hall, 2002.
Representation Theory of Finite Groups page 85
[4] J.-P. Serre, Linear representations of finite groups, Vol. 42, Springer-Verlag, New York,
1977.
Index
∼
=, 36 identity map, 39
identity matrix, 3
action of a group on a set, 14 image
affording representation, 15 of a CG-homomorphism, 33
algebra, 55 index of a subgroup, 76
inner product, 23
centraliser, 76
G-invariant, 19
centre, 41
invariant of similitude, 68
centre of G, 76
isomorphic CG-modules, 36
CG, 55, 58
CG-submodule, 19 ker(ρ), 8
CG-automorphism, 31 kernel
CG-endomorphism, 31 of a CG-homomorphism, 33
CG-homomorphism, 30 of a module, 19
CG-isomorphism, 31, 36
character linear transformation, 3
of a group, 69 linear extension, 17
afforded by a representation, or a module, linear transformation
69 matrix of a linear transformation, 4
degree, 69
faithful, 74 Mn (C), 3
inner product, 75 matrix
irreducible, 69 change of basis, 4
kernel, 74 inverse, 3
linear, 69 invertible, 3
orthogonality, 75 module
regular, 69 CG-module, 14
trivial, 69 cyclic, 60
character table, 80 irreducible, 19
characteristic polynomial, 39 permutation module, 18
class function, 70 reducible, 19
common composition factor, 57
commutator subgroup, 66 natural basis, 18
complete set of nonisomorphic irreducible natural basis, 55, 59
CG-modules, 65 notation of CG-module, 15
completely reducible module, 27
complex nj -th root of unity, 43 orthogonal basis, 24
complex conjugate, 23 orthogonal complement, 24
composition factor, 57 orthogonal vectors, 24
conjugacy class, 70 orthonormal basis, 24
conjugate symmetry, 23
constituent, 79 permutation matrix, 7
positive definite, 23
derived subgroup, 66 projection map, 36
direct sum of modules, 20
domain, 4 range, 4
regular representation of G, 59
eigenvalue, 39 representation, 5
eigenvector, 39 ρλ1 ,...,λr , 43
equivalence relation, 36 degree of, 5
direct sum, 11
fundamental theorem of algebra, 39 equivalent, 9
faithful, 8
general linear group, 3
irreducible, 10
group algebra, 55, 58
kernel, 8
HomCG (V, W ), 46 permutation, 8
reducible, 10
In , 3 regular, 7
idempotent of CG, 78 trivial, 8
86
Representation Theory of Finite Groups page 87
scalar matrix, 41
similar matrices, 68
subrepresentation, 10
proper, 10
subspace
G-invariant, 10
Tr, 68
trace of a square matrix, 68
word, 6
zero map, 39