Representation Theory
Representation Theory
Konstantin Ardakov
October 2020
Notation 1.1. Throughout this course, k will denote a field, and all vector spaces
are understood to be k-vector spaces. Recall that the general linear group of a
k-vector space V is the group GL(V ) of invertible k-linear transformations V → V
with the operation of composition.
Definition 1.2. Let G be a finite group and let V be a finite dimensional vector
space over k. A representation of G on V is a group homomorphism
ρ : G → GL(V ).
The degree of the representation is dim V .
If we need to be more precise about the ground field of scalars, we will say “ρ is
a k-representation of G”. Here are some examples from Geometry.
Example 1.3.
(a) The cyclic group G = hgi of order 2 acts on V = k by negation: ρ(g) = −1
gives a representation of G of degree 1.
(b) Let G be the symmetry group of a triangle: G = D6 = {e, r, r2 , s, sr, sr2 } and
let k = R. Then G acts by R-linear transformations on the plane V = R2 . This
gives rise to a representation
ρ : G → GL(V )
where ρ(r) is the rotation through 2π/3 and ρ(s) is the reflection in the y-axis.
(c) The symmetry group of the regular n-gon G = D2n acts on V = R2 by R-linear
transformations, giving a natural representation of G of degree 2.
(d) Let k = R, let X ⊂ R3 be the set of vertices of a cube centred at the origin,
and let G be the stabiliser of X in the rotation group SO3 (R). Then from your
Prelims M1 course you know that G is isomorphic to the symmetric group S4 .
In this way, we obtain a degree 3 representation
ρ : S4 → GL(R3 )
of this symmetric group.
Definition 1.4. Let X be a finite set. The free vector space on X, is the set
( )
X
kX := ax x : ax ∈ X
x∈X
Definition 1.5. Let X be a finite set equipped with a left action of the finite group
G. Each g ∈ G defines a permutation ρ(g) : X → X, given by ρ(g)(x) = g · x. This
permutation extends uniquely to an invertible linear map ρ(g) : kX → kX:
!
X X
ρ(g) ax x = ax g · x.
x∈X x∈X
Since g · (h · x) = (gh) · x for all g, h ∈ G and all x ∈ X one checks easily that
ρ(g)ρ(h) = ρ(gh) in GL(kX) for all g, h ∈ G. Thus ρ : G → GL(kX) is a represen-
tation, called the permutation representation (associated with X).
Example 1.6. Let k = Q be the field of rational numbers and let F be a finite field
extension of k (so that dimk F < ∞). Let G be the group of all field automorphisms
of F . Then every g ∈ G is a k-linear map g : F → F , and one of the main results
of Galois Theory tells us that G is a finite group. Then the inclusion G ,→ GL(F )
gives a k-linear representation of G of degree dimk F .
√ 2πi
For example, F could be the splitting field Q( 2, e 3 ) of the polynomial x3 − 2.
Then G is the symmetric group S3 and this gives a Q-linear representation ρ : S3 →
GL(F ) of degree 6.
0 1 0
Similarly, one can calculate
0 1 0
ρX ((12)) = 1 0 0 .
0 0 1
Warning : in this course, permutations in Sn are multiplied from right to left,
in contrast to the conventions in Prelims M1. Thus, for example, we have in S3
(123) · (12) = (13).
This way of writing and multiplying permutations is necessary to ensure that the
natural action of Sn on {1, · · · , n} given by σ · n = σ(n) is a left action: it satisfies
the axiom g · (h · x) = (gh) · x as opposed to a right action (x · g) · h = x · (gh).
such that
σ(g) ◦ ϕ = ϕ ◦ ρ(g) for all g ∈ G.
We say that ϕ is an isomorphism if it is bijective.
Example 1.17. Suppose that k is the finite field Fp and let G = hgi be the cyclic
group of order p. Let ρ : G → GL2 (k) be the matrix representation given by
!
i 1 i
ρ(g ) = for each i ∈ Z.
0 1
( ! !)
1 0
Let v1 = , v2 = be the standard basis for V = k 2 . Then U := hv1 i
0 1
is a G-stable subspace, because ρ(g i )(v1 ) = v1 for all i. The subrepresentation
ρU : G → GL(U ) and quotient representation ρV /U : G → GL(V /U ) in this case
are both trivial. However of course ρ itself is not trivial!
Recall from Linear Algebra that this means that U + W = V and U ∩ W = {0}.
Theorem 1.21 (Maschke). Let G be a finite group and suppose that |G| = 6 0 in
k. Let U be a G-stable subspace of a finite-dimensional G-representation V . Then
U admits at least one G-stable complement W in V .
6
Proof. By picking a basis for U and extending it to a basis for V we can find some
linear complement Z for U in V :
V = U ⊕ Z.
This Z will certainly not be G-stable in general. We will now use an averaging
argument to replace Z with a G-stable one. To this end, let π : V → V be the
projection map along the decomposition V = U ⊕ Z, so that
here, we use the hypothesis that the number |G| is invertible in the field k. We
now check directly that ϕ is a homomorphism of representations. To do this, it is
helpful to use the following notation:
In this finite sum, make the substitution y −1 = x−1 g. As x runs over all elements
of G, so does y. Since x is then equal to gy, we obtain
X X
|G|ϕ(g · v) = (gy) · π(y −1 · v) = g · y · π(y −1 · v) = g · |G|ϕ(v).
y∈G y∈G
Remark 1.22. Maschke’s Theorem fails if the characteristic of our ground field k
happens to divide |G|, see Example 1.17.
7
Corollary 1.24. Let G be a finite group and suppose that char(k) - |G|. Then
every finite dimensional representation ρ : G → GL(V ) of G is completely reducible.
Proof. Proceed by induction on dim V , the case dim V = 0 being true by defini-
tion. Let U1 be a G-stable non-zero subspace of V of smallest possible dimension.
Clearly U1 is irreducible. Then U1 admits a G-stable complement W by Maschke’s
Theorem, Theorem 1.21. Now dim W < dim V so by induction, W = U2 ⊕ · · · ⊕ Um
for some G-stable irreducible subspaces U2 , · · · , Um . Hence V = U1 ⊕ U2 ⊕ · · · ⊕ Um
is also completely reducible.
In algebraic geometry and algebraic number theory, most rings one encounters
are commutative. However in this course rings are most definitely not commutative
in general!
Definition 2.2. Let G be a finite group. The group ring of G (with coefficients in
k) is the vector space kG from Definition 1.5, with the following multiplication:
! !
X X X X
ax x · by y = ax bx−1 g g.
x∈G y∈G g∈G x∈G
8
If in addition the ground field k contains a primitive n-th root of unity ζ, then
the polynomial T n − 1 factors into a product of distinct linear factors (T − 1)(T −
ζ)(T − ζ 2 ) · · · (T − ζ n−1 ). Then the Chinese Remainder Theorem implies that
kG ∼
= k[T ]/hT n − 1i ∼
=k × k × ··· × k.
| {z }
n times
Next, recall the following important definition from Part A Rings and Modules.
Important Convention: From now on, we will use the bijective correspon-
dence provided by Proposition 2.5 to freely convert a representation ρ : G → GL(V )
into the corresponding kG-module V , and vice versa. We will also sometimes say
“V is a representation of G” when we actually mean “V is a kG-module”.
Example 2.7. Let A be a ring. The free A-module of rank 1 is the abelian group
A equipped with the left-multiplication action of A:
a · b = ab for all a, b ∈ A.
Definition 2.9. Let A be a ring. We say that A is semisimple if the free A-module
of rank 1 is completely reducible.
Example 2.10. Let G be a finite group such that |G| 6= 0 in k. Then the group
ring kG is semisimple.
10
Lemma 2.13. Let V, W be simple A-modules. Then every non-zero A-linear map
ϕ : V → W is an isomorphism.
Proposition 2.14. Let A be a semisimple ring. Then A has only finitely many
simple A-modules, up to isomorphism.
Theorem 2.15. Let G be a finite group such that |G| 6= 0 in k. Then G has only
finitely many irreducible representations, up to isomorphism.
Definition 2.16. For a finite group G, we will write rk (G) to denote the number
of isomorphism classes of irreducible k-representations of G.
A big motivating question for us will be: how do we compute rC (G) effectively?
Definition 3.2. Let A be a ring and let V be an A-module. The endomorphism ring
of V , EndA (V ), is the set of all A-module homomorphisms ϕ : V → V , equipped
with pointwise addition of homomorphisms, and composition as multiplication.
Definition 3.3. Let A be a ring. The opposite ring to A, Aop , has A as the
underlying abelian group, but with the following new multiplication:
EndA (V ) = k1V .
12
Definition 3.7. Let A be a k-algebra and let V be an A-module with EndA (V ) = k1V .
Then by Theorem 3.6, every element z ∈ Z(A) must act on V by a scalar, which
we denote by zV . The ring homomorphism
Z(A) → k, z 7→ zV
is called the central character of V .
Using these tools, we will now focus on semisimple rings. Recall from Proposition
2.14 that such a ring admits only finitely many simple modules up to isomorphism.
Notation 3.8. Until the end of §3, A will denote a fixed semisimple ring, and
V1 , · · · , Vr will denote a complete list of representatives for the isomorphism classes
of simple A-modules. We also fix a decomposition
ni
r M
M
(3.1) A= Li,j
i=1 j=1
of the A-module A into a direct sum of simple left ideals Li,j , where Li,j ∼
= Vi for
each i and j.
Proof. By Schur’s Lemma, Theorem 3.6, we have EndA (Vi ) = k1Vi for all i =
1, · · · , r. So we can define a k-linear map ψ : Z(A) → k r by ψ(z) := (zV1 , zV2 , · · · , zVr ).
Suppose that ψ(z) = 0 for some z ∈ Z(A); then zVi = 0 for all i; we will show that
z = 0. Consider the decomposition of 1 ∈ A along the decomposition (3.1):
ni
r X
X
1= ei,j for some ei,j ∈ Li,j .
i=1 j=1
P ni
r P ni
r P
P
Then z = z1 = zei,j = zVi ei,j . But zVi = 0 for all i, so z = 0. So, ψ
i=1 j=1 i=1 j=1
r
is injective and dim Z(A) 6 dim k = r.
ni
L
Definition 3.10. For each i = 1, . . . , r, define Bi := Li,j .
j=1
13
Proof. It follows from (3.1) that A = B1 ⊕ · · · ⊕ Br and that each Bi is a left ideal
of A; it will therefore be enough to show that each Bi is also a right ideal in A.
Fix a ∈ A and consider Li,j ⊆ Bi . Let i0 6= i and 1 6 j 0 6 ni0 be another
pair of indices, and consider the projection ϕ : A Li0 ,j 0 along our decomposition
(3.1). The restriction of ϕ ◦ ra : A → Li0 ,j 0 to Li,j is an A-module homomorphism
from Li,j to Li0 ,j 0 . Because i0 6= i, these two modules are not isomorphic, so this
restriction must be the zero map by Lemma 2.13. Varying i0 and j 0 , we see that the
projection of Li,j a onto each Bi0 with i0 6= i is equal to zero. Therefore Li,j a ⊆ Bi .
But since Bi is equal to the sum of all of the Li,j , Bi a ⊆ Bi for all a ∈ A.
We have now related the number of isomorphism classes of simple modules over
our semisimple k-algebra to the centre of A, which motivates the question of cal-
culating dim Z(kG). Computing the centre of group rings is very easy.
Definition 3.14. For a finite group G, let s(G) denote the number of conjugacy
classes of G.
14
Then {C
c1 , · · · , C
cs } is a basis for Z(kG) as a vector space, so
Corollary 3.16. Let G be a finite group, and let k be an algebraically closed field
with |G| =
6 0 in k. Then rk (G) = s(G).
We were able to obtain this striking result without really understanding the
internal structure of the ring kG properly. With a little additional effort, this
structure will become completely transparent at least in the case where the ground
field k is algebraically closed and where Maschke’s Theorem applies.
We continue with the notation established at (3.8) above.
Lemma 3.17.
(a) Each Bi is a ring with identity element ei .
(b) A is isomorphic to the product of the rings (Bi , ei ):
A∼
= B1 × · · · × Br .
(c) Each Bi is itself a semisimple ring, with unique simple module Vi .
Proof. (a) Lemma 3.11 implies that Bi is an additive subring of A, stable under
multiplication. We saw in the proof of Lemma 3.12 that for any a ∈ A, aei = ei a is
the Bi -component of a along the decomposition A = B1 ⊕· · ·⊕Br . So aei = ei a = a
for all a ∈ Bi which says that ei is an identity element in Bi .
(b) The isomorphism sends a ∈ A to (ae1 , · · · , aer ) ∈ B1 × · · · × Br .
(c) Fix ` = 1, . . . , ni and suppose that U is a Bi -submodule of Li,` . Then
Mr
A·U = Bj · U 6 U
j=1
Proposition 3.18. Let B be a semisimple ring with exactly one simple module
n times
z }| {
∼
V , up to isomorphism. Suppose that B = V ⊕ · · · ⊕ V as a left B-module, and let
D := EndB (V ). Then there is a ring isomorphism
B∼
= Mn (Dop ).
Proof. By Lemma 3.17, we may assume that r = 1, so that A has exactly one
simple module V up to isomorphism. Then A ∼
= Mn (Dop ) where D := EndA (V ),
by Lemma 3.18. But D ∼= k by Schur’s Lemma, Theorem 3.6.
4. Multilinear algebra
Lemma 4.1. Let V be a vector space and let G × V → V be a G-action on the set
V . Then this extends to a kG-module structure on V if and only if the G-action
on V is linear, which means that the following condition holds:
g · (v + λw) = (g · v) + λ(g · w) for all g ∈ G, v, w ∈ V, λ ∈ k.
We will usually identify V and W with their images {(v, 0) : v ∈ V } and {(0, w) :
w ∈ W } inside V × W , respectively. The sum of these images is all of V × W and
their intersection is {(0, 0)}, so this notation V ⊕ W is consistent with the one
you’ve met in Linear Algebra.
Definition 4.6. Let V and W be two vector spaces, let {v1 , · · · , vm } be a basis
for V and let {w1 , · · · , wn } be a basis for W . The tensor product of V and W ,
V ⊗ W,
is the free vector space (see Definition 1.4) on the set of formal symbols
{vi ⊗ wj : 1 6 i 6 m, 1 6 j 6 n}.
m
P n
P
If v = λi vi and w = µj wj are elements of V and W respectively, we define
i=1 j=1
the elementary tensor
m X
X n
(4.1) v ⊗ w := λi µj (vi ⊗ wj ) ∈ V ⊗ W
i=1 j=1
Remarks 4.7.
(a) From the definition we see that dim V ⊗ W = (dim V )(dim W ).
(b) The elementary tensors span V ⊗ W .
(c) Not every element of V ⊗ W is an elementary tensor v ⊗ w.
It may appear that this definition depends on the choice of bases for V and W .
To address this, we have the following
X 0 := {vi0 ⊗ wj0 : 1 6 i 6 m, 1 6 j 6 n}
is a basis for V ⊗ W .
⊗ : V × W → V ⊗ W, (v, w) 7→ v ⊗ w
In fact, this map is universal with respect to this property, in the following sense.
18
g · (v ⊗ w) := (g · v) ⊗ (g · w) for all g ∈ G, v ∈ V, w ∈ W.
α : V ∗ ⊗ W → Hom(V, W )
Proof. (a) Let S2 := hσi be the cyclic group of order 2. Since char(k) 6= 2, the group
ring kS2 admits orthogonal idempotents e1 := 1+σ 2 ∈ kS2 and e2 := 1−σ 2 ∈ kS2 ,
which give rise to the ideal decomposition from Lemma 3.11
kS2 = kS2 e1 ⊕ kS2 e2 = ke1 ⊕ ke2 .
It follows that every kS2 -module M admits an even-odd decomposition
M = e1 M ⊕ e2 M = {m ∈ M : σm = m} ⊕ {m ∈ M : σm = −m}.
Now S2 acts linearly on V ⊗ V as follows:
σ · (v ⊗ w) = w ⊗ v for all v, w ∈ V.
20
Using similar ideas, it is possible to find proper kG-submodules of the tensor cube
V ⊗ V ⊗ V , one for each irreducible representation of kS3 . Similarly we obtain a
decomposition of the nth tensor power V ⊗n of V as a direct sum of kG-submodules
S λ (V ), one for each irreducible representation λ of the symmetric group Sn . This
construction V 7→ S λ (V ) is called a Schur functor.
21
5. Character theory
Recall from Corollary 3.16 that in fact r = rC (G) is equal to s = s(G), so the
character table is always square. If a representation ρ is known, computing its
character is usually straightforward.
Example 5.5.
(a) The character table of the cyclic group of order 3, G = {1, x, x2 } is
1 x x2
1 1 1 1
χ 1 ω ω2
χ2 1 ω2 ω
22
2πi
where ω := e 3 is a primitive cube root of unity.
(b) Let G = S3 . In addition to the trivial character, we have the sign character
: S3 → {±1} ⊂ C× , defined by
(
1 if σ is even,
(σ) =
−1 if σ is odd.
We also have the two-dimensional irreducible representation W of S3 from
Example 1.20. Since 12 + 12 + 22 = 6 = |S3 |, we have found all the characters,
and the character table of S3 is
1 (123) (12)
1 1 1 1
1 1 −1
χW 2 −1 0
We record some basic facts about characters that we used in Example 5.5.
Characters of degree 1 are called linear characters. We have the following im-
portant consequence of our study of the group ring CG from §3.
Proof. Suppose that the simple kG-module Vi affords the character χi . Then
χi (1) = dim Vi by Lemma 5.6(a); now use Corollary 3.20(b).
Definition 5.8. Let N be a normal subgroup of the finite group G and let ρ :
G/N → GL(V ) be a representation. The inflated representation of G
ρ̇ : G → GL(V )
Definition 5.9. Let G be a finite group. The derived subgroup, G0 , is the subgroup
of G generated by all commutators [x, y] := xyx−1 y −1 in G:
G0 := hxyx−1 y −1 : x, y ∈ Gi.
23
Example 5.11. Let G = A4 be the alternating group of order 12. We know that
A4 has a normal subgroup of order 4, called the Klein four-group
V4 := {1, (12)(34), (14)(23), (13)(24)}.
Since A4 /V4 has order 3, it must be a cyclic group of order 3, hence abelian, so
A04 6 V4 and |A04 | ∈ {1, 2, 4}. No subgroup of order 2 in V4 is normal in A4 . So, A04
has to be V4 : it cannot be the trivial group since A4 is non-abelian.
Conclusion: A4 admits 3 distinct linear characters, inflated from A4 /V4 ∼ = C3 .
Definition 5.12. Let G be a finite group. The inner product on class functions
h−, −i : C(G) × C(G) → C
is defined as follows:
1 X
hϕ, ψi := ϕ(g)ψ(g)
|G|
g∈G
It is routine to verify that this is indeed a complex inner product on C(G), which
means that the following properties are satisfied for all ϕ, ψ ∈ C(G) and λ ∈ C:
• hλϕ, ψi = λhϕ, ψi and hϕ, λψi = λhϕ, ψi
• h−, −i is additive in both variables,
• hϕ, ψi = hψ, ϕi,
• hϕ, ϕi ≥ 0 with equality if and only if ϕ = 0.
We can now state our next big theorem.
χV = a1 χ1 + · · · + ar χr
and now Theorem 5.13 tells us that we can recover ai from χV as follows:
r
X r
X
hχi , χV i = hχi , aj χj i = aj δi,j = ai .
j=1 j=1
Proof. We know that hχi , χj i = δi,j by Theorem 5.13, so the χi ’s are pairwise or-
thogonal elements of the inner product space C(G) by Lemma 5.3. Now dim C(G) =
s(G) = rC (G) = r by Corollary 3.16, so {χ1 , · · · , χr } forms a basis for C(G).
Let us see how to use Theorem 5.13 to compute character tables: we can complete
the analysis that we started in Example 5.11 above, but first, some notation.
CG (g) = {x ∈ G : gx = xg}.
Example 5.18. Let G = A4 . Then A04 = V4 and G has 3 distinct linear characters
by Example 5.11. The representatives for the conjugacy classes in A4 are 1, g2 :=
(12)(34), g3 := (123) and g4 := (132). So, A4 has exactly 4 conjugacy classes, and
therefore rC (G) = s(G) = 4 because the character table is square by Corollary 3.16.
25
g 1 g2 g3 g4
G
|g | 1 3 4 4
|CG (g)| 12 4 3 3
χ1 1 1 1 1
χ2 1 1 ω ω2
χ3 1 1 ω2 ω
χ4 d a b c
2πi
where ω = e 3 . Proposition 5.7 tells us that 12 + 12 + 12 + d2 = |G| = 12. Then
Lemma 5.6(a) implies that d = 3 is the only possibility. Now, by Theorem 5.13:
and
and
Solving these linear equations we see that a = −1 and b = c = 0. So, the full
character table of A4 is
g 1 g2 g3 g4
G
|g | 1 3 4 4
|CG (g)| 12 4 3 3
χ1 1 1 1 1
χ2 1 1 ω ω2
χ3 1 1 ω2 ω
χ4 3 −1 0 0.
Proposition 5.20 (Fixed Point Formula). Let G be a finite group and let V be a
finite dimensional CG-module. Then
1 X
dim V G = χV (g) = h1, χV i.
|G|
g∈G
26
1
g ∈ CG. Then ge = eg = e for all g ∈ G so e2 = e; this is
P
Proof. Let e := |G|
g∈G
why e is called the principal idempotent of CG. We have
V = e · V ⊕ (1 − e) · V.
e · V = V G.
Before we give the proof of Theorem 5.13, we need to compute the characters
of the representations obtained through various multilinear constructions from §4.
Observe that the vector space of class functions C(G) is in fact a commutative ring,
via pointwise multiplication of functions:
(a) Let {v1∗ , · · · , vn∗ } be the dual basis for V ∗ relative to {v1 , · · · , vn }. Then
g · vi∗ = λ−1 ∗
i vi for all i = 1, · · · , n.
27
But each λi is a root of unity by Problem Sheet 1, so g · vi∗ = λi vi∗ and hence
n
X
χV ∗ (g) = tr(gV ∗ ) = λi = tr(gV ) = χV (g).
i=1
n Xm n
! m
X X X
χV ⊗W (g) = λi µj = λi µj = χV (g)χW (g).
i=1 j=1 i=1 j=1
Isomorphic representations have the same character. Then (a) and (c) give
χHom(V,W ) = χV ∗ ⊗W = χV ∗ χW = χV χW .
Proof. (a) Let f ∈ Hom(V, W ). Then f is fixed by the G-action if and only if
gW ◦ f = f ◦ gV for all g ∈ G.
Proof of Theorem 5.13. Let V and W be the simple CG-modules whose characters
are ϕ = χV and ψ = χW , respectively. Since V and W are simple, Schur’s Lemma
(Theorem 3.6) and Lemma 2.13 together tell us that
(
1 if V ∼
=W
dim HomCG (V, W ) =
0 if V W.
1 X (dim V )2
hχV , χW i = ||χV ||2 = |χV (g)|2 ≥ >0
|G| |G|
g∈G
r
(
X |CG (g)| if g is conjugate to h,
χi (g)χi (h) =
i=1
0 otherwise.
Since hχi , χk i = δi,k by the Row Orthogonality theorem, Theorem 5.13, this means
that the r × r matrix X := (xi,j ) is unitary:
X · X T = I.
Divide both sides by cj ck : by Lemma 5.17 we have 1/c2j = |G|/|gjG | = |CG (gj )|.
Example 5.24. Let G be the symmetric group S4 . The conjugacy class represen-
tatives are g1 = 1, g2 = (12)(34), g3 = (123), g4 = (12), g5 = (1234), with conjugacy
classes of sizes 1, 3, 8, 6, 6 respectively. We know from Problem Sheet 0 that V4 is
a normal subgroup of G with S4 /V4 being isomorphic to the symmetric group S3 .
This gives us three irreducible characters 1 e, e
, χ
g W obtained by inflation from S3
using Example 5.5(b):
g 1 g2 g3 g4 g5
G
|g | 1 3 8 6 6
|CG (g)| 24 8 3 4 4
1e 1 1 1 1 1
e 1 1 1 −1 −1
χg W 2 2 −1 0 0
χ4 d4 α4 β4 γ4 δ4
χ5 d5 α5 β5 γ5 δ5
Since rC (G) = s(G) = 5 by Corollary 3.16, we are missing two irreducible char-
acters χ4 and χ5 of degrees d4 and d5 , say. Then d24 + d25 = 24 − 12 − 12 − 22 = 18
by Theorem 5.23 with g = h = 1 (or Corollary 3.20(b)). The only solution to this
30
6. Induced representations
The Row and Column orthogonality theorems, Theorems 5.13 and Theorem 5.23
are powerful tools but they are only good for completing character tables and not for
constructing them. We will now introduce a powerful new technique of constructing
representations of G using representations of its proper subgroups. This technique
is called induction.
We return to the setting of kG-modules over a general field k for now.
Definition 6.1. Let H be a finite group and let V be a kH-module. The vector
space of H-coinvariants VH of V is
VH := V /k{h · v − v : h ∈ H, v ∈ V }.
Corollary 6.3. Let H be a subgroup of the finite group G and let W be a kH-
module. Then the space of H-coinvariants (kG ⊗ W )H of kG ⊗ W with respect to
the right H-action is a kG-module.
IndG
H W := (kG ⊗ W )H
Proof. The first assertion is an easy exercise. By Lemma 4.9, there is a linear map
β : xkH ⊗ W → W given on elementary tensors by β(xh ⊗ w) = h · w for all h ∈ H
and w ∈ W . Now for any y ∈ H we have
β : (xkH ⊗ W )H → W, xh ⊗ w 7→ h · w.
IndG
H W = (x1 ⊗ W ) ⊕ (x2 ⊗ W ) ⊕ · · · ⊕ (xm ⊗ W ).
where xi ⊗ W ∼
= W for each i. Now take dimensions.
In categorical terms (see Part C Category Theory), our next result tells us that
“induction is left-adjoint to restriction”.
33
G
Proof. The map W → ResG H IndH W given by w 7→ 1 ⊗ w is kH-linear. Hence,
given a kG-linear α : IndGH W → U , we can view α as being a kH-linear map
G G G
ResH IndH W → ResH U by restriction, and then precompose it with w 7→ 1 ⊗ w
to obtain the kH-linear map Φ(α). This shows that Φ is well-defined.
We will now construct a map Ψ in the opposite direction. Take some kH-linear
G
map β : W → ResG H U and define Ψ(β) : IndH W → U by setting
To see that this is well-defined, we must show that gh · β(w) = g · β(h · w) for all
g ∈ G, h ∈ H, w ∈ W . But this follows immediately from the hypothesis that β is
kH-linear. We check that Ψ(β) is kG-linear as follows: for all g, x ∈ G, w ∈ W ,
Since everything in sight is linear, this defines the required linear map
G
Ψ : HomkH (W, ResG
H U ) → HomkG (IndH W, U ).
which shows Ψ(Φ(α)) = α, and similarly for any β ∈ HomkH (W, ResG
H U ) we have
Since there are now two groups in play, we will denote the inner product on C(G)
from Definition 5.12 by using a subscript:
Proof. Let U be the CG-module whose character is ψ, and let W be the CH-module
whose character is ϕ. Applying Proposition 5.22 twice, we have
hIndGH ϕ, ψiG = dim HomCG (IndG
H W, U ), and
hϕ, ResG
H ψiH = G
dim HomCH (W, ResH U ).
Now apply Proposition 6.10.
Example 6.13. Let U be a simple CG-module and consider the trivial character
H 1 is the character of the free CG-module of rank 1 by
1 of H := {1}. Then IndG
Example 6.8. Then Frobenius Reciprocity tells us that
Corollary 6.12 suggests that there is linear map in the opposite direction
IndG
H : C(H) → C(G)
and by part (a), ρ(g) permutes the direct summands xi ⊗ W in the same way as g
permutes the numbers {1, · · · , m} (or equivalently, in the same way as g permutes
the left cosets G/H). Writing down the matrix of ρ(g) with respect to a choice
of basis for IndG
H W which respects the direct sum decomposition, we see that the
block matrices on the diagonal corresponding to the indices i that are not fixed by
g are all zero. However if g · i = i then ρ(g) preserves xi ⊗ W , and its restriction to
this subspace has trace equal to the trace of the action of x−1i gxi ∈ H on W .
It turns out that there is a more invariant way of expressing the result of this
calculation, which does not depend on the choice of coset representatives.
Theorem 6.16. Let H be a subgroup of the finite group G and let W be a finite
dimensional CH-module. Then for all g ∈ G we have
1 X ◦ −1
(IndG
H χW )(g) = χW (x gx).
|H|
x∈G
Proof. Fix g ∈ G. Using the notation of Lemma 6.14, we note that i ∈ Fix(g) ⇔
gxi H = xi H ⇔ x−1
i gxi ∈ H. So, we can rewrite Lemma 6.14(b) as follows:
X m
X
(IndG
H χW )(g) = χW (x−1
i gxi ) = χ◦W (x−1
i gxi ).
i∈Fix(g) i=1
Remark 6.17. It could very well happen that g G ∩ H = ∅. In this case, Theorem
6.16 tells us that (IndG
H χW )(g) = 0.
Then IndG G
H : C(H) → C(G) is left adjoint to the map ResH : C(G) → C(H) which
sends ψ ∈ C(G) to its restriction ψ|H : H → C:
hIndG G
H ϕ, ψiG = hϕ, ResH ψiH for all ϕ ∈ C(G), ψ ∈ C(H).
Proof. The formula holds true when ϕ and ψ are characters of representations of
H and G, respectively, by Frobenius Reciprocity (Corollary 6.12) together with
Theorem 6.16. The result follows because characters of representations span class
functions, by Corollary 5.15.
g G ∩ H = hH H
1 ∪ · · · ∪ h` .
χW (x−1 gx) =
P P̀ P
|H|χIndG
H W
(g) = |CG (g)|χW (y)
x∈S i=1 y∈hH
i
|hH
P̀
= |CG (g)| i |χW (hi ).
i=1
ϕx : N → C, h 7→ ϕ(x−1 hx).
hIndG G G G
N ϕ, IndN ϕiG = hResN IndN ϕ, ϕiN .
m
G
But ResG x1
+ · · · + ϕxm by Proposition 6.21, where G =
S
N IndN ϕ = ϕ xi N . We
i=1
x
may assume that x1 = 1; our assumption ϕ 6= ϕ for all x ∈ G\N then implies that
ϕxi 6= ϕ for i ≥ 2. Because each ϕxi is irreducible, hϕ, ϕiN = 1 and hϕxi , ϕiN = 0
for i ≥ 2 by Theorem 5.13. Hence || IndG 2
N ϕ|| = 1 and now we can use Corollary
G
5.15 to deduce that IndN ϕ is irreducible.
Using Corollary 6.22, we can now find all irreducible characters of the dihedral
groups D2n . We only treat the case where n = 2m + 1 is odd, and leave the even
case as an exercise.
Proof. Write G = hr, s : srs−1 = r−1 i and note that N := hri is a normal cyclic
subgroup of G of order 2m + 1. By Lemma 5.10, N has 2m + 1 linear characters
{ϕi : 0 6 i 6 2m} where ϕ : N → C× sends r to some fixed primitive (2m + 1)th
root of unity ζ, say. We calculate
(ϕi )s (rj ) = ϕi (s−1 rj s) = ϕi (r−j ) = ζ −ij = ϕ2m+1−i (rj ) for all i, j = 0, · · · , 2m.
38
Corollary 3.20(b) says that we have found all of the irreducible characters of G.
From Part A Rings and Modules, we know that the set of algebraic numbers is
the union of all subfields of C of finite dimension as a Q-vector space.
Examples 7.2.
(a) Every integer a ∈ Z is an algebraic integer, being a root of t − a = 0.
(b) Every root of unity is an algebraic integer.
(c) If z is an algebraic number, then mz is an algebraic integer for some integer m.
(d) A rational number α ∈ Q is an algebraic integer if and only if α ∈ Z. To see
this, write α = r/s with r, s coprime integers and suppose that
r n r n−1 r
+ an−1 + · · · + a1 + a0 = 0
s s s
for some a0 , a1 , · · · , an−1 ∈ Z. Clearing denominators we obtain the equation
rn + an−1 rn−1 s + · · · + a1 rsn−1 + a0 sn = 0
and therefore s divides rn . Since s and r are coprime, this is only possible if
s = ±1 and then α is an integer.
{z ∈ C : zM ⊆ M } ⊂ A.
for some integers uij ∈ Z. Consider the characteristic polynomial g(t) := det(tI−U )
of the d × d matrix U = (uij ). Since all entries of U are integers, g(t) is a monic
polynomial with integer coefficients. We can rewrite (7.1) as a matrix equation
U v = zv
and consider the abelian subgroup M of C generated by the set {αi β j : 0 6 i <
m, 0 6 j < n}. The monic polynomial equations satisfied by α and β ensure that
αM ⊆ M and βM ⊆ M.
Next we return to representation theory and extract some algebraic integers from
the complex group ring of a finite group.
Lemma 7.5. Let χ be a character of the finite group G. Then χ(g) is an algebraic
integer for all g ∈ G.
Proof. We know that χ(g) is a sum of n-th roots of unity where n is the order of
g. Since these are all algebraic integers, we can now apply Theorem 7.3.
Lemma 7.6. Let G be a finite group and let C1 , · · · , Cr be the conjugacy classes
in G. Let S be the additive subgroup of CG generated by the conjugacy class sums
c1 , · · · , C
C cr . Then S is a subring of Z(CG).
40
for some aijk (z) ∈ C. Now actually aijk (z) = |{(x, y) ∈ Ci × Cj : xy = z}| ∈ N,
which implies that aijk (z) = aijk (z g ) for any z ∈ Ck and any g ∈ G. So aijk :=
aijk (z) does not depend on z. Hence
r
X
C
ci C
cj = ck ∈ S
aijk C for all i, j = 1, · · · , r.
k=1
We will now compute the value of the central characters of simple CG-modules
— see Definition 3.7 — on our conjugacy class sums. These values turn out to be
algebraic integers!
G·v =
|g G |χV (g)
gc ·v for all v ∈ V.
χV (1)
G
Proof. (a) Since V is a simple CG-module and since the conjugacy class sum z := gc
is central in CG, it acts by a scalar zV ∈ C on every simple CG-module by Schur’s
Lemma, Theorem 3.6. Now take the trace of this action to obtain
We will now apply our knowledge of the value of the central characters of irre-
ducible complex G-representations on conjugacy class sums to obtain an interesting
group-theoretic application of representation theory.
We begin the proof by recalling Sylow theory from Part A Group Theory.
Definition 7.10. Let G be a finite group and let p be a prime. Write |G| = pα m
where p - m. A Sylow p-subgroup of G is a subgroup P of G of order pα .
The first step in the proof of Burnside’s Theorem is the following Lemma.
Lemma 7.12. Let G be a group of order pα q β where p, q are distinct primes and
α, β ≥ 1. Let g be a central element of a Sylow p-subgroup P of G. Then |g G | is a
power of q.
We will actually prove the following general result about finite simple groups.
Theorem 7.13. Let G be a finite group and suppose that the size of a conjugacy
class of a non-central element g ∈ G is a power of q. Then G is not a simple group.
On the one hand, because each σ(ζj ) is a root of unity, we have |σ(ζj )| = 1 for all
j = 1, · · · , n and all σ ∈ G. Hence |σ(α)| 6 1 for each σ ∈ G, so
Y
|a| = |σ(α)| 6 1.
σ∈G
On the other hand, clearly a is fixed by the action of G, and Galois Theory tells us
that this means that a ∈ Q(ω)G = Q. By our hypothesis, α is an algebraic integer,
so σ(α) is also an algebraic integer for any σ ∈ G. So, a is again an algebraic integer
by Theorem 7.3. Hence a ∈ Z by Example 7.2(d).
These two facts force a ∈ {−1, 0, 1}. If α 6= 0 then |a| = 1, so |ζ1 + · · · + ζn | = n.
By the solution to Question 5(a) on Problem Sheet 3, α = ζ1 = · · · = ζn .
Proof of Theorem 7.13. The idea will be to examine all of the non-trivial irreducible
representations ρ2 , · · · , ρr of G and show that for at least one of these, ρi : G →
GL(Vi ) say, the linear map ρi (g) is a scalar multiple of the identity: g ∈ ρ−1 ×
i (C ).
Once we have done this, we can consider the following two normal subgroups of G:
ker ρi 6 ρ−1 ×
i (C ).
If all of degrees χi (1) were divisible by q, then we would deduce from Lemma 7.5
and Theorem 7.3 that −1/q ∈ A, which contradicts Example 7.2(d). The same
reasoning shows that χi (g) 6= 0 and q - χi (1) for at least one index i.
Since |g G | is a power of q, χi (1) is coprime to |g G |. By Bezout’s Lemma we can
find integers a, b such that a|g G | + bχi (1) = 1. Hence
|g G |χi (g) χi (g)
a + b χi (g) =
χi (1) χi (1)
43
Now |g G |χi (g)/χi (1) ∈ A by Theorem 7.7(b) and χi (g) ∈ A by Lemma 7.5, and
then Theorem 7.3 then forces χi (g)/χi (1) to also be an algebraic integer. Since it is
nonzero, Lemma 7.14 tells us that the eigenvalues of ρi (g) are all equal to some root
of unity ζ. Then ρi (g) = ζ1Vi is a scalar multiple of the identity as claimed.