You are on page 1of 99

Representation Theory

of Finite Groups
Notes of the course
Istituzioni di Algebra Superiore

Prof. Fernando R. Villegas

Academic Year 2015/2016

Notes edited by
Alessandro Rubin & Alessandro Giacchetto
alrubin@sissa.it
agiacche@mpim-bonn.mpg.de

March 2, 2018
Contents

1 Fundamental concepts of Representation Theory 1


1.1 Representations and actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 One-dimensional representations . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Irreducible representations of S3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Operations on representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.1 Tensor product representation . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.2 Dual representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4.3 Hom(V, W ) representation . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4.4 Symmetric and alternating powers . . . . . . . . . . . . . . . . . . . . . . 12
1.5 Complete reducibility theorem and Schur’s lemma . . . . . . . . . . . . . . . . . 14

2 Character Theory 17
2.1 The character . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Orthogonality relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 The Frobenius divisibility theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3 Introduction to representations of Sn 31
3.1 Tables of S4 and A4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.2 Partitions, Young diagrams and tables of S5 and A5 . . . . . . . . . . . . . . . . 34
3.3 Alternating powers of Sn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

4 Real and quaternionic representations 43


4.1 Hamiltonian quaternions and SU(2) . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.2 Bilinear forms fixed by G . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.3 The Schur indicator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

5 Finite subgroups of SO(3) and SU(2) 55


5.1 Finite subgroups of SO(3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.2 Finite subgroups of SU(2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

6 Induced representations 63
6.1 Definition and characterizations of IndG HV . . . . . . . . . . . . . . . . . . . . . . 63
S4
6.2 Tables of IndS3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.3 The Heisenberg group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

7 Symmetric functions and Pólya’s theory of counting 71


7.1 Bases for the ring of symmetric functions . . . . . . . . . . . . . . . . . . . . . . 73
7.1.1 Elementary symmetric functions . . . . . . . . . . . . . . . . . . . . . . . 73
7.1.2 Complete homogeneous symmetric functions . . . . . . . . . . . . . . . . 75
7.1.3 Powersum symmetric functions . . . . . . . . . . . . . . . . . . . . . . . . 76

i
Contents

7.1.4 Schur functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77


7.1.5 Orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
7.2 Characters of Sn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.3 Pólya’s theory of counting and Burnside lemma . . . . . . . . . . . . . . . . . . . 83

8 Representations of GL2 (Fq ) 89


8.1 Conjugacy classes of GL2 (Fq ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.2 Irreducible representations of GL2 (Fq ) . . . . . . . . . . . . . . . . . . . . . . . . 90

Bibliography 95

ii
Chapter 1

Fundamental concepts of
Representation Theory

In this chapter we will introduce the language of the representation theory and the basic results.
For completeness, we will start reminding some definitions about actions.
Definition 1. An action of a group G on X is a map

α : G × X −→ X

such that
1. α(1G , x) = x for all x ∈ X

2. α(g, α(h, x)) = α(gh, x) for all x ∈ X and g, h ∈ G.


If G acts on X we will write G y X and use the notation α(g, x) = g · x.
Definition 2. Let G y X.
1. For every x ∈ X, we define the stabilizer of x as

Stab(x) = { g ∈ G | g · x = x }

It is easy to see that the stabilizer is a normal subgroup of G.

2. The action is said to be faithful if the stabilizer of every point of X is trivial. In other
words, for all non-trivial elements g ∈ G there exists a non-fixed point x. The action is
free if every stabilizer is trivial.

3. The orbit of x ∈ X is the set

G[x] = { g · x | g ∈ G } .

This is the set of all the elements we can reach applying G to x. The action is said to be
transitive if there exists only one orbit.
Remark 1. Every group G acts on itself by conjugation:

G × G −→ G
(g, x) 7−→ gxg −1

The orbit of x under this action is called conjugacy class of x.

1
1.1. Representations and actions

Proposition 1. Let be G y X. Then the set of all orbits is a partition of X.


Let us now remember for a while how we got Lagrange’s theorem and it is consequences.
Let H be a subgroup of G; there is a bijection between right and left cosets. Furthermore the
collection of the right (left) cosets is a partition of G and everyone is in a 1-1 correspond with
H. If we define the index of H in G as

G
[G : H] = = { number of cosets }
H
we can finally remember the Lagrange’s theorem:
Theorem 2. Let G be a finite group. Then

= |G|

G
H |H|
In particular, this means that |H| divides |G|.
Corollary 3. Let G y X. If G is finite, then

|G[x]| = [G : Stab(x)].

In particular,
|G| = |G[x]| | Stab(x)|.

1.1 Representations and actions


Definition 3. A representation of a group G on X is a group homomorphism

ρ : G −→ Aut(X),

where Aut(X) = { f : X → X | f is bijective }.


Remark 2. In a certain way, a representation is the same thing as an action. Indeed, when we
fix a representation ρ, we naturally obtain an action of G on X:

α : G × X −→ X
(g, x) 7−→ ρ(g)(x).

In fact, using the fact that ρ is an homomorphism, we have that

α(1, x) = ρ(1)(x) = idX (x) = x

α(g, α(h, x)) = ρ(g) (ρ(h)(x)) = ρ(g) ◦ ρ(h) (x)


= ρ(gh)(x) = α(gh, x)
On the other hand, an action α of G on X give us the representation

ρ : G −→ Aut(X)
g 7−→ α(g, ·).

Example 1. If X is finite of n elements, then

ρ : G −→ Sn ' Aut(X).

The above isomorphism is not canonical.

2
1.1. Representations and actions

Example 2. When X = G, we have the regular representation

ρ : G −→ Aut(G)
g 7−→ (x 7→ g · x) .

Definition 4. Let X = V be a K-vector space of finite dimension dim V = n. A linear


representation is a representation ρ on V such that for every g ∈ G, ρ(g) respects the linear
structure, i.e. ρ is a group homomorphism

ρ : G −→ GL(V ).

From now on, we will take G finite and X = V as a finite dimensional C-vector space. In
particular, we will identify, whenever this cause not confusion, V itself as a particular represen-
tations, omitting the map ρ and writing

ρ(g)(x) = g · x.

There is another way to look at a representation, which leads to generalizations of the theory.
Firstly, let us recall the following

Definition 5. Let R be a ring and G a finite group. We set


 
X 
R[G] =

ag g ag ∈ R .
 
g∈G

If we define the following operations

ag g + bg g = (ag + bg )g
X X X

g∈G g∈G g∈G

    

bg g  = ag bh (gh) =
X X X X X
 ag g    ag bh  k.
g∈G g∈G g,h∈G k∈G gh=k

we obtain that (R[G], +, ·) is a ring, called group-ring.


Note that if R is commutative, then R[G] is commutative if and only if G is commutative.

Remark 3. If we have a representation of G on V , we can extend the action from G to C[G].


Indeed, we can define  

ag g  · v =
X X
 ag g · v.
g∈G g∈G

In such a way, V becomes a left C[G]-module. With this remark, Representation Theory can be
seen as the study of particular modules.

Remark 4. To every representation ρ : G → X, X finite, we can associate a linear representa-


tion. Considering the vector space
( )

V = hXi =
X
ax ex ax ∈ C ,


x∈X

3
1.1. Representations and actions

whose dimension is the cardinality of X, we have just to define


!
=
X X
g· ax ex ax eg·x ,
x∈X x∈X

where the dot on the left-hand side defines the linear representation, while dot on the right-hand
side comes from the representation of X. Such linear representation is called permutation
representation.

Thus, we have restricted the study of representations of finite groups to linear representations
only. From now on, we will drop the adjective “linear”, calling every homomorphism ρ : G →
GL(V ) simply representation.

Definition 6. Given two linear representations

ρ : G −→ GL(V ) τ : G −→ GL(W ),

a G-linear map is a linear map ϕ : V −→ W such that the diagram

ϕ
V W

ρ(g) τ (g)

ϕ
V W

is commutative for all g ∈ G (that is, ϕ is C[G]-module map). We set

HomG (V, W ) = { ϕ ∈ Hom(V, W ) | ϕ is a G-linear map } .

Furthermore, we say that ρ and τ are isomorphic representations if ϕ is an isomorphism,


writing (V, ρ) ' (W, τ ) or simply V ' W .

Definition 7. The degree of a representation

ρ : G −→ GL(V )

is defined as dim V .

Definition 8. A subrepresentation of a representation V is a vector subspace U ⊆ V which


is invariant under G, that is
g · U ⊆ U.

A representation V is called irreducible if it has only1 trivial subrepresentations, i.e. (0) and
V itself.

Remark 5. Note that every one-dimensional representation of a group G is irreducible.

Exercise 1. If ϕ : V −→ W is a G-linear map, prove that ker ϕ and Im ϕ are subrepresentations.


1
Every vector space has a lot of subspaces, however thay are not necessarily stable under G.

4
1.2. One-dimensional representations

1.2 One-dimensional representations


We want to study now the one-dimensional representations of a group. Let V be a vector space
of dim V = 1. Clearly
GL (V ) ' C× = C \ { 0 }
and the isomorphism is canonical. Indeed, since every linear map φ ∈ GL(V ) is diagonalizable
in C, φ = λ idC with λ ∈ C× . Thus, φ 7→ λ is the above isomorphism.

Surely, every group G posses the trivial representation, defined as the map

ρ : G −→ C×
g 7−→ 1

Is there any representation? If yes, how many representations there exists? We can assume
without loss of generality that G is abelian: in fact we know that for a general one-dimensional
representation, ρ must factorize as
ρ
G C×
π ρ̃

G
[G, G]

where [G, G] is the commutator subgroup2 , generated by the set


n o
[g, h] = ghg −1 h−1 g, h ∈ G .

This statement follows from the fact that

ρ̃([g]) = ρ(g)

is well defined since C× is abelian. Thus, being a projection π surjective, ρ is equivalent to the
representation ρ̃.
Furthermore we can assume without loss of generality that G is cyclic: indeed for the fun-
damental theorem of finitely generated abelian groups, we have that G is the product of a finite
number of finite cyclic groups
G ' Zm1 × · · · × Zmr
and the representation is determined on every factor. Suppose G = hσi with ord σ = m and let
us prove that the one-dimensional representations of G are m, in 1-1 correspondence with the
mth roots of unity. Every homomorphism defined on a cyclic group is determined by the value
assumed by the generator of the group; furthermore

1 = ρ(σ m ) = ρ(σ)m ,

so that ρ(σ) = ζ must be a mth root of the unity. Eventually, different representations are not
isomorphic: since C× is abelian, for

ρ(g) = ϕ−1 τ (g)ϕ = τ (g)ϕ−1 ϕ = τ (g).

Thus, isomorphic representations ρ and τ are actually equal.


2
We recall that [G, G] is a normal subgroup and G/[G, G] is abelian: actually [G, G] is the smallest normal
subgroup such that the previous quotient is abelian.

5
1.3. Irreducible representations of S3

Example 3. There are always two distinct one-dimensional representations of the permutation
group Sn for n ≥ 3, since in this case
Sn Sn
[Sn , Sn ] = An ' Z2 .

We can find

1. the trivial representation

ρ : Sn −→ C×
σ 7−→ 1,

2. the sign representation

ρ : Sn −→ C×
σ 7−→ sgn(σ).

The sign representation is different from the trivial one, thus they are the two representation we
were looking for.

1.3 Irreducible representations of S3


Here we will underline some properties of the irreducible representations of S3 that will be
debated more in general in Section 1.5. We know that

S3 = hσ, τ | σ 2 = τ 3 = στ στ = idi ,

which can be concretely realized choosing σ = (1 2), τ = (1 2 3). In the previous section, we
have found the trivial and sign representations, denoted U and U 0 respectively. In addition,
according to Remark 4, we have the action of S3 on C3 , that is

πei = eπ(i) π ∈ S3 .

For the generators,

τ (a1 e1 + a2 e2 + a3 e3 ) = a1 e2 + a2 e3 + a3 e1
σ(a1 e1 + a2 e2 + a3 e3 ) = a1 e2 + a2 e1 + a3 e3

and if we think at τ and σ as matrices (in the e1 , e2 , e3 basis), we have

0 0 1 0 1 0
   

τ = 1 0 0 σ = 1 0 0 .
   
0 1 0 0 0 1

Such representation C3 is not irreducible, because there exists the one-dimensional subrepresen-
tation
U = he1 + e2 + e3 i
which is actually the trivial representation, since every µ ∈ S3 acts as he identity. The comple-
mentary subspace n o
V = (z1 , z2 , z3 ) ∈ C3 z1 + z2 + z3 = 0

6
1.3. Irreducible representations of S3

is still a representation.
Note that the argument followed until now is valid for every Sn , n ≥ 3. The action of Sn
on Cn is a permutation representation but, in this case (that is Sn that acts on a vector space),
it is called the defining representation. From this action we can separate a line fixed by
the action, which turns out to be the trivial one (as before). The resulting representation V
obtained in this way, such that
Cn = U ⊕ V,
is called the standard representation of Sn and has dimension n − 1.

In the special case of S3 , we can construct explicitly all its irreducible representation. Pre-
cisely, we will prove that:

1. the standard representation is irreducible;

2. the trivial, the sign and the standard representations are the only irreducible one of S3 ;

3. any representation of S3 is a direct sum of copies of those three.

Proof. In order to prove points 1 and 2, let us see that there are no non-trivial subspaces of V
fixed by τ and σ. Let ζ be a primitive 3rd root of the unity and set:

v1 = (1, ζ 2 , ζ) v2 = (1, ζ, ζ 2 )

By direct calculation is easy to see that:

ζ2 + ζ + 1 = 0

and that v1 , v2 is a basis of V . Clearly,

τ v1 = (ζ, 1, ζ 2 ) = ζv1 τ v2 = (ζ 2 , 1, ζ) = ζ 2 v2

σv1 = (ζ 2 , 1, ζ) = ζ 2 v2 σv2 = (ζ, 1, ζ 2 ) = ζv1 .


Thus, v1 , v2 are the eigenvectors of σ and τ . In the basis v1 , v2 of V
! !
ζ 0 0 ζ
τ= σ= ,
0 ζ2 ζ2 0

and it is clear that they have not a common eigenspace.

Let us prove now point 3. Let


ρ : S3 −→ GL(W )
be a representation of S3 . Since τ has order 3, ρ(τ ) has order 3 too. We have now to use the
following lemma from linear algebra:
Exercise 2. Let A ∈ GL(n, C) be a matrix of finite order n. Then A is diagonalizable and the
eigenvalues of A are n-roots of the unity.
In this particular case, I deduce that the eigenvalues of ρ(τ ) can only be 1, ζ, ζ 2 . Denoting
the eigenspaces as
Wλ = { w ∈ W | ρ(τ )w = λw }

7
1.4. Operations on representations

and remembering that a linear map is diagonalizable if and only if its domain is the direct sum
of its eigenspaces, we have that
W = W1 ⊕ Wζ ⊕ Wζ 2 .
Now, from the relation στ −1 = τ σ, we obtain that for w ∈ Wλ
τ (σw) = (τ σ)w = (στ −1 )w
= σ(τ −1 w) = σλ−1 w
= λ−1 (σw).

This means that σ fixes W1 and that switches Wζ with Wζ 2 (since ζ −1 = ζ 2 ). If we fix w1 , . . . , wm
a basis of Wζ , then σw1 , . . . , σwm is a basis of Wζ 2 (since the map σ· is an isomorphism). From:

τ wi = ζwi τ σwi = ζ −1 σwi = ζ 2 σwi

we have that
hwi , σwi i ' V ∀i = 1, . . . , m.
and so
Wζ ⊕ Wζ 2 ' V ⊕m = |V ⊕ ...
{z
⊕ V} .
m times
On the other hand, τ acts trivially on W1 and σ fixes it. Remembering that σ has order two,
we deduce that it is diagonalizable and

W1 = W1+ ⊕ W1− ,

with
W1± = { w ∈ W1 | σw = ±w } .
This implies that
W1+ = U ⊕r W1− = U 0⊕s .
Eventually, we have proved that

W = W1 ⊕ Wζ ⊕ Wζ 2 = U ⊕r ⊕ U 0⊕l ⊕ V ⊕m .

1.4 Operations on representations


In this section we will show how to construct representations from known ones.

1.4.1 Tensor product representation


Definition 9. Let V, W be K-vector spaces. A tensor product of V and W is a couple (T, τ ),
with T a vector space and τ : V ×W → T a linear map, such that the following universal property
holds: for every bilinear map ϕ : V × W → U there exists a unique linear map ψ : T → U such
that the following diagram commutes.
ϕ
V ×W U

τ ψ

8
1.4. Operations on representations

The vector space T is usually indicated with the symbol:

V ⊗W

Actually, we can prove that this space is unique, up to isomorphism.


Proposition 4. Let (T, τ ), (T 0 , τ 0 ) be tensor products of V and W . Then there exists a unique
isomorphism φ : T → T 0 such that
τ 0 = φ ◦ τ.
Proof. Let us use the universal property of the tensor product: if we think to T 0 as the above
U , we obtain that there exists a unique linear map φ1 : T → T 0 such that τ 0 = φ1 ◦ τ ; if we think
now to T as the above U , we get another linear map φ2 : T 0 → T such that τ = φ2 ◦ τ 0 .
V ×W

τ τ0
φ1
T T0
φ2

Clearly
(φ2 ◦ φ1 ) ◦ τ = φ2 ◦ (φ1 ◦ τ ) = φ2 ◦ τ 0 = τ,
so, by the tensor product universal property, we have φ2 ◦ φ1 = idT . Similarly φ1 ◦ φ2 = idT 0 , so
φ1 and φ2 are isomorphisms. With φ = φ2 we have the thesis.

Now we should prove that such a space exists. Actually, we can even construct it. From
the cartesian product V × W , take the free vector space F (V × W ) over K. We define the
equivalence relation for v, v1 , v2 ∈ V , w, w1 , w2 ∈ W and µ ∈ K

(v1 , w) + (v2 , w) ∼ (v1 + v2 , w)


(v, w1 ) + (v, w2 ) ∼ (v, w1 + w2 )
µ(v, w) ∼ (µv, w) ∼ (v, µw).

Then we set T = V × W∼ and τ : V × W → T the projection. The operations are well-defined,
so that T = V ⊗ W is the tensor product. Eventually, let us suppose that e1 , . . . , en is a basis
for V and f1 , . . . , fm is a basis for W . Then the elements ei ⊗ fj form a basis of V ⊗ W .
Eventually, if V, W are finite dimensional it is easy to see that

T = { f : V ∗ × W ∗ → K | f is bilinear }

is a tensor product for V and W . The structure of this set suggests also that

dim V ⊗ W = dim Hom(V ∗ × W ∗ , K)


= dim(V ∗ × W ∗ )
= (dim V ∗ ) (dim W ∗ )
= (dim V ) (dim W ) .

Remark 6. Not all the elements of V ⊗W have the form, ei ⊗fj which are called pure tensors.
For example, in C ⊗ C the element

e1 ⊗ e2 + e2 ⊗ e1

is not a pure tensor.

9
1.4. Operations on representations

Let α : V → V 0 , β : W → W 0 be linear maps. We define the tensor product of α and β as


the map α ⊗ β : V ⊗ W → V 0 ⊗ W 0 setting

(α ⊗ β)(ei ⊗ fj ) = α(ei ) ⊗ β(fj ),

extended by linearity.
Example 4. Let A ∈ K n×n be the matrix associated to f with respect to the basis (ei ) and
B ∈ K m×m be the matrix associated to g with respect to the basis (fi ). We want to construct
explicitly the matrix A ⊗ B ∈ K nm×nm associated to f ⊗ g with respect to the basis ei ⊗ fj .
Writing
n m
Aei = Bfj =
X X
aik ek bjl fl ,
k=1 l=1
then we obtain
A ⊗ B (ei ⊗ fj ) = Aei ⊗ Bfj
n m
" ! !#
=
X X
aik ek ⊗ bjl fl
k=1 l=1
=
X
aik bjl ek ⊗ fl
k,l

In this way we have constructed the Kroneker product of the matrices:


Ab11 Ab12 ..
 
 Ab21 Ab22 .. 
A⊗B = .
..
: : .

Note that
tr(A ⊗ B) = tr(A) tr(B). (1.1)
Now if ρ : G → GL(V ) and τ : G → GL(W ) are representations of G, then we define the
tensor product representation to be

σ : G −→ V ⊗ W
g 7−→ ρ(g) ⊗ τ (g),

Without reference to the representations, we can simply write

g · (v ⊗ w) = g · v ⊗ g · w,

extended by linearity.
Exercise 3. Prove that σ is actually a representation.

1.4.2 Dual representation


Let us suppose that G is acting on a vector space V . We want to define a representation on:

V ∗ = Hom(V, C).

Let us consider the standard pairing

V ∗ × V −→ C
v ∗ , v 7−→ (v ∗ , v) = v ∗ (v).

10
1.4. Operations on representations

We want to define the action of G on V ∗ in such a way that the pairing would be G-stable.
This means that we are looking for a representation ρ∗ : G → GL(V ∗ ) such that for every g
(ρ∗ (g)v ∗ , ρ(g)v) = (v ∗ , v).
Using the fact that for every linear map ϕ : V → V there exists a linear map ϕT : V ∗ → V ∗ such
that (v ∗ , ϕv) = (ϕT v ∗ , v), we have
(ρ∗ (g)v ∗ , ρ(g)v) = (v ∗ , ρ∗ (g)T ρ(g)v).
Requiring that
(v ∗ , ρ∗ (g)T ρ(g)v) = (v ∗ , v) ∀v ∗ ∈ V ∗ ∀v ∈ V,
it is reasonable to set
ρ∗ (g)T ρ(g) = idV ,
that is
ρ∗ (g) = ρ(g −1 )T .
Thus, our requirement completely determines the dual representation. Without reference to the
map ρ, we define the dual representation as
(g · v ∗ )(v) = v ∗ (g −1 · v).
Exercise 4. Prove that ρ∗ is actually a representation.

1.4.3 Hom(V, W ) representation


Let V and W be representations of G. We will define the action of G on Hom(V, W ) via the
following
Lemma 5. There exists a canonical isomorphism
Hom(V, W ) ' V ∗ ⊗ W.
Proof. Let us prove that the map
φ : V ∗ ⊗ W −→ Hom(V, W )
v ∗ ⊗ w 7−→ (v 7→ v ∗ (v)w )
is an isomorphism. Let v1 , . . . , vn be a basis of V , v1∗ , . . . , vn∗ the dual basis and w1 , . . . , wm a
basis of W . Clearly vi∗ ⊗ wj is a basis of V ∗ ⊗ W . Furthermore
φ(vi∗ ⊗ wj )(vk ) = vi∗ (vk )wj = δik wj ,
that is φ(vi∗ ⊗ wj ) has zero in every entries except for the element φij :
i
 .. 
.
φ(vi∗ ⊗ wj ) = 1 ··· 
 
  j.
 

Since φ(vi∗ ⊗ wj ) generates Hom(V, W ), φ is surjective. Eventually


dim Hom(V, W ) = dim(V ∗ ⊗ W )
.

11
1.4. Operations on representations

We are ready to describe the action of G on Hom(V, W ). Let us say that ϕ ∈ Hom(V, W )
corresponds to v ∗ ⊗ w. Then
(g · ϕ)(v) = (g · (v ∗ ⊗ w)) (v) = (g · v ∗ ⊗ g · w) (v)
= (g · v ∗ )(v)g · w
= v ∗ (g −1 · v)g · w
 
= g · v ∗ (g −1 · v)w
= g · ϕ(g −1 · v).
Definition 10. If V is a representations of G, we set

VG ={v ∈V |g·v =v ∀g ∈ G } ,

that is, the G-stable elements of V .


Exercise 5. Prove that V G is a subrepresentation of V .
Proposition 6. The G-stable linear maps are the same as the G-linear maps:

HomG (V, W ) = Hom(V, W )G .

Proof. We know that ϕ ∈ Hom(V, W )G if and only if

(g · ϕ)(v) = ϕ(v) ∀g ∈ G, ∀v ∈ V.

By definition of Hom(V, W ) representation, this means that

g · ϕ(g −1 · v) = ϕ(v) ∀g ∈ G, ∀v ∈ V.

Since g −1 · is a bijection, this is equivalent to

g · ϕ(u) = ϕ(g · u) ∀g ∈ G, ∀u ∈ V,

that is ϕ ∈ HomG (V, W ).

1.4.4 Symmetric and alternating powers


Let V be a vector space of dimension n and consider the tensor product:

V ⊗k = V
|
⊗ ·{z
· · ⊗ V} .
k times

We are interested in these two subspaces of V ⊗k :


Dn oE
S= v1 ⊗ · · · ⊗ vk − vσ(1) ⊗ · · · ⊗ vσ(k) σ ∈ Sk

A = h{ v1 ⊗ · · · ⊗ vk | vi = vj for some i 6= j }i

We can define two quotient vector spaces.


Definition 11. We define the kth symmetric power V as
⊗k
Symk V = V S

and the kth alternating power of V as


^k ⊗k
V = V A.

12
1.4. Operations on representations

Concretely, in the symmetric power we have identified the elements of the tensor product
space which are the same up to permutations of the “components”, while in the alternating
power we have identified with zero the elements which have two equal “components”.
Introducing the canonical projections

V ⊗k −→ Symk V
v1 ⊗ · · · ⊗ vk 7−→ e1 · · · · · ek
^k
V ⊗k −→ V
v1 ⊗ · · · ⊗ vk 7−→ e1 ∧ · · · ∧ ek

we have that a basis for Symk V is given by:

ei1 · · · · · eik i1 ≤ · · · ≤ ik

and a basis for V is given by:


Vk

ei1 ∧ · · · ∧ eik i1 < · · · < ik .

Proposition 7. The dimensions of the symmetric and alternating powers are given by

n+k−1
!
dim Sym V = k
,
k
 !
 n

k ≤ n,

^k
dim V = k

0

k > n.

Proof. In order to calculate the dimension of Symk V , we can think at it as as the space of
homogeneous polynomials in the variables e1 , . . . , en of degree k. For what concerns the wedge
space, we have to choose k different elements in a set of n.

Remark 7. In V the wedge product is alternating3 , in fact


Vk

0 = v1 ∧ · · · (vi + vj ) · · · (vi + vj ) · · · ∧ vk
= v1 ∧ · · · vi · · · vi · · · ∧ vk + v1 ∧ · · · vi · · · vj · · · ∧ vk
+ v1 ∧ · · · vj · · · vi · · · ∧ vk + v1 ∧ · · · vj · · · vj · · · ∧ vk
= v1 ∧ · · · vi · · · vi · · · ∧ vk + v1 ∧ · · · vj · · · vj · · · ∧ vk .

On the other hand, the dot product in Symk V satisfy v1 · · · · · vk = vσ(1) · · · · · vσ(k) by definition
and thus is symmetric. Actually this reflects something deeper, that is a universal property.

3
Remember that alternating is equivalent to antisymmetric, provided that charK 6= 2.

13
1.5. Complete reducibility theorem and Schur’s lemma

The k-linear symmetric map V k → Symk V obtained from the above projection is such that
for every k-linear symmetric map ϕ : V k → U , there exists a unique linear map ψ : Symk V → U
such that the following diagram commutes.
ϕ
Vk U
ψ

Symk V

The k-linear antisymmetric map V k → k V obtained from the above projection is such that for
V

every k-linear alternating map ϕ : V k → U there exists a unique linear map ψ : k V → U such
V

that the following diagram commutes.


ϕ
Vk U
ψ

Vk
V

Remark 8. In general V ⊗k is never the direct sum of Symk V and V , except for the case
Vk

k = 2. Explicitly
^2
V ⊗ V ' Sym2 V ⊕ V (1.2)
Indeed we can define the map
^2
φ : V ⊗ V −→ Sym2 V ⊕ V
v ⊗ w 7−→ (v · w, v ∧ w),

extended by linearity. This map is surjective, since for every element (v · w, u ∧ z), we can define
v⊗w+w⊗v u⊗z−z⊗u
x= + .
2 2
Then, φ(x) = (v · w, u ∧ z). Since the elements of the form (v · w, u ∧ z) generates Sym2 V ⊕ V,
V2

we have the surjectivity. Eventually, the domain and the codomain have same dimension:

n+1 n(n + 1) n(n − 1)


! !
^2 n
dim Sym V + dim
2
V = + = + = n2
2 2 2 2
= dim V ⊗ V.

1.5 Complete reducibility theorem and Schur’s lemma


We have seen in Section 1.3 that every representation of S3 is decomposable into the direct sum
of its irreducible representations. This is actually the case for every C-linear representation of
finite dimension of a finite group. In order to prove this, let us see that every subrepresentation
has a “complementary” one.
Theorem 8 (Maschke’s theorem). Let V be a representation of G and W ⊆ V a subrepresen-
tation. There exists a subrepresentation W 0 ⊆ V such that

V = W ⊕ W 0.

14
1.5. Complete reducibility theorem and Schur’s lemma

Proof. The idea is to define an Hermitian product, such that W 0 is the orthogonal of W with
respect to it. Let (·, ·)0 be the standard Hermitian product:
n
(v, w)0 =
X
ai bi
i=1

for some basis of V . We define a new inner product averaging (·, ·)0 over the group:
1 X
(v, w) = (gv, gw)0 .
|G| g∈G

This is an Hermitian inner product. Indeed (v, v) ≥ 0 since it is a sum of positive numbers and
(v, v) = 0 if and only if gv = 0 for all g ∈ G, so that v = 0 with g = id. The other requirements
are trivial. In addition, this product is G-stable since (gv, gw) is just a rearrangement of the
sum (v, w).
Assuming this, the orthogonal subspace

W 0 = { v ∈ V | (v, w) = 0 ∀w ∈ W }

is still a representation. Indeed, if v ∈ W 0 we have

(gv, w) = (v, g −1 w) = 0,

since g −1 w ∈ W , that is gv ∈ W 0 . Furthermore, W ∩ W 0 = 0, so that V = W ⊕ W 0 .

Remark 9. The definition of an Hermitian product on V fixed by G is usually referred as


Weyl’s unitary trick. It shows that every finite dimensional representation of a finite group can
be restricted to unitary matrices:

ρ : G −→ U(V ) ⊂ GL(V ).

Note that above proof does not work if G is infinite, because the sum of the inner products
becomes a series which may not be convergent. However, the machinery still works with the
appropriate notions for compact Lie groups, where the sum is substituted by an integral.
Corollary 9 (Complete reducibility theorem). Any representation V of G is decomposable in
a direct sum of irreducible representations.
Proof. If V is irreducible, we have nothing to prove. Otherwise, there exists an irreducible
non-trivial subrepresentation W . Applying the same argument to the complementary subrepre-
sentation and noting that dim W 0 < dim V , we obtain the thesis.

One of the most useful tool in Representation Theory is the following


Theorem 10 (Schur’s lemma). Let V, W be irreducible representations of G and let ϕ : V → W
be a G-linear map. Then ϕ is either an isomorphism or is the null map. Further, if V = W ,
then there exists λ ∈ C such that ϕ = λ idV .
Proof. We know that ker ϕ ⊆ V is a subrepresentation. Since V is irreducible, we have that
either ker ϕ = (0) or ker ϕ = V . In the second case ϕ ≡ 0. In the first case, we have that ϕ is
injective. Furthermore, with the same argument on Im ϕ, we obtain the surjectivity, obtaining
that ϕ is an isomorphism.
Let now V = W . In the complex field, we know that ϕ has at least one eigenvalue λ. The
map ϕ − λ idV is G-linear and not injective (since λ is an eigenvalue): according to the previous
part of the statement, ϕ − λ idV ≡ 0.

15
1.5. Complete reducibility theorem and Schur’s lemma

Corollary 11. Let G be an abelian group. Any irreducible representation V has dimension 1.
Proof. Let h ∈ G. Since G is abelian, we know that
h· : V −→ V
is G-linear, indeed ρ(h) ◦ ρ(g) = ρ(hg) = ρ(gh) = ρ(g) ◦ ρ(h).
ρ(h)
V V

ρ(g) ρ(g)
ρ(h)
V V

By Schur’s lemma, h = λ idV and any subspace W of V is G-stable (since h · W = λW = W ).


If dim V > 1, any proper subspace would be a subrepresentation, which is absurd.

Corollary 12 (Isotypical decomposition). Any representation V of G is decomposable in a


direct sum of the form
V ' V1⊕a1 ⊕ · · · ⊕ Vr⊕ar ,
where
• Vi are non-isomorphic irreducible representations;
• the factors Wi = Vi⊕ai are canonical; they are called the isotypical components of V ;
• the values ai are unique; they are called the multiplicity of Wi .
Proof. In accordance with the complete reducibility theorem, we already know that such de-
composition exists. Let us suppose that
V ' V1⊕a1 ⊕ · · · ⊕ Vr⊕ar ' U1⊕b1 ⊕ · · · ⊕ Us⊕bs
The key of the proof is that there exists an index j1 such that V1 ' Uj1 as representations. In
fact, if it is not so, then for every j = 1, . . . , s the map
⊕bj
V1 ,→ V1⊕a1 ,→ V  Uj  Uj
would be the null map, according to Schur’s lemma. However, this is absurd: if we take a
non-zero element of V1 , then we have a non-zero vector v in V . Since V ' U1⊕b1 ⊕ · · · ⊕ Us⊕bs ,
there exists an index j such that the projection of v onto Uj is non-zero. This contradicts the
above statement.
Without loss of generality, we can suppose that V1 ' U1 . Then a1 = b1 , because the irreducible
components are assumed to be distinct for different indices. With the same argument, we find
V2 ' U2 an so on, obtaining ai = bi , r = s and Vi ' Ui for every index.

Remark 10. Every isotypical component is the direct sum of copies of an irreducible represen-
tation, which is not uniquely defined.
For example, if G = { 1 } is trivial, than it acts on every vector space V . The only isotypical
component will be V itself, which can be written (if dim V = n > 1) as the direct sum of different
one-dimensional irreducible representations: for every v1 , . . . , vn basis of V
V = hv1 i ⊕ · · · ⊕ hvn i .
Exercise 6. Let V be an irreducible representation of G. Prove that an Hermitian product on
V fixed by G is unique up to scalars.

16
Chapter 2

Character Theory

We have already discussed that the aim of the course is to classify, as far as we can, all the
representations of a finite group. In this chapter we will develop and study a total invariant:
the character. Our first guess, since we are dealing with matrices, is to consider the matrices
invariants and it turns out that the trace is one of the main tools in dealing with representations.
Remember that in the construction of the isotypical decomposition of S3 , we crucially made
use of the eigenvalues of the representation. Further, in Section 1.5 we proved Schur’s lemma
thanks to an eigenvalue of the G-isomorphism. This suggests that the eigenvalues of a repre-
sentation assume a very important role in the theory. With Remark 12 we will understand how
knowing the trace of every element of G can allow us to know every eigenvalue.

2.1 The character


Definition 12. Let ρ : G → GL(V ) be a representation. We define the character of the
representation as the function:

χ : G −→ C
g 7−→ tr ρ(g),

It is clear that χ is a homomorphism on the product if and only if dim V = 1. Furthermore,


in this case
χV (g) = tr ρ(g) = ρ(g),
since ρ(g) ∈ C. Let us now underline some properties that will be very important.

Remark 11. The character is invariant under isomorphism of representations:

tr(ϕ ρ(g) ϕ−1 ) = tr ρ(g),

since conjugate matrices have same trace.

Lemma 13. The character is constant on conjugacy classes.

Proof. Easily:
   
χ(hgh−1 ) = tr ρ(hgh−1 ) = tr ρ(h)ρ(g)ρ(h)−1
= tr ρ(g) = χ(g).

17
2.1. The character

We are now ready to investigate in which way the character behave on the representations
described in Section 1.4.

Proposition 14. The following formulae hold true.

1. χ(1) = dim V

2. χU ⊕V = χU + χV

3. χV ⊗W = χV · χW

4. χV ∗ (g) = χV (g) = χV (g −1 )

Furthermore, if V is a 1-dimensional representation, then χV (g −1 ) = χ−1


V (g).

Proof. For what concerns point (1), we have by definition

χ(1) = tr In = n = dim V.

For points (2) and (3) we have just to remember that trace is additive and Formula 1.1 respec-
tively. Eventually, for point (4), we know that

χV (g) = λ1 + · · · + λn ,

where the λi are the eigenvalues of ρ(g) (counted without multiplicity). Clearly:
   
χV ∗ (g) = tr ρ(g −1 )T = tr ρ(g)−1 = λ−1 −1 −1
1 + · · · + λn = χV (g )

Furthermore, for finite groups, ρ(g) is of finite order, which means that |λi | = 1. This means
that λ−1
i = λi so
χV ∗ (g) = λ1 + · · · + λn = χV (g).

Proposition 15. The following formulae holds true:

1  1 
χ∧2 V (g) = χV (g)2 − χV (g 2 ) χSym2 V (g) = χV (g)2 + χV (g 2 ) .
2 2

Proof. Let g ∈ G be a linear map and e1 , . . . , en a basis of eigenvectors. The dot and wedge
product of this elements is still an eigenvector for the symmetric and wedge product respectively:

g · (ei1 · · · · · eik ) = (g · ei1 ) · · · · · (g · eik )


= λi1 · · · λik (ei1 · · · · · eik ),
g · (ei1 ∧ · · · ∧ eik ) = (g · ei1 ) ∧ · · · ∧ (g · eik )
= λi1 · · · λik (ei1 ∧ · · · ∧ eik ).

In particular, since these products form a basis of the two spaces, we have that

χSymk V (g) = χ∧k V (g) =


X X
λi1 · · · λik , λ i1 · · · λ ik .
i1 ≤···≤ik i1 <···<ik

18
2.1. The character

In the particular case of k = 2 we have:

n
!2 n
1 X 1X
χ∧2 V (g) = λi1 λi2 =
X
λi − λ2
i1 <i2
2 i=1 2 i=1 i
1 1
= χ2V (g) − χV (g 2 )
2X 2 X
χSym2 V (g) = λ i1 λ i2 = λi1 λi2 +
X
λ i1 λ i2
i1 ≤i2 i1 <i2 i1 =i2
n
= χV2 V (g) +
X
λ2i
i=1
1 
= χV (g)2 − χV (g 2 ) + χV (g 2 )
2
1 1
= χ2V (g) + χV (g 2 )
2 2

Remark 12. It can be easily seen that the element χ∧k V (g) is the kth coefficient of the char-
acteristic polynomial of g (formally, this involves the theory of the symmetric polynomials of
Chapter 7): for example, let us fix dim V = 3 and let us suppose that λ1 , λ2 and λ3 are the
eigenvalues of the map g. The characteristic polynomial of g is

pg (x) = (λ1 − x)(λ2 − x)(λ3 − x)


= −x3 + (λ1 + λ2 + λ3 )x2 − (λ1 λ2 + λ1 λ3 + λ2 λ3 )x + λ1 λ2 λ3 .

In the above proposition, we wrote χ∧2 (g) in terms of the trace of some powers of g (precisely, g
itself and g 2 ). As soon as the number of the alternating powers increases, we need to know the
trace of higher powers of g in order to calculate χ∧k (g). This are some of the other formulae

1h 3 i
χ∧3 V (g) = χV (g) − 3χV (g) χV (g 2 ) + 2χV (g 3 )
6
1 h 4 i
χ∧4 V (g) = χV (g) − 6χV (g 2 )χ2V (g) + 3χ2V (g 2 ) + 8χV (g)χV (g 3 ) − 6χV (g 4 ) .
24
Actually, since every element g of a finite group G has finite order, knowing the trace of every
powers of g means knowing the characteristic polynomial of g and so its eigenvalues.

Proposition 16. Let V be a permutation representation of G. Then χV (g) is the number of


fixed points of g (and thus is a non-negative integer).

Proof. For a permutation representation we have that g is a permutation matrix, whose columns
are the vectors of the standard basis. The trace of this matrix is exactly the number of fixed
points.

We will prove that every representation is determined by its character. For this reason, we
will use the so-called character table: according to Lemma 13, we just have to calculate the
values of the character of the representation on the conjugacy classes.

Example 5. It can be proved that the conjugacy classes of Sn are determined by the cycle
structure. For S3 we have three type of cycle structure:

19
2.1. The character

cycle structure elements in the conj. classes


(·) (·) (·) id
(· ·) (·) (1 2), (1 3), (2 3)
(· · ·) (1 2 3), (1 3 2)

Note that for the trivial representation U


χU (C) = 1 ∀C conjugacy class,
while for the sign representation U 0
χU 0 (C) = sgn C,
where sgn C is the sign of a representative of the conjugacy class. Eventually, in the standard
representation V we have ! !
ζ 0 0 ζ
τ= σ= .
0 ζ2 ζ2 0
Thus,
χV (C(1 2) ) = tr σ = 0
χV (C(1 2 3) ) = tr τ = ζ + ζ 2 = −1.
The character table will be

conjugacy classes id (1 2) (1 2 3)
elements in the conjugacy class 1 3 2
U 1 1 1


0
irreducible representations U 1 −1 1
2 0

V −1

We have proved in Section 1.3 that the above three representations are the the only irre-
ducible one of S3 . We have also seen that every other representation is sum of copies of these: let
us now try to understand how to use the character table to calculate the isotypical components
of a representation. In general this problem can be very hard; in the next section we will prove
some results that will simplify a lot this calculation. Meanwhile, let us work with the particular
case of Sym2 V .

We know that
V = { (z1 , z2 , z3 ) | z1 + z2 + z3 = 0 } .
A basis for V is
e1 = (1, −1, 0) e2 = (0, 1, −1)
and a basis for Sym2 V is
e21 e1 · e2 e22 .
A first calculation gives
τ e 1 = e2
τ e2 = (−1, 0, 1) = −e1 − e2
σe1 = (−1, 1, 0) = −e1
σe2 = (1, 0, −1) = e1 + e2 .

20
2.2. Orthogonality relations

Eventually,

τ e21 = τ e1 τ e1 = e22
τ e1 · e2 = e2 · (−e1 − e2 ) = −e1 e2 − e2
τ e22 = (−e1 − e2 )2 = e21 + 2e1 · e2 + e22
σe21 = e21
σe1 · e2 = −e21 − e1 · e2
σe22 = (e1 + e2 )2 = e21 + 2e1 · e2 + e2 .

This means that, in this basis, we have

0 0 1 1 −1 1
   

τ = 0 −1 2 , σ = 0 −1 2 .
   
1 −1 1 0 0 1

Thus χSym2 V (τ ) = 0 and χSym2 V (σ) = 1. If we go back to the character table

id (1 2) (1 2 3)
1 3 2
U 1 1 1
U0 1 −1 1
V 2 0 −1
Sym2 V 3 1 0

it is easy to see that χSym2 V = χU + χV . Assuming the total invariance of the character, then

Sym2 V ' U ⊕ V.

Exercise 7. Compute explicitly the character of W = V ⊗V and deduce from that the isotypical
decomposition of W (assuming again the total invariance of the character).

2.2 Orthogonality relations


Definition 13. A class function is a map

α : G −→ C

constant on conjugacy classes. We will denote the set of class functions on G as C`(G).
The set of class functions is a vector space of dimension dim C`(G) = #conjugacy classes
and it contains the set of characters of G. We can define a Hermitian product as
1 X
(α, β) = α(g)β(g).
|G| g∈G

Note that
1 X
(α, β) = #C α(C)β(C)
|G| C conj.
classes
where with α(C), β(C) we mean the functions valued in an element of the conjugacy class.
With this scalar product, we can now characterise the irreducible representations of a group.

21
2.2. Orthogonality relations

Theorem 17 (Orthonormality of irreducible characters). The characters of the irreducible


representations are orthonormal, that is
1 if V ' W
(
(χV , χW ) =
0 if V 6' W .

In order to prove this theorem, we need the following


Lemma 18 (First projection formula). Let V be a representation of G. The map:
1 X
π= g· : V −→ V G
|G| g∈G

is a G-linear projection of V in V G . Furthermore,


1 X
dim V G = χ (g). (2.1)
|G| g∈G V

Proof. Let us prove that π is G-linear. If h ∈ G, then


1 X
hπh−1 = hgh−1
|G| g∈G

Setting k = hgh−1 , we obtain


1 X
hπh−1 = k = π.
|G| k∈G

Let us prove that Im π = V G . If we take v ∈ V G , then we have


1 X 1 X 1
π(v) = g·v = v= |G|v = v.
|G| g∈G |G| g∈G |G|

On the other hand, let w ∈ V and π(w) ∈ Im π. We have


1 X 1 X
h · π(w) = hg · w = k · w = π(w).
|G| g∈G |G| k∈G

For what concerns the dimension of V G , we have just to observe that, being π the projection on
V G , dim V G = tr π and
1 X 1 X
tr π = tr g = χ (g).
|G| g∈G |G| g∈G V

Proof of Theorem 17. In accordance with Schur’s lemma, if V and W are not isomorphic irre-
ducible representations, the only G-linear map is the null-map. Furthermore, if V ' W , we
have that two G-linear maps φ and ψ differ just by a scalar factor, applying Schur’s lemma to
φ−1 ◦ ψ. Thus:
1 if V ' W
(
dim HomG (V, W ) =
0 if V 6' W .
Now, thanks to Formula 2.1 and Proposition 6, we have:
1 X
dim HomG (V, W ) = χ (g).
|G| g∈G Hom (V,W )

22
2.2. Orthogonality relations

On the other hand, Hom(V, W ) ' V ∗ ⊗ W and χV ∗ ⊗W = χV · χW . Hence:

1 X
dim HomG (V, W ) = χ (g) · χW (g) = (χV , χW ).
|G| g∈G V

The above theorem has important corollaries, which underline the value of Character Theory.

Remark 13. If V is irreducible, then

1 if V is trivial
(
dim V G
=
0 if V is not trivial.

Indeed, from Formula 2.1,


1 X
dim V G = χ (g) = (χU , χV ),
|G| g∈G V

where U is the trivial representation. From the orthonormality of irreducible characters, we


obtain the statement.

Corollary 19. The number of irreducible representations of G is less or equal to the number
of conjugacy classes.

Proof. Since the characters of the irreducible representations are orthonormal, they are linearly
independent. Thus

#irreducible representations ≤ dim C`(G) = #conjugacy classes.

Corollary 20. The character of a representation V is a linear combination of the the irreducible
characters χVi :
χV = a1 χV1 + · · · + ar χVr , (2.2)
where ai = (χV , χVi ) is the multiplicity of the ith isotypical component.

Proof. We know that V always decomposes as

V = V1⊕a1 ⊕ · · · ⊕ Vr⊕ar .

Then the decomposition


χV = a1 χV1 + · · · + ar χVr ,
follows from the formula for the character of a sum. On the other hand, taking the Hermitian
product of χV and χVi and using the above orthogonality relation, we find
r r
(χV , χVi ) = aj (χVj , χVi ) = aj δij = ai .
X X

j=1 j=1

Corollary 21 (Total invariance of the character). Two representations are isomorphic if and
only if they have the same character.

23
2.2. Orthogonality relations

Corollary 22. For every representation V of G,


r
(χV , χV ) =
X
a2i ,
i=1

where ai is the multiplicity of the ith isotypical component of V . In particular, V is irreducible


if and only if (χV , χV ) = 1.
Proof. The first formula follows from the decomposition Formula 20 and the orthonormality
relation of irreducible characters. For the second part, one implication is Theorem 17. On the
other hand, let us assume that (χV , χV ) = 1. Then we have

a2i = 1,
X

with ai ∈ N. The exists a unique index j such that aj = 1, while ai = 0 for all i 6= j. Thus
V ' Vj is irreducible.

Exercise 8. Let V and W be irreducible representations of G and suppose W is one-dimension.


Then V ⊗ W is irreducible.
Exercise 9. Let V be a representation of G. Then V is irreducible if and only if V ∗ is irreducible.
Proposition 23. Let Vreg be the regular representation of G. Then it contains every irreducible
representation Vi of G, with multiplicity dim Vi .
Proof. Since Vreg is a permutation representation, we have that χVreg (g) is the number of point
fixed by g. Thus,
|G| if g = 1
(
χVreg (g) =
0 if g =
6 1.
Taking the Hermitian product of χVreg and χVi , we find

1 X
(χVreg , χVi ) = χ (g) · χVi (g)
|G| g∈G Vreg
1
= |G| χVi (1) = dim Vi .
|G|

Corollary 24. Let us call Irr (G) = { V irreducible representation of G }. Then

|G| = (dim V )2 .
X

V ∈Irr(G)

In particular, dim V ≤ |G|.


Proof. We know that
V ⊕ dim V .
M
Vreg '
V ∈Irr(G)

Taking the dimension on both sides, we have

|G| = dim Vreg = (dim V )2 .


X

V ∈Irr(G)

24
2.2. Orthogonality relations

Remember that in the first part of the section we had examined some links between class
functions and characters. Let us deepen now such connection.
Lemma 25. Let α : G → C be a function and V a representation of G. The map

ϕα,V = α(g)g : V −→ V
X

g∈G

is G-linear if and only if α is a class function. In other words, we have a bijection

C`(G) ←→ Z(C[G])
X
α ←→ α(g)g
g∈G

Proof. If α is a class function, then

ϕα,V (h · v) = α(g)g · (h · v) = α(hgh−1 ) (hgh−1 )h · v


X X

g∈G g∈G
X 
= α(k) hk · v = h ·
X
α(k) k · v
k∈G k∈G
= h · ϕα,V (v)

On the other hand, suppose that there exist a representation V such that ϕα,V is G-linear and
α is not a class function, that is there exist ḡ, h̄ such that

α(ḡ) 6= α(h̄ḡ h̄−1 ).

Restricting it to an irreducible subrepresentation, we can suppose V irreducible. But the reg-


ular representation contains all the irreducible representations, so that we can take V = Vreg .
Evaluating h̄−1 ϕα,Vreg (h̄), we obtain

h̄−1 ϕα,Vreg (h̄) = α(g)h̄−1 g h̄ = α(h̄g h̄−1 )g


X X

g∈G g∈G

On the other hand, the map is G-linear, so that h̄−1 ϕα,Vreg (h̄) = ϕα,Vreg (1). Thus,

(α(h̄g h̄−1 ) − α(g)) g = 0.


X

g∈G

Since (g)g∈G is a basis for Vreg , α(h̄g h̄−1 ) = α(g) for all g. With g = ḡ we have an absurd.

Theorem 26. The irreducible characters form a orthonormal basis for C`(G).
Proof. We already know that the irreducible characters form an orthonormal set of C`(G). Let
α be a class function such that

(χV , α) = 0 ∀V ∈ Irr(G).

and let us prove that α ≡ 0. Let V be an irreducible representation of G and consider the
G-linear map
ϕα,V =
X
α(g)g.
g∈G

According to Schur’s lemma, there exists λα,V ∈ C such that

ϕα,V = λα,V idV .

25
2.2. Orthogonality relations

If we take the trace of both sides we have

tr ϕα,V 1 X |G|
λα,V = = α(g)χV (g) = (χ ∗ , α) = 0
dim V dim V g∈G dim V V

since V irreducible implies V ∗ irreducible. Thus, ϕα,V ≡ 0 for all V irreducible. As a conse-
quence, ϕα,Vreg ≡ 0, since the regular representation contains all the irreducible representation.
But in Vreg the elements g are linearly independent. From ϕα,Vreg (1) we obtain α ≡ 0.

Corollary 27. The number of irreducible representation of G is equal to the number of conju-
gacy classes of G. In particular, the character tables are always squared.

Lemma 25 would suggest another projection formula.

Proposition 28 (Second projection formula). Let V be a representation of G and

V = V1⊕a1 ⊕ · · · ⊕ Vr⊕ar

the decomposition in isotypical component. Every map

dim Vi X
πi = χ (g)g
|G| g∈G Vi

is a projector of V onto Vi⊕ai .

Proof. According to Lemma 25, every map of this type is G-linear. Let us consider the restriction
ϕij = πi |Vj . Since Vj is irreducible, we have g · Vj ⊆ Vj , so

ϕij : Vj −→ Vj ,

We can now apply Schur’s lemma: there exists λij ∈ C such that ϕij = λij idVj . Hence,

tr ϕij = λij dim Vj .

On the other hand, by definition of πi ,

dim Vi X
tr ϕij = χ (g) tr g|Vj
|G| g∈G Vi
dim Vi X
= χ (g)χVj (g)
|G| g∈G Vi
= dim Vi (χVi , χVj ) = δij dim Vi .

Thus, λij = δij and


r
⊕a
πi = δij idVj j .
M

j=1

26
2.3. The Frobenius divisibility theorem

2.3 The Frobenius divisibility theorem


In this chapter we will count how many ways there are so that an element g −1 of G can be
written as product of commutators. This knowledge will have an important consequence in
representation theory.

Let g ∈ G and set

Nk (g) = # { x1 , . . . , xk , y1 , . . . , yk ∈ G | [x1 , y1 ] · · · [xk , yk ]g = 1 } ,

Theorem 29. Nk (g) is a class function. Precisely, it is Fourier expansion1 is


2k−1
|G|

Nk (g) =
X
χ(g).
χ∈Irr(G)
χ(1)

Proof. Let us define the class function

1 g=1
(
δ(g) =
0 g=
6 1.

It is clear that
Nk (g) = δ ([x1 , y1 ] · · · [xk , yk ]g) .
X

x1 ...xk
y1 ...yk

Let us simplify the formula. Taking the scalar product of δ and χ, we find
1 X 1
(δ, χ) = δ(g)χ(g) = χ(1).
|G| g∈G |G|

Since the irreducible characters form an orthonormal basis for C`(G), we have
χ(1)
δ= (δ, χ)χ =
X X
χ.
χ∈Irr(G) χ∈Irr(G)
|G|

Thus, we obtain
χ(1)
Nk (g) = χ ([x1 , y1 ] · · · [xk , yk ]g)
X X

x1 ...xk χ∈Irr(G) |G|


y1 ...yk
χ(1) X
= χ ([x1 , y1 ] · · · [xk , yk ]g)
X

χ∈Irr(G)
|G| x1 ...xk
y1 ...yk

and all we need to prove is the identity


2k
|G|

χ ([x1 , y1 ] · · · [xk , yk ]g) =
X
χ(g)
x1 ...xk χ(1)
y1 ...yk

for χ irreducible. Let us do it by induction on k.


Suppose k = 1. Call V the vector space associated to χ and set

ϕy = xyx−1 .
X

x∈G
1
With this name we mean the fact that we wrote a class function as a linear combination of a basis of C`(G).

27
2.3. The Frobenius divisibility theorem

We have ϕy ∈ End(V ). Further, it is G-linear:

zϕy z −1 = zxyx−1 z −1 = zxy(zx)−1


X X

x∈G x∈G

= wyw = ϕy .
X

w∈G

By Schur’s lemma, ϕy = λy idV and taking the trace tr ϕy = λy dim V = λy χ(1). On the other
hand, by definition of ϕy ,

tr ϕy = χ(xyx−1 ) = χ(y) = |G|χ(y).


X X

x∈G x∈G

We conclude that
|G|
λy = χ(y)
χ(1)
and
|G|
xyx−1 =
X
χ(y)idV .
x∈G
χ(1)

Let us introduce the map ψ multiplying on the right by y −1 and summing over y:

ψ= xyx−1 y −1 = [x, y].


X X

x,y∈G x,y∈G

From the above formula, we find

|G| X
ψ= χ(y)y −1 .
χ(1) y∈G

It is easy to see that ψ is a G-linear map and then, for Schur’s lemma, ψ = λ idV . Taking the
trace as we did before,

|G| X
λ dim V = λ χ(1) = χ(y)χ(y −1 )
χ(1) y∈G
|G| X |G|2 |G|2
= χ(y)χ(y) = (χ, χ) = .
χ(1) y∈G χ(1) χ(1)

In the last step we have used the fact that (χ, χ) = 1, since it is irreducible. We conclude that
2
|G|

λ= .
χ(1)

Thus: 2
|G|

−1 −1
ψ= = idV .
X
xyx y
x,y∈G
χ(1)

Eventually, multiplying by g and taking the trace leads to the thesis


2
|G|

χ([x, y]g) =
X
χ(g).
x,y∈G
χ(1)

28
2.3. The Frobenius divisibility theorem

Exercise 10. Prove the inductive step: k =⇒ k + 1.


Definition 14. The number z ∈ C is said algebraic integer if it is the root of a monic
polynomial p ∈ Z[x].
Proposition 30. An irrational number is an algebraic integer if and only if it is an integer.
Proof. Clearly every integer is an algebraic integer. On the other side, let r = m
n be a root of
the polynomial
xk + ak−1 xk−1 + · · · a0 ai ∈ Z
and m, n ∈ Z such that gcd(m, n) = 1. By definition
k k−1
m m
 
+ ak−1 + · · · a0 = 0
n n
that is
mk + ak−1 mk−1 n + · · · + a0 nk = 0
In other words  
mk = −n ak−1 mk−1 + · · · + a0 nk−1
and so n divides mk . Since gcd(m, n) = 1, we conclude that n = ±1.

Proposition 31. The set of algebraic integers is a subring of C. Furthermore, the complex
conjugate of an algebraic integer is an algebraic integer.
Lemma 32. Let G be a finite group. The character χ(g) is an algebraic integer for all g ∈ G.
Proof. Since G is finite, every map g· has finite order. According to exercise 2, each of them is
diagonalizable and, above all, its eigenvalues are root of the unity and thus algebraic integer.
Eventually, the character of g is the sum of the eigenvalues of g· and the above proposition ends.

We are finally ready to state the main result of this section.


Theorem 33. The dimension of an irreducible representation divides |G|.
Proof. According to Theorem 29, the Fourier coefficients of Nk (g) are
2k−1
|G| 1 X

= (χ, Nk ) = χ(g)Nk (g).
χ(1) |G| g∈G

We know that Nk (g) is a (non-negative) integer, while χ(g) is an algebraic integer so g∈G χ(g)Nk (g)
P

is an algebraic integer. On the other hand, it is equal to the rational number


2k−1
|G|

|G|
χ(1)
An algebraic integer which is also a rational number is an integer:
2k−1
|G|

|G| ∈ Z ∀k ≥ 1.
χ(1)
It can be proved that this implies
|G|
∈ Z.
χ(1)

29
Chapter 3

Introduction to representations of Sn

No group is of a greater importance than the symmetric group since, according to Cayley’s
theorem, any group is isomorphic to a subgroup of a symmetric group. In this chapter we will
start analyzing the representations of Sn in two particular cases.

3.1 Tables of S4 and A4


Let us try to determine all the irreducible representation of S4 . We already know the character
table for the trivial representation U and the sign one U 0 .

id (1 2) (1 2 3) (1 2 3 4) (1 2)(3 4)
1 6 8 6 3
U 1 1 1 1 1
U0 1 −1 1 −1 1

We can check that U and U 0 are not isomorphic:


1 
(χU , χU 0 ) = 1 · 1 + 6 · (−1) + 8 · 1 + 6 · (−1) + 3 · 1
24
1 
= 1−6+8−6+3 =0
24
Remembering that the table must be squared, which are the other three irreducible repre-
sentations? We know that, from the defining representation C3 , we can find the standard one
V such that
C3 = U ⊕ V.
Then, χV = χC3 − χU . The defining representation is a permutation representation, so that χV
is the number of fixed points minus 1:

id (1 2) (1 2 3) (1 2 3 4) (1 2)(3 4)
1 6 8 6 3
V 3 1 0 −1 −1

This representation is irreducible because


1 
(χV , χV ) = 1 · 32 + 6 · 12 + 8 · 02 + 6 · (−1)2 + 3 · (−1)2
24
1 
= 9 + 6 + 6 + 3 = 1.
24

31
3.1. Tables of S4 and A4

We can now looking for the other representations tensorizing the ones we just have.
Tesorizing a representation with the trivial one is useless, indeed
χR⊗U = χR · χU = χR · 1 = χR
thus R ⊗ U ' R.
On the other hand, from Exercise 8, we have that V 0 = V ⊗ U 0 is irreducible, with character

id (1 2) (1 2 3) (1 2 3 4) (1 2)(3 4)
1 6 8 6 3
V0 3 −1 0 1 −1
Now we miss a representation W : let us construct it explicitly. Thanks to Corollary 24, its
dimension is
q √
|S4 | − (dim U )2 − (dim U 0 )2 − (dim V )2 − (dim V 0 )2 = 24 − 1 − 1 − 9 − 9 = 2.
One possibility is to look at the action of S4 on X = { (1 2)(3 4), (1 3)(2 4), (1 4)(2 3) }, the set
of two cycles, by conjugation:
α : S4 × X −→ X
(σ, x) 7−→ σxσ −1 .
The action can be represented as follows.

(1 2)(3 4) (1 3)(2 4) (1 4)(2 3)


id
(1 2) (1 4)(2 3) (1 3)(2 4)
(1 2 3) (1 4)(2 3) (1 2)(3 4) (1 3)(2 4)
(1 2 3 4) (1 4)(2 3) (1 2)(3 4)
(1 2)(3 4)

Thus, we have the permutation representation hXi, with a fixed line


Y = h(1 2)(3 4), (1 3)(2 4), (1 4)(2 3)i .
Separating it from the space, we obtain hXi = Y ⊕ W . The character will be the number of
fixed points of the action α minus 1. Eventually, the character table of S4 will be

id (1 2) (1 2 3) (1 2 3 4) (1 2)(3 4)
1 6 8 6 3
U 1 1 1 1 1
U0 1 −1 1 −1 1
V 3 1 0 −1 −1
V0 3 −1 0 1 −1
W 2 0 −1 0 2
Remark 14. The set
K = { id, (1 2)(3 4), (1 3)(2 4), (1 4)(3 4) }
is a subgroup of S4 . It is isomorphic to Z2 × Z2 and it is called the Klein group (or Vier-
ergruppe). Further,
π : S4 −→ S4K ' S3 .
Composing π with the standard representation of S3 , we obtain the representation W of S4 .

32
3.1. Tables of S4 and A4

Let us now compute the character table of A4 . Neglecting the odd permutations, we loose
the sign representation, while we still have the trivial and the standard representations.

id (1 2 3) (1 3 2) (1 2)(3 4)
1 4 4 3
U 1 1 1 1
V 3 0 0 −1

Since |A4 | = 12, we miss two representations, say Y and Z, of dimension one. Indeed

(dim Y )2 + (dim Z)2 = |A4 | − (dim U )2 − (dim V )2 = 2.

and the only couple whose squares sum to 2 is (1,1). Thanks to the argument discussed in
Section 1.2, we just need to compute the one-dimensional representations of the abelianization
of A4 . It can be proved that the commutator subgroup is the Klein group:

[A4 , A4 ] = { id, (1 2)(3 4), (1 3)(2 4), (1 4)(2 3) } ' K.

Thus, we have the exact sequence

1 −→ K −→ A4 −→ A4[A , A ] ' Z3 −→ 1.
4 4

Since Z3 is cyclic of order 3, we find

Z3 0 1 2
U 1 1 1
Y 1 ζ ζ2
Z 1 ζ2 ζ

Composing the projection and the non-trivial representations of Z3 , we obtain the missing
representation of A4 . For example, Y is given by

[id], [(1 4)(2 3)] 7−→ 0 7−→ 1


[(1 2 3)] 7−→ 1 7−→ ζ
[(1 3 2)] 7−→ 2 7−→ ζ 2 .

Here [ · ] represents the conjugacy class. Eventually, the character table of A4 is given by

id (1 2 3) (1 3 2) (1 2)(3 4)
1 4 4 3
U 1 1 1 1
Y 1 ζ ζ2 1
Z 1 ζ2 ζ 1
V 3 0 0 −1

Note that Y and Z are not self-dual, while Z ' Y ∗ .

33
3.2. Partitions, Young diagrams and tables of S5 and A5

3.2 Partitions, Young diagrams and tables of S5 and A5


We have seen that a central point in the calculation of character tables of Sn is its conjugacy
classes. We already said that the conjugacy classes are in 1-1 correspondence with the cycle
structure. One implication is immediate: two conjugate permutations have the same cycle
structure. Indeed

n elements
τ (·) · · · (·) (· ·) · · · (· ·) · · · ( · · ) · · · (· · · · ·) τ −1
z }| {
···
| {z }| {z } | {z }
m1 m2 mn
−1
= · · · τ (· · · · ·)τ τ · · · τ −1 (· · · · ·)τ −1
n elements
z }| {  
= · · · τ (·) · · · τ (·) · · · τ (·) · · · τ (·)
| {z }
mn

and the cycle structure is preserved. The other implication requires some work.

On the other hand, the cycle structure is in 1-1 correspondence with the partitions of n,
i.e. the ways of writing n as a sum of integers. We will indicate a partition of n with

λ = 1m1 · 2m2 · · · nmn ,

where n = |λ| = m1 + m2 · 2 + · · · + mn · n. Alternatively, we will write

λ = (λ1 , . . . , λs ) with λ1 ≥ λ2 ≥ · · · ≥ λs ,

so that |λ| = λ1 + λ2 + · · · + λs .

Remark 15. The first notation is recovered from the second one setting

mk = # { r | λr = k } ,

while from the first notation to the second one we have just to set

λ = (n, . . . , n, . . . , 1, . . . , 1).
| {z } | {z }
mn times m1 times

The partitions of n can be visually represented with Young diagrams. For a partition
λ = (λ1 , . . . , λs ), we will draw |λ| boxes, arranged in left-justified rows, such that the the first
rows have λ1 boxes, the second one have λ2 boxes an so on. Thus, we have the 1-1 correspondence

{ partitions of n } ←→ { Young diagrams with n boxes } .

As an example, the conjugacy classes of S5 , the partitions of 5 and the corresponding Young
diagrams are listed in the following table.

34
3.2. Partitions, Young diagrams and tables of S5 and A5

Young
cycle structure representative partition
diagram

(·)(·)(·)(·)(·) id 15

(· ·)(·)(·)(·) (1 2) 2 · 13

(· ·)(· ·)(·) (1 2)(3 4) 22 · 1

(· · ·)(·)(·) (1 2 3) 3 · 12
(· · ·)(· ·) (1 2 3)(4 5) 3·2
(· · · ·)(·) (1 2 3 4) 4·1
(· · · · ·) (1 2 3 4 5) 5

The problem now is to find a way to count the number of elements in each conjugacy class of
Sn . Let us call Cλ the conjugacy class associated to the partition λ. Since Sn acts on itself via
conjugation, we have that
#Cλ = [Sn : Stab(λ)].
Thus, we just have to compute the stabilizer of a cycle structure. It can be proved that, for
λ = 1m1 · 2m2 · · · nmn , we have1
n
Stab(λ) = Zm
Y
k
k o Smk .
k=1

The group can be seen as follows: for σ ∈ Cλ of the form


n elements
σ = (·) · · · (·) (· ·) · · · (· ·) · · · ( · · ) · · · (· · · · ·)
z }| {
···
| {z }| {z } | {z }
m1 m2 mn

an element in Stab(λ) will acts on every set of k-cycles shifting the elements inside the cycles
with Zk and swapping the cycles themselves with Smk . In particular,
n
Zλ = | Stab(λ)| = k mk mk !
Y

k=1

and we obtain the important result that the number of elements in Cλ will be
n!
#Cλ = .

With this formula, we can now compute the character table of S5 . We have the trivial
representation U , the sign U 0 , the standard V and V 0 = V ⊗ U 0 , with character
1
Let (G, ∗), (H, ∗) be two groups and let G acting on H. If we define on G × H the operation

(g, h) ∗ (g 0 , h0 ) = (g ∗ g 0 , (g · h) ∗ h0 ),

we obtain that (G × H, ∗) is a group. It is indicated as G o H and is called semidirect product. Note that if
the representation is the trivial one, we obtain the natural product (g, h) ∗ (g 0 , h0 ) = (g ∗ g 0 , h ∗ h0 ).

35
3.2. Partitions, Young diagrams and tables of S5 and A5

id (1 2) (1 2 3) (1 2 3 4) (1 2 3 4 5) (1 2)(3 4) (1 2 3)(4 5)
1 10 20 30 24 15 20
U 1 1 1 1 1 1 1
U0 1 −1 1 −1 1 1 −1
V 4 2 1 0 −1 0 −1
V0 4 −2 1 0 −1 0 1
The standard representation is irreducible, since
1  2 
(χV , χV ) = 4 + 10 · 22 + 20 + 24 + 20 · (−1)2
120
1  
= 16 + 40 + 64 = 1.
120
Thus, V 0 = V ⊗U 0 is irreducible too, thanks to Exercise 8. We claim now that V is irreducible.
V2

Its character can be computed with


1 
χ∧2 V (g) = χV (g)2 − χV (g 2 ) .
2
after having calculated the conjugacy class of every elements g 2 . On the whole

id (1 2) (1 2 3) (1 2 3 4) (1 2 3 4 5) (1 2)(3 4) (1 2 3)(4 5)
1 10 20 30 24 15 20
6 0 0 0 1 0
V2
V −2
and we can check that it is irreducible:
1  2 
(χ∧2 V , χ∧2 V ) = 6 + 24 + 15 · (−2)2
120
1  
= 36 + 24 + 60 = 1.
120
The tensor product of 2 V and the sign representation is useless, since 2 V ⊗ U 0 ' 2 V (look
V V V

at the character table). We can see that two missing representations must have dimension 5.
Indeed, calling a and b their dimensions, we have
a2 + b2 = |S5 | − 1 − 1 − 42 − 42 − 62 = 50.
The couples whose squares sums to 50 are (5, 5) and (1, 7). According to Theorem 33, the latter
must be discarded since 7 - |S5 |. Thus, the spaces has both dimension 5. We can see that they
comes, in a certain way, from Sym2 V . Its character can be computed via
1 
χSym2 V (g) = χV (g)2 + χV (g 2 ) .
2
id (1 2) (1 2 3) (1 2 3 4) (1 2 3 4 5) (1 2)(3 4) (1 2 3)(4 5)
1 10 20 30 24 15 20
Sym2 V 10 4 1 0 0 2 1

We can see that Sym2 V contains a copy of U and V :


(χSym2 V , χU ) = 1
(χSym2 V , χV ) = 1.

The remaining space W has dimension dim Sym2 V − dim U − dim V = 5, which is exactly the
dimension of the space we are looking for. Its character is

36
3.2. Partitions, Young diagrams and tables of S5 and A5

id (1 2) (1 2 3) (1 2 3 4) (1 2 3 4 5) (1 2)(3 4) (1 2 3)(4 5)
1 10 20 30 24 15 20
W 5 1 −1 −1 0 1 1

The last irreducible representation is simply W 0 = W ⊗ U 0 . Eventually, the character table of


S5 will be

id (1 2) (1 2 3) (1 2 3 4) (1 2 3 4 5) (1 2)(3 4) (1 2 3)(4 5)

1 10 20 30 24 15 20
U 1 1 1 1 1 1 1
U0 1 −1 1 −1 1 1 −1
V 4 2 1 0 −1 0 −1
V0 4 −2 1 0 −1 0 1
6 0 0 0 1 0
V2
V −2
W 5 1 −1 −1 0 1 1
W0 5 −1 −1 1 0 1 −1

Remark 16. In general, there is no correspondence between irreducible representations and


conjugacy classes. But in the particular case of Sn we will find a correspondence between
representations and Young diagrams. For S5 , the correspondence will be

U, U 0 ←→ , W, W 0 ←→ ,

^2
V, V 0 ←→ , V ←→ .

Note the “primed duality” and the flipped diagrams. We will investigate the correspondence in
Chapter 7, after a digression about symmetric functions.

We can compute now the irreducible representations of A5 . We still have the trivial repre-
sentation, the standard representation and the 5-dimensional representation W .

id (1 2 3) (1 2)(3 4) (1 2 3 4 5) (2 1 3 4 5)
1 20 15 12 12
U 1 1 1 1 1
V 4 1 0 −1 −1
W 5 −1 1 0 0

We miss two representations, say Y and Z, of dimension a and b, such that

a2 + b2 = |A5 | − 1 − 42 − 52 = 18

Thus, we must have a = b = 3. Now, noting that 2 V is no longer an irreducible representation,


V

it is easy to see that it not contains a copy of U . Indeed,


1
(χ∧2 V , χU ) = (6 + 12 + 12 − 30) = 0.
60

37
3.3. Alternating powers of Sn

Thus, due to dimension reasons, 2 V must contain two irreducible representations of dimension
V

3. They cannot be the same, otherwise the character of 2 V would be even. Therefore
V

χY + χZ = χ∧2 V .

Setting

id (1 2 3) (1 2)(3 4) (1 2 3 4 5) (2 1 3 4 5)
1 20 15 12 12
Y 3 α β ϕ ψ
Z 3 −α −2 − β 1−ϕ 1−ψ

and imposing the orthogonality between Y and the previous representations we find
 
3 + 20α + 15β + 12ϕ + 12ψ = 0 α = 0

 

12 + 20α − 12ϕ − 12ψ = 0 =⇒ β = −1
 
15 − 20α + 15β = 0
 ϕ + ψ = 1.

Supposing ϕ real and imposing the orthogonality between Y and Z, we obtain

0 = 60 (χY , χZ ) = 9 + 15 · (−1)2 + 12 ϕ(1 − ϕ) + 12(1 − ϕ)ϕ


 
= 24 1 + ϕ − ϕ2 .

The solution is the gold ratio



1± 5
ϕ= .
2
We can take the positive solution (the other one corresponds to interchanging Y and Z), ob-
taining the table character of A5 .

id (1 2 3) (1 2)(3 4) (1 2 3 4 5) (2 1 3 4 5)
1 20 15 12 12
U 1 1 1 1√
1 √
Y 3 0 −1 1+ 5
2√
1− 5
2√
Z 3 0 −1 1− 5
2
1+ 5
2
V 4 1 0 −1 −1
W 5 −1 1 0 0

3.3 Alternating powers of Sn


In the following, we will use this fact. Let V and U vector spaces. Then

^k k ^ ^k−l 
M l
V ⊕U ' V ⊗ U .
l=0

Theorem 34. Let V be the standard representation of Sn . The alternating powers V are
Vk

irreducible, with 0 ≤ k ≤ n − 1.

38
3.3. Alternating powers of Sn

Proof. Let W be the defining representation of Sn . We know that W ' U ⊕ V , where U is the
trivial and V is the standard representation. According to the above lemma
^k ^ ^0  ^ ^1 
k k−1
W ' V ⊗ U ⊕ V ⊗ U .

since i U = 0 for every i ≥ 2. In particular, since of every representation is U and of


V V0 V1

every representation is itself, we have that


^k ^  ^ 
k k−1 ^k ^k−1
W ' V ⊗U ⊕ V ⊗U ' V ⊕ V.

In order to show that (χ∧k V , χ∧k V ) = 1, we have just to prove that

(χ∧k W , χ∧k W ) = 2 k = 1, . . . , n.

Indeed, V = U is irreducible, while if V is irreducible, then


V0 Vk−1

2 = (χ∧k V , χ∧k V ) + (χ∧k−1 V , χ∧k V ) + (χ∧k V , χ∧k−1 V ) + (χ∧k−1 V , χ∧k−1 V )


= (χ∧k V , χ∧k V ) + 2a + 1,

where a is the multiplicity of V in V . From the disequation


Vk−1 Vk

0 ≤ (χ∧k V , χ∧k V ) = 1 − 2a,

and from the fact that a is a natural number, I obtain that a = 0 and so the thesis.

So let e1 , . . . , en be a basis of W and ei1 ∧ · · · ∧ eik , with i1 < · · · < ik , be a basis of W.


Vk

Let us take
B ⊆ { 1, . . . , n } , #B = k.
We define [B] as the set B with ordered elements:

[B] = (i1 , . . . , ik ), where i1 < · · · < ik .

We have to compute the trace of the permutation matrices

g · e[B] = ±e[gB] ,

where the sign is given by the reordering of the elements of gB. Since we are interest in
computing the trace, note that diagonal elements are non zero only for gB = B and in that case
the diagonal elements will be given by g · e[B] = sgn(g|B )e[B] . Hence,

χ∧k W (g) = sgn(g|B ).


X

B⊆{ 1,...,n }
#B=k, gB=B

Thus, we obtain
1 X 2
(χ∧k W , χ∧k W ) = χ k (g)
n! g∈S ∧ W
n

1 X
= sgn(g|B ) sgn(g|C )
X X
n! g∈S B⊆{ 1,...,n } C⊆{ 1,...,n }
n
#B=k, gB=B #C=k, gC=C
1
= sgn(g|B ) sgn(g|C ).
X X
n! B,C gB=B
#B=#C=k gC=C

39
3.3. Alternating powers of Sn

Since g fixes B and C, it also fixes B ∩ C and ¬(B ∪ C). Setting


l = #B ∩ C m = #¬(B ∪ C) k = #B = #C,
we have two cases.
• If k = l, then B = C and sgn · sgn = sgn2 = 1. Then
1 = |{z} (n − k)!
X
k!
g∈Sn
| {z }
elements elements that
gB=B that fix B fix ¬B

an we find
1 X 1 n
!
1= k!(n − k)! = 1.
X
n! B=C gB=B
n! k
#B=k

• If k < l, then B 6= C. Splitting g as g = abcd (see the figure), we have that the actions of
a, b, c and d commute, since they act on disjoint sets.

c
b
a

B C

Noting that g|B = ab e g|C = ac, we find

sgn(ab) sgn(ac) = sgn2 (a) sgn(b) sgn(c)


X X
| {z }
a,b,c,d a,b,c,d 1
= sgn(b) sgn(c)
X

a,b,c,d

= 1 sgn b sgn c 1.
X X X X

a∈Sl b∈Sk−l c∈Sk−l d∈Sm


| {z } | {z }
l! m!

The sums of the signs are zero, except for k − l = 1 where the sum is 1. Thus,

l! m! k = l + 1
(
sgn(ab) sgn(ac) =
X

a,b,c,d 0 k=6 l + 1.

Let us call
S = { B, C ⊂ { 1, . . . , n } | #B = #C = k, #B ∩ C = k − 1 } .
The only contribution to the initial sum for k < l is
1 1
!
n
#S l! m! = (n − k + 1) (n − k) (k − 1)!(n − k − 1)! = 1.
n! n! k − 1 | {z } | {z }
| {z } choices for choices for
choices B \ (B ∩ C) C \ (B ∩ C)
for B ∩ C

40
3.3. Alternating powers of Sn

The total contribution for (χ∧k W , χ∧k W ) comes from both cases k = l and k < l, so that
(χ∧k W , χ∧k W ) = 1 + 1 = 2.

Remark 17. Note that in the proof we had demonstrated something more: V and
Vk Vk−1
V
are distinct for every k.
Corollary 35. The standard V = V is always irreducible.
V1

We can now prove a duality for the alternating powers of the standard representation. The
statement will be clarified in Chapter 7 in connection with the Young diagrams.
Proposition 36. Let V be the standard and U 0 the sign representation of Sn . Then
^k ^n−k−1
V ' V ⊗ U 0.

Proof. First of all, note that eB is a basis of eigenvectors for W if and only if e¬B is a basis
Vk

of eigenvectors for n−k W . Thus,


V

χ∧k W (g) = sgn(g|B) χ∧n−k W (g) = sgn(g|¬B).


X X

gB=B g¬B=¬B

Since sgn(g|B) sgn(g|¬B) = sgn(g), we have


χ∧n−k W (g) = sgn(g)χ∧k W (g)
that is
n−k k
W = W ⊗ U0 (3.1)
^ ^

Let us now suppose that the the thesis holds true for k − 1. This means that
 
k k k−1 k n−(k−1)−1
W = V = V ⊗ U 0
^ ^ ^ ^ ^
V ⊕ V ⊕

k n−k
!
= V ⊗ U0
^ ^
V ⊕

Noting that U 0 ⊗ U 0 = U , if I tensorize both sides with U 0 I obtain


k k n−k
! !
0 0
W ⊗U =
^ ^ ^
V ⊗U ⊕ V ⊗U

that is (according to 3.1)


n−k k n−k
!
0
W =
^ ^ ^
V ⊗U ⊕ V

On the other hand


n−k n−k n−k−1
W =
^ ^ ^
V ⊕ V
thus
n−k−1 k
V = V ⊗ U0
^ ^

If I tensorize again both sides with U 0 I obtain the thesis.

It must be said that the symmetric power case is quite different. Indeed, Symk V is almost
never irreducible.

41
Chapter 4

Real and quaternionic


representations

In the previous chapter we have defined the character as a complex-valued function. The complex
field is fundamental, since it is algebraically closed and this fact leads us to the demonstration
of Schur’s lemma. However, the examples examined until now had shown quite only real-valued
characters. The following question naturally arises:
If the character is real valued, is the representation real in the sense that ρ : G → GLn (R)?
And what about vice versa?
Let us start this analysis observing that every action G on a real vector space V0 extends to
an action of G on
V = V0 ⊗R C
setting
g · (v ⊗ λ) = (g · v) ⊗ λ.

Definition 15. Let V be a complex vector space and G a finite group. A representation of G
on V is said to be real if there exist a real vector space V0 and a homomorphism G → GL(V0 )
such that
V0 ⊗R C ' V
as representations.

Since tensoring with C over R is just extending scalars, the character of any real represen-
tation is real. We may naively say that also the converse holds true. Unfortunately, this is not
the case. We will show a theoretical example in the following section.

4.1 Hamiltonian quaternions and SU(2)


Let
H = R[i, j, k] = { x = a + bi + cj + dk | a, b, c, d ∈ R } ,
which is a linear space over R. Defining

i2 = j 2 = k 2 = −1

ij = −ji = k,

43
4.1. Hamiltonian quaternions and SU(2)

we obtain a non-commutative division algebra. The elements of H are called quaternions. We


define the conjugate quaternion as

x = a − bi − cj − dk

and the norm as


n(x) = xx = a2 + b2 + c2 + d2 .
If x 6= 0, then the inverse quaternion is
x
x−1 = .
n(x)
We can also define the trace of a quaternion as

tr x = x + x̄ = 2a.

An important subset of the quaternions is

H1 = { x ∈ H | n(x) = 1 } ,

which forms a group with the multiplication.


Proposition 37. There is a bijection
H1 ' S 3 .
Thus, the 3-sphere has a (non-commutative) group structure.
Proof. Simply note that
n o n o
a + bi + cj + dk a2 + b2 + c2 + d2 = 1 ' (a, b, c, d) ∈ R4 a2 + b2 + c2 + d2 = 1 .

Another important subgroup of the quaternions is

Q8 = { ±1, ±i, ±j, ±k } ,

which is called the quaternion group.


Proposition 38. The following isomorphism holds true:
Q8
Z(Q8 ) ' Z2 × Z2 .
Proof. Follows from the fact that Z(Q8 ) = { ±1 }.

Remark 18. Q8 is an example of non-commutative group with only normal subgroups. In


order to calculate the subgroups, we just have to compute the powers of the generators

h1i = { 1 }
h−1i = { 1, −1 }
hii = { 1, i, −1, −i } = h−ii
hji = { 1, j, −1, −j } = h−ji
hki = { 1, k, −1, −k } = h−ki
hi, ji = { 1, i, −1, −i, j, −j, k, −k } = hi, ki = hj, ki

It is easy to see that gHg −1 ∈ H for all g ∈ Q8 .

44
4.1. Hamiltonian quaternions and SU(2)

Proposition 39. Every U ∈ SU(2) has the form


!
a b
U= a, b ∈ C.
−b a

In particular, every representation G −→ SU (2) has a real-valued character.


Proof. Let !
a b
U=
c d
Imposing U ∈ SU(2), we obtain
( T
U = U −1
! !
a c d −b
=⇒ =
det U = 1 b d −c a.

Therefore, a = d and −b = c. Eventually tr U = a + a ∈ R.

Lemma 40. SO(2) is an abelian group.


Proof. Following the proof of the above lemma, it is easy to see that every O ∈ SO(2) has the
form !
cos α − sin α
O= .
sin α cos α
By direct calculation, we can see that this matrices commute. Another way to prove the state-
ment is to observe that
SO(2) ' S 1 ,
with the rotation determined by an angle which can be identified as an arch of S 1 .

Lemma 41. The following isomorphism holds true:

SU(2) ' H1 .

Proof. We can think of the quaternions as

H = C ⊕ Cj = { a + bj | a, b ∈ C } .

In such a way, the product will be

(a + bj)(c + dj) = ac − bd + (ad + bc)j.

On the other hand, the product in SU(2) is


! ! !
a b c d ac − bd ad + bc
=
−b a −d c −ad − bc ac − bd

and the above isomorphism will be


!
a b
a + bj 7−→ .
−b a

Lemma 42. There exists a finite and non-abelian subgroup of SU(2).

45
4.1. Hamiltonian quaternions and SU(2)

Proof. Q8 is a finite non-abelian subgroup of H1 . The image of Q8 through the above isomor-
phism is the required subgroup. Just for completeness,
! !
1 0 i 0
1 7−→ id = i 7−→ iσ3 =
0 1 0 −i
! !
0 1 0 i
j 7−→ iσ2 = k 7−→ iσ1 = ,
−1 0 i 0

where σi are the Pauli matrices.

We are ready to give the example of a real-valued character associated to a non-real repre-
sentation.

Proposition 43. Let G be a non-abelian finite subgroup of SU(2). Then every 2-dimensional
representation
G ,→ SU(2) ⊂ GL2 (C)
cannot be real.

Proof. Let us suppose that the representation is real. Then

G ,→ SO(2) ⊂ SU(2).

However, SO(2) is abelian, while G is not.

Definition 16. Let V be a representation of G. We say that V is a quaternionic represen-


tation if there exists a G-antilinear map ψ : V → V such that ψ 2 = − idV . A representation V
which is neither real nor quaternionic is said to be complex.

Remark 19. According to the definition, a quaternionic representation V has the structure of
a H-module, where i ∈ H acts as left multiplication by i ∈ C, j acts as ψ and k acts as the
composition of j and i:

i : v 7−→ i v
j : v 7−→ ψ(v)
k : v 7−→ iψ(v).

Example 6. The quaternionic group is the prototype of the quaternionic representations. Let us
compute the character table of Q8 . The conjugacy classes are { 1 } , { −1 } , { ±i } , { ±j } , { ±k }.
We already know the trivial representation U and the SU(2) representation, with characters

1 −1 ±i ±j ±k
U 1 1 1 1 1
SU(2) 2 −2 0 0 0

With the usual calculation, we can find that the three missing representations are all one-
dimensional. On the other hand, we have the exact short sequence

1 −→ [Q8 , Q8 ] = { ±1 } −→ Q8 −→ Q8[Q , Q ] ' Z2 × Z2 −→ 1.


8 8

The Klein group Z2 × Z2 has non-trivial elements of order 2, so it is easy to see that the only
(one-dimensional) representations are

46
4.1. Hamiltonian quaternions and SU(2)

Z2 × Z2 (0, 0) (1, 0) (0, 1) (1, 1)


U 1 1 1 1
X 1 −1 −1 1
Y 1 −1 1 −1
Z 1 1 −1 −1
The above exact sequence gives us the remaining representations. For example, X is given by
±1 7−→ (0, 0) 7−→ 1
±i 7−→ (1, 0) 7−→ −1
±j 7−→ (0, 1) 7−→ −1
±k 7−→ (1, 1) 7−→ 1.
Eventually, the character table of the quaternionic group is
1 −1 ±i ±j ±k
U 1 1 1 1 1
X 1 1 −1 −1 1
Y 1 1 −1 1 −1
Z 1 1 1 −1 −1
SU(2) 2 −2 0 0 0
Remark 20. We have seen that the character is a total invariant for representations of a
group G. Is it still true if I change the group? That is, does the character determine the group?
Unfortunately turns out that Q8 and D8 (the dihedral group of order 8) have the same character
table, but the 2-dimensional representations are not isomorphic.
So, let us calculate the character table of the dihedral group of order 8, that is the set of
symmetries of the square. It has a presentation
D8 = hσ, τ | σ 2 = τ 4 = 1, στ σ = τ −1 i ,
where σ is a reflection an τ a rotation of π2 . We have the trivial representation U and the
2-dimensional representation O(2), as we can see geometrically. The one-dimensional represen-
tations can be found from the exact sequence
n o
1 −→ id, τ 2 ' { ±1 } −→ D8 −→ D8{ id, τ 2 } ' Z2 × Z2 −→ 1

The final result will be


id τ2 σ τ στ
1 1 2 2 2
U 1 1 1 1 1
X 1 1 −1 −1 1
Y 1 1 −1 1 −1
Z 1 1 1 −1 −1
O(2) 2 −2 0 0 0
We can see that the character table is the same as that of Q8 , but with a concrete difference:
O(2) ⊂ GL2 (R) is a real representation, while SU(2) is not (Proposition 43). We will see
in the following section that this avoids the existence of a G-linear isomorphism between the
representations.
We just have seen two types of structure for a representation, i.e real or quaternionic. In
the following sections, we will characterize these types of representations.

47
4.2. Bilinear forms fixed by G

4.2 Bilinear forms fixed by G


Let V be a vector space and let us define

Bil(V ) = { B : V × V → C | B is bilinear }

Definition 17. A bilinear form B : V × V → C is said non-degenerate if

B(u, v) = 0 ∀v ⇔ u = 0.

Lemma 44. Let V a representation of G. The map

Φ : Hom(V, V ∗ ) −→ Bil(V )
ϕ 7−→ Bϕ (u, v) = ϕ(u)v

is a canonical isomorphism. Furthermore

1. ϕ ∈ Hom is G-stable if and only if Bϕ is G-stable.

2. ϕ ∈ Hom is an isomorphism (of representations) if and only if Bϕ is non-degenerate.

Proof. Since dim Hom(V, V ∗ ) = dim Bil(V ), we have just to prove that Φ is surjective. However
this is trivial: if we fix a bilinear map B : V × V −→ C we can define the homomorphism

V −→ V ∗
v 7−→ B(v, ·).

For what concerns the first point, B fixes G since ϕ do it:

B(g · u, g · v) = ϕ(g · u)g · v = g −1 · ϕ(u)g · v


 
= ϕ(u) g −1 gv = ϕ(u)v
= B(u, v).

and the other implication is similar.


For what concerns the second points, let us note that if ϕ is an isomorphism then :

Bϕ (u, v) = 0 ∀v ∈ V ⇒ ϕ(u)v = 0 ∀v ∈ V ⇒ ϕ(u) = 0 ⇒ u = 0.

that is Bϕ is non-degenerate. On the other hand, if exists u such that ϕ(u) = 0 then
Bϕ (u, v) = ϕ(u)v = 0 for all v ∈ V . But the bilinear form is non-degenerate so u = 0.

Proposition 45. A representation V has real-valued character if and only if there exists a
non-degenerate bilinear form on V fixed by G. In addition, if V is irreducible, the bilinear form
is symmetric or skew-symmetric and unique up to scalars.

Proof. V has a real-valued character if and only if χV = χV . Since χV = χV ∗ , this happens if


and only if exists a G-linear isomorphism ϕ : V −→ V ∗ . The above lemma concludes.
Let us now prove the second statement. Assuming V irreducible, for Schur’s lemma, ϕ : V →
V ∗ is unique up to scalars. Indeed, if ψ is another G-stable linear map, the composition

ϕ ψ −1
V V∗ V

48
4.2. Bilinear forms fixed by G

is of the form  idV for some  ∈ C, that is ϕ = ψ. Since ϕ is unique up to scalars, also Bϕ is
unique up to scalars.
Consider now the non-degenerate bilinear form on V fixed by G given by

C(u, v) = B(v, u).

According to above lemma, there exists ψ such that C(u, v) = ψ(u)v. Then

B(u, v) = C(v, u) = 2 B(u, v).

Hence, 2 = 1 and B is symmetric if  = 1, skew-symmetric if  = −1.

We will see that the existence of a symmetric or skew-symmetric G-stable bilinear form
characterize the different types of representations. In the following, it will be crucial the existence
of a Hermitian product H fixed by G.

Lemma 46. Let V be a representation with real-valued character. There exists a G-antilinear
isomorphism ψ : V → V . In particular, ψ 2 : V → V is a G-linear isomorphism.

Proof. First of all, let us define the antilinear map

η : V 7−→ V ∗
u 7−→ H(u, ·).

In particular, η is a G-stable isomorphism. Indeed,

η(g · u)(v) = H(g · u, v) = H(u, g −1 · v)


= η(u)(g −1 · v).

Further, η is injective since H is non-degenerate

η(u) = 0 ⇔ H(u, v) = 0 ∀v ∈ V ⇔ u = 0,

and we conclude that it is bijective since V and V ∗ has the same dimension.
Now, suppose V with real-valued character and let ϕ : V → V ∗ be a G-stable isomorphism.
The composition
ψ = η −1 ◦ ϕ : V −→ V
is the wanted G-antilinear isomorphism.

Proposition 47. An irreducible representation V of G is real if and only if there exists a


non-degenerate symmetric bilinear form on V preserved by G.

Proof. Say V = V0 ⊗ C. Let S0 be a scalar product on V0 and set


1 X
B(u, v) = S0 (gu, gv),
|G| g∈G

which is a positive-definite symmetric bilinear form on V0 that is preserved by G. Let us now


extend B to V0 ⊗ C setting
B(u ⊗ λ, v ⊗ µ) = λµ B(u, v)
It is easy to see that B verifies the desired properties.

49
4.2. Bilinear forms fixed by G

Suppose now that we have a bilinear form as in the hypotheses. Then V has real-valued
character and there exists ψ as in the previous lemma. For Schur, ψ 2 = λ idV . The map

B : V × V 7−→ C
(u, v) 7−→ H(ψ(u), v)

is a G-stable non-degenerate bilinear form. For Proposition 45 such bilinear form is unique, so
that B must be symmetric. Then

B(ψ(u), u) = H(ψ 2 (u), u) = λH(u, u),

while for symmetry


B(ψ(u), u) = B(u, ψ(u)) = H(ψ(u), ψ(u)).
If u 6= 0,
H(ψ(u), ψ(u))
λ= > 0,
H(u, u)
that is λ is real and positive. Replacing ψ with √ψλ , we get ψ 2 = idV . So ψ is diagonalizable
with eigenvalues ±1 and
V = V + ⊕ V −,
where V ± is the eigenspace associated to ±1. We can see that V ± = iV ∓ , since ψ is antilinear.
Indeed, if v ∈ V ±
ψ(iv) = −iψ(v) = −i(±v) = ∓iv.
In particular, V = V + ⊕ iV + . Viewing V as a real vector space, ψ is R-linear (the conjugation
does not count) and dimR V + = dimR V − = dimC V , so that

V = V + ⊗R C.

Proposition 48. An irreducible representation V of G is quaternionic if and only if there exists


a non-degenerate skew-symmetric bilinear form on V preserved by G.
Proof. Suppose V quaternionic and let ψ be the G-antilinear map such that ψ 2 = − idV . Let H
be a Hermitian form on V fixed by G and define

B(u, v) = H(ψ(u), v).

The map B is bilinear. Indeed, it is linear in the second variable, while for the first one

B(λu + µv, ·) = H(ψ(λu + µv), ·) = H(λψ(u) + µψ(v), ·)


= λH(ψ(u), ·) + µH(ψ(v), ·)
= λB(u, ·) + µB(v, ·).

Let us see that B is non-degenerate. If B(u, v) = 0 for all v ∈ V , then H(ψ(u), v) = 0 for all
v ∈ V . However H is non-degenerate, so ψ(v) = 0. Since ψ is invertible (its inverse is ψ 3 ), we
conclude that v = 0. Similarly, the bilinear form C defined as

C(u, v) = B(v, u)

is a G-stable non-degenerate bilinear form. For Schur’s lemma, that B and C differ by a scalar
λ, that is
B(u, v) = λB(v, u).

50
4.3. The Schur indicator

Let us prove that λ = −1. From the definition of B, we have

B(u, v) = λC(v, u) = λ2 B(u, v).

Thus λ2 = 1. To conclude, observe that taking v = ψ(u)

H(ψ(u), ψ(u)) = λH(ψ 2 (u), u) = −λH(u, u).

However H is positive definite, so λ is real and negative, that is λ = −1.


Suppose now that we have a bilinear form as in the hypotheses. Then V has real-valued
character and there exists ψ as in the previous lemma. For Schur, ψ 2 = λ idV . The map

B : V × V 7−→ C
(u, v) 7−→ H(ψ(u), v)

is a G-stable non-degenerate bilinear form. For Proposition 45 such bilinear form is unique, so
that B must be skew-symmetric. Then

B(ψ(u), u) = H(ψ 2 (u), u) = λH(u, u),

while for antisymmetry

B(ψ(u), u) = −B(u, ψ(u)) = −H(ψ(u), ψ(u)).

If u 6= 0,
H(ψ(u), ψ(u))
λ=− < 0,
H(u, u)

that is λ is real and negative. Replacing ψ with √ψ , we get ψ 2 = − idV , that is V is quater-
|λ|
nionic.

Summarizing, an irreducible representation V of G is

• real if and only if there exists a symmetric G-stable bilinear form,

• quaternionic if and only if there exists a skew-symmetric G-stable bilinear form,

• complex if and only if there are no G-stable bilinear forms.

4.3 The Schur indicator


In this section, we will introduce a tool which characterize the different types of representations.

Definition 18. Let V be a representation of G. The Schur’s indicator for V is defined as

1 X
(V ) = χ (g 2 ).
|G| g∈G V

In order to involve the bilinear forms we discussed before, let us remember Formula 1.2. It easy
to see that
 G ^2 G
(V ∗ ⊗ V ∗ )G ' Sym2 V ∗ ⊕ V∗ .

51
4.3. The Schur indicator

Lemma 49. The following dimension formulae holds true:


 G 1h i
dim Sym2 V ∗ = dim (V ∗ ⊗ V ∗ )G + (V ) (4.1)
2
^ 2 1h G i
dim dim (V ∗ ⊗ V ∗ )G − (V ) .
V∗ = (4.2)
2
Proof. Let us use the first projection formula. For the first relation,
 G 1 X χ(g)2 + χ(g 2 ) 1 X (V )
dim Sym2 V ∗ = = χ(g)2 + ,
|G| g∈G 2 2|G| g∈G 2

while for the second one


^ 2 G 1 X χ(g)2 − χ(g 2 ) 1 X (V )
dim V∗ = = χ(g)2 − .
|G| g∈G 2 2|G| g∈G 2

Thus, we have
 G ^ 2 G 1 X
dim (V ∗ ⊗ V ∗ )G = dim Sym2 V ∗ + dim V∗ = χ(g)2
|G| g∈G

and this concludes the proof.

Theorem 50. An irreducible representation V of G is


• real if and only if (V ) = 1,
• quaternionic if and only if (V ) = −1,
• complex if and only if (V ) = 0.
Proof. Suppose first that χV is real valued. From Proposition 45 we know that

dim (V ∗ ⊗ V ∗ )G = 1,

since the bilinear form is unique up to scalars.


• According to Proposition 47, V is a real representation if and only if the bilinear form is
symmetric. This means that
 G ^ 2 G
dim Sym2 V ∗ =1 dim V∗ = 0.

Using Formula 4.1 we have


1 (V )
1= + ,
2 2
that is (V ) = 1.
• According to Proposition 48, V is a quaternionic representation if and only if the bilinear
form is skew-symmetric. This means that
 G ^ 2 G
dim Sym2 V ∗ =0 dim V∗ = 1.

Using Formula 4.2 we have


1 (V )
1= − ,
2 2
that is (V ) = −1.

52
4.3. The Schur indicator

Suppose now χV is not real valued. According to Proposition 45, there is no non-degenerate
bilinear form on V fixed by G. This means that
^2 G  G
dim V∗ = dim Sym2 V ∗ = dim (V ∗ ⊗ V ∗ )G = 0,

that is (V ) = 0.

The following table summarizes the results for an irreducible representation V of G.

Type of representation χV (g) non-degenerate bilinear (V )


form fixed by G?
Real real yes, symmetric 1
Quaternionic real yes, skew-symmetric −1
Complex not all real no 0

Exercise 11. All the representations of G are real if and only if every element of G is conjugate
to its inverse.

Exercise 12. Let V be an irreducible representation. Then V ⊗ V ∗ is a real representation.

Exercise 13. Let V, W be quaternionic or real representations. Then V ⊗ W is a real repre-


sentation.

Exercise 14. Let V be a real representation. Then V is a real representation. Let V be a


V2

quaternionic representation. Then k V is


V

1. real if k is even,

2. quaternionic if k is odd.

Exercise 15. Let G be a group of odd order. Every non-trivial irreducible representation is
complex.

Exercise 16. Every 1-dimensional representation cannot be quaternionic.

53
Chapter 5

Finite subgroups of SO(3) and SU(2)

The aim of this section is to classify all finite subgroups of SO(3) and, as a consequence, of
SU(2). This classification will concretely realize some real representations of S4 , A4 and A5
thanks to the five platonic solids:

Figure 5.1: Form left to right: tetrahedron, cube, octahedron, icosahedron and dodecahedron.

5.1 Finite subgroups of SO(3)


Theorem 51. Let SO(3) be the group of the rotations of R3 and let G ⊆ SO(3) be a finite
group (with |G| = n). We have the following possibilities:

• G is the symmetry group of a regular n-gon in R2 (that is isomorphic to the cyclic group
Cn of order n).

• The order of G is even and G is the symmetry group of a regular n/2-gon in R3 (that is
the diedral group of order n).

• G is the symmetry group T of the tetrahedron (that is isomorphic to A4 ).

• G is the symmetry group O of the octahedron (that is the same group of the cube and is
isomorphic to S4 ).

• G is the symmetry group I of the icosahedron (that is the same of the dodecahedron and
is isomorphic to A5 ).

Proof. In order to prove the statement, we will proceed in stages.

1. We will show that G acts on the set of poles of S2 fixed by the rotations of G.

2. We will prove that the action must have either 2 or 3 distinct orbits (if G is not trivial).

55
5.1. Finite subgroups of SO(3)

3. The two-orbits case will lead us to the cyclic group, while the three-orbits case will lead
to four subcases, corresponding to the remaining groups.

Let us demonstrate the above points.


1. Consider an element g 6= id in G. It acts on R3 as a rotation along an axis through the
origin. Such line intersects the unit sphere S2 in two points, the poles of g.
g

We define P to be the set of poles of the elements of G \ { id }. Take now p ∈ P, a pole of h,


and an element g ∈ G. Then we have

g(p) = g(h(p)) = (ghg −1 )(g(p)),

so that g(p) is fixed by ghg −1 . Since h =


6 id, then ghg −1 6= id and g(p) is a pole of ghg −1 . In
particular, this defines an action G × P → P.
2. Let n=|G| and np the number of elements of G which fix p. Then

1X
(np − 1) = n − 1,
2 p∈P

since the number of elements of G which fixes p (rather than the identity) summed up for every
poles and divided by two counts the number of elements in G \ { id }. The factor 21 is due to
the fact that the poles are double counted. Let now N be the number of distinct orbits of the
action and choose a pole pi for each orbit. We will have

npi = | StabG (pi )|.

On the other hand, the cardinality of the orbit of pi is


n
#G[pi ] = [G : StabG (pi )] = ,
np i

so that the above formula becomes


N
1X n
(npi − 1) = n − 1,
2 i=1 npi

that is
N
1 1
  !
2 1− = 1− (5.1)
X
.
n i=1
np i

If G is trivial, then it is cyclic (G = hidi). If G is not trivial, then n ≥ 2 and

N N
1 1
  !
1≤2 1− = 1− 1=N
X X
<
n i=1
npi i=1

56
5.1. Finite subgroups of SO(3)

so that N ≥ 2. Furthermore, we have npi ≥ 2 since every stabilizer contains the identity and at
least a non-trivial rotation. This implies that 12 ≤ 1 − n1p and
i

N
1 1
!
N
 
1− =2 1− <2
X

2 i=1
np i n

thus N < 4.
3. Let us consider the two-orbits case. Equation 5.1 becomes

2 1 1
= + ,
n np 1 np 2

so that n = np1 = np2 . In this case, there is a single axis of rotation. In particular, we have
a rotation of the plane through the origin perpendicular to the axis. In this case, G ⊂ SO(2).
The finite subgroups of SO(2) are cyclic, generated by a rotation of 2πn .
For N = 3, Equation 5.1 becomes

2 1 1 1
1+ = + + . (5.2)
n np 1 np 2 np 3

We can suppose np1 ≤ np2 ≤ np3 . For the reason we discussed before, we have np1 ≥ 2. On the
other hand, the left hand side of Equation 5.2 is greater that 1, so that np1 ≤ 2. Thus, np1 = 2
and the equation becomes
1 2 1 1
+ = + .
2 n np 2 np 3
In order to determine np2 , note that

1 1 2 1 1 2
< + = + ≤ .
2 2 n np 2 np 3 np 2

Hence, np2 can be 2 or 3. If np2 = 2, than n = 2np3 . In particular, every np3 ≥ 2 gives a
solution. Eventually, if np2 = 3 the equation becomes

1 1 2 1
< + = .
6 6 n np 3

Thus, 2 ≤ np3 ≤ 5.

The last step of the demonstration is the concrete realization of the subgroups of SO(3) with
the above properties (indeed, we have only found some constraints on npi , but no one assures
us that such groups exist).
For N = 2, as said before, we have the rotations of 2π n around an axis through the origin,
that is the rotations of a regular n-gon in R around its centre. They form the cyclic group Cn .
3

For what concerns the case N = 3, from equation 5.2 we know that the order of the group is
!−1
1 1 1
n=2 + + −1
np 1 np2 np 3

In particular, according to the constraints on npi , the groups we are going to build are summa-
rized in the following table.

57
5.1. Finite subgroups of SO(3)

np 1 np 2 np3 n G
2 2 m 2m D2m
2 3 3 12 A4
2 3 4 24 S4
2 3 5 60 A5

Let us construct the action of such groups on R3 .


a) Consider a regular m-gon with centre in the origin. We define

• σ to be the rotation of π around an axis passing through two opposite edges,

• τ to be the rotation of 2π
m around the axis orthogonal to the plane containing the m-gon
and passing through the origin.

Then it is clear that σ 2 = id, τ m = id and στ σ = τ −1 . Then the dihedral group

D2m = hσ, τ |σ 2 = τ m = id, στ σ = τ −1 i ,

is a finite subgroup of SO(3).

b) Consider a tetrahedron with centre in the origin. For example, take as vertices the points

1 1
       
−1 −1
v1 = 1 , v2 = −1 , v3 = −1 , v4 =  1  ,
       
1 −1 1 −1

so that = 0.
P4
i=1 vi
z
v4

v1

y
x v2
v3

We have now two kinds of rotations.

• The rotations RE of π around an axis passing through the middle points of two opposite
edges. They are three and of order 2.

• The rotations RV of 2π3 around an axis passing through a vertex and the centre of the
opposite face. They are four and of order 3.

58
5.1. Finite subgroups of SO(3)

Further, choosing two specific rotations RE and RV , we have (RE RV )3 = id. Setting

T = hRV , RE | RE
2
= RV3 = (RE RV )3 = idi ,

we obtain a finite subgroup of SO(3). The group permutes the four vertices of the tetrahedron,
preserving the parity. Indeed, the rotation around the axis through the edges v1 v2 and v3 v4
corresponds to
RE = (1 2)(3 4),
while the rotation around the vertex v1 is

RV = (1 2 3).

With this identifications, it follows that T ' A4 .


c) Consider a cube with centre in the origin. For example, take as vertices the points

1 1 1 1
       

v1 = 1 , v2 =  1  , v3 = −1 , v4 = −1 ,


       
1 −1 1 −1
       
−1 −1 −1 −1
v5 =  1  , v6 =  1  , v7 = −1 , v8 = −1 ,
       
1 −1 −1 1

so that = 0.
P4
i=1 vi

v8 z
v5
v4 2
3
v1 4
y
x v7 1
v6
v3
v2

We have three kinds of rotations.


• The rotations RE of π around an axis passing through the middle points of two opposite
edges. They are six and of order 2.

• The rotations RV of 2π3 around an axis passing through a vertex and the opposite one.
They are four and of order 3.

• The rotations RF of π
2 around the centres of a face and of the opposite one. They are
three and of order 4.
Further, choosing three specific rotations RE , RV and RF , we have RE RV RF = id. Setting

O = hRE , RV , RF | RE
2
= RV3 = RF4 = RE RV RF = idi ,

we obtain a finite subgroup of SO(3). The group permutes the four diagonals of the cube.
Indeed, with the labelling in figure, the rotation around the axis through the edges v1 v2 and
v7 v8 corresponds to
RE = (1 2),

59
5.2. Finite subgroups of SU(2)

the rotation around the vertices v4 is

RV = (1 2 3),

while the rotation centred in the face v1 v2 v3 v4 correspond to

RF = (1 2 3 4).

With this identifications, it follows that O ' S4 .


d) In a similar way, it can be show that the group of symmetry of the icosahedron corresponds
to the last case, for which I ' A5 . The objects which are permuted depends on the way we
embed the solid.

Eventually, the fact that for S4 and A5 the same group of symmetry corresponds to two different
solids follows from the duality shown in the picture below.

Remark 21. We can see from the pictures above that the tetrahedron is obtained from the cube
neglecting some of the vertices. As a consequence, we obtain the immersion A4 ⊂ S4 . Further,
we have constructed real 3-dimensional representations of the groups D2m , A4 , S4 and A5 . In
particular, we have the standard representations of A4 and S4 and one of the two 3-dimensional
irreducible representations of A5 , depending on the way we embed the icosahedron.

5.2 Finite subgroups of SU(2)


Thanks to the classification of finite subgroups of SO(3), we are able to classify the finite
subgroups of SU(2).

Theorem 52. Let Ĝ be a finite subgroup of SU(2). Then Ĝ is isomorphic either to

• a cyclic group Cn ,

• a binary dihedral group (or dicyclic group) D̂2m ,

• the binary tetrahedron group T̂ ,

• the binary octahedron group Ô,

• the binary icosahedron group Î.

60
5.2. Finite subgroups of SU(2)

Further, we have the 1-1 correspondence


n o
finite subgroups Ĝ of SU(2) containing − id ←→ { finite subgroup G of SO(3) }

via the isomorphism G → Ĝ{ ± id }.


In order to prove the theorem, we need the following
Lemma 53. The only element of SU(2) of order 2 is − id.
 
a b
Proof. According to Proposition 39, let us take U = −b a
∈ SU(2). If we impose
! !
a2 − |b|2 ab + ab 1 0
U = 2
=
−ab − ab a2 − |b|2 0 1

then a2 − |b|2 = a2 − |b|2 = 1, that is a2 = a2 . Then a = ±a, so that a ∈ R or a ∈ iR. However


a∈/ iR since, in this case
1 = a2 − |b|2 = −|a|2 − |b|2 = − det U = −1,
I deduce that a ∈ R \ { 0 } and
0 = ab + ab = ab + ab.
thus ab = 0, which implies b = 0. Eventually a2 = 1 and we obtain U = ± id. Since the identity
has order 1, we are left with − id.
Proof of Theorem 52. Let us consider the double cover

1 { ± id } SU(2) π
SO(3) ' SU(2){ ± id } 1.

If Ĝ is a finite subgroup of SU(2), then G = π(Ĝ) is a finite subgroup of SO(3).


• Let us suppose that |Ĝ| is odd. Then − id ∈
/ Ĝ by Lagrange’s theorem. As a consequence,
Ĝ ' G is a finite subgroup of SO(3) of odd order. From the previous theorem, the only
finite subgroups of SO(3) of odd order are the cyclic ones.
• Let us suppose that |Ĝ| is even. Then Ĝ contains an element of order 2. Indeed, the
identity is the only element of order 1. The remaining odd number of elements can be
paired with its inverse. Thus, an element must be paired with itself, that is it is of order
2. By the previous lemma, − id ∈ Ĝ and

G ' Ĝ{ ± id }.

In order to conclude the proof, we must show that Ĉn are cyclic groups of even order. Take
  

Ĝ = ei k = 1, . . . , 2n ,

n 0

0 e−i n

which is a finite subgroup of SU(2). Clearly, it is a cyclic group of order 2n. Further, the
isomorphism
G ' Ĝ{ ± id }.
gives us ( ! )
2kπ 2kπ
cos n
sin n
0
G' k = 1, . . . , n ,

− sin 2kπ
n
cos 2kπ
n
0

0 0 1
which is a cyclic group of order n.

61
Chapter 6

Induced representations

Until now we have always worked with representations of a group and we never dealt with its
subgroups. If we have a representation ρ : G → V and H ⊆ G a subgroup, we can define a
representation on H simply restricting the domain of ρ:

ρ|H : H −→ V.

It is a convention to define the restriction representation as

ResG
H V = ρ(H).

On the other hand, suppose to have H ⊆ G a subgroup and to know a representation


ρ : H → W . Can we extend it to a representation on G? The answer is yes, it can be done and the
representation is called “induced”. Due to the fact that it is often easier to find representations
of smaller groups, the operation of forming induced representations is an important tool to
construct new representations.

6.1 Definition and characterizations of IndG


HV

Definition 19. Let W be a representation of H ⊆ G and let

G = { σ = gσ H | gσ ∈ G } .
H

The induced representation of W is

IndG
HW =
M
Wσ ,
σ∈G/H

where Wσ is a copy of W labelled by the coset. It is clear that dim IndG


H W = [G : H] dim W .

In order to define the action of G on V = IndG


H W , we have to consider the left action of G
on G/H:

g · (gσ H) = (g · gσ )H. (6.1)


Actually, this is a permutation representation of G of dimension [G : H]. We know that

gσ ∈ gσ H ∀σ ∈ GH

63
6.1. Definition and characterizations of IndG
HV

and, since the cosets are a partition of G/H, for every g there exists gτ such that ggσ ∈ gτ H.
In particular there exists also h ∈ H such that

ggσ = gτ h.

Now we have just to define the action of G on Wσ (since it extends by linearity to V ). This is:

g · wσ = (h · w)τ .

The formula has to be interpreted in this way: take a vector w ∈ W and think of it as an element
wσ ∈ Wσ . There exists a gτ ∈ G and h ∈ H such that ggσ = gτ h. Since we know how h acts on
W , you have just to consider the vector h · w and think of it as an element (h · w)τ ∈ Wτ .
Remark 22. According to the above notation we have

g · Wσ = (h · W )τ = Wτ ,

since W is H-stable. This means that the action of G permutes the component of V in accordance
with the action of G on G/H.
Example 7. If H = { 1 }, the permutation representation of G is the regular representation.
Theorem 54 (Frobenius reciprocity). IndG H is the left-adjoint functor of the functor ResH : if
G

W is a representation of H, U is a representation of G, then


   
HomH W, ResG
H U ' HomG IndH W, U .
G

In particular, holds true


(χW , χResG U )H = (χIndG W , χU )G . (6.2)
H H
 
Proof. Let ϕ ∈ HomH W, ResG
H U . Define ψ as on every Wσ as

−1
gσ ϕ gσ
ψ : Wσ W U U.

The ψ doesnot depend on gσ since ϕ is H-linear. Further, ψ is G-linear. On the contrary, take
ψ ∈ HomG IndG H W, U . Then

ResG
HU
 
ϕ = ψ|W ∈ HomH W, ResG
HU .
   
The maps are one the inverse of the other, so that HomH W, ResG H U ' HomG IndH W, U .
G

For the scalar product formula, we just have to prove it for irreducible representations due
to linearity. Taking the dimensions on both sides of the previous equation, we obtain the
statement.

Exercise 17. Prove the universal property for Ind. Let W be a representation of H ⊆ G and
U a representation of G. For every H-module homomorphism ϕ : W → ResG H U , there exists
a unique G-module homomorphism ψ : IndH W → U such that ψ|W ≡ ϕ and the following
G

diagram commute.
ψ
IndG
HW U

i ϕ

64
6.2. Tables of IndSS43

Exercise 18. Prove that

IndG
H (W1 ⊕ W2 ) = IndH (W1 ) ⊕ IndH (W2 ) .
G G

The following exercise characterize the induced representations and prove that the above
construction is independent of any choice of the coset representative.

Exercise 19. Prove the following characterizations of IndG


H W.

• Define the induced representation as IndG


H W = C[G] ⊗C[H] W and prove that the C[G]-
module structure is the same as the above one.

• Set
IndG
H W = { f : G → W | f (hx) = hf (x) ∀h ∈ H } ,
with the action
g · f (x) = f (x · g).

Let us work out the character of an induced representation IndG H W . The contribution to
the trace of g will come only from those cosets such that gσ = σ. Thus,

χIndG W (g) = χW (h) = χW (gσ−1 ggσ ). (6.3)


X X
H
σ : ggσ =gσ h σ : gσ=σ

In particular, if H is a normal subgroup of G, then χIndG W (g) = 0 unless g ∈ H.


H

Exercise 20. Let Cg be a conjugacy class of G. Then Cg ∩H decomposes into conjugacy classes
D1 , . . . , Dr of H. Prove that
r
[G : H] X
χIndG W (g) = Di χW (Di ).
H #Cg i=1

Deduce that for W = U the trivial representation of H,

[G : H]
χIndG U (g) = #(Cg ∩ H).
H #Cg

6.2 Tables of IndSS43


Let us study the induced representations of H = S3 in G = S4 . We can choose as cosets
representatives
id (1 4) (2 4) (3 4).
The elements of G act on them in the following way.

1. The element g = (1 2) fixes the cosets id and (3 4) and switches the other two. Indeed

g gσ ggσ gτ h
(1 2) id (1 2) id (1 2)
(1 2) (1 4) (1 4 2) (2 4) (1 2)
(1 2) (2 4) (1 2 4) (1 4) (1 2)
(1 2) (3 4) (1 2)(3 4) (3 4) (1 2)

65
6.2. Tables of IndSS43

2. The element g = (1 2 3) fixes the coset id and permutes the other three. Indeed

g gσ ggσ gτ h
(1 2 3) id (1 2 3) id (1 2 3)
(1 2 3) (1 4) (1 4 2 3) (2 4) (1 2 3)
(1 2 3) (2 4) (1 2 4 3) (3 4) (1 2 3)
(1 2 3) (3 4) (1 2 3 4) (1 4) (1 2 3)

3. The element g = (1 2 3 4) permutes all the cosets. Indeed

g gσ ggσ gτ h
(1 2 3 4) id (1 2 3 4) (1 4) (1 2 3)
(1 2 3 4) (1 4) (2 3 4) (2 4) (2 3)
(1 2 3 4) (2 4) (1 2)(3 4) (3 4) (1 2)
(1 2 3 4) (3 4) (1 2 3) id (1 2 3)

4. The element g = (1 2)(3 4) switches id with (3 4) and (1 4) with (2 4). Indeed

g gσ ggσ gτ h
(1 2)(3 4) id (1 2)(3 4) (3 4) (1 2)
(1 2)(3 4) (1 4) (1 3 4 2) (2 4) (1 3 2)
(1 2)(3 4) (2 4) (1 2 3 4) (1 4) (1 2 3)
(1 2)(3 4) (3 4) (1 2) id (1 2)

We are now ready to study the character tables of the induced representations from S3 . For a
more clear exposition, let us remember the character table of S4 :

id 1 1 1 3 3 2
(1 2) 6 1 −1 1 −1 0
(1 2 3) 8 1 1 0 0 −1
(1 2 3 4) 6 1 −1 −1 1 0
(1 2)(3 4) 3 1 1 −1 −1 2

1) Let W = ⊕ be the defining representation of S3 . We know that


4!
dim IndSS43 W = [S4 : S3 ] dim W = · 3 = 12.
3!
Let us label the elements A1 , A2 , A3 , . . . , D2 , D3 . Remembering that the character is the number
of fixed point, we find:

1 (1 4) (2 4) (3 4) χ S
IndS4 W
3

id A1 A2 A3 B1 B2 B3 C1 C2 C3 D1 D2 D3 12
(1 2) A2 A1 A3 C2 C1 C3 B2 B1 B3 D2 D1 D3 2
(1 2 3) A2 A3 A1 D2 D3 D1 B2 B3 B1 C2 C3 C1 0
(1 2 3 4) D2 D3 D1 A2 A3 A1 B1 B3 B2 C2 C1 C3 0
(1 2)(3 4) D2 D1 D3 C2 C3 C1 B3 B1 B2 A2 A1 A3 0

66
6.2. Tables of IndSS43

Using the above character table of S4 we can calculate the molteplicity of the irreducible repre-
sentations of S4 in IndSS43 .
1
(χInd , χ )= (12 · 1 · 1 + 2 · 6 · 1) = 1
24
1
(χInd , χ ) = (12 · 1 · 1 − 2 · 6 · 1) = 0
24

1
(χInd , χ )= (12 · 1 · 3 + 2 · 6 · 1) = 2
24
1
(χInd , χ )= (12 · 1 · 3 − 2 · 6 · 1) = 1
24

1
(χInd , χ )= (12 · 1 · 2 − 2 · 6 · 0) = 1
24
Eventually we find that
⊕2
IndSS43 ⊕ = ⊕ ⊕ ⊕ .

2) Take now U = as the trivial representation of S3 . We know that


4!
dim IndSS43 U = [S4 : S3 ] dim U = · 1 = 4.
3!
Let us label the elements A, B, C, D. We have:

1 (1 4) (2 4) (3 4) χ S
IndS4 U
3

id A B C D 4
(1 2) A C B D 2
(1 2 3) A D B C 1
(1 2 3 4) D A B C 0
(1 2)(3 4) D C B A 0

Following the above steps we get

IndSS43 = ⊕ .

3) Eventually, take the standard representation V = of S3 . In order to compute IndSS43 ,


we can use the Frobenius reciprocity formula. The character of is

S3 id (1 2) (1 2 3)
2 0 −1

Identifying the linear space with the character, we find


     
IndSS43 , = , ResSS43 = , = 0.

On the other hand    


IndSS43 , = , ResSS43 .

However the character of is

67
6.3. The Heisenberg group

S4 id (1 2) (1 2 3) (1 2 3 4) (1 2)(3 4)
3 1 0 −1 −1

so that the character of ResSS43 is

S3 id (1 2) (1 2 3)
ResSS43 3 1 0

It can be easily seen that


ResSS43 = ⊕ .
Thus
   
IndSS43 , = , ResSS43
 
= , ⊕
   
= , + ,
= 0 + 1 = 1.

With the same procedure, it can be shown that

IndSS43 = ⊕ ⊕ .

Remark 23. In this example it is easy to see that there is a trick to built the induced repre-
sentations: you have just to add a box in every possibile position of the Young diagram of the
representation you are inducing and then sum what you get. Actually, this is not a case, but
the general procedure to construct induced representations of Sn+1 from Sn .

6.3 The Heisenberg group


As an application, let us study the irreducible representations of the so-called Heisenberg
group: let Fq be the finite field with q elements and set
n 1 a c o
Heis = 1 b a, b, c ∈ Fq ,

1

which is a group with the matrix product of order q 3 . We will indicate the elements of Heis
simply as (a, b, c). The product will be

(a, b, c)(x, y, z) = (a + x, b + y, ay + z + c).

Thus, the commutator will be

[(a, b, c), (x, y, z)] = (0, 0, ay − bx),

so that the commutator subgroup (which coincides with the centre) will be

Z = { (0, 0, c) | c ∈ Fq } ' Fq .

It fits the exacts sequence

1 −→ Z −→ Heis −→ HeisZ ' Fq ⊕ Fq −→ 1,

68
6.3. The Heisenberg group

where in the isomorphism HeisZ ' Fq ⊕ Fq we forget about the last component (and the
product turns out to be the usual sum in Fq ). Thus, we have q 2 one-dimensional representations
U1 , . . . , Uq2 of Heis, coming from those of Fq ⊕ Fq . On the other hand, it can be shown that the
number of conjugacy classes of Heis is q 2 + q − 1, so that we miss q − 1 irreducible representation.
Those representations arise as induced representations.
Note that dim IndGH W is inversely proportional to |H|. In our case, we could try to use Z
to induce a representation on Heis, but we immediately see that it is too small:
q3
dim IndHeis
Z W = dim W = q 2 dim W.
q
On the other hand, it is convenient to choose an abelian subgroup of Heis, so that the irreducible
representations of it are easily computable. The correct choice turns out to be
N = { (0, b, c) | b, c ∈ Fq } ,
which is an abelian subgroup of order q 2 . Let us consider the non-trivial irreducible representa-
tions of N defined by
N −→ Wζ = C×
(0, b, c) 7−→ ζ c ,
where ζ is a non-trivial qth root of 1. We will have q − 1 non-trivial irreducible representations
of Wζ , one for every non-trivial choice of ζ. We define the Schrödinger representation to be
Vζ = IndHeis
N (Wζ ).

Let us prove that these are the missing irreducible representations of Heis. Since N is a normal
subgroup of Heis, from Formula 6.3 it follows that χVζ (g) = 0 unless g ∈ N . On the other hand,
the elements of Heis/N are of the form
(x, 0, 0), x ∈ Fq .
The corresponding conjugate element of g = (0, b, c) ∈ N is
(x, 0, 0)−1 g(x, 0, 0) = (0, b, −bx + c),
so that
   
χVζ (g) = χWζ (x, 0, 0)−1 g(x, 0, 0) = χWζ (0, b, −bx + c)
X X

x∈Fq x∈Fq
c
if b = 0
(

= ζ −bx+c = ζ c ζ −bx =
X X

x∈Fq x∈Fq
0 if b 6= 0.

As a consequence we have
1 X 1 X −c 0c
(χVζ , χVζ 0 ) = χV (g)χ V 0
(g) = qζ qζ
q 3 g∈Heis ζ ζ q 3 c∈F
q
1 X

 1=1 if ζ = ζ 0
1 X ζ0 c  q c∈Fq
  

= =
q ζ 1 X 00c
ζ =0 if ζ 6= ζ 0 .

c∈Fq 

q

c∈Fq

This prove that the Schrödinger representations are irreducible and distinct for different choices
of ζ. In particular, the character table will be

69
6.3. The Heisenberg group

(0, 0, c) (a, b, c)
with (b, c) 6= (0, 0)
Ur 1 ∗
Vζ ζc 0

which shows that the one-dimensional representations are distinct from the Schrödinger ones.
Thus, the character table is complete.
The generalization of the above discussion to infinite-dimensional representation is known
as the Stone-von Neumann theorem. Historically, this result was significant in Quantum Me-
chanics because it was a key step in proving that Heisenberg matrix mechanics, which presents
quantum mechanical observables and dynamics in terms of operators, is unitarily equivalent to
Schrödinger’s wave mechanical formulation of the theory.

70
Chapter 7

Symmetric functions and Pólya’s


theory of counting

The group Sn acts on Z[x1 , · · · , xn ] (the ring of polynomials with coefficients in Z) in a natural
way: fixed p ∈ Z[x1 , · · · , xn ] we can set
σ · p(x1 , . . . , xn ) = p(xσ(1) , . . . , xσ(n) ) σ ∈ Sn .
We define
Λn = Z[x1 , . . . , xn ]Sn ,
the set of symmetric polynomial with integer coefficients. It is a graded algebra, with grading

Λn = Λkn , Λkn = { p ∈ Λn | deg p = k } .
M

k=1

The group Λkn can be seen as the eigenspace of the map xi 7→ axi , where a is a non-zero scalar,
associated to the eigenvalue ak . Note that the above construction can be done in every unitary
commutative ring.
In applications of symmetric functions, the number of variables is usually irrelevant. So let us
try to construct an algebraic structure for “polynomials” in infinitely many variables. Let λ be
any partition of the form
λ = (λ1 , λ2 , . . . ), with λ1 ≥ λ2 ≥ . . . .
We define the length of the partition as
l(λ) = # { r | λr > 0 } .
If l(λ) ≤ n, then λ = (λ1 , . . . , λn ) and using the multi-index notation, the polynomial
mλ (x1 , . . . , xn ) =
X
xσ(λ)
σ∈Sn

is clearly symmetric. Then


M = { mλ | l(λ) ≤ n, |λ| = k }
is a Z -basis of Λkn . Let now m ≥ n and consider the map
ρm,n : Λm −→ Λn
if i ≤ n,
(
xi
xi 7−→
0 if i > n.

71
We have that ρm,n is surjective, since

mλ (x1 , . . . , xn ) if l(λ) ≤ n,
(
ρm,n (mλ (x1 , . . . , xm )) =
0 if l(λ) > n.

The restrictions
ρkm,n : Λkm −→ Λkn
are surjective homomorphism for m ≥ n, which are bijective if n ≥ k. We can take the inverse
limit
Λk = lim Λkn ,
← −
n

so that an element f ∈ Λk is a sequence f = (fn ) with

fn = fn (x1 , . . . , xn ) ∈ Λkn

and ρkm,n (fm ) = fn whenever m ≥ n. The elements of Λk are no longer polynomials, but formal
infinite sums of monomials. Thus, the name symmetric functions. The projections

ρkn : Λk −→ Λkn

are isomorphisms for n ≥ k. Thus, Λk has a Z -basis consisting of all monomial symmetric
functions mλ (with |λ| = k) defined by

ρkn (mλ ) = mλ (x1 , . . . , xn )

Definition 20. We define the ring of symmetric functions as



Λ= Λk .
M

k=0

Here Λ0 = Z. It has a Z -basis defined by all monomial symmetric functions mλ , with λ running
through all partitions. For convenience, we set m0 = 1.
Example 8. If λ = (2, 0, . . . ), then

mλ (x1 ) = x21
mλ (x1 , x2 ) = x21 + x22
..
.
mλ = x21 + x22 + x23 + · · ·

Take λ = (2, 1, 0, . . . ). Then

mλ (x1 , x2 ) = x1 x22 + x21 x2


mλ (x1 , x2 , x3 ) = x1 x22 + x1 x23 + x21 x2 + x21 x3 + x2 x23 + x22 x3
..
.
mλ = x1 x22 + x1 x23 + · · · + x21 x2 + x21 x3 + · · · + x2 x23 + x2 x24 + · · ·

Remark 24. Note that Λ̃ = limn Λn 6= Λ. Indeed,


←−
f= (1 + xr ) = 1 + (x1 + x2 + · · · ) + (x1 x2 + x1 x3 + · · · ) + · · · ∈ Λ̃ \ Λ.
Y

r≥1

72
7.1. Bases for the ring of symmetric functions

7.1 Bases for the ring of symmetric functions


We already seen the monomial basis (mλ ) for Λ. In this section, we will discuss the bases of ele-
mentary functions, complete homogeneous functions, powersum functions and Schur functions.

7.1.1 Elementary symmetric functions


For every r ≥ 0, we define the elementary symmetric function er as
er = m1r =
X
xi1 · · · xir .
i1 <···<ir

For a partition λ = (λ1 , λ2 , . . . ), we set


eλ =
Y
e λr .
r≥1

In order to explore the relation between monomial functions and elementary functions, we need
to work with Young diagrams.
Definition 21. Let λ be a partition of n, which is in 1-1 correspondence with a Young diagram.
We define the dual partition λ0 to be the partition associated to the Young diagram obtained
from the previous one with a reflection along the diagonal NW-SE.
Example 9.
λ= = 3 · 1, λ0 = = 2 · 12 .

Definition 22. We define the natural order among the set of partitions of n as

λ ≤ µ1
 1


λµ ⇔ λ1 + λ2 ≤ µ1 + µ1
..

.

It is not a total order, unless n ≤ 5.


Example 10. Take n = 6. Then 3 · 13 and 23 are not comparable.
Pr
i=1 λi
6
λ Young diagrams
r= 1 2 3 4 5 6
5·1
6 6 6 6 6 6 6
5·1 5 6 6 6 6 6
4·2
4·2 4 6 6 6 6 6
4 · 12 4 5 6 6 6 6
4 · 12
3·2·1 3 5 6 6 6 6
3·2·1
3 · 13 3 4 5 6 6 6
23 2 4 6 6 6 6 3 · 13 23
22 · 12 2 4 5 6 6 6
22 · 12
2 · 14 2 3 4 5 6 6
2 · 14
16 1 2 3 4 5 6
2 · 14

73
7.1. Bases for the ring of symmetric functions

Lemma 55. Let λ, µ be partitions of n. Then λ  µ if and only if we can fill the boxes of µ
following the rules below:
• put λ1 1’s in the first row of µ;
• put λ2 2’s in the first row (if blank boxes are left) and in the second row of µ;
• put λ3 2’s in the first and second rows (if blank boxes are left) and in the third row of µ;
..
.

Theorem 56 (Fundamental theorem of symmetric functions). Let λ be a partition of n. Then


eλ0 = mλ +
X
aλµ mµ ,
µ≺λ

with aλµ ∈ Z.
Proof. Let eλ0 = eλ01 eλ02 · · · , where

eλ01 =
X
xi1 · · · xiλ0
1
i1 <···<iλ0
1

eλ02 =
X
xj1 · · · xjλ0
2
j1 <···<jλ0
2

..
.
Multiplying out the product eλ0 , we obtain a sum of monomials of the form
xµ = (xi1 · · · xiλ0 )(xj1 · · · xjλ0 ) · · · .
1 2

If we now enter the numbers i1 , . . . , iλ01 in order down the first column of the diagram of λ, the
numbers j1 , . . . , jλ02 again order down the second column and so on, it is clear that numbers ≤ r
so entered in the diagram of λ must occur in the top r rows. Hence, for every r ≥ 1
µ1 + · · · + µ r ≤ λ 1 + · · · + λ r ,
i.e. µ  λ. It follows that
eλ0 =
X
aµν mµ ,
µλ
with aµν non-negative integers. From the argument above, we can see that the monomial function
mλ occurs exactly once, so that aλλ = 1 and the statement is proved.

Corollary 57. The elementary symmetric functions eλ with |λ| = k form a Z -basis of Λk .
Thus,
Λ = Z[e1 , e2 , . . . ].
Proof. It follows from the invertibility of the matrices
1
 

 1 ∗ 

 .. 

 . .

0 1
 
 
1

74
7.1. Bases for the ring of symmetric functions

In the following, it will be useful to consider the generating function1 for the sequence
(er )r∈N .We have

E(t) =
X
er tr .
r=0
Exercise 21. Prove that
E(t) = (1 + xn t).
Y

n≥1

7.1.2 Complete homogeneous symmetric functions


For every r ≥ 0, we define the complete homogeneous symmetric function hr as
hr =
X
mλ .
|λ|=r

For a partition λ = (λ1 , λ2 , . . . ), we set


hλ =
Y
hλr .
r≥1

In this case, the generating function is



H(t) =
X
hr tr .
r=0

Exercise 22. Prove that


H(t) = (1 − xn t)−1 .
Y

n≥1
Deduce the relation E(t)H(−t) = 1, so that e1 h1 = 1 and
n
(−1)r er hn−r = 0, for n ≥ 1. (7.1)
X

r=0

The symmetry between the elementary functions and the complete homogeneous functions
suggests the definition of
ω : Λ −→ Λ
er 7−→ hr ,
extended so that it becomes a ring homomorphism.
Lemma 58. The homomorphism ω is an involution: ω 2 = id.
Proof. Applying ω to Equation 7.1, we find that for all n ≥ 1
n
(−1)r hn−r ω(hr ) = 0.
X

r=0

For n = 0, we already know that ω 2 (e0 ) = ω 2 (1) = 1. By induction on n, with the previous
relation, we find that
ω(hn ) = en ,
so that ω 2 = id.
1
For a sequence (an )n∈N , the generating function is defined as the formal power series in the variable t

X
G(t) = an tn .
n=0

75
7.1. Bases for the ring of symmetric functions

Corollary 59. The complete homogeneous symmetric functions hλ with |λ| = k form a Z -basis
of Λk . Thus,
Λ = Z[h1 , h2 , . . . ].

7.1.3 Powersum symmetric functions


For every r ≥ 0, we define the powersum symmetric function pr as

pr = mr =
X
xri .
i=1

For a partition λ = (λ1 , λ2 , . . . ), we set

pλ =
Y
p λr .
r≥1

Exercise 23. Prove that the generating function for (pr )r∈N is

xi d
P (t) = = log H(t).
X

i=1
1 − xi t dt

Deduce the relations


P (t)H(t) = H 0 (t) P (−t)E(t) = E 0 (t).
Proposition 60 (Newton’s formulae). The following formulae holds true:
n
nhn =
X
pr hn−r ,
r=1
n
nen = (−1)r−1 pr en−r .
X

r=1

Corollary 61. Setting ΛQ = Λ ⊗Z Q, we have that

ΛQ = Q[p1 , p2 , . . . ].

Corollary 62. The following formulae holds true:


1
hn =
X
pλ ,
|λ|=n
Z λ
X λ
en = pλ , where λ = (−1)|λ|−l(λ) .
Z
|λ|=n λ

As an application of the formula P (−t)E(t) = E 0 (t), let us see a characterization of nilpotent


matrices.
Proposition 63. Let A be an n × n complex matrix. Then A is nilpotent if and only if

tr Ar = 0 ∀r = 1, . . . , n.

Proof. Let A be nilpotent. Then every eigenvalue is zero and tr Ar = 0. On the contrary, the
formula can be rewritten as


X pr
E(−t) = (1 − xn t) = exp −
Y
tr .
n≥1 r≥1
r

76
7.1. Bases for the ring of symmetric functions

If x1 , . . . , xn are the eigenvalues of A, then tr Ar = pr = 0 and


n
tr Ar r
!
det(id −tA) = exp − = 1.
X
t
r=1
r

Then the characteristic polynomial is tn :

det(A − t id) = (−t)n det(id −tA) = (−t)n .

For Cayley-Hamilton, A is nilpotent.

7.1.4 Schur functions


Let us start discussing the skew-symmetric polynomials in Z[x1 , . . . , xn ]. Take α = (α1 , . . . , αn ).
Antisymmetrizing the monomials xα , we obtain

aα (x1 , . . . , xn ) = sgn(σ) xσ(α) .


X

σ∈Sn

Note that
α
aα (x1 , . . . , xn ) = det(xi j )i,j .
Then σ · aα (x1 , . . . , xn ) = sgn(σ) aα (x1 , . . . , xn ). It is clear that aα (x1 , . . . , xn ) = 0, unless the
αr are all distinct. Indeed, if α1 = α2 , then

(1 2) · aα (x1 , . . . , xn ) = aα (x1 , . . . , xn )
= −aα (x1 , . . . , xn ).

Hence, we can assume α1 > · · · > αn ≥ 0. We define now δ = (n − 1, n − 2, . . . , 1, 0) and


λ = α − δ. Then λ is partition of length l(λ) ≤ n. Noting that aλ+δ (x1 , . . . , xn ) vanishes for
xi = xj and i 6= j, we obtain that aλ+δ (x1 , . . . , xn ) is divisible by the Vandermonde determinant

aδ (x1 , . . . , xn ) = (xi − xj ).
Y

i<j

Since both polynomials are skew-symmetric, we have that

aλ+δ (x1 , . . . , xn )
sλ (x1 , . . . , xn ) =
aδ (x1 , . . . , xn )

is a symmetric polynomial, called the Schur polynomial in the variables x1 , . . . , xn correspond-


ing to the partition λ of length l(λ) ≤ n. Note that sλ (x1 , . . . , xn ) is homogeneous of degree |λ|,
so that sλ (x1 , . . . , xn ) ∈ Λn .

Example 11. For n = 2, we have the partitions λ = 2 = (2, 0), so that α = (3, 0), and
µ = 12 = (1, 1), so that β = (2, 1). The associated Schur functions are

x3
1 1
1
3
x2 x31 − x32

s2 (x1 , x2 ) = = = x21 + x1 x2 + x22 ,
x1 − x2 x1 − x2

x2 x1
1
2
x2 x2 x21 x2 − x22 x1

s12 (x1 , x2 ) = = = x1 x2 .
x1 − x2 x1 − x2

77
7.1. Bases for the ring of symmetric functions

We define now An to be the set of skew-symmetric polynomials in the variables x1 , . . . , xn .


It is clear that An is a Λn -module. The map

An −→ Λn
αλ+δ (x1 , . . . , xn ) 7−→ sλ (x1 , . . . , xn )

is a Λn -module isomorphism, with inverse

Λn −→ An
sλ (x1 , . . . , xn ) 7−→ aλ+δ (x1 , . . . , xn ).

In particular,
S = { sλ (x1 , . . . , xn ) | l(λ) ≤ n, |λ| = k }

forms a Z -basis for Λn . Thus, Λk has a Z -basis consisting of all Schur functions sλ (with
|λ| = k) defined by
ρkn (sλ ) = sλ (x1 , . . . , xn ).

As for the other basis, we can find some formulae for the change of basis. In the following
section, we will apply the identity below.

Theorem 64 (Jacobi-Trudy identity). For any λ partition of k with l(λ) ≤ n,

sλ = det(hλi −i+j ).

7.1.5 Orthogonality
Another useful tool in the theory of symmetric functions is given by inner products on Λ. Let
us define a scalar product imposing the orthonormality condition to the bases (hλ ) and (mλ )

hhλ , mµ i = δλµ ,

extend by Z -bilinearity on Λ.

Exercise 24. Show that

hpλ , pµ i = Zλ δµν , hsλ , sµ i = δµν ,

where
Zλ = k mk mk !
Y

k≥1

Deduce that h·, ·i : ΛQ × ΛQ → Q is symmetric and positive definite.

Corollary 65. The involution ω is an isometry: hω(u), ω(v)i = hu, vi.

Further results can be found in (Macdonald, 1995).

78
7.2. Characters of Sn

7.2 Characters of Sn
Let σ be a permutation in Sn . Then the cycle structure of σ determines a unique partition
ρ(σ) = 1m1 · · · nmn of n, in accordance with the diagram
n elements
σ = (·) · · · (·) (· ·) · · · (· ·) · · · ( · · ) · · · (· · · · ·)
z }| {
···
| {z }| {z } | {z }
m1 m2 mn

We define the map

ψ : Sn −→ Λ
σ 7−→ pρ(σ) .

Composing it with the immersion Sm × Sn ,→ Sm+n , we obtain

ψ(σ × τ ) = ψ(σ)ψ(τ ).

Indeed, the immersion takes (σ × τ ) to a permutation of m + n objects of cycle type ρ(σ) ∪ ρ(τ ),
letting them act on { 1, . . . , m } and { m + 1, . . . , m + n }. Thus

ψ(σ × τ ) = pρ(σ)∪ρ(τ ) = pρ(σ) pρ(τ ) = ψ(σ)ψ(τ )

and we obtain the above statement.


A rich algebraic structure coming from the representations of Sn is given by the ring of
virtual representations.

Definition 23. We define Rn as the free Z -module generated by the irreducible characters of
Sn . The elements of Rn are called virtual representations and assume the form

f=
X
aλ χλ ,
χλ ∈Irr Sn

with aλ ∈ Z. Note that the irreducible characters of Sn are indexed by partitions λ of n. The
sum in Rn is defined for χ, θ ∈ Irr(Sn ) as the usual sum of characters. We set

R=
M
Rn .
n=0

Here R0 = Z. We can define on R a ring structure and a scalar product.

• For χ ∈ Irr(Sm ), θ ∈ Irr(Sn ), we define the multiplications as


S
χ · θ = IndSm ×Sn (χ × θ)
m+n

and then we extend it to R.

• For χ, θ ∈ Irr(Sn ), we define the scalar product as


1 X
hχ, θiSn = χ(σ)θ(σ).
n! σ∈S
n

and then we extend it to R.

79
7.2. Characters of Sn

Thus, R becomes a graded, commutative, unitary ring with a scalar product. Observe that the
definition agrees with the Hermitian product defined for general class functions on a group G,
since for every character χ of Sn

χ(σ) = χ(σ −1 ) = χ(σ),

noting that σ and σ −1 have the same cycle structure.

The connection between the virtual representations and the map ψ defined above is given in
the following

Definition 24. The characteristic map of Frobenius is defined as

ch : R −→ ΛC = Λ ⊗Z C
R 3 fn 7−→ ch(fn ) = hfn , ψi
n

and then extended by linearity. The map is well defined, since ψ(σ) ∈ ΛZ ⊂ ΛC and, for every
f ∈ R, f (σ) ∈ C ⊂ ΛC .

Lemma 66. If f ∈ Rn , then


X 1
ch(f ) = f (ρ)ψ(ρ),
|ρ|=n

where ρ is a partition of n, f (ρ) and ψ(ρ) are f and ψ evaluated in any conjugacy class of Sn
identified by the partition ρ.

Proof. It follows from the fact that f and ψ are class functions and that

#Cρ 1
= .
n! Zρ

Proposition 67. The map characteristic map of Frobenius is an isometry with respect to the
scalar products defined on Λ and R:

hch(f ), ch(g)i = hf, gi .

Proof. We can suppose f, g ∈ Rn , due to linearity of ch. Then


1
hch(f ), ch(g)i = f (ρ) g(ρ0 ) hpρ , pρ0 i
X

|ρ|=|ρ0 |=n
Z ρ Z ρ0

1
= f (ρ) g(ρ0 ) Zρ δρρ0
X
Z Z
|ρ|=|ρ0 |=n ρ ρ
0

X 1
= f (ρ) g(ρ)
|ρ|=n

1 X
= f (σ)g(σ) = hf, gi .
n! σ∈S
n

Theorem 68. The map characteristic map of Frobenius is a ring isomorphism.

80
7.2. Characters of Sn

Proof. The map is Z -linear by definition. On the other hand, if f ∈ Rm and g ∈ Rn ,


S S
ch(f · g) = ch(IndSm+n
m ×Sm
(f × g)) = hIndSm ×Sm (f × g), ψiS
m+n
m+n
Sm+n
= hf × g, ResSm ×Sm (ψ)iSm ×Sn

due to Frobenius reciprocity. Thus,


S
ch(f · g) = hf × g, ResSm ×Sm (ψ)iS
m+n
m ×Sn
1
= f (σ)g(τ )ψ(σ)ψ(τ )
X
m!n! σ×τ ∈S ×S
m n

= hf, ψiSm hg, ψiSn ,

so that ch is a ring homomorphism.


Let us call now ηn the character of the trivial representation of Sn . Then
X 1
ch(ηn ) = pρ = hn ,
|ρ|=n

where in the last equation we have used Corollary 62. If λ is a partition of n, we define
ηλ = ηλ1 · ηλ2 · · · , which is the character of Sn induced by the trivial character of Sλ1 × Sλ2 × · · · .
Since ch is a ring homomorphism,

ch(ηλ ) = ch(ηλ1 ) · ch(ηλ2 ) · · · = ηλ1 · ηλ2 · · · = ηλ .

We define now for a partition λ of n

χλ = det(ηλi −i+j ),

which are possibly virtual characters of Sn . Due to Jacobi-Trudy identity and the fact that ch
is an homomorphism (which implies commutativity between ch and det), we obtain

ch(χλ ) = sλ ,

the Schur function associated to λ. But ch is an isometry too, so that

hχλ , χµ i = hsλ , sµ i = δλµ .

This means that (χλ ) are an orthonormal basis of Rn . Since the number of partitions of n equals
the number of irreducible representations, we have that ch is an isomorphism.

Theorem 69. The characters χλ are the irreducible representations of Sn .

Proof. We have proved that (χλ ) form an orthonormal basis of Rn . But a Z -module admits a
unique Z -basis, up to sign: if (eλ ) is an orthnormal basis and f = λ an en is normalized, then
P

1=
X
|aλ |2 ,
λ

which implies that there exists a partition µ such that aµ = ±1 and aλ = 0 for λ 6= µ. In
particular, f = ±eµ . Since the irreducible characters of Sn form an orthonormal basis of Rn ,
we have that χλ are irreducible characters up to a sign. To show the irreducibility, it suffices to
show that χλ (id) ≥ 1. For a complete proof, see (Macdonald, 1995).

81
7.2. Characters of Sn

Corollary 70. Let us set

χλρ = χλ (σ), σ with cycle type ρ.

Then (χλρ ) is the character table of Sn . In particular, they are the coefficients in the expansion
X 1
sλ = χλρ pρ .
Z
|ρ|=n ρ

Proof. The fact that (χλρ ) is the character table of Sn follows from the previous theorem. On
the other hand, we know that sλ = ch(χλ ). Eventually,
X 1
sλ = ch(χλ ) = χλρ pρ .
Z
|ρ|=n ρ

Example 12. Let us check the results computing the character table of S2 . We know from a
previous example that

s2 (x1 , x2 ) = x21 + x1 x2 + x22 ,


s12 (x1 , x2 ) = x1 x2 .

On the other hand,

p2 (x1 , x2 ) = m12 (x1 , x2 ) = x21 + x22 ,


p12 (x1 , x2 ) = m21 (x1 , x2 ) = (x1 + x2 )2 .

Thus, we obtain
1  1 1
s2 (x1 , x2 ) = p12 (x1 , x2 ) + p2 (x1 , x2 ) = p12 (x1 , x2 ) + p2 (x1 , x2 ),
2 Z 12 Z12

so that χ212 = χ22 = 1, and similarly

1  1 1
s12 (x1 , x2 ) = p12 (x1 , x2 ) − p2 (x1 , x2 ) = p12 (x1 , x2 ) − p2 (x1 , x2 ),
2 Z12 Z 12
2 2
so that χ112 = −χ12 = 1. The character table will be

S2 12 2
χ2 1 1
2
χ1 1 −1

As an application, we can also explain the “primed” notation used in the previous chapters
used to representations tensored with the sign one. It follows from a simple application of the
exercise below.

Exercise 25. Prove that

ω(pλ ) = λ pλ ,
ω(sλ ) = sλ0 .

82
7.3. Pólya’s theory of counting and Burnside lemma

We have that (1n )0 = n, so that

s1n = ω(sn ) = ω(hn ) = en .

On the other hand,


X ρ
en = pρ ,
|ρ|=n

n
so that χ1ρ = ρ is the sign character. In general, applying ω to sλ we find
X 1
sλ0 = ω(sλ ) = χλρ ω(pρ )
Z
|ρ|=n ρ
X ρ
= χλρ pρ ,
Z
|ρ|=n ρ
0 0
obtaining χλρ = ρ χλρ . Without reference to the values on the conjugacy classes, χλ =  ⊗ χλ .

7.3 Pólya’s theory of counting and Burnside lemma


The aim of this section is to develop a tool very useful in counting objects with fixed properties
up to permutations (thus, we will see how algebra participates in combinatorial problems). Let
us consider the following guide example: we want to know how many necklaces can be formed
with 6 beads (up to rotations and reflections of the necklace) such that every beads can take 2
different colours.

Since we have a low number of cases, let us count them. If we suppose that the two colours
are black and white, there are only2 these possibilities:

no black beads 1 black bead 2 black beads 3 black beads

No. of cases 1·2=2 1·2=2 3·2=6 3

Therefore there are 13 possibilities. In order to count all the cases in a more general way, we
have to encode correctly the information.

Let X be a finite set (the location of the beads) and G be the group of symmetries of X
acting on X. Let C be the set of the colours; we define a colouring of X as a function

ϕ : X −→ C.
2
When we count the cases with 0, 1, 2 black beads, we must remember that there is also the dual case with
0, 1, 2 white beads (the case with 3 black beads is the same of that one with 3 white beads).

83
7.3. Pólya’s theory of counting and Burnside lemma

The number possible colourings is clearly

#C X = # { ϕ : X → C } .

G acts on C X in the following way

(g · ϕ)(x) = ϕ(g −1 · x).

Two colourings ϕ1 , ϕ2 are the same up to isomorphism if there exists g ∈ G such that ϕ1 = g ·ϕ2 ,
that is if they are in the same orbit. Eventually, we have translated the initial problem in
counting the orbits of the group G acting on the set Ω = C X . A formula to count all possible
colourings is given by the following

Lemma 71 (Burnside lemma). Let G be a finite group acting on a finite set Ω. Then the
number of orbits is given by
1 X
|Ω/G| = # Fix(g).
|G| g∈G

Proof. The action of G on Ω gives a linear representation V of G. Clearly

|Ω/G| = dim V G = (χU , χV ),

where U is the trivial representation. Now we have just to remember that χV (g) is the number
of points fixed by g.

Example 13. If G acts with no fixed points (except for the identity), Burnside formula gives
1
|Ω/G| = #Ω,
|G|

that is the case where all the orbits have the same size.

Let us now set #X = n and #C = k. Further, we will indicate with ρ(g) the cycle type of
the permutation g· of X as in the previous section.

Definition 25. The cycle indicator is the polynomial in the variables t1 , . . . , tn


1 X ρ(g)
ZG (t1 , . . . , tn ) = t .
|G| g∈G

Proposition 72 (Pòlya enumeration theorem). The number of possible ways to colour a set X
(up to symmetries) with k colours is given by

ZG (k, . . . , k),

where G is the group of symmetries of X.

Proof. Let us take g ∈ G. To understand the number of elements fixed by g, we must look at
the cycle structure. A colouring is fixed by the action of g· if and only if each cycle consists of
elements x of the same colour. Thus, for every cycle we have m possible choices and the number
of elements fixed by g is therefore

k m1 +···+kn = k m1 · · · k mn .

According to Burnside lemma, we have the statement.

84
7.3. Pólya’s theory of counting and Burnside lemma

Example 14. Let us analyse with this new tool the above case of the necklace. We can identify

2 3
X = { 1, · · · , 6 } ' 1 4 C = { ◦, • } , G ' D6 .
6 5

Let us study the cycle structure. Indicate with τ the rotation of angle 3,
π
with σi the reflection
along the ith vertex and with σij the reflection along the ij edge.

g ∈ D6 action on X cycle structure tρ(g)


id id 16 t61
τ, τ −1 (1 2 3 4 5 6) 6 t16
τ 2 , τ −2 (1 3 5)(2 4 6) 32 t23
τ3 (1 4)(2 5)(3 6) 23 t32
σ1 , σ2 , σ3 (1 3)(4 6) 22 12 t21 t22
σ12 , σ23 , σ34 (1 2)(3 6)(4 5) 23 t32

In particular, this means that


1 6
ZG = (t + 2t6 + 2t23 + 4t32 + 3t21 t22 ).
12 1
If we replace every ti with the number of colours k, we get
1 6
# necklaces = (k + 3k 4 + 4k 3 + 2k 2 + 2k).
12
Note that in the particular case of k = 2 (that is the case we started from) we get exactly 13.

Example 15. Let us count now how many graphs can be drawn with 4 vertices. The idea is to
take X as all the edges connecting 4 vertices and the colours as white (for the absent line) and
black (for the present line):

4 3

X = { (1, 2), (1, 3), . . . } ' C = { · · · , —– } , G ' S4 .

1 2

g∈G #Cg action on X cycle structure tρ(g)


id 1 id 16 t61
(1 2) 6 (2 4)(3 5) 22 · 12 t21 t22
(1 2 3) 8 (1 4 2)(3 5 6) 32 t23
(1 2 3 4) 6 (1 4 6 3)(2 5) 4·2 t2 t4
(1 2)(3 4) 3 (2 5)(3 4) 22 · 12 t21 t22

In particular, we get
1 6
ZG = (t + 9t21 t22 + 8t23 + 6t2 t4 ).
24 1
Setting ti = 2, we have a total of 11 different graphs. They are the following ones.

85
7.3. Pólya’s theory of counting and Burnside lemma

Exercise 26. Do the same with the cube, where G = S4 . In particular, compute the table

g∈G #Cg cycle structure tρ11 tρ22 · · ·


id 1 16 t61
RF , RF−1 6 4 · 12 t21 t4
RF2 3 22 · 12 t21 t22
RE 6 23 t32
RV , RV−1 8 32 t23

so that
1 6
ZG = (t + 6t21 t4 + 3t21 t22 + 6t32 + 8t23 ).
24 1
Setting ti = 2, we have that we can colour the vertices of a cube in 10 different ways.

We can generalize now the above result to a weighted case.

Definition 26. Let ϕ : X → C be a colouring. The weight function is the polynomial

wϕ (u1 , . . . , um ) =
Y
uϕ(x) .
x∈X

Here every variable ui represents the weight of a colour.

Example 16. Let us compute the weights for the previous example, where u corresponds to ◦
and v to •.

u6 u5 v u4 v 2 u3 v 3 v6 uv 5 u2 v 4 u3 v 3

In order to prove a weighted version of Pólya enumeration theorem, we have to generalize


the Burnside lemma. Let us set Ω/G the set of orbits for the action of G on Ω.

Lemma 73 (Burnside lemma, weighted version). Let G be a finite group acting on a finite set
Ω. Let w : Ω → C be a function constant on each orbit. Then
1 X X
w(ϕ) =
X
w(ϕ).
ϕ∈Ω/G
|G| g∈G ϕ∈Fix(g)

86
7.3. Pólya’s theory of counting and Burnside lemma

Proof. The action of G on Ω gives a linear representation V = hC X i of G. We define the map


ω : V −→ V
ϕ 7−→ w(ϕ)ϕ,
extended on V by linearity. We have that ω is G-linear, since w is constant on orbits:
ω(g · ϕ) = w(g · ϕ)g · ϕ = w(ϕ)g · ϕ = g · ω(ϕ).
Restricting and corestricting it to V G and taking the trace, we find
G
tr(ω|VV G ) =
X
w(ϕ).
ϕ∈Ω/G

On the other hand, the corestriction of ω to V G can be be obtained via composition with the
projection
1 X
π= g : V −→ V G .
|G| g∈G
In particular, π ◦ ω : V → V G has the form
1 X
π ◦ ω(ϕ) = w(ϕ) g · ϕ.
|G| g∈G

Restricting it to V G and taking the trace, we find


1 X X
tr(π ◦ ω|V G ) = w(ϕ)
|G| g∈G ϕ∈Fix(g)

and the statement is proved.


Theorem 74 (Pólya enumeration theorem, weighted version).
wϕ (u1 , . . . , um ) = ZG (t1 , . . . , tn ),
X

ϕ∈C X /G

where ti = uic .
P
c∈C
Proof. Let us take g ∈ G. A colouring is fixed by the action of g· if and only if each cycle
consists of elements x of the same colour. Thus, the colouring must have the form
m1 m2
z }| {z }| {
g = (·) · · · (·) (· ·) · · · (· ·) · · ·
ϕ: c1 · · · cm1 c01 · · · c0m2 · · ·
so that the sum of weights over fixed points will be
wϕ (u1 , . . . , un ) =
X X
uc1 · · · ucm u2c0 · · · u2c0m · · ·
1 1 2
ϕ∈Fix(g) c1 ...cm1
c01 ...c0m2
...
!
=
X X
uc uc2 · · · ucm u2c0 · · · u2c0m · · ·
1 1 2
c∈C c2 ...cm1
c01 ...c0m2
...
!m1
=
X X
uc u2c0 · · · u2c0m u3c00 · · · u3c00m · · ·
1 2 1 3
c∈C c01 ...c0m2
c00
1 ...cm3
00
...
!m1 !mn
=
X X
uc ··· unc .
c∈C c∈C

87
7.3. Pólya’s theory of counting and Burnside lemma

Summing over G and applying Burnside lemma, we obtain the statement.

Remark 25. In the weighted versions of Burnside lemma and Pólya enumeration theorem,
setting ui = 1 we obtain the previous results.

88
Chapter 8

Representations of GL2(Fq )

In this chapter, we will compute the irreducible representations of G = GL2 (Fq ), where Fq is
the finite field of q = pn elements, q odd. To count the elements g of G, we can choose the first
column between every pairs of numbers except (0 0), so that we have q 2 − 1 possibilities. For
the second column, we have q 2 , minus the proportional columns which are q. Thus, the order
of G is
|G| = (q 2 − 1)(q 2 − q).

8.1 Conjugacy classes of GL2 (Fq )


Let us compute the conjugacy classes of G. The conjugacy classes are of four different types:
! !
x 0 x 1
ax = with x ∈ F×
q , bx = with x ∈ F×
q ,
0 x 0 x
! !
x 0 x y
cx,y = with x, y ∈ F×
q ,x 6= y dx,y = with y ∈ F×
q ,
0 y y x

where  is a non-square element of F× q . Here q must be odd, otherwise all elements of Fq are
squares. Note that and are conjugate by 0 1 , while d
x,y and dx,−y are conjugate by

cx,y c y,x −1 0
0 − .

1 0
Let us count now the number of elements in each conjugacy class. For ax we will have one
element per class, since ax is conjugate to itself only. For bx , we will have

#Cbx = [G : StabG (bx )].

On the other hand, a b ∈ StabG (bx ) if and only if ad − cb 6= 0 and



c d
!−1 ! ! !
a b x 1 a b x 1
= .
c d 0 x c d 0 x

We can rewrite the second condition as


! ! ! !
x 1 a b a b x 1
= ,
0 x c d c d 0 x

which leads to
c=0
! ! (
ax + c bx + d ax a + bx
= ⇔
cx dx cx c + dx a = d.

89
8.2. Irreducible representations of GL2 (Fq )

Thus,
StabG (bx ) = a 6= 0

a b

0a

has order q(q − 1), so that #Cbx = q 2 − 1. The same procedure for cx,y and dx,y leads to

StabG (cx,y ) = ad 6= 0 , StabG (dx,y ) = (a, b) 6= 0 ,


 a 0  a b 
0d b a

whose orders are (q −1)2 and q 2 −1 respectively, so that #Ccx,y = q(q +1) and #Cdx,y = q(q −1).
Counting the number of possible classes for each type, we can summarize the informations
as in the following table.

Types of C #C No. of classes


ax 1 q−1
bx −1
q2 q−1
(q−1)(q−2)
cx,y q(q + 1) 2
q(q−1)
dx,y q(q − 1) 2

The Jordan theorem assures us that these are the conjugacy classes. As a check, we find that
(q − 1)(q − 2) q(q − 1)
#C = (q − 1) + (q 2 − 1)(q − 1) + q(q − 1) + q(q + 1)
X

C
2 2
= q(q − 1)(q 2 − 1) = |G|.

Here we have encountered some important subgroups of G:

B= ad 6= 0 ,
 a b
0d
D= ad 6= 0 ,
 a 0
0d
K= (a, b) 6= 0 .
 a b 
b a

The first one is called the Borel subgroup, the second one is the group of invertible diagonal
matrices, while the third one is the image of the immersion


q 2 ,→ G
a + b 7→ a b

b a .

8.2 Irreducible representations of GL2 (Fq )


For n ≥ 1 and for every field K, the commutator subgroup is SLn (K), so that the abelianization
of GLn (K) is K × as a multiplicative group with the projection

det : GLn (K) → K × .

In our case, the irreducible one-dimensional representations of G are given by α : F×


q → C via
×

composition:
α ◦ det : G → C× .
We call these representations Uα . They are |F×
q | = q − 1, one for each choice of α. The character
will be

ax bx cx,y dx,y
Uα α(x)2 α(x)2 α(x)α(y) α(x2 − y 2 )

90
8.2. Irreducible representations of GL2 (Fq )

Consider now the projective line


n o
P1 (Fq ) = (x : y) (x, y) ∈ (Fq )2 \ { (0, 0) } ,

which is a group of order q + 1. We have that G acts transitively on P1 (Fq ):

G × P1 (Fq ) −→ P1 (Fq )
a b , (x : y) 7−→ (ax + by : cx + dy).

c d

Thus, we have a a permutation representation G → Aut(P1 (Fq )). Removing the trivial repre-
sentation, we obtain the q-dimensional representation V . We set Vα = V ⊗ Uα . Note that

g Fix(g) χV (g) = # Fix(g) − 1


ax P1 (Fq ) q
bx (1 : 0) 1
cx,y (1 : 0), (0 : 1) 0
dx,y ∅ −1

Hence, we have the character

ax bx cx,y dx,y
Vα q α(x)2 0 α(x)α(y) −α(x2 − y 2 )

Exercise 27. Show that Vα is irreducible and not isomorphic for different α.
To construct other irreducible representations, we will induce representations from the irre-
ducible ones of the Borel subgroup.
Exercise 28. Prove that the commutator subgroup of B is

N=

1 b

b ∈ Fq ,
01

so that the abelianization of the Borel subgroup is BN ' D.


The one-dimensional representations of B are given by

ψα,β : B −→ D −→ C×
a b 7−→ a 0 7−→ α(a)β(d),
 
0d 0d

where α, β : F×
q → C . We set Wα,β = IndB (ψα,β ), whose dimension is
× G

q(q − 1)(q 2 − 1)
dim Wα,β = [G : B] = = q(q + 1).
(q − 1)2
In order to compute the character of Wα,β , we could use Formula 6.3:

χWα,β (g) = χψα,β (gσ−1 ggσ ).


X

σ : gσ=σ

Note that the action on GB is equivalent to that of P1 (Fq ). Indeed, for gσ representative of
σ ∈ GB , we have the map !
a b
gσ = 7→ (a : c),
c d

91
8.2. Irreducible representations of GL2 (Fq )

which is well defined, since for every other element g in σ


! ! !
a b a0 b0 a0 a ∗
g= = 7→ (a0 a : a0 c) = (a : c).
c d 0 d0 a0 c ∗

Thus, we can use what we have discovered about the fixed point of the action on P1 (Fq ).

• The elements ax fix all P1 (Fq ). However, foe every g = a b



c d ∈ G,

g −1 ax g = g −1 x id g = x id = ax ,

so that
χWα,β (ax ) = χψα,β (ax ) = (q + 1)α(x)β(x).
X

P1 (F q)

• The elements bx fix the point ∞ = (1 : 0) only. A representative corresponding to it is


!
1 0
g∞ = ,
0 1

so that g∞ x ∞ = bx and
−1 b g

χWα,β (bx ) = χψα,β (bx ) = α(x)β(x).

• The elements cx,y fix the points 0 = (0 : 1) and ∞ = (1 : 0). A representative corresponding
to 0 is !
0 1
g0 = ,
1 0
so that
! ! !
0 1 x 0 0 1
g0−1 cx,y g0 =
1 0 0 y 1 0
! !
0 y 0 1
=
x 0 1 0
!
y 0
= .
0 x

Thus, the character will be


! !
x 0 y 0
χWα,β (cx,y ) = χψα,β + χψα,β = α(x)β(y) + α(y)β(x).
0 y 0 x

• The elements dx,y fix nothing, so that

χWα,β (dx,y ) = 0.

The character table for Wα,β will be

ax bx cx,y dx,y
Wα,β (q + 1) α(x)β(x) α(x)β(x) α(x)β(y) + α(y)β(x) 0

92
8.2. Irreducible representations of GL2 (Fq )

Exercise 29. Prove that

Wα,β ' Wβ,α , Wα,α ' Vα ⊗ Uα ,

while in the other cases the representations are irreducible and not isomorphic. Thus, we have
(q−1)(q−2)
2 irreducible representations of type Wα,β .

The last type of irreducible representations are the more difficult ones and can be constructed
from the one-dimensional representations of K:

φ : K 7−→ C× .

The representations IndG


K (φ) have dimension [G : K] = q(q − 1) and the characters are

ax bx cx,y dx,y
√ √
IndG
K (φ) q(q − 1) φ(x) 0 0 φ(x + y) + φq (x + y)

q(q−1)
Note that IndG K (φ) ' IndK (φ ), since x = x, so that we have
G q q
2 representations of type
IndK (φ). However, these representations are not irreducible. A deeper analysis of the character
G

reveals that the representations Xφ identified by

χXφ = χV ⊗Wα,1 − χWα,1 − χIndG (φ) ,


K

where α = φ|F× q
, are the missing irreducible representations. The character table of GL2 (Fq )
turns out to be

ax bx cx,y dx,y
Uα α(x)2 α(x)2 α(x)α(y) − y 2 )
α(x2
Vα q α(x)2 0 α(x)α(y) −α(x2 − y 2 )
Wα,β (q + 1) α(x)β(x) α(x)β(x) α(x)β(y) + α(y)β(x) 0
√ √
Xφ (q − 1) φ(x) −φ(x) 0 −φ(x + y) − φq (x + y)

Exercise 30. Prove that the representations Xφ are irreducible and not isomorphic for φ 6= φq .

Just to check our result, we see that

(q − 1)(q − 2) q(q − 1)
dim(V )2 = (q − 1) · 12 + (q − 1) · q 2 + · (q + 1)2 + · (q − 1)2
X

Irr(G)
2 2
1 1
 
= (q − 1) 1 + q 2 + (q 3 + 2q 2 + q − 2q 2 − 4q − 2) + (q 3 − 2q 2 + q)
2 2
 
= (q − 1) q 3 − q = q(q − 1)(q − 1)2 = |G|.

Further results can be found in (Serre, 1977)

93
Bibliography

Fulton, W. and J. Harris. Representation Theory. Vol. 129. Graduate Texts in Mathematics.
Springer-Verlag, 1991.
Macdonald, I. G. Symmetric Functions and Hall Polynomials. Oxford Mathematical Mono-
graphs. Clarendon Press, 1995.
Serre, J.-P. Linear Representations of Finite Groups. Vol. 42. Graduate Texts in Mathematics.
Springer-Verlag, 1977.

95

You might also like