1 Representation theory of finite groups

1.2 Formalities

1.2.1 Subrepresentations and irreducibility

Definition 1.14.

A subrepresentation of a representation (ρ,V) of G is a subspace WV such that ρ(g)wW for all wW.

Then (ρW,W) is also a representation of G, where ρW(g)w=ρ(g)w.

This has a block matrix interpretation: if we choose a basis for W and extend it to a basis for V, then the matrix of ρ(g) will have the following block matrix form:

(ρW(g)𝟎ρV/W(g).).

Here ρV/W is also a representation of G, the quotient representation.

Definition 1.15.

A representation V of G is irreducible if it is nonzero and has no subrepresentations except for {0} and V itself.

Example 1.16.

Let V be the permutation representation of S4 acting on the faces of the cube. Remember that we can see a vector in V as a way of writing a complex number on each face. Then we can spot three subrepresentations:

  • the subrepresentation where the numbers on all faces are equal

  • the subrepresentation where the numbers on opposite faces sum to zero

  • the subrepresentation where the numbers on opposite faces are equal and all the labels sum to zero.

Note that in lectures we did the related example of edges of the tetrahedron.

Definition 1.17.

Suppose that (ρ,V) is a representation of G and that W1,W2 are two subrepresentations such that every element of V can be written uniquely as the sum of an element of W1 and an element of W2. Then we say that V is the (internal) direct sum of W1 and W2 and write V=W1W2.

This also has a block matrix interpretation: if we choose bases for W1 and W2, then the matrix of ρ(g) will have the following block matrix form:

(ρ1(g)𝟎𝟎ρ2(g).).

We can generalize this definition to a finite number of subrepresentations W1,,Wk. If V is the direct sum of subrepresentations that are irreducible, then we will say that V is decomposable. We will see that complex representations of finite groups are always decomposable.

Example 1.18.

Consider the permutation representation of S3. This is the representation on V=3 for which σ(ei)=eσ(i) where e1,e2,e3 are the standard basis vectors. It is not irreducible!

Let W0=e1+e2+e3. Then σ(e1+e2+e3)=e1+e2+e3 for all σ, so W0 is a subrepresentation (and the action of G is trivial).

Let W1={(x,y,z):x+y+z=0.}. Then since S3 permutes the coordinates of a vector, it doesn’t change their sum; so σ(W1)W1, and W1 is a subrepresentation (of dimension two). Since W0W1={0} and their dimensions add to dimV=3, we have V=W0W1.

I claim that W1 is irreducible. Indeed, suppose UW1 is a nonzero subrepresentation; we have to show that U=W1. Let (x,y,z)U be nonzero. As x=y=z can’t happen, we can apply an element of G to permute the coordinates so that xy. Then applying (12), we have (y,x,z)U. Taking the difference, (x-y,y-x,0)U; scaling, (1,-1,0)U. Applying (23), we have (1,0,-1)U. But these vectors span W1 (e.g. because they are linearly independent and dimW1=2), so U=W1 as required.

Example 1.19.

If (ρ,V) is a finite-dimensional representation of G=, with T=ρ(1), then T has an eigenvector v which spans a one-dimensional subrepresentation of V. Thus V is reducible unless dimV=1. The irreducible subrepresentations of V are exactly the lines spanned by T-eigenvectors, and V is decomposable if and only if it has a basis of T-eigenvectors. This is equivalent to T being diagonalizable. As you know from Algebra II (or Linear Algebra I?) this is not always the case! For example,

T=(1101).

The next definition is optional: we will avoid using quotient representations.

Definition 1.20.

Suppose that (ρ,V) is a representation of G and that WV is a subrepresentation. Then the quotient vector space V/W has a representation ρV/W of G defined by

ρV/W(g)v¯=ρ(g)v¯.

Remember that v¯ is the coset v+W. You should check that this is well-defined, i.e. that if v¯=v¯1 then ρ(g)v¯=ρ(g)v1¯.

1.2.2 Homomorphisms and isomorphisms

Definition 1.21.

Suppose that (ρ,V) and (σ,W) are representations of G. Then a G-homomorphism (or homomorphism of representations of G, or map of representations of G, or if we are being lazy just a homomorphism) VW is a linear map ϕ:VW such that

ϕ(ρ(g)v)=σ(g)ϕ(v)

for all vV, gG.

In other words, ϕ ’commutes’ with the action of G: gϕ(v)=ϕ(gv). We write HomG(V,W) for the (vector space) of G-homomorphisms from V to W.

There is another word that is sometimes used for G-homomorphism: intertwiner, or G-intertwiner.

Definition 1.22.

A G-isomorphism (or just an isomorphism) is a bijective G-homomorphism.

If there is a G-isomorphism VW then we write VW.

Lemma 1.23.

Suppose that V and W are representations of G.

  1. 1.

    If THomG(V,W) is an isomorphism, then T-1HomG(W,V).

  2. 2.

    Suppose dimV=dimW=n, and choose bases for them, so that ρV(g) and ρW(g) are (invertible) n×n matrices. Then ρVρW if and only if there exists TGLn() such that

    TρV(g)T-1=ρW(g)

    for all gG.

Proof.

Exercise. ∎

Lemma 1.24.

Given ϕHomG(V,W), then ker(ϕ)V and im(ϕ)W are subrepresentations.

Proof.

We know that they are subspaces, so we just have to show that they are preserved by the action of G.

For the kernel: suppose that vker(ϕ) and gG. Then ϕ(v)=0, and

ϕ(gv)=gϕ(v)=g0=0,

so gvker(ϕ). So ker(ϕ) is G-stable as required.

For the image: suppose wim(ϕ). Then w=ϕ(v) for some vV. Then

gw=gϕ(v)=ϕ(gv)

is also in the image of ϕ. ∎

Example 1.25.

Let G=D3S3 where the isomorphism takes r(123) and s(23). Let (ρ,V) be the permutation representation of G, so in the basis e1,e2,e3 we have

ρ(r)=(001100010),ρ(s)=(100001010).

Then let G act on the equilateral triangle centred at 0 with vertex at (0,1), giving a two dimensional representation σ on W=2 such that

σ(r)=(1/2-3/23/21/2),σ(s)=(-1001).

Let v1, v2, v3 be the vertices labeled anticlockwise from the top and let T:VW be the linear map taking ei to vi. I claim that T is a G-homomorphism.

The matrix of T is

(0-3/23/21-1/2-1/2)

and now we just have to check that

(0-3/23/21-1/2-1/2)(001100010)=(-1/2-3/23/2-1/2)(0-3/23/21-1/2-1/2),

which is true, and a similar equation coming from s.

In fact we could see this without calculation; by the way we defined the isomorphism D3S3, we have that

σ(g)T(ei)=σ(g)vi=vgi=T(egi)=T(ρ(g)ei)

for all gS3.

The kernel of the homomorphism T is the subspace spanned by e1+e2+e3, and in fact T defines an isomorphism from the subrepresentation

{(a,b,c)V:a+b+c=0}V

to W.

Remark 1.26.

If V1 and V2 are representations of G then we may form their (external) direct sum with underlying vector space V1V2 such that

g(v1,v2)=(gv1,gv2)

for all gG, v1V1, v2V2. If V1 and V2 happen to be subrepresentations of some other representation V, then saying V is the direct sum of V1 and V2 is the same as saying that the map

V1V2 V
(v1,v2) v1+v2

is an isomorphism.

1.2.3 Change of group

We haven’t yet covered this section; we will come back to it.

Definition 1.27.

If H is a subgroup of G and (ρ,V) is a representation of G, then we get a representation (ρ|H,V), the restriction of ρ to G, which is given by

ρ|H(h)=ρ(h)

for hH.

This representation is also written V|H, ResHGρ, or ResHGV.

If ρ is an irreducible representation of G, then ρ|H need not be. As an extreme example, if H={e} then ρ|H is just dim(V) copies of the trivial representation!

Example 1.28.

Let ρ be the irreducible two-dimensional representation of S3, realised as the space

V={(x,y,z):x+y+z=0}

of the permutation representation. Let H=123S3. Then ρ|H is reducible (as it must be: all irreducible representations of HC3 are one-dimensional). Indeed, let v=(1,ω,ω2) and w=(1,ω2,ω) where ω=e2πi/3. Then v and w are H-subrepresentations of V with V=vw.

The action of H on v is through the character taking (123) to ω-1, and the action on w is through the character (123)ω.

Exercise: what is the restriction of ρ to {e,(12)}?

If K is a normal subgroup of G, then G/K is a group. If (ρ,V) is a representation of G/K then we define the lift (or inflation) ρ~ of ρ to be the homomorphism GGL(V) defined by ρ~(g)=ρ(gK):

ρ~:GG/K𝜌GL(V).

Then (ρ~,V) is a representation of G.

Example 1.29.

Take G=S4 and take K={e,(12)(34),(13)(24),(14)(23)}. Then there is an isomorphism

G/KS3

defined as follows: label the three non-identity elements of K as {a,b,c}. Then G acts by conjugation on this set of three elements, giving us a homomorphism GS3. Explicitly, we have

(12)(bc)

and

(123)(acb).

Then the homomorphism is surjective with kernel K (check these statements!) and so defines an isomorphism

G/KS3.

If (ρ,V) is any representation of G, then its kernel is

ker(ρ)={gG:ρ(g)=I}.

This is just the kernel of the homomorphism ρ:GGL(V), and so it is a normal subgroup. If the kernel of ρ is trivial, then we say that ρ is faithful. In this case, ρ determines an embedding of G inside GLn()!

Lemma 1.30.

If KG is a normal subgroup, and ρ is a representation of G, then the following are equivalent:

  1. 1.

    ker(ρ) contains K;

  2. 2.

    ρ is isomorphic to the inflation of a representation of G/K.

Proof.

Exercise! ∎

Corollary 1.31.

If ρ is a representation of G/K then ρ~ is irreducible if and only if ρ is.

Proof.

We prove the equivalent statement that ρ~ is reducible if and only if ρ is. If ρ is reducible, so has a nonzero proper subrepresentation W, then so is ρ~ since W is also a subrepresentation of ρ~. Conversely, if ρ~ has a nonzero proper subrepresentation σ, then Kker(ρ)ker(σ) and so, by the lemma, σ=ρ~1 for a representation ρ1 of G/K that is then a nonzero proper subrepresentation of ρ. ∎