Fourier Analysis On Finite Non-Abelian Groups: Terence Tao
Fourier Analysis On Finite Non-Abelian Groups: Terence Tao
TERENCE TAO
1. Hilbert spaces
In these notes, when we say that V is a Hilbert space, we mean that V is a finite
dimensional complex vector space equipped with an inner product ⟨v, w⟩ which is
Hermitian (linear in v, anti-linear in w) and positive definite.
Given any linear transformation T : V → V from a Hilbert space to itself, we
can define its trace tr(T ) by the formula
∑
tr(T ) := ⟨e, T e⟩
e
where e ranges over some orthonormal basis of V ; it is easy to see that this definition
is independent of the choice of basis. Observe that trace is linear and that tr(ST ) =
tr(T S).
If V, W are finite dimensional complex Hilbert spaces, we define HS(V → W ) to
be the space of all Hilbert-Schmidt operators from V to W ; since V, W are finite-
dimensional, HS(V → W ) is just the same as Hom(V → W ). We observe that
HS(V → W ) is also a Hilbert space, with the Hilbert-Schmidt inner product
⟨S, T ⟩HS(V →W ) := tr(ST ∗ ).
where S ∗ is of course the adjoint of S. We abbreviate HS(V → V ) as HS(V ).
We use IV : V → V to denote the identity operator on V ; note that tr(IV ) =
dim(V ).
We use U (V ) ⊂ HS(V ) to denote the space of unitary operators on V , i.e. those
operators T ∈ HS(V → V ) such that T T ∗ = T ∗ T = IV . This forms a group with
identity IV .
2. Finite groups
Let G be a finite group with group operation ◦, identity id, and cardinality |G|.
We give this group normalized counting measure dx:
∫
1 ∑
f (x) dx := f (x),
G |G|
x∈G
3. Representations
By a (unitary) representation ρ of G, we mean a Hilbert space Vρ , together with
a homomorphism ρ : G → U (V ). In other words, for each group element x we
assign a unitary operator ρ(x) : Vρ → Vρ such that ρ(x)ρ(y) = ρ(xy).
Examples of representations include the zero representation 0, with V0 = {0}
and 0(x) := 0; the trivial representation ρ0 , with Vρ0 = C and ρ0 (x) = IC ; and the
regular representation τ , with Vτ := L2 (G) and τ (x) := τx .
A morphism ϕ : ρ1 → ρ2 between two representations is any unitary map ϕ :
Vρ1 → Vρ2 such that we have the intertwining relationship
ρ2 (x)ϕ = ϕρ1 (x) for all g ∈ G.
If the morphism has an inverse, we say it is an isomorphism.
We say that two representations ρ1 , ρ2 are isomorphic, and write ρ1 ≡ ρ2 , if
there is an isomorphism from ρ1 to ρ2 . This is clearly an equivalence relation.
1In the case of infinite groups, the relevant algebra would be L1 (G) rather than L2 (G). For
finite groups, of course, these spaces are the same.
FOURIER THEORY OF FINITE GROUPS 3
= ⟨τ (f )δx , τ (f )δx ⟩ dx
∫G
= ∥f ∗ δx ∥2L2 (G) dx
∫G
= ∥f ∥2L2 (G) dx
G
= ∥f ∥2L2 (G)
as desired.
From (3) we have τ (f ∗ g) = τ (f )τ (g) (this is just a fancy way of saying convo-
lution is associative). Thus the map f 7→ τ (f ) is not only a Hilbert space isometry,
4 TERENCE TAO
5. Irreducibility
If V and W are Hilbert spaces, then V ⊕ W := {(v, w) : v ∈ V, w ∈ W } is also
a Hilbert space, and if T : V → V and S : W → W are Hilbert-Schmidt, we can
define the direct sum T ⊕ S : V ⊕ W → V ⊕ W as (T ⊕ S)(v, w) := (T v, Sw).
Given two representations ρ1 and ρ2 , we can define their direct sum ρ1 ⊕ ρ2 by
defining Vρ1 ⊕ρ2 := Vρ1 ⊕ Vρ2 and (ρ1 ⊕ ρ2 )(x) := ρ1 (x) ⊕ ρ2 (x). By linearity, we
thus have that
(ρ1 ⊕ ρ2 )(f ) = ρ1 (f ) ⊕ ρ2 (f ) (5)
Also observe that ρ1 ⊕ ρ2 ≡ ρ2 ⊕ ρ1 .
We call a representation ρ reducible if it can be written as ρ ≡ ρ1 ⊕ ρ2 for some
non-zero representations ρ1 , ρ2 , and irreducible if it is non-zero and not reducible;
we consider the zero representation to be neither reducible nor irreducible. We will
use Γ to index the irreducible representations of G, up to isomorphism; thus every
ξ ∈ Γ gives rise to an irreducible representation ρξ , and ρξ ≡ ρξ′ if and only if
ξ = ξ ′ . We shall index Γ so that ρ0 is the trivial representation. We shall also
abbreviate Vρξ as Vξ .
Because we are requiring all our representations to be finite dimensional, every
representation ρ can be decomposed as the sum of a finite number of irreducible
representations,
⊕ ⊕c
ρ≡ ρξ ξ
ξ∈Γ
where the cξ are some non-negative integers (called the multiplicity of ρξ in ρ),
⊕c
and ρξ ξ is the direct sum of cξ copies of ρξ . There is an issue as to whether the
coefficients cξ are unique; they are, but we will prove this later when we derive an
explicit formula for them.
In particular, the regular representation can be decomposed as
⊕ ⊕m
τ≡ ρξ ξ (6)
ξ∈Γ
for some non-negative integers mξ . Note that since τ is finite dimensional, at most
finitely many of the mξ are non-zero.
We now fix the mξ (if there is failure of uniqueness, we pick mξ arbitrarily; later
on we will compute mξ explicitly and show that one does have uniqueness, but it
FOURIER THEORY OF FINITE GROUPS 5
is not needed for this present discussion). Now let f ∈ L2 (G). By (6) and (5) we
see that ∑
∥τ (f )∥2HS(L2 (G)) = mξ ∥ρξ (f )∥2HS(Vξ ) .
ξ∈Γ
We rewrite this by defining a measure dξ on Γ by
∫ ∑
f (ξ) dξ := mξ f (ξ).
Γ ξ∈Γ
but this leads of course into the field of representation theory, which we will not
detail here.)
In order to address these two questions we need a fundamental tool in represen-
tation theory, namely Schur’s lemma.
6. Schur’s lemma
A basic fact about irreducible representations is that they have no non-trivial
invariant spaces:
Lemma 6.1. Let ξ ∈ Γ, and let W be a subspace of Vξ such that ρξ (x)W ⊆ W for
all x ∈ G (i.e. W is an invariant space of ρξ ). Then W = Vξ or W = {0}.
Proof Suppose W is a proper subspace of Vξ . Then W ⊥ , the orthogonal comple-
ment of W in Vξ , is also a proper subspace. Since ρξ (x) is unitary and leaves W
invariant, it also leaves W ⊥ invariant (here we are using the finite dimensionality).
Thus ρξ ≡ ρξ |W ⊕ ρξ |W ⊥ , contradicting irreducibility.
This has the following important consequence:
Lemma 6.2 (Schur’s lemma). If ξ, ξ ′ ∈ Γ and ϕ : ρξ → ρξ′ is an morphism, then
either ϕ = 0, or ϕ is an isomorphism. If ξ = ξ ′ , then ϕ is a constant multiple of
the identity IVξ .
Proof Observe that ϕ(Vξ ) is an invariant subspace of Vξ′ , and hence is either equal
to Vξ′ or {0}. Similarly, ker(ϕ) is an invariant subspace of Vξ , and is hence equal
to either Vξ or {0}. Considering all the possibilities we see that ϕ is either 0 or an
isomorphism.
Now suppose ξ = ξ ′ . Then by the first part of Schur’s lemma, we know that
ϕ − λIVξ (which is also a morphism) is either an isomorphism or zero for every
λ ∈ C. Since ϕ must have at least one eigenvalue, the claim follows.
As a consequence, we can obtain the following ergodic theorem.
Lemma 6.3. Let ξ, ξ ′ ∈ Γ, and let T∫ ∈ HS(Vξ → Vξ′ ). Let ⟨T ⟩ ∈ HS(Vξ → Vξ′ )
denote the averaged operator ⟨T ⟩ := G ρξ′ (x)T ρξ (x)−1 dx.
• If ξ ̸= ξ ′ , then
⟨T ⟩ = 0. (13)
′
• If ξ = ξ , then
tr(T )
⟨T ⟩ = IV . (14)
dim(Vξ ) ξ
Proof By a change of variables z := y ◦ x we observe the morphism property
∫
ρξ′ (y)⟨T ⟩ = ρξ′ (y ◦ x)T ρξ (x)−1 dx
∫G
= tr(T ) dx
G
= tr(T )
tr(T ) I as claimed.
we must have tr(⟨T ⟩) = dim(V ξ)
Vξ
Γ
HS(Vξ ) dξ, and mξ = dim(Vξ ).
˜ by
Proof We define the modified measure dξ
∫ ∑
˜ :=
f (ξ) dξ f (ξ) dim(Vξ );
Γ ξ∈Γ
˜
our second task is to show ∫that dξ = dξ.
Let F be any element of Γ HS(Vξ ) dξ, so that each F (ξ) is an element of HS(Vξ ),
and let f ∈ L2 (G) be the modified inverse Fourier transform
∫
˜
f (x) := ⟨F (ξ), ρξ (x)⟩ dξ.
Γ
We claim that F = fˆ; this will prove surjectivity, and by comparing the above
˜ (recall that
formula against the Fourier inversion formula (11) we see that dξ = dξ
F is arbitrary).
Now we show F = fˆ. Expanding this, we see that we have to show that
∫ ∫
F (ξ ′ ) = ˜ dx
⟨F (ξ), ρξ (x)⟩ρξ′ (x) dξ
G Γ
′
for all ξ ∈ G. By Fubini’s theorem, it will suffice to show that
∫ { 1
dim(Vξ ) F (ξ) if ξ = ξ ′
⟨F (ξ), ρξ (x)⟩ρξ′ (x) dξ dx = (15)
G 0 if ξ ̸= ξ ′ .
To do this, we use Lemma 6.3, but with T : Vξ → Vξ′ specialized to a rank 1
operator of the form
Ta,a′ v := ⟨v, a⟩a′ for all v ∈ Vξ
for some a ∈ Vξ , a′ ∈ Vξ′ to be chosen later. By definition, we have
∫
⟨Ta,a′ ⟩v = ρξ′ (x)a′ ⟨ρξ (x)−1 v, a⟩ dx.
G
Specializing to the case v := F (ξ)a, and then summing a over an orthonormal basis,
we obtain
∑ ∫
⟨Ta,a′ ⟩F (ξ)a = tr(ρξ (x)−1 F (ξ))ρξ′ (x)a′ dx.
a G
8 TERENCE TAO
for all a′ ∈ Vξ′ , which proves the second half of (15). When ξ = ξ ′ , we see from
(14) that
∫ ∑ tr(Ta,a′ )
tr(ρξ (x)−1 F (ξ))ρξ (x)a′ dx = F (ξ)a.
G a
dim(Vξ )
But tr(Ta,a′ ) = ⟨a′ , a⟩, so we can rewrite the right-hand side as
1 ∑
F (ξ) ⟨a′ , a⟩a.
dim(Vξ ) a
∑ ′
Since a ranges over an orthonormal basis, a ⟨a , a⟩a = a′ , and this proves the first
half of (15).
One particular consequence of the above is
Theorem
∫ 7.2 (Peter-Weyl theorem). The convolution algebra L2 (G) is isomorphic
to Γ HS(Vξ ) dξ, as an algebra, as a Hilbert space, and as a representation of G
(this last fact follows from (12); in fact the two spaces are isomorphic for G acting
either on the left or on the right).
As a particular consequence of this isomorphism we see from dimension count
that
∑
|G| = dim(Vξ )2 ; (16)
ξ∈Γ
in particular, we
∫ see that Γ is finite.
The space Γ HS(Vξ ) dξ can be thought of as a space of block-diagonal ma-
trices, with each block having dim(Vξ ) rows and columns (and having weight
mξ = dim(Vξ ). This realizes our earlier stated goal of representing the convolution
operators in HS(L2 (G)) as block-diagonal matrices.
8. Conjugation-invariant functions
The group G acts on L2 (G) by conjugation, f (x) → f (y ◦ x ◦ y −1 ), or in other
words f 7→ τy f τy−1 . The group G similarly
∫ acts on HS(Vξ ) by conjugation, F (ξ) 7→
−1
ρξ (y)F (ξ)ρξ (y) , and hence acts on Γ HS(Vξ ) dξ as well. From the Peter-Weyl
theorem we know that these actions are intertwined by the Fourier transform.
In particular, a function f is invariant under conjugation if and only if its Fourier
transform fˆ(ξ) is invariant under conjugation as well. But by Schur’s lemma (or
(14)) we see that an operator in HS(Vξ ) is invariant under conjugation by the
ρξ (y) if and only if it is a constant multiple of the identity IVξ . Thus the Fourier
transform can be restricted to an isomorphism ∫ between the conjugation invariant
functions L2 (G)G of L2 (G), and the space Γ CIVξ dξ.
Conjugation-invariant functions in L2 (G)G are sometimes called class functions
since they are constant on each conjugacy class. The Fourier analysis of class
functions is simpler than that of general functions, because each block HS(Vξ )
in the matrix representation has been replaced by a scalar. To compute more
effectively we introduce the notion of a character.
FOURIER THEORY OF FINITE GROUPS 9
and hence
χξ (x) = tr(ρξ (x)).
∫
Since the δξ are an orthonormal basis of Γ CIVξ dξ, the characters χξ form an
orthonormal basis of L2 (G)G ; in particular we have
∫
f = ⟨f, χξ ⟩χξ dξ
Γ
and in particular
⟨f, χξ ⟩
fˆ(ξ) = IV .
dim(Vξ ) ξ
We can in fact define a character χρ ∈ L2 (G) for every representation ρ, by the
same formula χρ (x) := tr(ρ(x)). From (5) we have χρ1 ⊕ρ2 = χρ1 + χρ2 , so every
character is a linear combination of the irreducible characters. Also, from the linear
independence of the χξ we thus see that the multiplicity of the irreducible represen-
tation ρξ in any other representation ρ is well-defined; in fact, by orthonormality of
the characters, this multiplicity is equal to ⟨χξ , χρ ⟩. Thus one can identify the space
of representations of G (modulo equivalence) with the positive integer combinations
of the characters χξ .
We now specialize to the case where G is abelian. In this case L2 (G)G = L2 (G),
which implies that HS(Vξ ) = CIVξ for all ξ ∈ Γ; in other words, all irreducible
representations ρξ must be one-dimensional. Thus we may as well set Vξ := C, and
the characters χξ are essentially the same as the representations ρξ . In particular,
since ρξ is unitary, we have |χξ (x)| = 1 for all x ∈ G.
Now observe that if ρξ and ρξ′ are one-dimensional representations, then ρξ ρξ′
is another one-dimensional representation, which must of course be irreducible. We
can thus define an addition operation on Γ by setting ρξ+ξ′ := ρξ ρξ′ . This can
easily be verified to turn Γ into an abelian group, sometimes called the dual group
G∗ of G; from (16) we thus see that |Γ| = |G|. This theory then collapses to the
familiar theory of the Fourier transform on abelian groups.
The trick of multiplying two representations ρξ and ρξ′ to form a third repre-
sentation ρξ ρξ′ is not restricted to one-dimensional representations; it is easy to
take two higher-dimensional representations ρ and ρ′ and form a tensor product
10 TERENCE TAO