Chapter 4: Introduction To Representation Theory: Preprint Typeset in JHEP Style - HYPER VERSION
Chapter 4: Introduction To Representation Theory: Preprint Typeset in JHEP Style - HYPER VERSION
Gregory W. Moore
2. Basic Definitions 4
2.1 Representation of a group 4
2.2 Matrix Representations 5
2.3 Examples 6
2.3.1 The fundamental representation of a matrix Lie group 6
2.3.2 The determinant representation 6
2.3.3 A representation of the symmetric groups Sn 6
2.3.4 Z and Z2 7
2.3.5 The Heisenberg group 7
3. Unitary Representations 8
3.1 Invariant Integration 9
3.2 Unitarizable Representations 10
3.3 Unitary representations and the Schrödinger equation 11
8. Schur’s Lemmas 22
–1–
11.5 Application: Bloch’s Theorem in Solid State Physics 37
11.6 The Heisenberg group extension of Ŝ × S for an abelian group S 39
18. Algebras 71
18.1 Coalgebras, Bialgebras, and Frobenius algebras 72
18.2 When are two algebras equivalent? Introduction to Hochschild cohomology 73
–2–
21. Applications of the Projection Operators 89
21.1 Decomposition of a representation into its isotypical parts 90
21.2 Block diagonalization of Hermitian operators 92
21.2.1 Projecting quantum wavefunctions 92
21.2.2 Finding normal modes in classical mechanics 93
23. Symmetric groups and tensors: Schur-Weyl duality and the irreps of
GL(d, k) 99
23.1 Free fermions on a circle and Schur functions 103
23.1.1 Schur functions, characters, and Schur-Weyl duality 108
23.2 Bosons and Fermions in 1+1 dimensions 111
23.2.1 Bosonization 111
∀g ∈ G 7→ U (g) : H → H (1.1)
where U (g) is a unitary operator. Moreover, these unitary operators should “act in the
same way as the physical symmetry.” Mathematically, this means that we have the operator
equation:
Moreover, if we have a symmetry of the physical system we should have the same kind
of time-development of two systems related by the symmetry, so
–3–
2. Basic Definitions
T : g 7→ T (g)
(2.1) eq:repii
G → GL(V )
commutes. Equivalently,
for all g ∈ G.
for all g ∈ G.
Familiar notions of linear algebra generalize to representations:
1. The direct sum ⊕, tensor product ⊗ etc. of representations. Thus, the direct sum of
(T1 , V1 ) and (T2 , V2 ) is the rep (T1 ⊕ T2 , V1 ⊕ V2 ) where the representation space is
V1 ⊕ V2 and the operators are:
(T1 ⊕ T2 )(g) v1 , v2 := T1 (g)v1 , T2 (g)v2 (2.5)
–4–
2. Similarly, for the tensor product, the carrier space is V1 ⊗ V2 and the group elements
are represented by:
(T1 ⊗ T2 )(g) v1 ⊗ v2 := (T1 (g)v1 ) ⊗ (T2 (g)v2 ) (2.6)
T : G → GL(n, κ) (2.8)
Given a representation, and an ordered basis {~v1 , . . . , ~vn } for V we can associate to a
representation a matrix representation. The point is, with a basis we can identify GL(V ) ∼ =
GL(n, κ). Specifically, we get the matrix from the linear transformation by:
n
X
T (g)~vk = T (g)jk~vj (2.9)
j=1
so the T (g)jk are the matrix elements of an element of GL(n, κ). We will sometimes denote
this matrix as T (g) if it is understood we are using an ordered basis for V .
Now, if we change basis
n
X
~vi = Sji~vj0 (2.10)
j=1
–5–
If we choose a basis vi for V then this is the matrix representation for the dual repre-
sentation in the dual basis v̂i .
b.) If T is a complex matrix representation wrt basis vi then the complex conjugate
representation with respect to v̄i is: g → T (g)∗ .
c.) If T is a real representation, then there exists an S ∈ GL(n, C) such that for all
g ∈ G:
T ∗ (g) = ST (g)S −1 (2.13)
Warning: The matrix elements T (g)ij of a real representation can of course fail to be
real numbers!
2.3 Examples
2.3.1 The fundamental representation of a matrix Lie group
GL(n, κ), SL(n, κ), O(n, κ), U (n) are all matrix representations of themselves! In the first
three examples V = κn . In the fourth V = Cn . These are called “the fundamental
representation.” Note that when κ = Cn and n > 2 the fundamental representation is not
equivalent to its complex conjugate 1 so the fundamental representation is not the same as
the “minimal-dimensional nontrivial representation.”
and thus the matrix representation has matrix elements which are just 0’s and 1’s with a
single nonzero entry in each row and column.
1
This is easily proven using characters, see below.
–6–
Explicitly for n = 2 we have, for examples:
!
01
(12) → (2.18) eq:symmrep
10
2.3.4 Z and Z2
–7–
Let V be the “shift operator:”
0 1 0 ··· 0
! 010 0 0 1 ··· 0
01
V = V = 0 0 1 V = · · · ··· ··· ··· ··· (2.22)
10
100 · · · ··· ··· ··· 1
1 0 ··· ··· 0
3. Unitary Representations
sec:sUR
Of particular importance in physics are the unitary representations. By Wigner’s theorem
described in section **** above we know that symmetry transformations will act as unitary
or anti-unitary transformations. The basic reason for this is that these preserve norms and
hence probability amplitudes.
T : G → U (n) (3.2)
Exercise
a.) Show that if T (g) is a rep on an inner product space then T (g −1 )† is a rep also.
b.) Suppose T : G → GL(V ) is a unitary rep on an inner product space V . Let {~vi }
be an ON basis for V . Show that the corresponding matrix rep T (g)ij is a unitary matrix
rep.
c.) Show that for a unitary matrix rep the transpose-inverse and complex conjugate
representations are equivalent.
–8–
Definition 3.2. If a rep (T, V ) is equivalent to a unitary rep then such a rep is said to be
unitarizable.
Example. A simple example of non-unitarizable reps are the detµ reps of GL(n, k)
described in section 2.3.4.
Important Remark: The notion of averaging over the group can be extended to a much
larger class of groups than finite groups. It is given by invariant integration over the group
which is the analog of the operation
1 X
f→ f (g) (3.3)
|G| g
on functions.
In general we replace Z
1 X
f (g) → f (g)dg (3.4)
|G| g G
R
where G f (g)dg should be regarded as a rule such that:
R
1. G f (g)dg is a complex number depending linearly on f ∈ RG .
2. It satisfies the left invariance property:
Z Z
f (hg)dg = f (g)dg (3.5)
G G
for all h ∈ G. This is the generalization of the rearrangement lemma. (We can define
right invariance in a similar way. Compact groups admit integration measures which are
simultaneously left- and right- invariant. )
Examples:
Z Z +2π
dθ
f (g)dg ≡ f (θ)
G=U (1) 0 2π
Z X
f (g)dg ≡ f (n) (3.6) eq:intgroup
G=Z n∈Z
Z Z +∞
f (g)dg ≡ dxf (x)
G=R −∞
The notion of invariant integration also extends to all compact groups. For the impor-
tant case of G = SU (2) we can write it as follows. First, every element of SU (2) can be
written as: ♣Definition of β
here is backwards
from what we use
when we describe
reps using
homogeneous
polynomials below.
♣
–9–
!
α β
g= (3.7)
−β ∗ α∗
for 2 complex numbers α, β with
|α|2 + |β|2 = 1. (3.8)
In this way we identify the group as a manifold as S 3 . That manifold has no globally
well-defined coordinate chart. The best we can do is define coordinates that cover “most”
of the group but will have singularities are some places. (It is always important to be
careful about those singularities when using explicit coordinates!) One way to do this is to
write
1
α = ei 2 (φ+ψ) cos θ/2
1
(3.9)
β = iei 2 (φ−ψ) sin θ/2
Proof for finite groups: If T is not already unitary with respect to the inner product
(·, ·)1 then we can define a new inner product by:
1 X
hv, wi2 ≡ hT (g)v, T (g)wi1 (3.12)
|G|
g∈G
This proof generalizes using a left- and right- invariant Haar measure, which always exists
for a compact group.
2
We have chosen an orientation so that, with a positive constant, this is Tr2 (g −1 dg)3 ).
– 10 –
Exercise
a.) Show that
hT (g)v, T (g)wi2 = hv, wi2 (3.13)
and deduce that T (g) is unitary with respect to h·, ·i2 .
b.) Show that if T (g) is a finite-dimensional matrix rep of a finite group then it is
equivalent to a unitary matrix rep.
~2
− ∆ψ(~x) + V (~x)ψ(~x) = Eψ(~x) (3.14) eq:shro
2m
and suppose V is rotationally invariant:
Now define
VE = {ψ(~x) : ψsolves(3.14) and is normalizable} (3.16)
Normalizable, or square integrable means:
Z
2
kψk = | ψ(~x) |2 d3 x < ∞ (3.17)
R3
so VE ⊂ L2 (R3 ).
Claim: VE is a representation space of O(3): T : O(3) → GL(VE ) given, as usual, by:
Must check:
1. ψ solves (3.14) ⇒ T (g)ψ solves (3.14).
2. k ψ k2 < ∞ ⇒k T (g)ψ k2 < ∞
Check of 1: If x0i = (g −1 )ji xj then
3 3
X ∂2 X ∂2
∆0 = = =∆ (3.19)
i=1
∂x0i 2 i=1
∂x2i
and therefore:
∆ T (g)ψ (x) = ∆(ψ(x0 )) = ∆0 (ψ(x0 ))
(3.20)
Check of 2: x0 = g −1 x gives
0
dxi
d3 x0 =| det | d3 x = d3 x (3.21)
dxj
– 11 –
so:
k ψ k2 =k T (g)ψ k2 (3.22)
cal states of a theory form unitary reps of the symmetry group. Since the group commutes with the Hamiltonian we can diag
Thus, for example, the wavefunctions in a spherically symmetric potential have quan-
tum numbers, |E, j, m, · · · i.
∆u(~x) + k 2 u(~x) = 0
3
X ∂2 (3.23) eq:hlmhltz
∆=
i=1
∂x2i
Let
– 12 –
Example Returning to our clock and shift operators, we can define a projective repre-
sentation of Zn × Zn by
0
(s̄, s̄0 ) → U s̄ V s̄ (4.3)
but these only satisfy the group law up to a power of ω. In fact, what we have is a
representation of Heis(Zn × Zn ).
1. If G is a left-action on X then
2. If G is a left-action on X then
3. If G is a right-action on X then
4. If G is a right-action on X then
Example: Consider a spacetime S. With suitable analytic restrictions the space of scalar
fields on S is Map(S, κ), where κ = R or C for real or complex scalar fields. If a group
G acts on the spacetime, there is automatically an induced action on the space of scalar
fields. To be even specific, suppose X = M1,d−1 is d-dimensional Minkowski space time, G
is the Poincaré group, and Y = R. Given one scalar field Ψ and a Poincaré transformation
g −1 · x = Λx + v we have (g · Ψ)(x) = Ψ(Λx + v).
Similarly, suppose that X is any set, but now Y is a G-set. Then again there is a
G-action on Map(X, Y ):
– 13 –
We can now combine these two observations and get the general statement: We assume
that both X is a G1 -set and Y is a G2 -set. We can assume, without loss of generality, that
we have left-actions on both X and Y . Then there is a natural G1 ×G2 -action on Map(X, Y )
defined by:
φ((g1 , g2 ), Ψ)(x) := g2 · (Ψ(g1−1 · x)) (5.6) eq:GenAction
note that if one writes instead g2 · (Ψ(g1 · x)) on the RHS then we do not have a well-defined
G1 × G2 -action (if G1 and G2 are both nonabelian). In most applications X and Y both
have a G action for a single group and we write
This is a special case of the general action (5.6), with G1 = G2 = G and specialized to the
diagonal ∆ ⊂ G × G.
Example: Again let X = M1,d−1 be a Minkowski space time. Take G1 = G2 and let
G = ∆ ⊂ G × G be the diagonal subgroup, and take G to be the Poincaré group. Now
let Y = V be a finite-dimensional representation of the Poincaré group. Let us denote the
action of g ∈ G on V by ρ(g). Then a field Ψ ∈ Map(X, Y ) has an action of the Poincaré
group defined by
g · Ψ(x) := ρ(g)Ψ(g −1 x) (5.8)
This is the standard way that fields with nonzero “spin” transform under the Poincaré
group in field theory. As a very concrete related example, consider the transformation of
electron wavefunctions in nonrelativistic quantum mechanics. The electron wavefunction
is governed by a two-component function on R3 :
!
ψ+ (~x)
Ψ(~x) = (5.9)
ψ− (~x)
Now that we have seen several examples of representations we would like to introduce a
“universal example,” of a representation of G, called the regular representation. We will
see that from this representation we can learn about all of the representations of G, at
least when G is compact.
– 14 –
Recall once again the principle of section 5: If X has G-action then the space FX,Y of
all functions X → Y also has a G-action. In particular, if Y is a vector space then FX,Y is
a vector space and we have a natural source of representations of G. In particular, we can
make the simplest choice Y = C, so the space of complex-valued functions on X, FX,C is
a very natural source of representations of G. ♣THIS IS OLD.
SHOULD MAKE IT
If G acts on nothing else, it certainly acts on itself as a group of transformations of A REPRESENTA-
TION OF G × G
X = G with left or right action. Applying the above general principle we see that the FROM THE
START. ♣
space of complex-valued functions on G:
φ 7→ L(g) · φ (6.2)
φ 7→ R(g) · φ (6.4)
Exercise
Check that
L(g1 )L(g2 ) = L(g1 g2 ) (6.6)
and that
R(g1 )R(g2 ) = R(g1 g2 ) (6.7)
– 15 –
Now let us assume | G |< ∞. Then RG is a finite dimensional vector space. we can
label a function f by its values f (g).
Exercise
If G is a finite group, then show that
Exercise
a.) Let δ0 , δ1 , δ2 be a basis of functions in the regular representation of Z3 which are
1 on 1, ω, ω 2 , respectively, and zero elsewhere. Show that ω is represented as
010
L(ω) = 0 0 1 (6.12)
100
b.) Show that
L(h) · δg = δh·g
(6.13)
R(h) · δg = δg·h−1
and conclude that for the left, or right, regular representation of a finite group the repre-
sentation matrices in the δ-function basis have entries which are simply zeroes or ones in
the natural basis.
– 16 –
6.2 RG as a unitary rep: Invariant integration on the group
When G is a finite group we can turn the vector space RG into an inner product space by
the defining an inner product using averaging over the group:
1 X ∗
hφ1 , φ2 i := φ1 (g)φ2 (g) (6.14) eq:innprod
|G|
g∈G
Theorem. With the inner product (6.14) the representation space RG is unitary for both
the RRR and the LRR.
If we choose a basis wµ for V then the operators ρ(g) are represented by matrices:
X
ρ(g) · wν = D(g)µν wµ (6.19)
µ
If we take T = eνµ to be the matrix unit in this basis then ΨT is the function on G given by
the matrix element D(g −1 )µν . So the ΨT ’s are linear combinations of matrix elements of
– 17 –
the representation matrices of G. The advantage of (6.18) is that it is completely canonical
and basis-independent.
Note that ι : T 7→ ΨT “commutes with the G × G action.” What this means is that
(The reader should check this carefully.) Such a map is said to be equivariant. Put
differently, denoting by ρEnd(V ) the representation of G × G on End(V ) and ρReg.Rep. the
representation of G × G on Map(G, C) we get a commutative diagram:
End(V )
ι / Map(G, C) (6.21)
ρEnd(V ) ρReg.Rep.
End(V )
ι / Map(G, C)
left-right-invariant measure on G. 3
7.1 Definitions
Sometimes reps are “too big” and one wants to reduce them to their “essential parts.”
Reducing a representation to smaller representations is closely analogous to diagonalization
of matrices.
For example – we will see that the natural representation of a group - its regular
representation - is in fact highly reducible.
3
In order for this to be completely correct we need to assume that G is a compact group. Then we
introduce a left-right-invariant measure on G and replace the LHS by L2 (G).
– 18 –
Definition. Let W ⊂ V be a subspace of a group representation T : G → GL(V ). Then
W is invariant under T if ∀g ∈ G, w ∈ W
T (g)w ∈ W (7.1)
Note:
• Given any nonzero vector v ∈ V , the linear span of {T (g)v}g∈G is an invariant
subspace. In an irrep this will span all of V .
• If W is an subrepresentation of V then T descends to a representation on V /W .
Let us see how this works in terms of matrix representations. Suppose T is reducible.
Then we can choose a basis
{v1 , . . . vk } for W
and
{v1 , . . . vk , vk+1 , . . . vn } for V
In this basis the matrix representation of T looks like:
!
T1 (g) T12 (g)
T (g) = (7.2) eq:redrep
0 T2 (g)
where
T1 (g) ∈ M atk×k
T12 (g) ∈ M atk×n−k
T2 (g) ∈ M at(n−k)×(n−k)
Writing out T (g1 )T (g2 ) = T (g1 g2 ) we see that T1 is the representation on W and T2
is the representation on V /W . T12 transforms in a more complicated way.
Definition. A representation T is called completely reducible if it is isomorphic to
W 1 ⊕ · · · ⊕ Wn (7.3)
– 19 –
where the Wi are irreducible reps. Thus, there is a basis in which the matrices look like:
T1 (g) 0 0 ···
0 T (g) 0
2 ···
T (g) = (7.4)
eq:comredrep
···
0 0 T3 (g)
··· ··· ··· ···
Examples
• G = Z2
!
10
1→
01
! (7.5)
01
(12) →
10
function on G nµ function on G
z }| { X z}|{
R(g) · Tij = Tsj (g) Tis (7.9) eq:mtxeltinvt
| {z }
s=1
matrix element for RG
– 20 –
7.2 Reducible vs. Completely reducible representations
subsec:ssRedC
Irreps are the “atoms” out of which all reps are made. Thus we are naturally led to study
the irreducible reps of G. In real life it can and does actually happen that a group G
has representations which are reducible but not completely reducible. Reducible, but not
completely reducible reps are sometimes called indecomposable.
of GL(n, R). As we will see, the Poincare group has indecomposable reps.
R_G
W^\perp
y ∈ W ⊥ ⇔ ∀x ∈ W, hy, xi = 0 (7.11)
Let g ∈ G, y ∈ W ⊥ . Compute
T (g)y, x = y, T (g)† x
(7.12)
= y, T (g −1 )x
– 21 –
therefore: T (g)y ∈ W ⊥
therefore: W ⊥ is an invariant subspace. ♠
In particular, by section 4:
1. Finite dimensional reps of finite groups are completely reducible.
2. More generally, finite dimensional reps of compact Lie groups are completely re-
ducible.
It follows that for a finite group the regular representation RG is completely re-
ducible. In the next several sections we will show how to decompose RG in terms of
the irreducible representations of G.
It contains the representation µ with the correct degeneracy in V . It is called the isotypical
component belonging to µ. Note that it can be written as
8. Schur’s Lemmas
♣This section needs
improvement:
When thinking about irreducible representations it is important to understand how they Schur’s lemma says
that the algebra of
are related to each other. A key technical tool is Schur’s lemma. It is usually stated as intertwiners of an
irrep is a division
two separate statements, each of which is practically a triviality. algebra. So R, C, H.
What is here is
correct, but not the
best viewpoint. ♣
– 22 –
Proof: kerA and Im A are both invariant subspaces. Of course, if V and V 0 are
inequivalent then A = 0. ♠.
Proof: Since we are working over the complex field A has a nonzero eigenvector Av =
λv. The eigenspace C = {w : Aw = λw} is therefore not the zero vector space. But it is
also an invariant subspace. Therefore, it must be the entire carrier space. ♠.
Remarks:
• Consider the isotypical decomposition of a completely reducible representation. By
Schur’s lemma, the intertwiners of V (µ) with itself are of the form K ⊗ 1 where K ∈
End(Caµ ) is arbitrary.
• Schur’s lemma is used in quantum mechanics very often. Note that a Hamiltonian
invariant under some symmetry group U (g)HU (g)−1 = H is an intertwiner. Therefore,
if we decompose the Hilbert space of states into irreps of G then H must be scalar on
each irrep. In particular this leads to important selection rules in physics. For example, if
∆H is an invariant operator which commutes with some symmetry G of a physical system
then the matrix elements of ∆H between states in different irreducible representations of
G must vanish.
Let G be a finite group. We are interested in finding the reduction of RG to its irreducible
representations. Our next technical tool - very useful in a variety of contexts - are the
orthogonality relations of matrix elements of irreps.
Label the distinct, i.e. inequivalent, irreducible representations by
T (µ) µ = 1, 2, . . . , r (9.1)
and let nµ = dimT (µ) . (We will prove later that r < ∞ and nµ < ∞.) In order not to
overburden the notation we will denote the carrier space and the homomorphism by the
same name T (µ) , relying on context to tell which is intended. Choose bases for T (µ) and
consider these to be matrix irreps.
Let B ∈ M atnµ ×nν , and define a matrix A ∈ M atnµ ×nν by: ♣SHOULD GIVE A
MORE
CONCEPTUAL
1 X (µ) DESCRIPTION OF
A≡ T (g)BT (ν) (g −1 ) (9.2) eq:orthi
WHAT IS GOING
|G| ON HERE USING
g∈G P ∨
g T (g) ⊗ T (g).
♣
We claim that A is an intertwiner, that is, ∀h ∈ G:
– 23 –
Proof:
1 X (µ)
T (µ) (h)A = T (h)T (µ) (g)BT (ν) (g −1 )
|G|
g∈G
1 X (µ)
= T (hg)BT (ν) (g −1 )
|G|
g∈G
1 X (µ) 0 (9.4)
= T (g )BT (ν) ((h−1 g 0 )−1 )
|G| 0
g ∈G
1 X (µ)
= T (g)BT (ν) ((g)−1 )T (ν) (h)
|G|
g∈G
(ν)
= AT (h)
In the third line let g 0 = hg, and use the rearrangement lemma.
Therefore, by Schur’s lemma:
(µ)
X
(ν) −1
Ti` (g)Tms (g ) = |G|λδµν δis (9.6) eq:orthiii
g∈G
(µ)
X
Tm` (g −1 g) = |G|λnµ
g∈G
|G|δ`m = |G|λnµ
(9.7)
δ`m
⇒λ=
nµ
Putting it all together we get the main result of this section:
Theorem The orthogonality relations for the group matrix elements are:
1 X (µ) (ν) 1
Ti` (g)T̂sm (g) = δµν δis δ`m (9.8) eq:orthogmatrx
|G| nµ
g∈G
– 24 –
where the Haar measure [dg] is normalized to unit volume. Recall that
(ν) −1,tr
T̂sm (g) = (T (ν) (g))sm (9.10)
1 X (µ) ∗
φi` (g)(φ(ν)
sm (g)) = δµν δis δ`m (9.12) eq:onrels
|G|
g∈G
(µ)
Interpretation: The vectors φij form an orthonormal set of vectors in RG .
Below we will apply this to the decomposition of the regular representation into irreps.
– 25 –
Let us define
µ n (µ)
W := ⊕µ ⊕i=1 Ri ⊂ RG
z nµ }|times
{ (10.4)
W ∼ ⊕
= µ T µ
⊕ · · · ⊕ T µ
Note:
(µ)
• W = Span{Tij }.
(µ)
• By the orthogonality relations the Tij are linearly independent. Thus we really do
have a direct sum .
Exercise
(µ)
Show that the set of functions Tij with µ, j held fixed forms a copy of the irrep T (µ)
under the left regular rep. More on this below.
At this point the possibility remains that W is a proper subspace of RG . We will now
(µ)
show that in fact W = RG exactly, by showing that Tij span RG .
To do this recall the vector space RG is an inner product space by the rule
1 X ∗
(φ1 , φ2 ) ≡ φ1 (g)φ2 (g) (10.5) eq:innproda
|G|
g∈G
Recall:
1. RG is a unitary representation, wrt this product.
2. The normalized matrix elements φµij are orthonormal wrt this inner product. This
follows from (9.13).
By the proposition 7.2.1 we can use the inner product to decompose
RG ∼
= W ⊕ W⊥ (10.6)
– 26 –
Tsiµ (h)fs (g)
X
(R(h)fi )(g) =
s
⇒
Tsiµ (h)fs (g)
X
fi (gh) =
s
(10.8)
⇒
Tsiµ (g −1 g 0 )fs (g)
X
fi (g 0 ) =
s
µ µ 0
X
= Tsk (g −1 )Tki (g )fs (g)
s,k
That is, fi is a linear combination of the functions φµij . But this contradicts the
assumption that
{fi } ⊂ W ⊥ (10.10)
so W ⊥ = 0. ♠.
Thus, we have arrived at the decomposition of the RRR into its irreducible pieces:
z nµ }|
times
{
∼ µ
RG = ⊕µ T ⊕ · · · ⊕ T µ
(10.11) eq:rrrep
But we are not quite done. To arrive at the most beautiful statement we should study
RG as a representation under the the left regular rep. So we consider the linear subspaces
where we hold µ, j fixed and consider the span:
(µ) (µ)
Lj ≡ Span{φij }i=1,...,nµ (10.12)
– 27 –
forms the rep T̂ (µ) ⊗ T (µ) of GL × GR . Finally, we can identify
T̂ (µ) ⊗ T (µ) ∼
= End(T (µ) ) (10.15)
−1
T (gL , gR ) · Φ = T (µ) (gL )ΦT (µ) (gR ) (10.16)
Theorem 10.1 Let G be a finite group. The matrix elements φµij of the distinct irreps of
G decompose RG into irreps of GLeft × GRight :
RG ∼
= ⊕µ T̂ (µ) ⊗ T (µ) ∼
= ⊕µ End(T (µ) ) (10.17)
Moreover, if we use matrix elements of unitary irreps in an ON basis then the φµij form an
ON basis for RG and
RG ∼= ⊕µ T (µ) ⊗ T (µ) (10.18)
Remarks:
• This theorem generalizes beautifully to all compact groups and is known as the
Peter-Weyl theorem.
• This leaves the separate question of actually constructing the representations of the
finite group G. Until recently there has been no general algorithm for doing this.
Recently there has been a claim that it can be done. See
Vahid Dabbaghian-Abdoly, ”An Algorithm for Constructing Representations of Finite
Groups.” Journal of Symbolic Computation Volume 39, Issue 6, June 2005, Pages 671-688
Exercise
Show that the number of distinct one-dimensional representations of a finite group G
is the same as the index of the commutator subgroup [G, G] in G. 4
************************************ ***************************************
MATERIAL FROM MATHMETHODS 511 2015:
For now, we just content ourselves with the statement of the theorem for G a finite
group:
4
Hint: A one-dimensional representation is trivial on [G, G] and hence descends to a representation of
the abelianization of G, namely the quotient group G/[G, G].
– 28 –
Theorem: Let G be a finite group, and define an Hermitian inner product on L2 (G) =
Map(G, C) by
1 X ∗
(Ψ1 , Ψ2 ) := Ψ (g)Ψ2 (g) (10.20)
|G| g 1
Then let {Vλ } be a set of representatives of the distinct isomorphism classes of irreducible
(λ)
unitary representations for G. For each representation Vλ choose an ON basis wµ , µ =
λ (g) defined by
1, . . . , nλ := dimC Vλ . Then the matrix elements Dµν
nλ
X
ρ(g)wν(λ) = λ
Dµν (g)wµ(λ) (10.21)
µ=1
1 λ1 ,λ2
(Dµλ11 ν1 , Dµλ22 ν2 ) = δ δµ1 ,µ2 δν1 ,ν2 (10.22) eq:PW-2
nλ
Idea of proof : The proof is based on linear algebra and Schur’s lemma. The normalization
constant on the RHS of (10.22) is easily determined by setting λ1 = λ2 and ν1 = ν2 = ν
and summing on ν, and using the hypothesis that these are matrix elements in a unitary
representation. The relation to (6.22) is obtained by noting that the linear transformations
(λ)
T = eνµ , given by matrix units relative to the basis wµ form a basis for End(Vλ ). ♠
– 29 –
a generator ω = exp[2πi/n]. Since G is abelian all the representation matrices can be
simultaneously diagonalized so all the irreps are one-dimensional. They are:
V = C and ρm (ω) = ω m where m is an integer. Note that m ∼ m + n so the set of
irreps is again labeled by Z/nZ and in fact, under tensor product the set of irreps itself
forms a group isomorphic to Z/nZ.
The matrix elements in the irrep (ρm , V ) are
mj
D(m) (ω j ) = ω mj = e2πi n (10.27)
The decomposition of a function Ψ on the group G is known as the discrete Fourier trans-
form.
Remark: The theorem applies to all compact Lie groups. For example, when G = U (1) =
{z||z| = 1} then the invariant measure on the group is just −i dz dθ iθ
z = 2π where z = e :
Z 2π
dθ
(Ψ1 , Ψ2 ) = (Ψ1 (θ))∗ Ψ2 (θ) (10.29)
0 2π
Now, again since G is abelian the irreducible representations are 1-dimensional and the
unitary representations are (ρn , Vn ) where n ∈ Z, Vn ∼
= C and
ρn (z) := z n (10.30)
Now, the orthonormality of the matrix elements is the standard orthonormality of einθ and
the Peter-Weyl theorem specializes to Fourier analysis: An L2 -function Ψ(θ) on the circle
can be expanded in terms of the matrix elements of the irreps:
X
Ψ= Ψ̂n D(n) (10.31)
Irreps ρn
When applied to G = SU (2) the matrix elements are known as Wigner functions or
monopole harmonics. They are the matrix elements
j
Dm L ,mR
(g) := hmL |ρj (g)|mR i (10.32)
in the standard ON basis of the unitary spin j representation diagonalizing the diagonal
subgroup of SU (2). So
1 3
j = 0, , 1, , . . . , mL , mR ∈ {−j, −j + 1, . . . , j − 1, j} (10.33)
2 2
Recall that SU (2) ∼= S 3 as a manifold. Using the standard volume form, with unit vol-
ume we can define L2 (SU (2)). The entire theory of spherical harmonics and Legendre
polynomials is easily derived from basic group theory. ♣MAKE GOOD
ON THIS CLAIM
BELOW. ♣
************************************ ***************************************
– 30 –
11. Fourier Analysis as a branch of Representation Theory
By the same reasoning the space of 1-dimensional unitary irreps of a group is a group.
Note that in this case (T )−1 (g) = T ∗ (g) and T (g) is always represented by a phase.
Definition The space of unitary irreps of a group G is called the unitary dual and is
denoted by Ĝ.
In the case when G is abelian, the unitary dual Ĝ is itself a group. It is sometimes
called the Pontryagin dual group.
Examples
• Finally, let us consider Zn thought of as multiplicative nth roots of unity. For any
integer s we can define a representation by taking
where ω is any nth root of unity. These are clearly unitary irreps. Note that
– 31 –
and
T (s1 ) ⊗ T (s2 ) ∼
= T (s1 +s2 ) (11.8)
so the dual group is another copy of Zn :
cn ∼
Z = Zn (11.9)
n = 12 + 12 + · · · + 12 (11.10)
• The theory we are discussing in this section extends to locally compact abelian groups
such as G = R. The unitary irreps are specified by a “momentum” k ∈ R.
This is a rep because T (k) (x + y) = T (k) (x)T (k) (y). It is unitary for k real.
Now notice that
0 0
T (k) ⊗ T (k ) = T (k+k ) (11.12)
so we conclude:
b∼
R =R (11.13)
• G = Z. Unitary irreps of Z are labelled by a real number θ:
Note that
T (θ) = T (θ+2π)
0 0 (11.15) eq:unirri
T (θ) ⊗ T (θ ) = T (θ+θ )
b∼
Z = U (1) (11.16) eq:dlgrp
• G = R/Z ∼
= U (1). Now the unitary irreps are labelled by n ∈ Z:
and we check:
0 0
T (n) ⊗ T (n ) = T (n+n ) (11.18) eq:unirriii
so
[
U (1) ∼
=Z (11.19) eq:dlgrpp
– 32 –
• G = Λ ⊂ Rd , a lattice. Given a vector ~k ∈ Rd we can define a unitary irrep by:
~ ~
T (k) : ~n → e2πik·~n (11.20)
b∼
Λ = Rd /Λ∗ (11.23) eq:torirep
thus generalizing Z
b = U (1).
• . G = Rd /Λ is a d-dimensional torus.
G =G (11.24)
bb
Clearly:
f0 + f1 (0) f0 − f1 (1)
f= T + T (11.27)
2 2
Let us generalize this:
•
G = Zn ∼ = {1, ω, ω 2 , . . . , ω n−1 } where ω = e2πi/n . We worked out the unitary irreps
above.
– 33 –
The orthogonality relations on the irreducible matrix elements (11.6) flow from:
1 X (j) ` 1 X 2πi(j`−k`)/n
T (ω )(T (k) (ω ` ))∗ = e
n n (11.28)
` `
= δj,k
n−1
X
f= fˆj T (j) (11.29)
j=0
Z Z +2π
dθ
f (g)dg ≡ f (θ)
G=U (1) 0 2π
Z X
f (g)dg ≡ f (n) (11.31) eq:intgroupp
G=Z n∈Z
Z Z +∞
f (g)dg ≡ dxf (x)
G=R −∞
With this in mind the above Theorem 10.1 becomes the following statements:
• Let us begin with G = (R, +) then Ĝ = (R, +). Unitary irreps are labelled by k ∈ R:
1
T (k) (x) = √ e2πikx (11.32)
2π
The orthogonality relations are:
Z
dxT (k) (x)(T (`) (x))∗ = δ(k − `) (11.33)
R
5
A Hausdorff space is called locally compact if every point has a compact neighborhood. All the abelian
groups we will meet in this course are locally compact. **** GIVE AN EXAMPLE OF A non-locally-
compact abelian group ***
– 34 –
The statement that any L2 function on the group R can be expanded in the matrix
elements of its unitary irreps is that statement that any L2 function f (x) can be expanded:
Z
f (x) = dk fˆ(k)T (k) (x) (11.34)
R
b
X
(T (θ1 ) (n))∗ T (θ2 ) (n) = δperiodic (θ1 − θ2 )
n∈Z
X (11.39) eq:orthogrelsa
≡ δ(θ1 − θ2 − 2πm)
m∈Z
– 35 –
Remarks
• The functional analysis of L2 (G) is often called “nonabelian harmonic analysis.” For
more about this point of view see,
A.A. Kirillov, Elements of the Theory of Representations, Springer-Verlag, 1976.
Suppose f ∈ S(R) (the Schwarz space of functions of rapid decrese for | x |→ ∞).
Define the Fourier transform:
Z ∞
fˆ(k) = f (y)e−iky dy . (11.42)
−∞
Then:
Remarks
• If it possible to give other proofs of this formula. We will content ourselves with its
being a consequence of the orthogonality relations.
• The rapid decrease condition can be relaxed to
const
| f (x) | + | f 0 (x) | + | f 00 (x) | ≤ (11.45)
1 + x2
Remark The above result has a beautiful generalization to arbitrary lattices in Euclidean
space Rd . The generalized Poisson summation formula states that:
X X
e2πi~v·~x = δ (d) (~x − ~l) (11.46)
~v ∈Λ ~l∈Λ∗
or - in terms or functions
– 36 –
X Z X Z
~ ~t ~
X
−2πim·
f (~n) = e f (t)dt = e+2πim·
~ t
f (t)dt (11.47) eq:poissonsumm
Rd Rd
n∈Z d
~ m∈Z
~ d m∈Z
~ d
That is:
X X
f (~v ) = fˆ(~l) . (11.48)
~v ∈Λ ~l∈Λ∗
Remarks
• Since people have different conventions for the factors of 2π in Fourier transforms it is
hard to remember the factors of 2π in the PSF. The equation (11.44) has no factors of 2π.
One easy way to see this is to integrate both sides from t = −1/2 to t = +1/2.
• One application of this is the x-ray crystallography: The LHS is the sum of scattered
waves. The RHS constitutes the bright peaks measured on a photographic plate.
• Another application is in the theory of elliptic functions and θ-functions described later.
Exercise
a.) Show that
r
X
−πan2 +2πibn 1 X − π(m−b)2
e = e a (11.49) eq:formtwo
a
n∈Z m∈Z
b.) If τ is in the upper half complex plane, and θ, φ, z are complex numbers define the
theta function
2 +2πi(n+θ)(z+φ)
X
ϑθ,φ z = eiπτ (n+θ) (11.50) eq:thefun
n∈Z
θ −z −1 2 −φ
ϑ[ ]( | ) = (−iτ )1/2 e2πiθφ eiπz /τ ϑ[ ](z|τ ) (11.51) eq:esstmn
φ τ τ θ
Λ ⊂ R3 (11.52)
and an electron (in the single-electron approximation) satisfies the Schrödinger equation
~2 2
− ∇ ψ + V (~x)ψ = Eψ. (11.53)
2m
– 37 –
Now, V (~x) will be complicated in general, but we do know that
∀~a ∈ Λ. Therefore the energy eigenspace VE is some representation of the lattice translation
group Λ.
As we have seen, the one dimensional reps of Λ are
~
~a → e2πik·~a (11.55)
and are labeled by ~k ∈ Λ̂ = Rn /Λ∗ . Therefore, VE can be written as a sum of reps labeled
by ~k:
~
ψ(~x + ~a) = e2πik·~a ψ(~x) (11.57)
~
that is, ψ(~x) = e2πik·~x u(~x) where u is periodic. ~k, the “crystal momentum,” is only defined
modulo Λ∗ . It is properly an element of a torus.
Since we are working with infinite groups the sum on ~k in (11.56) in general must be
interpreted as an integral.
E*
1/a
k1 k2 k k k
3 4
Figure 3: Four irreps occur at energy E∗ in this one-dimensional band structure. For d = 2, 3
dimensional bandstructures there are infinity many k’s forming a surface. The surface at the Fermi
energy is the “Fermi surface.” fig:bandstruct
Remarks
– 38 –
~
• If we substitute ψ(~x) = e2πik·~x u(~x) into the Schrödinger equation we obtain an
elliptic eigenvalue problem on a compact space. It will therefore have a tower of discrete
eigenvalues. This leads to the band structure illustrated schematically in 3.
• In solid state physics the dual lattice is referred to as the reciprocal lattice. Physicists
generally choose a fundamental domain for the torus Rd /Λ∗ and refer to it as a Brillouin
zone.
• If the lattice Λ has a symmetry group then it will act on the torus. At fixed points
there is enhanced symmetry and bands will cross.
χ2 (s1 )
s (s1 , χ1 ), (s2 , χ2 ) = . (11.59) eq:cocyclechoi
χ1 (s2 )
This gives
There are two very natural representations of this group. First we consider H = L2 (S).
First of all L2 (S) is a representation of S. After all for s0 ∈ S we can define the
translation operator:
On the other hand, H is also a representation of the Pontrjagin dual group of characters,
denoted Ŝ. If χ ∈ Ŝ is a character on S then we define the multiplication operator Mχ on
H via
Note that H is not a representation of S×Ŝ. This simply follows from the easily-verified
relation
– 39 –
(z, (s, χ)) → zTs Mχ (11.64) eq:heishom
where z ∈ U (1). Equation (11.64) defines a homomorphism into the group of invertible
operators on L2 (S).
Notice the complete symmetry in the construction between S and Ŝ (since the double-
dual gives S again). Thus, we could also provide a representation from Ĥ = L2 (Ŝ). The
two representations are in fact equivalent under Fourier transform:
Z
ψ̂(χ) := dsχ(s)∗ ψ(s) (11.65) eq:svnthmi
In fact, a theorem, the Stone-von Neumann theorem guarantees the essential unique-
ness of the unitary representations of Heisenberg groups:
while the dual group χt̄ has character χt̄ (s̄) = e2πist/n acts by multiplication.
b∼
Example 2 S = R and Ŝ = R. Denote elements q ∈ R and p ∈ R = R.
Example 3 S = Z and Ŝ = U (1) ... This generalizes to the duality between lattices and
tori.
Remarks
• Most free field theories can be interpreted in this framework, with proper regard
for the infinite-dimensional aspects of the problem. There is the interesting exception of
self-dual field theories which are more subtle.
– 40 –
12. Induced Representations
♣MATERIAL
HERE IMPORTED
Induced representations are certain representations defined by taking sections of associated FROM GMP 2010
ASSUMES
bundles. They form a unifying theme for a broad class of representations in physics: KNOWLEDGE OF
BUNDLES. NEED
1. Finite groups TO EDIT THOSE
PARTS AWAY. ♣
2. Compact groups: The Borel-Weil-Bott Theorem
3. Lorentz groups: The Wigner construction.
4. Representations of Heisenberg groups and coherent states
5. Representations of Loop Groups
Suppose H ⊂ G is a subgroup.
Suppose ρ : H → GL(V ) is a representation of H with representation space V .
Note that
H → G
↓π (12.1)
G/H
is a principal H bundle with total space P = G.
Therefore, according to what we have said above, using (ρ, V ) we can form the associ-
ated vector bundle
G ×H V (12.2)
Now, the key point is that Γ(G ×H V ) is also a representation of G. This representation
is called the induced representation and sometimes denoted:
Γ(G ×H V ) = IndG
H (V ) (12.3)
Using the grand tautology (??) we can equivalently define the induced representation
as:
Definition Let (V, ρ) be a representation of H. As a vector space, IndG
H (V ) is the set
of all smooth functions ψ : G → V satisfying the equivariance condition:
for all h ∈ H, g ∈ G.
Note that, because G is a group we can multiply both on the left, and on the right.
This allows us to define a representation of G in a way that does not spoil the equivariance
condition:
Let us check that this is indeed a representation. If ψ ∈ IndG
H (ρ), and g ∈ G we define
T (g) · ψ as the new function whose values are given in terms of the old one by
– 41 –
However, the action (12.5) manifestly does not interfere with the equivariance property
of ψ! In detail: if h ∈ H then
(T (g) · ψ)(g 0 h) = ψ(g −1 g 0 h) = ρ(h−1 )ψ(g −1 g 0 ) = ρ(h−1 )((T (g) · ψ)(g 0 )) (12.7) eq:explicit
(We denoted this by IndGH before but H, G will be fixed in what follows so we simplify the
notation.) On the other hand, there is an obvious functor going the other way, since any
G-rep W is a foriori an H-rep, by restriction. Let us denote this “restriction functor”
How are these two maps related? The answer is that they are “adjoints” of each other!
This is the statement of Frobenius reciprocity:
We can restate the result in another way which is illuminating because it helps to
answer the question: How is IndG
H (V ) decomposed in terms of irreducible representations
of G? Let Wα denote the distinct irreps of G. Then Schur’s lemma tells us that
IndG ∼ G
H (V ) = ⊕α Wα ⊗ HomG (Wα , IndH (V )) (12.11) eq:schurlemma
– 42 –
IndG ∼
H (V ) = ⊕α Wα ⊗ HomH (R(Wα ), V ) (12.12) eq:frobrecipii
where the sum runs over the unitary irreps Wα of G, with multiplicity one.
The statement (12.12) can be a very useful simplification of (12.11) if H is “much
smaller” than G. For example, G could be nonabelian, while H is abelian. But the
representation theory for abelian groups is much easier! Similarly, G could be noncompact,
while H is compact. etc.
In order to prove (12.12) we note that it is equivalent (see the exercise below) to the
statement that the character of IndG
H (V ) is given by
X
χ(g) = χ̂(x−1 gx) (12.13) eq:charind
x∈G/H
where x runs over a set of representatives and χ̂ is the character χV for H when the
argument is in H and zero otherwise.
On the other hand, (12.13) can be understood in a very geometrical way. Think of the
homogeneous vector bundle G ×H V as a collection of points gj H, j = 1, . . . , n with a copy
of V sitting over each point. Now, choose a representative gj ∈ G for each coset. Having
chosen representatives gj for the distinct cosets, we may write:
Geometrically, this is a section whose support is located at the point gi H. The equivariant
function is then given by
ψi,a (gj h) := ρ(h−1 )va δi,j (12.16)
Now let us compute the action of g ∈ G in this basis:
Fortunately, we are only interested in the trace of this G-action. The first key point is
that only the fixed points of the g-action on G/H contribute. Note that the RHS above
is supported at j = g · i, but if we are taking the trace we must have i = j. But in
– 43 –
(23)
g H
1
(123) (123)
(13) (12)
g H (23) g H
3 2
(123)
(12) (13)
Figure 4: The left action of G = S3 on G/H. In fact, this picture should be considered as a
picture of a category, in this case, a groupoid. fig:indrepi
this case ggi = gi h(g, i) and hence g −1 gi = gi h(g, i)−1 so for fixed points we can simplify
h(g −1 , i) = h(g, i)−1 , and hence when we take the trace the contribution of a fixed point
ggi H = gi H is the trace in the H-rep of h(g, i) = gi−1 ggi , as was to be shown ♠
Remark: The Ind map does not extend to a ring homomorphism of representation
rings.
Example A simple example from finite group theory nicely illustrates the general idea.
Let G = S3 be the permutation group. Let H = {1, (12)} ∼ = Z2 be a Z2 subgroup. G/H
consists of 3 points. The left action of G on this space is illustrated in (12.5).
There are two irreducible representations of H, the trivial and the sign representation.
These are both 1-dimensional. Call them V (), with = ±. Accordingly, we are looking
at a line bundle over G/H and the vector space of sections of G ×H V () is 3-dimensional.
A natural basis for the space of sections is given by the functions which are “δ-functions
supported at each of the three points”:
si (gj H) = δij
g1 H = (13)H = {(13), (123)}
(12.18) eq:basis
g2 H = (23)H = {(23), (132)}
g3 H = (12)H = {1, (12)}
These sections correspond to equivariant functions on the total space. The space of all
functions F : G → R is a six-dimensional vector space. The equivariance condition:
– 44 –
F (12) = F (1)
F (123) = F (13) (12.19) eq:equicond
F (132) = F (23)
(12)g1 = g2 (12)
(12)g2 = g1 (12)
(12)g3 = g3 (12)
(13)g1 = g3 (12)
(13)g2 = g2 (12) (12.20) eq:multout
(13)g3 = g1 (12)
(23)g1 = g1 (12)
(23)g2 = g3 (12)
(23)g3 = g2 (12)
!
−1 0
ρ2 (12) =
0 1
√ ! (12.23) eq:rhotwo
−√12 23
ρ2 (123) =
− 23 − 12
– 45 –
As H = Z2 representations, we have W () ∼
= V () and
W2 ∼
= V (+1) ⊕ V (−1) (12.24)
Therefore
To find out how to decompose the representation IndSS32 (V ()) in terms of G = S3 irreps
it suffices to compute the character for g = (12) and g = (123). Now, g = (12) has exactly
one fixed point, namely g3 H and h(g, 3) = (12) for this element. Therefore,
On the other hand, g = (123) clearly has no fixed points, and therefore the character is
zero. It follows immediately that we have the decomposition (12.26).
*********************************************
*********************************************
INCORPORATE:
Now let us turn to induced representations:
Let G be a group and H a subgroup. Suppose that ρ : H → End(V ) is a representation
of the subgroup H. Then, as we have seen Map(G, V ) is canonically a G × H-space. To
keep the notation under control we denote a general function in Map(G, V ) by Ψ. Then
the left-action of G × H defined by declaring that for (g, h) ∈ G × H and Ψ ∈ Map(G, V )
the new function φ((g, h), Ψ) ∈ Map(G, V ) is the function G → V defined by:
for all g0 ∈ G. Now, we can consider the subspace of functions fixed by the action of 1 × H.
That is, we consider the H-equivariant functions which satisfy
– 46 –
for every g ∈ G and h ∈ H. Put differently: There are two natural left-actions on
Map(G, V ) and we consider the subspace where they are equal. Note that the space of such
functions is a linear subspace of Map(G, V ). We will denote it by IndG H (V ). Moreover, it
is still a representation of G since if Ψ is equivariant so is (g, 1) · Ψ.
The subspace IndG H (V ) ⊂ Map(G, V ) of H-equivariant functions, i.e. functions satis-
fying (12.30) is called the induced representation of G, induced by the representation V of
the subgroup H. This is an important construction with a beautiful underlying geometrical
interpretation. In physics it yields:
Example: Let us take V = C with the trivial representation of H, i.e. ρ(h) = 1. Then
the induced representation is the vector space of functions on G which are invariant under
right-multiplication by H. This is precisely the vector space of C-valued functions on the
j
homogeneous space G/H. For example, the invariant Wigner functions Dm L ,mR under
j
right-action by the diagonal U (1) subgroup of SU (2) are DmL ,0 (g). These descend to
functions on SU (2)/U (1) ∼= S 2 known (for j integral) as the spherical harmonics. The case
of V a trivial representation generalizes in a beautiful way: When (ρ, V ) is nontrivial the
induced representation is interpreted not as a space of functions on G/H but rather as a
vector space of sections of a homogeneous vector bundle over G/H determined by the data
(ρ, V ). See **** below.
Exercise
Prove (6.20).
– 47 –
Take H to be the trivial group and V the trivial representation and explain the relation
of (12.12) to the Peter-Weyl theorem.
**************************************************
**************************************************
Let us now consider the space of C∞ sections Γ(Lk ) as a representation of SU (2). The GMP 2010 BUT
ALL MATERIAL
ASSUMING
right-action by U (1) is given by INDUCED REPRE-
3
g · eiχ := ge−iχσ (13.1) SENTATIONS,
BUNDLE
THEORY, AND
where on the RHS we have ordinary matrix multiplication. According to our general BOREL-WEIL-
BOTT IS
principle, the sections of Vk are the same as equivariant functions ψ : SU (2) → C such that SUPPRESSED ♣
3
ψ(ge−iχσ ) = e−ikχ ψ(g) (13.2) eq:eqvfunct
In order to analyze (13.2) we must recall the Peter-Weyl theorem for SU (2):
L2 (SU (2)) ∼
= ⊕j∈ 1 Z+ Dj ⊗ Dj (13.3)
2
j
Here Pm L ,mR (x) is an associated Legendre polynomial and we have chosen the slightly
– 48 –
SU (2)
IndU (1) (Vk ) ∼
= ⊕j 0 ≥ 1 |k| Dj
0
(13.6) eq:indrepsut
2
The RHS is one-dimensional exactly for those values of j 0 given by 2j 0 − |k| ∈ 2Z+ .
• In the special case k = 0 we are talking about sections of the trivial line bundle.
But these are just complex valued L2 functions on S 2 . Indeed, in this case mR = 0
and the functions become independent of ψ and hence well-defined functions on S 2 rather
than sections of nontrivial line bundles. In this case the functions are known as spherical
harmonics and usually denoted:
`
Dm,0 = Y`,m (θ, φ) (13.9)
!
u
(g · ψ` )(u, v) = ψ` (g −1 )
v
(13.11) eq:mtrxelem
= ψ` ((detg)−1 (g22 u − g12 v, −g21 u + g11 v))
= (detg)−k (g22 u − g12 v)` (−g21 u + g11 v)k−`
By expanding out using the binomial theorem and collecting terms one obtains explicit
matrix elements. We will do this momentarily.
– 49 –
What we have learned so far is that the holomorphically induced representation from
B to SL(2, C) for Vk is the k + 1-dimensional representation which can canonically be
viewed as the space of homogeneous polynomials of degree k. [*** Redundant??? ***]
GL(2,C)
Now, since SU (2) ⊂ GL(2, C) we can restrict the functions in HolIndB (Vk ) to
∼
SU (2). Since SU (2) ∩ B = T = U (1) they will be functions in the induced representation
SU (2)
IndU (1) (Vk ). The monomial basis functions become
we continue the computation of the matrix elements (13.11) for this representation of
SU (2):
functions on SU (2) by expanding out the two factors in (13.15) using the binomial theorem
and collecting terms:
j
X j + m j − m s j−m−t j+m−s
D̃m 0 m (g) = ᾱ α β̄ (−β)t (13.17) eq:xplcttldee
0
s t
s+t=j+m
Warning: This differs from standard Wigner functions by normalization factors. These
are important when dealing with unitary representations. We return to unitarity below.
j j
For the moment we merely remark that D̃m L m is proportional to DmL m .
– 50 –
SU (2)
Now, we found earlier that IndU (1) (Vk ) is spanned by matrix elements with mR =
−k/2. In equation (13.17) this means we should put m = −j which forces s = 0 and hence
j j+m 2j
D̃m,−j = (−1) αj−m β j+m (13.18)
j+m
Up to a normalization constatn this is the explicit basis (13.12) we set out to identify.
Remarks
• If we consider a charged particle confined to a sphere and coupled to a Dirac monopole
of charge k we will see that quantization of the system leads to a Hilbert space
SU (2)
Γ(Lk ) = IndU (1) (Vk ). (13.19)
Since SU (2) is a global symmetry of the quantum system the Hilbert space must be a
representation of SU (2). Above we have given the decomposition of this Hilbert space in
terms of irreps.
• SU (2)/U (1) is also a Kahler manifold and therefore we will be able to introduce the
supersymmetric quantum mechanics with this target space with N = 2 supersymmetry.
If the particle is coupled to a gauge field on a line bundle of first chern class k (i.e. if
the particle is charged, with charge k) then the above representation has the physical
interpretation as the space of BPS states.
************************
***********************
for any X ∈ su(2). Taking X = X3 = 2i σ3 and evaluating the definition of the function on
the LHS evaluated on (u, v) we get:
1 1 i ∂ ∂
f (u − i u, v − i v) = f − u −v f
2 2 2 ∂u ∂v
so we conclude:
1 ∂ ∂
T (X3 ) · f = −i u −v f
2 ∂u ∂v
and, in particular,
iT (X 3 ) · f˜j,m = mf˜j,m
Similarly we have
i 1
f + T (X 1 ) · f = T (e 2 σ ) · f
i
∂ ∂
(13.20)
=f − v +u f
2 ∂u ∂v
– 51 –
i 1
since for g = e 2 σ we have α = 1 + O(2 ), β = − 2i . In just the same way we get:
i 2
f + T (X 2 ) · f = T (e 2 σ ) · f
1
∂ ∂
(13.21)
=f − v −u f
2 ∂u ∂v
These act as
∂
J + · f˜j,m = u f˜j,m
∂v
= (j − m)f˜j,m+1
(13.22)
∂
J − · f˜j,m = v f˜j,m
∂u
= (j + m)f˜j,m−1
Or, as matrices:
0 1 0 0 ···
0
0 2 0 ···
0 0 0 3 ···
J+ = (13.23)
eq:sutwomtrx
· · · ··· ··· ··· ···
0 0 0 ··· 2j
0 0 0 ··· 0
0 0 0 0 ···
2j
0 0 0 ···
0 2j − 1 0 0 ···
J− = (13.24)
· · · · · · ··· ··· ···
0 0 0 ··· 0
0 0 ··· 1 0
Exercise
Check that this is a representation of su(2). It is different from what one finds in most
textbooks on quantum mechanics because it is not a unitary matrix representation. See
the next section.
*************************
*************************
– 52 –
13.1 Unitary structure
In Chapter 7 we observed that we can define an Hermitian metric on the holomorphic line
bundle H k such that the norm-squared of a section is
1
k s k2 = |s(z)|2 (13.25)
(1 + |z|2 )k
where
i dzdz̄
vol (CP 1 ) = (13.27)
2π (1 + |z|2 )2
is the unit volume form. We claim that the action of SU (2) on H 0 (CP 1 ; H k ) we described
above is unitary with respect to this metric.
To see this let us first note that we can identify the space Hk of homogeneous polyno-
mials in two variables of degree k with the space Pk of polynomials in a single variable of
degree ≤ k as follows:
If ψ(u, v) ∈ Hk then we map it to
1 |β̄z + ᾱ|2
= (13.31)
(1 + |z 0 |2 ) (1 + |z|2 )
shows that
hg · s, g · si = hs, si (13.32)
In this way H 0 (CP 1 ; H k ) becomes a Hilbert space on which SU (2) acts unitarily.
The ON basis corresponding to f˜m is
s
(2j + 1)!
ψj,m = z j−m (13.33)
(j + m)!(j − m)!
– 53 –
j
and can be taken to be |j, mi in our definition above of the Wigner functions Dm L ,mR .
Exercise
j j
Find the proportionality factor between D̃m L ,mR and DmL ,mR .
************************
***********************
UNITARIZE REP OF J + AND J −
*************************
*************************
– 54 –
with s
(2j)! ζ j+m
um (ζ) = (13.38)
(j + m)!(j − m)! (1 + |ζ|2 )j
******* IS SQRT CORRECT? ********
2. The inner product of two such polynomials is
¯ 2j
(1 + ξη)
hξ|ηi = (13.39)
((1 + |ξ|2 )j (1 + |η|2 )j
(1 + ζz)2j
f (z) = (13.40)
(1 + |ζ|2 )j
Examples:
1. The wavefunction of the oxygen molecule!
2. Wavefunctions of electrons in the presence of a magnetic monopole will take values
in this bundle.
3. Nice application of this coherent state formalism to the QHE in paper by Haldane.
Remarks:
1. The coherent states are good for taking the semiclassical limit of large j. Then the
support of |ζi is sharply peaked at a point on CP 1 . **** SHOW *****
2.
3. Relate infinite dimensional induced rep to the Schwinger representation in terms of
oscillators. *** What is this??? ***
Note the A−1 in the argument of Ψ in the second equality. This is necessary to get a
left-action of the group on the space of fields. If we use Ax then we get a right-action. ♣For the case of
complex scalar fields
Now, quite generally, if V is a representation space for G, and O ∈ End(V ) is an when A is in a
component of
invariant linear operator, i.e. an operator which commutes with the action of G, O(1, 1) that reverses
the orientation of
time we have the
ρ(g)O = Oρ(g) (14.2) choice of whether to
include complex
conjugation in the
action. This would
be adding a “charge
conjugation” action.
♣
– 55 –
then any eigenspace, say {v ∈ V |Ov = λv} will be a sub-representation of G.
Consider the operator
∂ µ ∂µ = −∂02 + ∂12 (14.3)
acting on the space of scalar fields. This is an example of an invariant operator, as one
confirms with a simple computation. It can be made manifest by writing
∂ µ ∂µ = −4∂+ ∂− (14.4)
where
∂ 1
∂± := ±
= (∂0 ∓ ∂1 ) (14.5)
∂x 2
The Klein-Gordon equation for a complex or real scalar field Ψ(x0 , x1 ) is
∂ µ ∂µ + m2 Ψ = 0
(14.6)
The space of fields satisfying the KG equation is a representation space of O(1, 1), by our
general remark above.
Now we relate the orbits of O(1, 1) to the representations furnished by solutions of the
KG equation:
For field configurations which are Fourier transformable we can write
Z
0 1
0 1
Ψ(x , x ) = dk0 dk1 ψ̂(k)eik0 x +ik1 x (14.7)
that is:
(k 2 + m2 )ψ̂(k) = 0 (14.9)
Thus, the support in Fourier space is an orbit. In physics the orbit with k0 > 0 is
called the mass-shell.
For example, suppose that m2 > 0. Then on the orbit with k 0 > 0 we have k =
m(cosh θ, sinh θ) and we can write the general complex-valued solution as
where the amplitude ā should be regarded as a complex-valued function on the orbit. Then
a complex-valued solution on the KG equation is generated by such a function on the orbit
by:
Z
Ψ(x) = d2 kδ(k 2 + m2 )āeik·x
Z
dk
= p 1 ā(k1 )eik·x (14.11) eq:MomentumMod
2 2
R 2 k1 + m
Z ∞
1 0 1
= dθā(θ)eim(x cosh θ+x sinh θ)
2 −∞
– 56 –
where we chose a convention designation of the argument of the function ā on the orbit.
Thus, we can identify a space of solutions to the KG equation with a space of functions
on an orbit.
Now recall that the space of functions on an orbit is a special case of an induced
representation. As we will soon see, induced representations are the right concept for
generalization to other representations of the Lorentz group.
Remark 1: The components of the Lorentz group: If we use a particular orbit to generate
representations we only represent the subgroup of the Lorentz group which preserves that
orbit. Thus, for the orbit with m > 0 and k0 > 0 we can include the parity-reversing com-
ponent, but not the time-reversing component. Of course, there is a similar representation
of SO0 (1, 1) q P · SO0 (1, 1) given by the hyperbola with k0 < 0. If we wish to represent the
full O(1, 1) then we must include both orbits. Note that, if we wish to use real solutions of
the KG equation we must include both hyperbolae and there is a reality condition relating
the Fourier modes: (ā(k))∗ = a(−k)
Remark 2: The representation can be made into a unitary representation. See below.
Remark: In the massless case there is a “more primitive” relativistic wave equation
which is first order, and whose solutions are always solutions of the massless KG equa-
tion. Namely, we can consider the separate equations
∂+ Ψ = 0 (14.13) eq:anti-self-d
∂− Ψ = 0 (14.14) eq:self-dual
Note that these equations are themselves Lorentz invariant, even though the operators
∂± are not invariant differential operators. Solutions to (14.13) are right-moving scalar
– 57 –
fields and solutions to (14.14) are left-moving scalar fields. Such on-shell scalar fields
are also known as chiral scalar fields. Equations (14.13) and (14.14) are notable in part
because they play an important role in string theory and conformal field theory. It is
also interesting to note that it is quite subtle to write an action principle that leads to
such equations of motion. They are also called the anti-self-dual and self-dual equations
of motion, respectively because, if we choose the orientation dx ∧ dt then the Hodge star
operation is ∗dt = dx and ∗dx = dt and hence ∗(dx± ) = ±dx± . Therefore if ∂± Ψ = 0 its
“fieldstrength” F = dΨ satisfies ∗F = ∓F .
The action of P, T ∈ O(1, 1) on a real scalar field is given by:
(P · Ψ)(x, t) := Ψ(−x, t)
(14.15)
(T · Ψ)(x, t) := Ψ∗ (x, −t)
The KG equation is separately P and T invariant, but the (anti)-self-dual equations are
not. Nevertheless, the latter equations are P T invariant. This is a special case of the
famous CPT theorem:
Roughly speaking, mathematical consistency implies that if a physical theory is in-
variant under group transformations in the neighborhood of the identity then it must be
invariant under the transformations in the full connected component of the identity. But
this does not mean the theory is invariant under disconnected components such as the com-
ponents containing P and T . As a matter of fact, Nature chooses exactly that option in
the standard model of the electroweak and strong interactions. However, if we also assume
that there is a certain relation to “analytically continued” equations in Euclidean signature
then, since P T is in the connected component of the identity of O(2), such theories must
in fact be P T invariant. ♣There is much
more to discuss
about the nature of
14.3 The case of d dimensions, d > 2 the analytic
continuation
implied here. In
Now consider Minkowski space M1,d−1 with d > 2. The nature of the orbits is slightly what sense is SO(2)
and “analytic
different. continuation” of
SO(1, 1)? It does
not do simply to
1. For λ2 > 0 we can define make the boost pure
imaginary unless
x± → z, z̄. ♣
O+ (λ) = {x|(x0 )2 − (~x)2 = λ2 & sign(x0 ) = sign(λ)} (14.16)
– 58 –
Figure 5: Illustrating orbits of the connected component of the identity in O(1, 3). In (a) the top
and bottom hyperboloids are separate orbits, and if we include time-reversing transformations the
orbits are unions of the two hyperboloids. In (b) there are three orbits shown with x0 > 0 x0 < 0
(the future and past, or forward and backward light cones), and the orbit consisting of the single
point. In (c), once x2 has been specified, there is just one orbit, for d > 2. fig:LorentzOrb
3.
O± = {x|x2 = 0 & sign(x0 ) = ±1} (14.20)
Vectors in this orbit are of the form (x0 , |x0 |n̂) where n̂ ∈ S d−2 ⊂ Rd−1 and the sign
of x0 is invariant under the action of the identity component of O(1, 3). (Show this!).
Note that, for d = 2 the sphere S 0 has two disconnected components, leading to
left- and right-movers. But for d > 2 there is only one component. We can think of
n̂ ∈ S d−2 as parametrizing the directions of light-rays. That is, the point where the
light ray hits the celestial sphere. In one spatial dimension, a light ray either moves
left or right, and this is a Lorentz-invariant concept. In d − 1 > 1 spatial dimensions,
we can rotate any direction of light ray into any other. See Figure 5(b). One can
show that these orbits too are homogeneous spaces: 6
O± ∼
= SO0 (1, d − 1)/I (14.21)
6
The isotropy group of a light ray is I ∼ = ISO(d − 2), where ISO(d − 2) is the Euclidean group
on Rd−2 . The easiest way to show this is to use the Lie algebra of so(1, d − 1) and work with light-cone
coordinates. Choosing a direction of the light ray along the xd−1 axis and introducing light-cone coordinates
x± := x0 ± xd−1 , and transverse coordinates xi , i = 1, . . . , d − 2 if the lightray satisfies x− = 0 then we
have unbroken generators M +i and M ij .
– 59 –
4. The final orbit is of course {x = 0}.
However, for scalar fields, there is no analog of the left- and right-chiral boson. There
are analogs involving interesting first order equations such as the Dirac equation and the
(anti-) self-dual equations for fields with spin.
Quite generally, we can define an inner product on the space of complex-valued solu-
tions of the KG equation such that the action of the Lorentz group is unitary. Observe
that, given any two complex-valued solutions Ψ1 , Ψ2 the current
∂ µ jµ = 0 (14.24)
is conserved. Therefore, if we choose a spatial slice with normal vector nµ and induced
volume form vol the inner product
Z
(Ψ1 , Ψ2 ) := nµ jµ vol (14.25)
Σ
is independent of the choice. So, fixing a Lorentz frame and taking Σ to be the slace at a
fixed time we have
Z
(Ψ1 , Ψ2 ) := −i (Ψ∗1 ∂0 Ψ2 − (∂0 Ψ1 )∗ Ψ2 ) dd−1 ~x (14.26)
Rd−1
This is clearly not positive definite on the space of all solutions but does become positive
definite when restricted to the space of complex solutions associated with a single orbit.
Indeed, substituting the expansion in momentum space (14.22) we get
dd−1~k
Z
(Ψ1 , Ψ2 ) = p (ā1 (~k))∗ (ā2 (~k)) (14.27)
Rd−1 2 ~k 2 + m2
Having phrased things this way, it is clear that there is an interesting generalization:
We can choose other representations of H = SO(d−1) and consider the induced representa-
tions of the Lorentz group. This indeed leads to the unitary representations, corresponding
to particles with nontrivial spin. ♣Incorporate some
of the following
remarks: ♣
– 60 –
15. Characters and the Decomposition of a Representation to its Irreps
Given a group G there are some fundamental problems one wants to solve:
I am not aware of any systematic procedure for deriving the list of irreps of any group.
However, if - somehow - the list is known then the answers to 2 and 3 are readily provided
through the use of characters.
Exercise
Show that for a unitary representation
A second motivation for looking at characters is the following: matrix elements are
complicated. Sometimes one can extract all the information one is seeking by considering
this simpler and more basic set of functions on G associated to a representation.
Recall that for any element of g ∈ G the set
Definition A function on G that only depends on the conjugacy class is called a class
function. In particular, characters are class functions. The space of class functions is a
subspace of RG .
– 61 –
Theorem. The χµ form a basis for the vector space of class functions in RG .
Proof : This is a corollary of the previous section: Any function can be expanded in
X µ µ
F (g) = F̂ij Tij (g) (15.5)
Therefore, it suffices to compute the average of Tijµ (g) over a conjugacy class:
1 X (µ)
(Tijµ )average (g) ≡ Tij (hgh−1 ) (15.6)
|G|
h∈G
Expanding this out and doing the sum on h using the ON relations shows that
F average = F (15.8)
so
F̂ijµ nµ χµ (g)
X
F (g) = (15.9)
ij
Theorem The number of conjugacy classes of G is the same as the number of irreducible
representations of G.
Since the number of representations and conjugacy classes are the same we can define
a character table:
m 1 C1 m 2 C2 ··· ··· mr Cr
χ1 χ1 (C1 ) ··· ··· ··· χ1 (Cr )
χ2 χ2 (C1 ) ··· ··· ··· χ2 (Cr )
(15.10)
··· ··· ··· ··· ··· ···
··· ··· ··· ··· ··· ···
χr χr (C1 ) ··· ··· ··· χr (Cr )
Here mi denotes the order of Ci .
As we will see, character tables are a very useful summary of important information
about the representation theory of a group.
Exercise
– 62 –
In a previous exercise we computed the average number of fixed points of a finite group
G acting on a finite set X.
Rederive the same result by interpreting the number of elements of X fixed by g as a
character χ(g) of the representation given by the functions on X.
1 X
χν (g −1 )χµ (g) ≡ (χν , χµ ) = δµν (15.11)
|G|
g∈G
T ∼
= a1 T (1) ⊕ a2 T (2) ⊕ · · · ⊕ ar T (r) = ⊕µ aµ T (µ) (15.12) eq:reducible
The ai are nonnegative integers, they measure the number of times a rep appears and
are referred to as the “multiplicities.”
From the orthogonality relation we get:
Theorem
Two representations are equivalent iff they have the same character.
If you have two reps and you want to see whether they are inequivalent then an efficient
test is to compute the characters.
The orthogonality relations on characters can be stated more beautifully if we use the
fact that characters are class functions:
1 X
mi χµ (Ci )χν (Ci )∗ = δµν (15.14) eq:orthogrel
|G|
Ci ∈C
X |G|
χµ (Ci )∗ χµ (Cj ) = δij (15.15) eq:orthogrelii
µ
mi
– 63 –
Proof : Let us first prove (15.14). From unitarity we have χµ (g −1 ) = (χµ (g))∗ . Since
χµ are class functions the sum over the elements in the group can be written as a sum over
the conjugacy classes Ci ∈ C.
The equation (15.14) can be interpreted as the statement that the r × r matrix
r
mi
Sµi := χµ (Ci ) µ = 1, . . . , r i = 1, . . . , r (15.16)
|G|
satisfies
r
X
∗
Sµi Sνi = δµν (15.17)
i=1
Therefore, Sµi is a unitary matrix. The left-inverse is the same as the right-inverse, and
hence we obtain (15.15). ♠
Exercise
Show that right-multiplication of the character table by a diagonal matrix produces a
unitary matrix.
Exercise
Using the orthogonality relations on matrix elements, derive the more general relation
on characters:
1 X δµν (ν)
χµ (g)χν (g −1 h) = χ (h) (15.18) eq:convolchar
|G| nµ
g∈G
Exercise
A more direct proof of (15.15) goes as follows. Consider the operator L(g1 ) ⊗ R(g2 )
acting on the regular representation. We will compute
in two bases.
First consider the basis φµij of matrix elements. We have:
– 64 –
so the trace in this basis is just:
X
TrRG [L(g1 ) ⊗ R(g2 )] = χµ (g1 )∗ χµ (g2 ) (15.21)
µ
On the other hand, we can use the delta-function basis: δg . Note that
So in the delta function basis we get a contribution of +1 to the trace iff g = g1 gg2−1 , that
is iff g2 = g −1 g1 g, that is iff g1 and g2 are conjugate, otherwise we get zero.
Now show that
16.1 S3
The group S3 has three obvious irreducible representations:
1+ . 1 → 1 (ij) → 1 (123) → 1, etc.
1− . 1 → 1 (ij) → −1 (123) → +1, etc. This is the sign representation, or the
homomorphism to Z2 mentioned in lecture 1.
2. Symmetries of the triangle.
!
10
1→
01
!
−1 0
(23) → (16.1)
0 1
! √ !
cos(2π/3) sin(2π/3) −1/2 3/2
(123) → = √
− sin(2π/3) cos(2π/3) − 3/2 −1/2
These reps are irreducible. Note that by applying group elements in the third repre-
sentation to any vector we obtain a set of vectors which spans R2 .
Indeed we can check:
6 = 12 + 12 + 22 (16.2)
to conclude that there are no other irreps.
It is now straightforward to write out the character table for S3 .
– 65 –
1
2 3
Figure 6: The symmetries of the equilateral triangle give the 2-dimensional irrep of S3 . fig:triangle
Now let us see how we can use the orthogonality relations on characters to find the decom-
position of a reducible representation.
Example 1 Consider the 3 × 3 rep generated by the \ action of the permutation group S3
on R3 . We’ll compute the characters by choosing one representative from each conjugacy
class:
100 010 010
1 → 0 1 0 (12) → 1 0 0 (132) → 0 0 1 (16.4)
001 001 100
1 3 2
a1+ = (χ1+ , χ) = 3 + 1 + 0 = 1 (16.6)
6 6 6
1 3 2
a1− = (χ1− , χ) = 3 + (−1) · 1 + 0 = 0 (16.7)
6 6 6
1 3 2
a2 = (χ2 , χ) = 3 · 2 + 0 · 1 + (−1) · 0 = 1 (16.8)
6 6 6
Therefore:
χV = χ1+ + χ2 (16.9)
– 66 –
Example 2 Consider S3 acting by permuting the various factors in the tensor space
V ⊗ V ⊗ V for any vector space V . Now, if dimV = d then we have
χ([1]) = d3
χ([(ab)]) = d2 (16.10) eq:vcubed
χ[(abc)] = d
So we can compute
1 3 2 1
a1+ = (χ1+ , χ) = d3 + d2 + d = d(d + 1)(d + 2) (16.11)
6 6 6 6
1 3 2 1
a1− = (χ1− , χ) = d3 + (−1) · d2 + d = d(d − 1)(d − 2) (16.12)
6 6 6 6
1 3 3 2 1
a2 = (χ2 , χ) = 2d + 0 · d + (−1) · d = d(d2 − 1)
2
(16.13)
6 6 6 3
Thus, as a representation of S3 , we have
d(d + 1)(d + 2) d(d − 1)(d − 2) d(d + 1)(d − 1)
V ⊗3 ∼
= 1+ ⊕ 1− ⊕ 2 (16.14)
6 6 3
Note that the first two dimensions are those of S 3 V and Λ3 V , respectively, and that the
dimensions add up correctly.
We are going to develop this idea in much more detail when we discuss Schur-Weyl
duality in section ****
Remarks
• Notice that this does not tell us how to block-diagonalize the matrices. We will see
how to do that later.
Exercise
Repeat example 2 to decompose V ⊗2 as a representation of S2 .
Exercise
Write out the unitary matrix Sµi for G = S3 .
Exercise
a.) Suppose we tried to define a representation of S3 by taking (12) → 1 and (23) → −1.
What goes wrong?
b.) Show that for any n there are only two one-dimensional representations of Sn
– 67 –
16.2 Dihedral groups
The irreps of the dihedral group Dn are easily written down.
Recall that Dn is the group generated by x, y subject to the relations:
x2 = 1
yn = 1 (16.15) eq:rels
xyx−1 = y −1
We can define two-dimensional complex representations T (j) by:
!
ωj
T (j) : y →
ω −j
! (16.16) eq:twodee
01
x→
10
– 68 –
where j = 1, . . . , k.
For n = 2k even we have
MISSING TABLE
where j = 1, . . . , k − 1
**** SHOULD apply these tables to solve some problem ***
It contains the representation µ with the correct degeneracy in V . It is called the isotypical
component belonging to µ. Note that it can be written as
Now suppose:
V1 = ⊕aµ T µ V2 = ⊕bν T ν (17.5)
then
V1 ⊗ V2 = ⊕µ,ν aµ bν T µ ⊗ T ν (17.6)
– 69 –
µν N λ times
z }| {
T (µ) ⊗ T (ν) (λ)
= ⊕λ T ⊕ · · · ⊕ T (λ)
(17.7) eq:decomp
λ
= ⊕λ Nµν T (λ)
Exercise
Show that
χT1 ⊗T2 (g) = χT1 (g)χT2 (g) (17.8)
Now let us put this observation to use and solve our problem. Taking the trace of
(17.7) we get:
X
λ
χµ (g)χν (g) = Nµν χλ (g) (17.9) eq:chardec
λ
λ
Nµν = (χλ , χµ χν ) (17.10) eq:fusion
1 X
λ
Nµν = χµ (g)χν (g)χλ (g −1 ) (17.11) eq:fusioni
|G|
g∈G
This can be written in a different way, which we will explain conceptually in the next
section. For simplicity choose unitary irreps, and recall that the orthogonality relations on
characters is equivalent to the statement that
r
mi
Sµi := χµ (Ci ) µ = 1, . . . , r i = 1, . . . , r (17.12)
|G|
– 70 –
18. Algebras
♣First part with
definitions was
We first need the general abstract notion: Definition An algebra over a field k is a vector copied into chapter
2. No need to
repeat here. ♣
space V over k with a notion of multiplication of two vectors
V ×V →V (18.1)
denoted:
v1 , v2 ∈ V → v1 · v2 ∈ V (18.2)
which has a ring structure compatible with the scalar multiplication by the field. Con-
cretely, this means we have axioms:
i.) (v1 + v2 ) · v3 = v1 · v3 + v2 · v3
ii.) v1 · (v2 + v3 ) = v1 · v2 + v1 · v3
iii.) α(v1 · v2 ) = (αv1 ) · v2 = v1 · (αv2 ), ∀α ∈ k.
The algebra is unital, i.e., it has a unit, if ∃1V ∈ V (not to be confused with the
multiplicative unit 1 ∈ k of the ground field) such that:
iv.) 1V · v = v · 1V = v
If, in addition, the product satisfies:
– 71 –
Exercise
a.) Write out a basis and structure constants for the algebra Mn (k).
b.) Show that the set of n × n matrices over an associative algebra, A, denoted Mn (A),
is itself an associative algebra.
Exercise
a.) If A is an algebra, then it is a module over itself, via the left-regular representation
(LRR). a → L(a) where
L(a) · b := ab (18.7)
Show that if we choose a basis ai then the structure constants
ai aj = cij k ak (18.8)
– 72 –
DIAGRAMS.
A bialgebra is an algebra which is simultaneously a coalgebra so that the structures are
all compatible, that is, the comultiplication ∆ and the counit are algebra homomorphisms.
Finally, a Hopf algebra is a bialgebra with an extra bit of structure S : H → H. It
is a generalization of the algebra of functions on a group and is a starting point for the
theory of quantum groups.
Exercise
a.) In the literature one sometimes sees a Frobenius algebra defined as an algebra A
with a nondegenerate bilinear form
such that σ(xy, z) = σ(y, zx). Show that this is equivalent to our definition.
b.) Show that if (A, θ) is a Frobenius algebra then the dual algebra A∗ is a left A-
module which is isomorphic to A as a left A-module.
The group algebra is a useful tool for addressing the fundamental questions at the beginning
of section ****
1. Useful for producing explicit reduction of matrix reps by the method of projection
operators.
2. Useful for producing the representations of the symmetric group Sn .
3. Important and beautiful mathematical structure.
The regular representation is an algebra. Indeed, the space of complex-valued functions
F un(X) on any space X is an algebra under pointwise multiplication:
Note that this is a commutative algebra. In particular, we can apply this remark to X = G.
Now, the group ring, usually denoted Z[G] is, as an abelian group the free abelian
group generatred by G. That is, it is the space of formal linear combinations
X
Z[G] := {x = xg g : xg ∈ Z} (19.2)
g∈G
– 73 –
The ring structure is defined by taking
X X group multiplication
X z}|{
x(g)g ∗2 y(g)g := x(g)y(h) g·h
g∈G g∈G g,h∈G
" # (19.4) eq:groupalgmul
X X
= x(g)y(h) k
k∈G g,h∈G:gh=k
Replacing Z → C (more properly, taking the tensor product with C) gives the group
algebra C[G].
As a vector space C[G] is naturally dual to the regular representation: RG ∼
= C[G]∗ .
The natural pairing is X
hf, xi := xg f (g) (19.5)
g∈G
Of course, there is a natural basis for C[G], namely the elements g themselves, and the
dual basis is the basis of delta functions δg . Thus the isomorphism takes f : G → C to
P
g f (g)g.
Now we see that there are two algebra structures on RG and C[G].
First of all, pointwise multiplication defines an algebra structure on C[G] since
which implies
(
g1 g1 = g2
g1 ∗1 g2 = (19.7) eq:pointwiseii
0 g1 6= g2
However, the multiplication (19.7) is not what is usually understood when one speaks
of the group algebra, rather what is meant is the multiplication (19.4).
Under the isomorphism RG ∼ = C[G] the product ∗2 must correspond to a second algebra
structure on RG , and indeed there is one - called the convolution product.
X
(f1 ∗2 f2 )(g) := f1 (h)f2 (h−1 g) (19.8) eq:covolution
h∈G
– 74 –
Example Z2 = {1, σ}. The group algebra is
RZ2 = {a · 1 + b · σ : a, b ∈ C} (19.9)
These are the “double-numbers,” a lesser known cousin of the complex numbers.
Exercise
Which elements of RZ2 are invertible?
for x ∈ C[G].
We know from the Peter-Weyl theorem that this representation is highly reducible.
We can now construct explicit projection operators which project onto the irreducible
(µ)
components. We assume that the distinct irreps Tij are somehow known.
Let us define:
by
nµ X (µ)
Pijµ := T̂ij (g)g (19.13) eq:projops
|G|
g∈G
Recall that the dual representation is the transpose inverse of the original one so
(µ) (µ)
T̂ij (g) = Tji (g −1 )
These elements satisfy the simple product law:
0
Pijµ Piµ0 j 0 = δµµ0 δji0 Pijµ0 (19.14) eq:prodlaw
They are therefore orthogonal projection operators in the group algebra acting on itself.
Moreover, again using similar manipulations one can show that:
(µ)
g · Pijµ = µ
X
Tsi (g)Psj (19.15) eq:geerep
s
(µ)
Pijµ · g −1 = T̂sj (g)Pisµ
X
(19.16) eq:geerepii
s
– 75 –
and thus the projection operators Pijµ explicitly decompose C[G] into the irreps T (µ) ⊗
T̂ (µ) of Gleft ×Gright . This will be useful below in showing how to decompose representations
into irreps.
Exercise
Prove (19.14)(19.15) and (19.16) using the orthogonality conditions The manipulations
are very similar to those used to prove theorem 10.1.
19.2 The center of the group algebra and the subalgebra of class functions
As we have noted, the group algebra C[G] is a noncommutative algebra. Correspondingly,
RG is a noncommutative algebra under the convolution product.
A little thought shows that the center of the group algebra Z[C[G]] is spanned by the
P
elements ci = g∈Ci g where we sum over the conjugacy class Ci of g. This simply follows
from X
gci g −1 = ghg −1 = ci (19.17)
h∈Ci
Correspondingly, the center of RG under the convolution product are the class functions.
Indeed, we can now interpret the exercise with equation (15.18) above as saying that
the characters satisfy:
δµν
χµ ∗2 χν = χµ (19.18)
nµ
The product is commutative, and in fact, diagonal in this basis. On the other hand, the
convolution product is nondiagonal in the conjugacy class basis:
X
δCi ∗2 δCj = Mijk δCk (19.19)
k
It is interesting to compare with the pointwise product, where the reverse is true
X
λ
χµ ∗1 χν = Nµν χλ (19.20)
λ
while
δCi ∗1 δCj = δij Ci (19.21)
So the unitary matrix Sµi is a kind of Fourier transform which exchanges the product for
which the characters of representations or characteristic functions of conjugacy classes is
diagonalized: s
X |G|
χµ = Siµ δCi (19.22)
mi
i
Exercise
– 76 –
Show that the structure constants in the basis δCi in the convolution product can be
written as:
∗ S∗ S
S0i S0j X Siµ jµ kµ
Mijk = (19.23) eq:convolusc
S0k µ nµ
Exercise
An algebra is always canonically a representation of itself in the left-regular represen-
tation. The representation matrices are given by the structure constants, thus
shows that the fusion coefficients are the structure constants of the pointwise multiplication
algebra of class functions.
Note that since the algebra of class functions is commutative the matrices L(χµ ) can
be simultaneously diagonalized.
Show that the unitary matrix Siµ is in fact the matrix which diagonalizes the fusion
coefficients.
The group algebra illustrates the notions of bialgebra and Frobenius algebra very nicely.
We would like to explain briefly the relation of these algebraic structures to two-dimensional
topological field theory.
The axioms of topological field theory give a caricature of what one wants from a
field theory: Spaces of states and transition amplitudes. By stripping away the many
complications of “real physics” one is left with a very simple structure. Nevertheless, the
resulting structure is elegant, it is related to beautiful algebraic structures which, at least
in two dimensions, which have surprisingly useful consequences. This is one case where
one can truly “solve the theory.”
Of course, we are really interested in more complicated theories. But the basic frame-
work here can be adapted to any field theory. What changes is the geometric category
under consideration.
– 77 –
Figure 7: A spacetime Xd = Y × R. Y is (d − 1)-dimensional space. fig:tfti
In a generic physical theory the vector space has a lot of extra structure: It is a Hilbert
space, there are natural operators acting on this Hilbert space such as the Hamiltonian.
The spectrum of the Hamiltonian and other physical observables depends on a great deal
of data. Certainly they depend on the metric on spacetime since a nonzero energy defines
a length scale
~c
L= (20.1)
E
In topological field theory one ignores most of this structure, and focuses on the de-
pendence of H(Y ) on the topology of Y . For simplicity, we will initially assume Y is
compact.
So: We have an association:
(d − 1)-manifolds Y to vector spaces: Y → H(Y ), such that homeomorphic manifolds
map to isomorphic vector spaces.
Now, we also want to incorporate some form of locality, at the most primitive level.
Thus, if we take disjoint unions
Note that (20.2) implies that we should assign to H(∅) the field of definition of our
vector space. For simplicity we will take H(∅) = C, although one could use other ground
fields.
– 78 –
∼ Y and (∂Xd )out ∼
so that there is a homeomorphism (∂Xd )in = = Y 0.
Here we will be considering oriented manifolds. Then an oriented cobordism from Y1
to Y2 must have
∂M = Y2 − Y1 (20.4)
where the minus sign indicates the opposite orientation on Y1 . (Recall that an orientation
on M determines one on ∂M . We will adopt the convention “outward normal first” - ONF
- “One Never Forgets.” ).
If Xd is an oriented cobordism from Y to Y 0 then the Feynman path integral assigns
a linear transformation
F (Xd ) : H(Y ) → H(Y 0 ). (20.5)
Again, in the general case, the amplitudes depend on much more than just the topology
of Xd , but in topological field theory they are supposed only to depend on the topology.
More precisely, if Xd ∼
= Xd0 are homeomorphic by a homeomorphism = 1 on the boundary
of the cobordism, then
F (Xd ) = F (Xd0 ) (20.6)
One key aspect of the path integral we want to capture - again a consequence of locality
- is the idea of summing over a complete set of intermediate states. In the path integral
formalism we can formulate the sum over all paths of field configurations from t0 to t2
by composing the amplitude for all paths from t0 to t1 and then from t1 to t2 , where
t0 < t1 < t2 , and then summing over all intermediate field configurations at t1 . We refer
to this property as the “gluing property.” The gluing property is particularly obvious in
the path integral formulation of field theories.
∂M = Y2 − Y1 (20.7)
∂M 0 = Y3 − Y2 (20.8)
– 79 –
What we are describing, in mathematical terms is a functor between categories.
The above axioms can be concisely summarized by defining a cobordism category whose
objects are homeomorphism classes of oriented (d − 1)-manifolds and whose morphisms are
oriented cobordisms, two morphisms being identified if they differ by a homeomorphism
which is an identity on the boundary. Let us call this Cob(d).
Definition A d-dimensional topological field theory is a functor from the category Cob(d)
to the category VECT of vector spaces and linear transformations.
Actually, these are both tensor categories and we want a functor of tensor categories.
At this point we begin to see how we can incorporate other field theories. We can
change the geometric category to include manifolds with Riemannian metric, spin structure,
etc.
exp[−T H] (20.12)
where H is the Hamiltonian, and T is the Euclidean time interval (in some presumed
metric).
Evidently, by the axioms of topological field theory, P 2 = P and therefore we can
decompose
– 80 –
All possible transitions are zero on the second summand since, topologically, we can
always insert such a cylinder. It follows that it is natural to assume that
One can think of this as the statement that the Hamiltonian is zero.
Let us now compose these cobordisms we get the identity map as in 11. It then follows
from some linear algebra that Q is a nondegenerate pairing, so we have an isomorphism to
the linear dual space:
H(−Y ) ∼
= H(Y )∗ , (20.17)
under which Q is just the dual pairing. Moreover δ(1) has the explicit formula:
X
δ:1→ φµ ⊗ φµ (20.18)
where we can let φµ be any basis for H(Y ), and then define the dual basis by Q(φµ , φν ) =
δµ ν .
Exercise
– 81 –
αµν φµ ⊗ φν and compute the process in 9.
P
Prove this formula. Write δ(1) =
Figure 12: Composing Q with δ gives the dimension: dimH(Y ) = Z(Y × S 1 ). fig:dimcirc
where β is the radius of the circle and H is the Hamiltonian. If in addition we enrich
the category to include spin structures then there will be several kinds of traces.
Exercise
Show that a 1-dimensional TFT is completely specified by choosing a single vector
space V and a vector in that space.
Exercise
Show that any cobordism of ∅ to Y defines a state in the space H(Y ). This is a
primitive version of the notion of the “Hartle-Hawking” state in quantum gravity. It is
also related to the state/operator correspondence in conformal field theory.
20.3 Two dimensional closed topological field theory and Commutative Frobe-
nius Algebras
Some beautiful extra structure shows up when we consider the case d = 2, due to the
relatively simple nature of the topology of 1-manifolds and 2-manifolds. To begin, we
restrict attention to closed (d − 1)-manifolds, that is, we consider a theory of closed strings.
– 82 –
In this case, the spatial (d − 1) manifolds are necessarily of the form S 1 ∪ S 1 ∪ · · · ∪ S 1 ,
i.e. disjoint unions of n S 1 ’s.
If we assign H(S 1 ) = C then
H(S 1 q S 1 q · · · q S 1 ) = C ⊗n (20.20)
F (Σ) : C ⊗n → C ⊗m (20.21)
where Σ is a Riemann surface. There are n ingoing circles and m outgoing circles.
Now, topology dictates that the vector space C in fact must carry some interesting
extra structure.
m:C⊗C →C (20.22)
θ:C→C (20.24)
– 83 –
Figure 15: the trace map fig:trace
The proof (which is not difficult) depends on Morse theory and will not be given here.
For a detailed discussion see, for example,
– 84 –
Figure 18: Associativity fig:associativ
G. Segal and G. Moore, “D-branes and K-theory in 2D topological field theory” hep-
th/0609042; To appear in Mirror Symmetry II, Clay Mathematics Institute Monograph,
by P. Aspinwall et. al.
∂X = Y0 ∪ Y1 ∪ ∂cstr X (20.25)
where ∂cstr X is the time-evolution of the spatial boundaries. We will call this the “con-
strained boundary.”
In d = 2, in this enlarged geometric category the initial and final state-spaces are
associated with circles, as before, and now also with intervals. The boundary of each
interval carries a label a, b, c, . . . from the set B0 .
– 85 –
Definition: We denote the space Oab for the space associated to the interval [0, 1]
with label b at 0 and a at 1.
In the theory of D-branes, the intervals are open strings ending on submanifolds of
spacetime. That is why we call the time-evolution of these boundaries the “constrained
boundaries” – because the ends are constrained to live in the D-brane worldvolume.
As in the closed case, the cobordism [0, 1] × [0, 1] defines Pab : Oab → Oab , and we can
assume WLOG that it is Pab = 1.
Now consider the cobordism in 21. This clearly gives us a bilinear map
As in the closed case we see that these maps satisfy an associativity law. Moreover, as in
the closed case, 1a is an identity for the multiplication.
Comparing with the definition of a category we see that we should interpret B0 as the
space of objects in a category B, whose morphism spaces Hom(b, a) = Oab . Note that the
morphism spaces are vector-spaces, so we have a C-linear category. In fact, this category
– 86 –
has a very special property as we learn from considering the S-shaped cobordism (the open
string analog of 11). We learn that θa : Oaa → C defines a nondegenerate inner product:
a θ
Oab ⊗ Oba → Oaa → C
(20.28) eq:dualpair
θ
Oba ⊗ Oab → Obb →b C
Remark: It is important to note that the argument for commutativity fails in the
open case: The algebras Oaa are in general noncommutative.
So, to give an open and closed TFT involves giving a Frobenius cateogry. But the
open and closed strings must be related to each other. The essential new information is a
pair of linear maps
ιa : C → Oaa
a
(20.30) eq:linmaps
ι : Oaa → C
– 87 –
Figure 25: Factorization of the open string loop on closed string exchange. Also known as the
“Cardy condition.” fig:cardycond
a
πb = ιb ◦ ι a . (20.36) eq:cardycon
Exercise
7
These are actually generalization of the conditions stated by Cardy. One recovers his conditions by
taking the trace. Of course, the factorization of the double twist diagram in the closed string channel is an
observation going back to the earliest days of string theory.
– 88 –
Draw pictures associated to the other algebraic conditions given above.
Theorem Open-Closed Sewing Theorem. The above conditions are the complete set
of sewing constraints on the algebraic data.
1
θV (ψ) = Tr(ψ) (20.37)
|G|
Exercise
Write out the full set of open-closed string data for the example of a finite group and
check the sewing conditions.
Let us return to the projection operators in C[G]. In this section we will assume - as we
may - that the irreps T µ are unitary so
(µ) (µ)
T̂ij (g) = (Tij (g))∗ (21.1)
Thus we have
nµ X (µ)
Pijµ = (Tij (g))∗ g (21.2) eq:projopsa
|G|
g∈G
0
Pijµ Piµ0 j 0 = δµµ0 δji0 Pijµ0 (21.3) eq:prodlawa
(µ)
g · Pijµ = µ
X
Tsi (g)Psj (21.4) eq:geerepa
s
(µ)
Pijµ · g −1 = (Tsj (g))∗ Pisµ
X
(21.5) eq:geerepiia
s
– 89 –
21.1 Decomposition of a representation into its isotypical parts
While characters tell us the irreps into which a representation can be decomposed, they do
not tell us how to decompose the representation explicitly. To do this, we need the method
(µ)
of projection operators. We assume that the matrix elements Tij are known.
For simplicity we consider finite groups | G |< ∞, but the same techniques extend to
compact groups.
Now, let T (g) be any representation on V . Extending by linearity we obtain a rep-
resentation of the algebra RG . Define operators on V by applying T to (19.12) to get
8
nµ X (µ)
Pijµ = T (Pijµ ) = (Tij (g))∗ T (g) (21.6)
|G|
gG
Now we imediately get the properties (19.14), (19.15), (19.16) for the operators Pijµ .
Therefore, if Pµlk~v = w
~ l 6= 0 then
n o
~ j } = Pµjk~v
{w (21.7)
j=1,···nµ
span an invariant subspace of V transforming in the rep T (µ) according to the representation
T µ . Using these operators we can in principle decompose a representation into its irreps.
111
1
P1+ = 1 1 1 P12+ = P1+
3
111
000
P1− = 0 0 0
000
(21.8)
0 0 0
P21,1 = 0 1/2 −1/2 (P21,1 )2 = P21,1
0 −1/2 1/2
√ √
0 1/ 3 −1/ 3
√ √
P22,1 = 0 −1/2 3 1/2 3
√ √
0 −1/2 3 1/2 3
8 (µ)
We choose Tij = unitary matrices for simplicity.
– 90 –
Therefore,
x 1
x+y+z
P1+ y = ( ) 1 (21.9)
3
z 1
so:
n o
V1 = Span V~1 = (1, 1, 1) = {(x, x, x) : x ∈ R} (21.10)
Similarly P2 projects onto the orthogonal subspace
x 0 x √ 2
1,1 1 2,1 3
P2 y = (y − z) 1 P2 y = (y − z) −1 (21.11)
2 2
z −1 z −1
So: n o
V2 = Span V~2 = (0, 1, −1)V
~3 = (2, −1, −1) (21.12)
are the two invariant subspaces. Check
1 1 1
S = 0 1 −1 (21.13)
2 −1 −1
Conjugates the rep into block form. Thus, the decomposition into irreps is:
R3 ∼
= V1+ ⊕ V2 (21.14) eq:expldec
An important shortcut: In general the characters of irreps are much easier to calculate
than the matrix elements. Therefore the following route to decomposition is an important
shortcut:
If
V ∼
= ⊕µ aµ T (µ) (21.15)
then we can easily project onto the isotypical component as follows: Note that for unitary
T (g), (Pµij )∗ = Pµij , so
nµ
Piiµ
X
P µ := (21.16) eq:charporj
i=1
form a system of orthogonal, Hermitian, projection operators:
(Pµ )2 = Pµ
(Pµ )† = Pµ (21.17) eq:projopsaa
Pµ Pν = 0 µ 6= ν
– 91 –
They project onto the invariant subspaces V (µ) .
The point is: from the definition (21.16) we get:
nµ X
Pµ = χµ (g)T (g) (21.18) eq:calcpmu
|G|
g∈G
∀g ∈ G [T (g), H] = 0 (21.20)
Then
[Pijµ , H] = 0 (21.21)
Then, the projection operators Pλ onto the different eigenspaces of H commute with the
different projection operators Pijµ and in particular with P µ .
Therefore, in this situation, by reducing CN to the irreps of G, we have partially block-
diagonalized H.
Of course, this is very useful in quantum mechanics where one wants to diagonalize
Hermitian operators acting on wavefunctions.
nµ X µ
ij
ψa,µ (x) ≡ (Tij (h))∗ ψa (h−1 · x) (21.22)
|G|
hG
– 92 –
nµ X µ
ψaij (g −1 · x) = (Tij (h))∗ ψa (h−1 · g −1 · x)
|G|
hG
nµ X µ −1 ∗
= (Tij (g h)) ψa (h−1 · x)
|G|
hG
nµ (21.23)
(Tisµ (g −1 ))∗ ψasj (x)
X
=
s=1
nµ
Tsiµ (g)ψasj (x)
X
= if T µ is unitary.
s=1
1
ψ 1 (x) = (ψ(x) + ψ(−x))
2 (21.24)
2 1
ψ (x) = (ψ(x) − ψ(−x))
2
transform acccording to the two irreps of Z2 .
~q = (q 1 , . . . , q n ) (21.25)
Suppose we have a generalized harmonic oscillator so that the kinetic and potential energies
are:
1 1
T = mij q̇ i q̇ j V = Uij q i q j (21.26)
2 2
where Uij is independent of q i . We can obtain solutions of the classical equations of motion
by taking
q i (t) = Re(γ i eiωt ) (21.27)
where
(−ω 2 mij + Uij )γ j = 0 (21.28)
If a group G acts as (T (g) · q)i = T (g)ij q j and
– 93 –
Figure 26: A system of beads and springs. fig:normalmode
The Lagrangian is
1 1 X i
L = m(q̇ i )2 − k (q − q i+1 )2 (21.30)
2 2
i
where
2 −1 0 0 ··· ··· −1
−1 2 −1 0 0 0 ···
A= 0 −1 2 −1 0 ··· 0 (21.32)
· · · ··· ··· ··· ··· ··· −1
−1 0 ··· 0 0 −1 2
Thus, we reduce the problem of finding normal modes to the problem of diagonalizing
this matrix.
The problem has an obvious Zn symmetry where a generator T (ω) takes
Here ei is the standard basis for Cn so T (ω ` )ei = ei+` and again the subscript is understood
modulo n.
We know the irreps of Zn : T µ (ω j ) = ω jµ , µ = 0, . . . , n − 1. The projection operators
are thus
n−1
µ 1 X −µ`
P = ω T (ω ` ) (21.34)
n
`=0
9
This alters the problem physically. We should add a constraint which says, roughly speaking that
(q i+1 − q i ) = 2π. We will ignore this constraint in what follows.
P
– 94 –
In terms of matrix units eij we have
1 X −µ`
Pµ = ω ei+`,i (21.35)
n
i,`
AP µ = (2 − ω −µ − ω µ )P µ
πµ 2 µ (21.38)
= (2 sin ) P
n
Similarly,
1 X 1 X X
P µ ~q = ω −µ` q s es+` = ω µs q s ω −µj ej (21.39)
n n s
`,s j
It is a general principle that when you have unexplained degeneracies you should search
for a further symmetry in the problem.
Indeed, in this case, we have overlooked a further symmetry:
q i → q n−i (21.43)
This leads to Dn symmetry of the problem which explains the degeneracy (21.42).
Exercise
Using the character tables above for Dn construct the normal modes which are in
representations of Dn .
– 95 –
22. Representations of the Symmetric Group
where νj is the number of cycles of length j. Clearly since we must account for all n letters
being permuted: X
n = ν1 + 2ν2 + · · · + nνn = jνj (22.2)
To every cycle decomposition we can associate a partition as follows:
λ1 := ν1 + ν2 + · · · + νn
λ2 := ν2 + · · · + νn
λ3 := ν3 + · · · + νn (22.3) eq:lambdaTOnu
······
λn := νn
Note
1. λ1 ≥ λ2 ≥ · · · ≥ λn ≥ 0
P
2. λi = n
3. Given a partition of n we can recover the cycle decomposition (j)νj since the
Q
classes, and hence the irreps of Sn , are in 1-1 correspondence with the partitions of n and hence in 1-1 correspondence with Y
– 96 –
λ = (2,0)
λ = (1,1)
λ = (3,0,0)
λ = (2,1,0)
λ = (5,3,2,1,1,0,0,0,0,0,0,0)
λ = (1,1,1)
Similarly, C(T ) consists of permutations which only permute numbers within each column
of T .
For a Young tableau T define the following elements in the group ring Z[Sn ]:
X
P := p (22.4)
p∈R(T )
X
Q := (q)q (22.5)
q∈C(T )
X
P (T ) := P Q = (q)pq (22.6)
p∈R(T ),q∈C(T )
– 97 –
dimR(T ) can also be characterized as the number of standard tableaux corresponding to
the underlying partition.
6. Another formula for dimR(T ) is the hook length formula. For a box in a Young
diagram define its hook length to be the number of squares to the right in its row, and
underneath in its column, counting that box only once. Then
n!
dimR(T ) = Q (22.8)
hooklengths
22.2.1 Example 1: G = S3
There are 3 partitions, 3 Young diagrams, and 4 standard tableaux shown in 28.
2
P
Clearly P (T1 ) = p∈S3 p and P (T1 ) = 6P (T1 ). Moreover Z[S3 ] · P (T1 ) is one-
dimensional. This is the trivial representation.
P
Similarly, P (T4 ) = (q)q spans the one-dimensional sign representation.
There are two standard tableaux corresponding to λ = (2, 1) with
– 98 –
One easily checks that the character is the same as that we computed above for the 2.
Of course, similar statements hold for P (T3 ), which the reader should check as an
exercise.
22.2.2 Example 2: G = S4
References
1. For further discussion of the above material see the books by Miller, Hammermesh,
Curtis+Reiner, Fulton+Harris Representation Theory.
2. There is a second, elegant method for constructing the irreps of the symmetric
group using induced representations. See the book by Sternberg for an account.
23. Symmetric groups and tensors: Schur-Weyl duality and the irreps of
GL(d, k)
V ⊗n ≡ V ⊗ V ⊗ · · · ⊗ V (23.1)
– 99 –
For example, suppose that V = k d is also the fundamental representation of GL(d, R)
or GL(d, C) for k = R or k = C. Given a basis {vi } for V a typical element can be expanded
in the basis as: X
v= ti1 ,i2 ,...,in vi1 ⊗ · · · ⊗ vin (23.4)
i1 ,i2 ,...,in
We will often assume the summation convention where repeated indices are automatically
summed. Under the action of GL(d, k):
g · vi = gji vj (23.5)
we therefore have:
(g · t)i1 ···in = tj1 ···jn gj1 i1 · · · gjn in (23.6) eq:tensor
Now, note that the actions of GL(d, k) and Sn commute, so that V ⊗n is a representation
of GL(d, k) × Sn .
Considered as a representation of Sn , we have complete reducibility:
V ⊗n ∼
= ⊕λ Vλ ⊗ Rλ (23.8)
Proof For complete proofs see Segal, Miller, Fulton-Harris. We will just sketch the
idea from Segal’s lectures.
We can see fairly easily why all the representations of U (d) must occur if we accept
the Peter-Weyl theorem. Note that the representations which we obtain from V ⊗d have
matrix elements which are algebraic expressions in the matrix elements aij . Indeed, they
lie in C[aij ] (the aij occur individually of course from V itself). But this forms a dense
subalgebra of L2 (U (d)).
– 100 –
Now let us prove the Vλ is irreducible. Recall that the center of a matrix algebra
End(W ) is just C. Now we note that
Since Gl(d) forms a dense open set in End(V ), the operators T (g) acting on V ⊗n generate
an algebra which is all of EndSn (V ⊗n ). However, this means that the operators that
commute with both G and Sn , i.e.
EndG×Sn (V ⊗n ) (23.11)
and
EndG×Sn (V ⊗n ) = ⊕λ EndG (Vλ (23.13)
Therefore EndG (Vλ must be contained in the center of End(Vλ ), but the latter is just C.
Therefore EndG (Vλ is just C. But, by Schur’s lemma, this must mean that Vλ is irreducible.
♠
P (T )V ⊗n ∼
= P (T 0 )V ⊗n (23.14)
if T and T 0 correspond to the same partition, i.e. the same Young diagram.
With some work it can be shown that - as representations of GL(d, k) we have an
orthogonal decomposition
V ⊗n = ⊕T P (T )V ⊗n (23.15)
where T runs over the standard tableaux of Sn . Of course, tableaux with columns of length
≥ d will project to the zero vector space and can be omitted.
Example
• If we take the partition λ = (n, 0, 0, . . . ) then P (λ) projects to totally symmetric
tensors and we get S n (V ). Weyl’s theorem tells us that if V = k d is the fundamental
representation of GL(d, k) then S n (V ) is an irreducible representation. We computed
before that its dimension is
n n+d−1
dimS (V ) = (23.16)
n
Note in particular that we have infinitely many irreducible representations.
– 101 –
• If we take the partition λ = (1, 1, . . . , 1) then P (λ)V ⊗n = Λn V is the subspace of
totally antisymmetric tensors of dimension
d
dimΛn (V ) = (23.17)
n
Note that this subspace vanishes unless dimV = d ≥ n.
• Let V be of dimension d and consider V ⊗3 as a representation of S3 . Using characters
we showed above that
d(d + 1)(d + 2) d(d − 1)(d − 2) d(d + 1)(d − 1)
V ⊗3 ∼
= 1+ ⊕ 1− ⊕ 2 (23.18)
6 6 3
Clearly, the first and second summands correspond to the totally symmetric and totally
antisymmetric tensors, respectively.
Applying the projector (22.9) to a tensor tijk vi ⊗ vj ⊗ vk produces a tensor with mixed
symmetry:
P (T2 )vi ⊗ vj ⊗ vk = vi ⊗ vj ⊗ vk + vj ⊗ vi ⊗ vk − vk ⊗ vj ⊗ vi − vk ⊗ vi ⊗ vj (23.19)
and therefore the components are of the form:
tijk + tjik − tkji − tjki (23.20)
for arbitrary tijk . This tensor space is the space of tensors t̃ijk which satisfy the identities:
– 102 –
23.1 Free fermions on a circle and Schur functions
n
Ψ~n (z1 , . . . , zN ) = det1≤i,j≤N zi j (23.29) eq:slater
For simplicity let us assume N is odd. Then there is a unique groundstate obtained
by filling up the states −nF ≤ n ≤ nF with
N −1
nF = (23.30) eq:fermil
2
(If N is even there are two groundstates.) This defines the Fermi sea. The ground
state wavefunction is
– 103 –
The Schur functions form a linear basis for Z[x1 , . . . , xN ]SN where basis elements are
associated with partitions. To define them mathematically let us return to a monomial
xα = xα1 1 · · · xαNN with the αi ≥ 0 and now try to skew-symmetrize:
X
(σ)σ · xα (23.33)
σ∈SN
Note that this function is totally antisymmetric, and moreover vanishes unless the αi are
all distinct. WLOG assume that
λ +j−1
∆λ (x) := det(xi j )
· · · xλ1 N +N −1
λ1 λ2 +1
x1 x1
· · · ·
(23.39) eq:deltsun
= det · · · ·
· · · ·
λ1 λ2 +1 λN +N −1
xN xN · · · xN
Exercise
Show that
∆λ (σ · x) = (σ)∆λ (x) (23.41)
– 104 –
where (σ) is the sign homomorphism.
∆λ (x)
Φλ (x) := (23.42) eq:schurf
∆0 (x)
This is a totally symmetric function. In fact, it is a polynomial in the xi . To see
this consider it as a meromorphic function of a complex variable x1 . Note that it is in
fact an entire function since the potential poles at x1 = xi are cancelled by zeroes of the
numerator. Next note that the growth at infinity is obviously xm 1 for an integer m.
Ψ~n ∆λ (z)
= (23.45) eq:fration
Ψ~0 ∆0 (z)
for
αj = nj + nF
, (23.46)
= λj + j − 1
that is
Remarks
• The second main theorem of symmetric polynomials is:
Theorem. The Schur functions Φλ (x) form a linear integral basis for Z[x1 , . . . , xN ]SN .
That is, any symmetric polynomial with integral coefficients can be written as a lin-
ear combination, with integral coefficients, of the Φλ (x). Note that we are not forming
polynomials in the Φλ .
• The module AN of skew-symmetric polynomials in x1 , . . . , xN is isomorphic to ΛN
via multiplication by ∆0 . Therefore the ∆λ form a a Z-basis for AN . This close relation
between completely symmetric and antisymmetric functions comes up in the theory of
matrix models – integrals over space of N × N matrices. It also suggests a relation between
bosons and fermions, at least in 1+1 dimensions. That indeed proves to be the case – there
– 105 –
is a nontrivial isomorphism known as bosonization which is an isomorphism of quantum
field theories of bosons and fermions in 1+1 dimensions.
•
There is an elegant expression for the Schur functions Φλ as a determinant of a matrix
whose entries involve the hk . The expression is
hλ1 hλ1 +1 hλ1 +2 hλ1 +3 hλ1 +4 ...
h
λ2 −1 hλ2 hλ2 +1 hλ2 +2 hλ2 +3 ...
Φλ = 0 0 1 h1 h2 ···
(23.50) eq:phitwoa
0 0 0 1 h1 ···
··· ··· ··· ··· ··· ···
= hλ1 hλ2 − hλ1 +1 hλ2 −1
With a little algebra one can confirm that this is indeed (23.44). Note, however, that
both sides of this equality make sense for N > 2.
• Define the Z × N matrix Ξ whose (p, q)th entry is hq−p where h0 = 1 and hk = 0 for
k < 0. Here −∞ < p < ∞ labels the rows, while q = 1, 2, . . . labels the columns. Thus the
matrix looks like 31.
Define the integers sk = k − λk . These eventually become sk = k. Then Φλ is the
determinant of the matrix formed from Ξ by keeping only the rows labeled by sk . This
leads to an upper triangular matrix is the determinant of the finite matrix above.
– 106 –
• The definition of Φλ only makes sense as a function of N such that λi = 0 for i > N .
However, the relations between Φλ and the hi are universal.
Exercise
Verify that (23.50) and (23.44) are equal for N = 2.
Exercise
a.) Let λ = (n, 0, 0, 0, . . . ). Show that Φλ = hn .
b.) Let λ = (1n , 0, 0, . . . ). Then Φλ = en .
Exercise
Using Cauchy’s determinant identity:
1 ∆0 (xi )∆0 (yj )
det =Q (23.51)
1 − x i yj i,j (1 − xi yj )
Show that X X
1
exp sk (x)sk (y) = Φλ (x)Φλ (y) (23.52)
k
k λ
Hint: Multiply both sides by ∆0 (x)∆0 (y). For the answer see Hammermesh, p. 195
– 107 –
23.1.1 Schur functions, characters, and Schur-Weyl duality
Now it is interesting to combine Schur functions with Schur-Weyl duality.
We have seen that irreducible representations of GL(d, C) (and of U (d) ) can be labeled
by Young diagrams with ≤ d rows. We called these Sλ above.
It turns out that the Φλ (x) are characters of the representations Sλ . That is if g ∈
GL(d, C) can be diagonalized to Diag{x1 , . . . , xd } then
This is known as the Weyl character formula and will be derived later.
The relation between the Φλ and the power functions sk (x) is a very nice application
of the Schur-Weyl duality theorem.
Suppose that a group element σ ∈ SN has cycle decomposition (1)`1 (2)`2 · · · where
P
j`j = N . As we have discussed this determines a conjugacy class C[`] as well as a
partition. Now, suppose g ∈ GL(d, C) is diagonalizable to Diag{x1 , . . . , xd } and let us
evaluate
Then:
X
s(`) = χλ (C[`]))Φλ (23.61) eq:frobenius
{λ}
– 108 –
acts on the groundstate wavefunction Ψgnd (x) to produce the quantum states:
X
O(`)Ψgnd (x) = χλ (C[`]))Ψ~n (23.63) eq:transamp
λ
where the ni are the fermion occupation numbers corresponding to the partition asso-
ciated to λ as in (23.47). Thus the characters of the symmetric group can be regarded as
“transition amplitudes” in the free fermion problem!
Indeed, we can now write the character table of the symmetric group SN :
χλ (C[`]) = ∆0 (x)s(`) (x) α (23.64) eq:charactsymm
1 ,...,αd
where the subscript means we extract the coefficient of xα1 1 · · · xαd d and we recall that
αj = λj + j − 1 (23.65)
We will see later that this is a special case of the Weyl dimension formula.
Hint: Put xj = ejt and let t → 0.
References:
– 109 –
There are a lot of beautiful connections here with quantum field theory. In particular
there are close connections to two-dimensional Yang-Mills theory and to the phenomenon
of bosonization in two dimensions. See
For much more about this see the references below.
1. Pressley and Segal, Loop Groups, Chapter 10.
2. M. Stone, “Schur functions, chiral bosons, and the quantum-Hall-effect edge states,”
Phys. Rev. B42 1990)8399 (I have followed this treatment.)
3. S. Cordes, G. Moore, and S. Ramgoolam, “Lectures on 2D Yang-Mills Theory,
Equivariant Cohomology and Topological Field Theories,” Nucl. Phys. B (Proc. Suppl
41) (1995) 184, section 4. Also available at hep-th/9411210.
4. M. Douglas, “Conformal field theory techniques for large N group theory,” hep-
th/9303159.
5. Papers of Jimbo, Miwa, ..... (Kyoto school) on Toda theory.
– 110 –
*****************************************
FOLLOWING MATERIAL SHOULD BE MOVED TO CHAPTER ON REPRESEN-
TATIONS. IT WOULD MAKE MORE SENSE TO TALK ABOUT REPS OF THE SYM-
METRIC GROUP FIRST.
23.2.1 Bosonization
Finally, we want to describe a truly remarkable phenomenon, that of bosonization in 1+1
dimensions.
Certain quantum field theories of bosons are equivalent to quantum field theories of
fermions in 1+1 dimensions! Early versions of this idea go back to Jordan 10 The subject
became important in the 1970’s. Two important references are
S. Coleman, Phys. Rev. D11(1975)2088
S. Mandelstam, Phys. Rev. D11 (1975)3026
The technique has broad generalizations, and plays an important role in string theory.
To get some rough idea of how this might be so, let us consider the loop group S 1 →
U (1). For winding number zero, loop group elements can be written as:
+∞
X
z → g(z) = exp[i jn z n ] (23.68)
n=−∞
z n → g(z)z n (23.69)
Note that there are two Fermi levels, and if N is large they are “far apart” meaning
that operators such as i zin for small n will not mix states near the two respective lev-
P
els. Let us therefore imagine taking N large and focussing attention on one of the Fermi
levels. Therefore we adjust our energy level so that the groundstate wavefunctions are
10
P. Jordan, Z. Phys. 93(1935)464.
– 111 –
1, z −1 , z −2 , . . . and we imagine that we have taken N → ∞. Moreover, let us extend the
LU (1) action to LC∗ . Then we can separately consider the action of
0
X
g− (z) = exp[i ϕn z n ] (23.71)
−∞
and
∞
X
g+ (z) = exp[i ϕn z n ] (23.72)
1
with a−j = a†j . These are realized as follows: a†k is multiplication by sk and ak is k ∂s∂k .
The state
(a†1 )`1 (a†2 )`2 · · · (a†N )`N |0i (23.75)
– 112 –
∞
X
exp[i jn a†n ]|0i (23.78) eq:coherenttwo
n=1
thus producing a nontrivial isomorphism between a bosonic and fermionic Fock space.
To make the isomorphism between bosonic and fermionic Fock spaces more explicit we
+
introduce a second-quantized formalism. B−n creates a fermionic state with wavefunction
n
z , Bn annihilates it, so we introduce:
X
Ψ(θ) = Bn einθ
n∈Z
X (23.79) eq:fermflds
† + −inθ
Ψ (θ) = B−n e .
n∈Z
The filled Fermi sea satisfies the constraints:
1 1
Ψ(θ) = ei(nF + 2 )θ b(θ) + e−i(nF + 2 )θ b̄(θ)
1 1
(23.81) eq:bcsys
Ψ† (θ) = e−i(nF + 2 )θ c(θ) + ei(nF + 2 )θ c̄(θ).
We introduce complex coordinates z = eiθ , and define the mode expansions:
X
b(z) = bn z n
n∈Z+ 12
X
c(z) = cn z n
n∈Z+ 12
X (23.82) eq:bcsysi
b̄(z) = b̄n z̄ n
n∈Z+ 12
X
c̄(z) = c̄n z̄ n
n∈Z+ 12
bn , cn , b̄n , c̄n are only unambiguously defined for | n |<< N . That is, we focus on
operators that do not mix excitations around the two Fermi levels. The peculiar half-
integral moding is chosen to agree with standard conventions in CFT. In terms of the
original nonrelativistic modes we have:
+
cn = B−nF −+n
bn = BnF ++n
(23.83) eq:mapmodes
c̄n = Bn+F +−n
b̄n = B−nF −−n
– 113 –
where = 12 , so that
(p)
Hbc = ⊕p∈Z Hbc (23.86) eq:grdspce
and the states obtained by moving a state from the Fermi sea to an excited level
correspond to the subspace of zero bc number:
(0)
Hchiral := Hbc (23.87) eq:chrlspc
c
[Ln , Lm ] = (n − m)Ln+m + n(n2 − 1)δn+m,0 (23.89) eq:virasoroalg
12
with c = 1.
Now let us consider the second quantized operators corresponding to z n . Before taking
the limit these are given by:
X Z
Υn = zin ←→ dθΨ† (θ) einθ Ψ(θ). (23.90) eq:upsop
Now let us take the large N limit and focus on terms that do not mix the two Hilbert
spaces of excitations around the two Fermi levels. Cross terms between barred and unbarred
fields involve operators that mix the two Fermi levels. Since we are only interested in the
case of the decoupled Fermi level excitations we may replace:
– 114 –
I I
−1−n
Υn → dz z c b(z) + dz̄ z̄ −1+n c̄ b̄(z̄)
X
= cn−m bm + c̄m−n b̄−m (23.91) eq:upsii
m
= αn + ᾱ−n
X
∂z φ(z) = i αm z m−1
m∈Z
(23.92) eq:bosalg
[αm , αn ] = [ᾱm , ᾱn ] = mδm+n,0
[αm , ᾱn ] = 0.
1X
Ln = αn−m αm (23.93) eq:virop
2
which satisfy (23.89), again with c = 1. Using the α we can define a vacuum αn | 0i = 0
for n ≥ 0 and a statespace
( )
Y
Hα = Span | ~ki ≡ (α−j ) kj
| 0i . (23.94) eq:achalph
(0)
Hα ∼
= Hbc (23.95) eq:bsnztin
We will not prove this but it can be made very plausible as follows. The Hilbert space
may be graded by L0 eigenvalue. The first few levels are:
At level L0 = n, the fermion states are labeled by Young diagrams Y with n boxes.
At level L0 = n, the Bose basis elements are labeled by partitions of n. We will label a
partition of n by a vector ~k = (k1 , k2 , . . . ) which has almost all entries zero, such that
P
j jkj = n. Bosonization states that the two bases are linearly related:
X
|Yi= h~k | Y i | ~ki (23.97) eq:linrel
~k∈Partitions(n)
– 115 –
2. Look up papers of the Kyoto school, Jimbo et. al. for perhaps helpful ways of
presenting this material.
***************
For much more about this see the references below.
1. Pressley and Segal, Loop Groups, Chapter 10.
2. M. Stone, “Schur functions, chiral bosons, and the quantum-Hall-effect edge states,”
Phys. Rev. B42 1990)8399 (I have followed this treatment.)
3. S. Cordes, G. Moore, and S. Ramgoolam, “Lectures on 2D Yang-Mills Theory,
Equivariant Cohomology and Topological Field Theories,” Nucl. Phys. B (Proc. Suppl
41) (1995) 184, section 4. Also available at hep-th/9411210.
4. M. Douglas, “Conformal field theory techniques for large N group theory,” hep-
th/9303159.
– 116 –
Next time – include
Verma modules – how the more standard treatment of SU(2) reps fits in.
basis
z j+m
fm (z) = p ↔ | j, m > (23.98)
(j + m)!(j − m)!
for −j ≤ m ≤ j
is orthonormal wrt:
Z
(2j + 1)! 1
(f, g) = d2 z f ∗ g(z) (23.99)
π (1+ | z |2 )2j+1 (z)
– 117 –