Representation Theory IV

3 SL2

In this section we discuss the finite-dimensional representation theory of the Lie algebra 𝔰𝔩2,. We then use that to study the representation theory of SL2(), 𝔰𝔩2,, SL2(), and SU(2), which are all closely related.

3.1 Weights

We fix the following ’standard’ basis of 𝔰𝔩2,:

H =(100-1),
X =(0100),
and
Y =(0010).

These satisfy the following commutation relations, which are fundamental (check them!):

[H,X] =2X,
[H,Y] =-2Y,
and
[X,Y] =H.

We decompose representations of 𝔰𝔩2, into their eigenspaces for the action of H. The elements X and Y will then move vectors between these eigenspaces, and this will let us analyze the representation theory of 𝔰𝔩2,.

Since SL2() is simply connected, we have

Proposition 3.1.

Every finite-dimensional representation of 𝔰𝔩2, is the derivative of a unique representation of SL2().

Note that we have not proved this. However, we will use the result freely in what follows. It is possible to give purely algebraic proofs of all the results for which we use the previous proposition, but it is more complicated.

Proposition 3.2.

Let (ρ,V) be a finite-dimensional complex-linear representation of 𝔰𝔩2,. Then ρ(H) is diagonalizable with integer eigenvalues.

Proof.

By the previous proposition ρ is the derivative of a representation ρ~ of SL2(). We can identify U(1) as a subgroup of SL2() by the following map:

f:eit(eite-it).

By Maschke’s Theorem for U(1), ρ((eite-it)) on V can be diagonalized. Taking the derivative, we see that ρ((i00-i)) can be diagonalized and hence so can ρ(H)=-iρ((i00-i)).

In fact, the classification of irreducible representations of U(1) shows that ρ(iH) has eigenvalues in i and so ρ(H) has eigenvalues in . ∎

Remark 3.3.

The proof of the proposition is a instance of Weyl’s unitary trick. We turned the action of H, which infinitesimally generates a non-compact one-parameter subgroup of SL2(), into the action of the compact group U(1) infinitesimally generated by iH. The action of this compact subgroup can be diagonalized.

The proposition does not hold for an arbitrary representation of the one-dimensional Lie algebra 𝔥=H generated by H. Namely, the map zH(1z01) cannot be diagonalized. It is implicitly the interaction of H with the other generators X and Y which makes the proposition work.

Let (ρ,V) be a finite-dimensional complex-linear representation of 𝔰𝔩2,. By Proposition 3.2 we get a decomposition

V=αVα (3.1)

where each Vα is the eigenspace for ρ(H) with eigenvalue α:

Vα={vV:ρ(H)v=αv}.
Definition 3.4.
  1. 1.

    Each α occurring in equation (3.1) is called a weight (more precisely, an H-weight) for the representation ρ.

  2. 2.

    Each Vα is called a weight space for ρ.

  3. 3.

    The nonzero vectors in Vα are called weight vectors for ρ.

Example 3.5.

The set of weights of the zero representation is empty, while the trivial representation has a single weight, 0.

Example 3.6.

Let V=2 be the standard representation. Write e1,e-1 for the standard basis. Then He1=e1 and He-1=-e-1. Thus the set of weights of V is {±1}.

Example 3.7.

We consider the adjoint representation ad of 𝔤=𝔰𝔩2, on itself. By the commutation relations, we see directly that adH has eigenvalues 0 ([H,H]=0), 2 ([H,X]=2X), and -2 ([H,Y]=-2Y), so the set of weights is

{-2,0,2}.

The non-zero weights 2 and -2 are called the roots of 𝔰𝔩2, and their weight spaces are the root spaces 𝔤2 and 𝔤-2. The weight vectors are called root vectors.

Thus we have the root space decomposition

𝔰𝔩2, =𝔤0𝔤2𝔤-2
=HXY.
Example 3.8.

We consider 22 where 2 is the standard representation. Then

H(e1e1)=(He1)e1+e1He1=2e1e1

and similarly

H(e1e-1)=H(e-1e1)=0,H(e-1e-1)=-2e-1e-1

so that the weights are {-2,0,0,2}. Note that this is a multiset — a set with repeated elements — and we say that the weight 0 has ‘multiplicity two’ (in general, the multiplicity of a weight is the dimension of the weight space).

Example 3.9.

Take V=Symk(2). A set of basis vectors is

{e1ae-1b:a,b0,a+b=k}.

We calculate

H(e1ae-1b) =a(He1)e1a-1e-1b+b(He-1)e1ae-1b-1
=(a-b)e1ae-1b.

Thus the weights are (writing b=k-a):

{-k,2-k,4-k,,k-4,k-2,k}.

We will soon see an explanation for this pattern.

3.1.1 Highest weights

The following is our first version of the fundamental weight calculation.

Lemma 3.10.

Let (ρ,V) be a complex-linear representation of 𝔰𝔩2,. Let α be a weight of V and let vVα. Then

X(v)Vα+2

and

Y(v)Vα-2.

Thus we have three maps:

H :VαVα
X :VαVα+2
Y :VαVα-2.
Proof.

We have, for vVα,

ρ(H)ρ(X)v =[ρ(H),ρ(X)]v+ρ(X)ρ(H)v
=ρ([H,X])v+ρ(X)ρ(H)v
=2ρ(X)v+αρ(X)v
=(α+2)ρ(X)v.

So ρ(X)vVα+2 as required.

The claim about the action of Y is proved similarly. ∎

Definition 3.11.

A vector vVα is a highest weight vector if it is a weight vector and if

Xv=0.

In this case we call the weight of v a highest weight.

Lemma 3.12.

Any finite-dimensional complex linear representation V of 𝔰𝔩2, has a highest weight vector.

Proof.

Indeed, let α be the numerically greatest weight of V (there must be one, as V is finite-dimensional) and let v be a weight vector of weight α. Then Xv has weight α+2 by the fundamental weight calculation, so must be zero as α was maximal. ∎

Example 3.13.

Let V=22. Then the highest weight vectors are e1e1 and e1e-1-e-1e1.

These are easily checked to be highest weight vectors — the first is killed by X since Xe1=0, the second becomes

e1e1-e1e1=0.

It is left to you to check that there are no further highest weight vectors.

3.2 Classification of representations of sl2,C.

The key point here is that a highest weight vector must have non-negative integer weight n, and it then generates an irreducible representation of dimension n+1 whose isomorphism class is determined by n and which has a very natural basis of weight vectors.

Let (ρ,V) be a complex-linear representation of 𝔰𝔩2,.

Lemma 3.14.

Suppose that V has a highest weight vector v of weight n. Then the subspace W spanned by the vectors

v,Y(v),Y2(v)=Y(Y(v)),

is an 𝔰𝔩2,-invariant subspace of V.

Moreover, n0, the dimension of W is n+1, v,Y(v),,Yn(v) are a basis for W, with Yn+1(v)=0.

Proof.

Let W be the span of the Yk(v). Since the Yk(v) are weight vectors, their span is H-invariant. It is clearly also Y-invariant. So we only need to check the invariance under the X-action. I claim that for all m1,

XYm(v)=m(n-m+1)Ym-1(v). (3.2)

The proof is by induction. The case m=1 is:

XY(v)=([X,Y]+YX)v=Hv+Y(Xv)=Hv=nv.

If the formula holds for m, then

XYm+1(v) =([X,Y]+YX)Ym(v)
=HYmv+Y(XYmv)
=(n-2m)Ymv+m(n-m+1)Ymv (induction hypothesis)
=(m+1)(n-m)Ymv

as required.

If n is not a nonnegative integer, then m(n-m+1)0 for all m. So

Ym-1v0XYmv0Ymv0,

whence Ymv0 for all m. As these are weight vectors with distinct weights, they are linearly independent and so span an infinite dimensional subspace.

If Yiv=0 for some 0<in, then

0=XYiv=i(n-i+1)Yi-1v

and so Yi-1v=0. Repeating gives that v=Y0v=0, a contradiction.

Now,

XYn+1(v)=(n+1)(n-n)v=0

and so Yn+1v is either zero or a highest weight vector of weight -(n+2)<0. We have already seen that the second possibility cannot happen, so Yn+1v=0.

Thus W is spanned by the (nonzero) weight vectors v,Yv,,Ynv with distinct weights, which are therefore linearly independent and so a basis for W. ∎

Remark 3.15.

If we didn’t assume that V was finite-dimensional, then the first part of the previous lemma would still be true.

Corollary 3.16.

In the situation of the previous lemma, W is irreducible.

Proof.

Suppose that WW is a nonzero subrepresentation. Then it has a highest weight vector, which must be (proportional to) Yiv for some 0in. But then

XYiv=i(n-i+1)Yi-1v0

if i>0 and so i=0, meaning that vW. But then YivW for all i, so W=W as required. ∎

The weights of W are illustrated in Figure 3.2.

Theorem 3.17.

Suppose V is an irreducible finite-dimensional complex-linear representation of 𝔰𝔩2,. Then:

  1. 1.

    There is a unique (up to scalar) highest weight vector, with highest weight n0.

  2. 2.

    The weights of V are n,n-2,,2-n,-n.

  3. 3.

    All weight spaces Vα are one-dimensional (we say that V is ‘multiplicity free’).

  4. 4.

    The dimension of V is n+1.

For every n0, there is a unique irreducible complex-linear representation of 𝔰𝔩2, (up to isomorphism) with highest weight n.

Proof.

Let V be an irreducible finite-dimensional complex-linear representation of 𝔰𝔩2,. Let vV be a highest weight vector. By Lemma 3.14 its weight is a nonnegative integer n, and by Corollary 3.16 the vectors v,Yv,,Ynv span an irreducible subrepresentation of V of dimension n+1, which must therefore be the whole of V. The claims all follow immediately.

Moreover, the actions of X, Y and H on this basis are given by explicit matrices depending only on n, so the isomorphism class of V is determined by n.

It remains only to show that a representation with highest weight n exists for all n0. Consider Symn(2). The vector e1n is a highest weight vector of weight n, so we are done. In fact, in this case Yae1n is proportional to e1n-ae-1a, so the irreducible representation generated by e1n is in fact the whole of Symn(2). ∎

In fact, if (ρ,V) is the irreducible representation with highest weight n, highest weight vector v, then the matrices of H, Y, and X with respect to the basis

{v,Y(v),,Yn(v)}

are respectively

ρ(H) =(nn-2-n),
ρ(Y) =(01010),
and
ρ(X) =(0n02(n-1)03(n-2)n0).
Theorem 3.18.

Every finite-dimensional irreducible complex-linear representation of SL2() or 𝔰𝔩2, is isomorphic to Symn(2), the symmetric power of the standard representation.

Proof.

We already proved this for 𝔰𝔩2,. If V is a representation of SL2(), then its derivative will be isomorphic to Symn(2). Since SL2() is connected, this implies that VSymn(2). ∎

3.3 Decomposing representations

Theorem 3.19.

Let V be any finite-dimensional complex-linear representation of 𝔰𝔩2,. Then V is completely reducible, that is, splits into a direct sum of irreducible representations.

Proof.

We will prove this later (Theorem 3.30) using the compact Lie group SU(2). ∎

It is easy to decompose a representation V of 𝔰𝔩2, into irreducibles by looking at the weights. Firstly, look at the maximal weight k of V. Then there must be a weight vector v of weight k, which is necessarily a highest weight vector, and so V must contain a copy of Symk(2) — namely, the subspace v,Yv,,Ykv. By complete reducibility we have

VSymk(2)W.

The weights of W are then obtained by removing the weights of Symk(2) from the weights of V, and we repeat the process.

In particular, this shows that a finite-dimensional representation of 𝔰𝔩2, is determined, up to isomorphism, by its multiset of weights.

Proposition 3.20.

If V, W are representations of 𝔰𝔩2, then:

  • {weights of VW}={weights of V}+{weights of W}.

  • {weights of Symk(V)}={sums of unordered k-tuples of weights of V}.

  • {weights of Λk(V)}={sums of unordered ‘distinct’ k-tuples of weights of V}.

Proof.

Omitted: try it yourself! For a similar result, see section 4.5 below. ∎

Example 3.21.

We should illustrate what is meant by ‘distinct’: it is ‘distinct’ as elements of the multiset. Suppose that the weights of V are {2,0,0,-2}. Then to obtain the weights of Λ2(V) we add together unordered, distinct, pairs of these in every possible way, getting:

{2+0,2+0,2+-2,0+0,0+-2,0+-2}={2,2,0,0,-2,-2}.
Example 3.22.

Let 2 be the standard representation of SL2() with weight basis e1, e-1. Consider V=Sym2(2)Sym2(2). Let v2=e12, v0=e1e-1 and v-2=e-12 be weight vectors in Sym2(2) corresponding to the weights 2, 0 and -2. Then the weights of V are

  • 4, multiplicity one, weight vector: v2v2.

  • 2, multiplicity two, weight space: v2v0,v0v2.

  • 0, multiplicity three, weight space: v2v-2,v0v0,v-2v2.

  • -2, multiplicity two, weight space: v0v-2,v-2v0.

  • -4, multiplicity one, weight vector: v-2v-2.

Decomposing
Figure 4: Decomposing Sym2(2)Sym2(2).

The weights of Sym2(2) are {-2,0,2} and so the weights of VV are

{-2,0,2}+{-2,0,2}={-4,-2,-2,0,0,0,2,2,4}.

This is the same as the set of weights of

Sym4(2)Sym2(2)

and so this is the required decomposition into irreducibles.

We can go further, and decompose V into irreducible *sub*representations. This means finding irreducible subrepresentations of V such that V is their direct sum.

The copy of Sym4(2) in V has highest weight vector v2v2. We can find a basis by repeatedly hitting this with Y (writing for ‘equal up to a nonzero scalar’):

Y(v2v2) =2(v2v0+v0v2)v2v0+v0v2
Y2(v2v2) v2v-2+4v0v0+v-2v2
Y3(v2v2) 6(v0v-2+v-2v0)v0v-2+v-2v0
Y4(v2v2) v-2v-2.

These vectors are a basis for the copy of Sym4(2) in V.

Next, we find the copy of Sym2(2) in V. We start by looking for a highest weight vector of weight 2:

v2v0-v0v2

does the trick. Hitting this with Y gives v2v-2-v-2v2, and doing so again gives v-2v0-v0v-2 (up to scalar). These vectors are a basis for the copy of Sym2(2) in V.

Finally, we find the trivial representation in V. We need only find a weight vector of weight 0 which is killed by X, and

v2v-2-2v0v0+v-2v2

does the job: this vector spans a copy of the trivial representation.

3.4 Real forms and complete decomposability

We use our understanding of the representation theory of 𝔰𝔩2, into an understanding of the representation theory of SL2() and SU(2), and prove complete reducibility.

Definition 3.23.

A real form of a complex Lie algebra 𝔤 is a real Lie algebra 𝔥𝔤 such that every element Z of 𝔤 can be written uniquely as X+iY for X,Y𝔥.

Remark 3.24.

In lectures, we didn’t give this definition, just doing the special case of 𝔰𝔲n𝔰𝔩n,.

For dimension reasons, necessary and sufficient conditions are that 𝔥i𝔥=0 and dim𝔥=dim𝔤.

Example 3.25.

The Lie algebra 𝔰𝔩n, is a real form of 𝔰𝔩n,.

Example 3.26.

The Lie algebra 𝔰𝔲n is a real form of 𝔰𝔩n,. Indeed, dim𝔰𝔲n=n2-1=dim𝔰𝔩n, and if X,iX𝔰𝔲n then

iX=-(iX)=iX=-iX

so X=0.

Explicitly, we may write A𝔰𝔩n, as X+iY with

X=12(A-A)

and

Y=-i2(A+A)

in 𝔰𝔲n.

Example 3.27.

Recall that

𝔰𝔬n={A:n,A+AT=0}.

Let

𝔰𝔬n,={A:n,A+AT=0}.

Since the defining equation A+AT=0 can be checked separately on the real and imaginary parts of A,

𝔰𝔬n,={B+iC:B,C𝔰𝔬n}

and so 𝔰𝔬n is a real form of 𝔰𝔬n,.

Proposition 3.28.

Let 𝔥 be a real form of 𝔤.

There is a one-to-one correspondence between representations of 𝔥 and complex-linear representations of 𝔤 under which irreducible representations correspond to irreducible representations.

Proof.

Given a -linear representation of 𝔤, it is a representation of 𝔥 by restriction. Conversely, if (ρ,V) is a representation of 𝔥, then it extends to a unique -linear representation of 𝔤 given by the formula (forced by -linearity)

ρ(X+iY)v=ρ(X)v+iρ(Y)v.

It is easy to see that this preserves the Lie bracket. The proposition follows (the final statement is left as an exercise). ∎

As a corollary we immediately obtain

Theorem 3.29.

The representation theories of 𝔰𝔩n, and 𝔰𝔲n are ‘the same’ as the complex-linear representation theory of 𝔰𝔩n,. All finite-dimensional irreducible representations of 𝔰𝔩2,R, SL2(), 𝔰𝔲2, or SU(2) are of the form Symn(2).

Proof.

The claims about Lie algebras follow from the above discussion. Every (finite-dimensional) irreducible representation of 𝔰𝔩2, is of the form Symn(2), and these clearly exponentiate to representations of SL2(), despite this not being a simply connected group! Similarly for SU(2) (which is simply connected). Since SL2() and SU(2) are connected, every representation of them is determined by its derivative, so we have a complete list of the irreducible representations. ∎

Theorem 3.30.

Every finite-dimensional complex-linear representation of 𝔰𝔩n, is completely reducible.

Proof.

Let V be a finite-dimensional complex-linear representation of 𝔰𝔩n, and let WV be a subrepresentation. Then W is an 𝔰𝔲n-subrepresentation. As SU(n) is simply-connected, V and W exponentiate to representations of SU(n). Since SU(n) is compact, by Maschke’s theorem there is a complementary SU(n)-subrepresentation W with

V=WW.

Then W is a 𝔰𝔲n-subrepresentation, and so a -linear 𝔰𝔩n,-subrepresentation, so that

V=WW

as representations of 𝔰𝔩n,. Complete reducibility follows. ∎

The argument in this theorem is called Weyl’s unitary trick. For a similar application of this idea, see the proof of Proposition 3.2.

3.5 SO(3) (non-examinable)

We skipped this section, due to the strike, so the material is non-examinable.

3.5.1 Classification of irreducible representations

We have already (see the first problem set this term) seen that 𝔰𝔲2 and 𝔰𝔬3 are isomorphic. We therefore have:

Theorem 3.31.

There is an irreducible two-dimensional representation V of 𝔰𝔬3 (coming from 𝔰𝔬3𝔰𝔲2𝔤𝔩2,) such that the irreducible complex representations of 𝔰𝔬3 are exactly

Symn(V)

for n0.

We want to know which of these representations exponentiate to an irreducible representation of SO(3). For this, we revisit the connection with SU(2).

Let , be the bilinear form on 𝔰𝔲2 defined by

X,Y=-tr(XY).
Lemma 3.32.

The form , is a positive definite bilinear form preserved by the adjoint action of SU(2).

Proof.

The bilinearity is clear. To see that it is positive definite, if X𝔰𝔲2 then

X,X=-tr(X2)=tr(XX).

Direct calculation shows tr(XX) is the sum of the squares of the entries of X, which is strictly positive for nonzero X.

We have, for gSU(2),

AdgX,AdgY=-tr(gXg-1gYg-1)=-tr(gXYg-1)=-tr(XY)

which shows that this inner product is SU(2)-invariant. ∎

Corollary 3.33.

There is an isomorphism

SU(2)/{±I}SO(3).
Proof.

Choosing an orthonormal basis for 𝔰𝔲2 with respect to , identifies the set of linear maps 𝔰𝔲2𝔰𝔲2 preserving the inner product with SO(3). We therefore have a homomorphism

SU(2)SO(3)

whose kernel is {gSU(2):Adg=Ad}={±I}. The derivative of this homomorphism is injective, otherwise there would be a one-parameter subgroup in the kernel of the group homomorphism, and since the Lie algebras have the same dimension we get an isomorphism of Lie algebras. The group homomorphism is therefore surjective, since exponentials of elements of 𝔰𝔬3 generate the group SO(3). ∎

Remark 3.34.

Let ,𝒥,𝒦 be the elements of 𝔰𝔲2 given by

12(01-10),12(0ii0),12(i00-i)

respectively. Up to scalar, these are an orthonormal basis for 𝔰𝔲2. Since the derivative of the above homomorphism SU(2)SO(3) is the adjoint map, computed with respect to this basis, we see that the induced isomorphism 𝔰𝔲2𝔰𝔬3 takes:

Jx =(00000-1010)
𝒥Jy =(001000-100)
𝒦Jz =(0-10100000).

It will be useful to know where this isomorphism takes the elements H,X,Y of 𝔰𝔩2,=𝔰𝔲2,. For example, as H=-2i𝒦, we see that it goes to -2iJz. Or, for the lowering operator Y, we have

Y=(0010)=-(+i𝒥)-(Jx+iJy)

and similarly XJx-iJy.

Corollary 3.35.

For each 0, there is a unique irreducible representation V() of SO(3) of dimension 2+1.

The derivative of this representation is isomorphic to the representation Sym2(V) of 𝔰𝔬3.

This gives the complete list of irreducible representations of SO(3) up to isomorphism.

Proof.

We simply have to work out which representations Symk(V) of 𝔰𝔬3 exponentiate to a representation of SO(3). Since we have

SU(2)/{±I}SO(3)

and each Symk(V) exponentiates to a unique representation of SU(2) — which we also call Symk(V) — this is equivalent to asking for which k the centre {±I} of SU(2) acts trivially on Symk(V). But we see that -I acts as (-1)k, so the answer is: for even k only.

Thus Symk(V) exponentiates to a representation of SO(3) if, and only if, k=2 is even, and we obtain the result. ∎

We can consider the weights of these representations. Under the isomorphism

𝔰𝔲2,=𝔰𝔩2,𝔰𝔬3,

the element H=-2i𝒦 maps to -2iJz. Since the H-weights of V() are -2,-2(-1),,2(-1),2, we must divide these by -2i to find the weights of Jz acting on V():

-i,-i(-1),,i(-1),i.
Example 3.36.

Consider the standard three-dimensional representation of SO(3) on 3. The weights of Jz are simply its eigenvalues as a 3×3 matrix, which are -i,0,i. We see that this representation is isomorphic to V(1).

3.5.2 Harmonic functions

We can use our understanding of the representation theory of SO(3) to shed light on the classical theory of spherical harmonics.

We let 𝒫 be the subspace of [x,y,z] consisting of homogeneous polynomials of degree . This has an action of SO(3) given by

(gf)(𝐱)=f(gT𝐱)

where 𝐱=(xyz) is a vector in 3. We therefore get a representation of SO(3).

Lemma 3.37.

The elements Jx,Jy, and Jz of SO(3) act on 𝒫 according to the following formulae:

Jx =zy-yz
Jy =xz-zx
Jz =yx-xy.
Proof.

Exercise. In fact, prove that the action of A=(aij)𝔰𝔬3 is given by

i,jaijxixj

by considering

ddtf(exp(tAT)𝐱)=ddtf((I+tAT)𝐱)

at t=0 (where we rewrite x,y,z as x1,x2,x3). ∎

It will be useful in what follows to recall

Lemma 3.38.

(Euler’s formula) If f is a homogeneous polynomial of degree , then

xxf+yyf+zzf=f.

The representation 𝒫 is not irreducible. Let r2[x,y,z] be the polynomial

r2=x2+y2+z2.

Note that r2 is clearly invariant under the action of SO(3).

Lemma 3.39.

The map 𝒫𝒫+2 defined by

fr2f

is an injective homomorphism of SO(3)-representations.

Proof.

We have, for gSO(3),

g(r2f)=g(r2)g(f)=r2g(f)

as required. ∎

Next, we consider the Laplace operator:

Δf=2x2+2y2+2z2.

This is map from 𝒫𝒫-2.

Lemma 3.40.

The map Δ is a map of SO(3) representations.

Proof.

We must show that, for g=(gij)SO(3),

(Δf)(gT𝐱)=Δ(gf)(𝐱)).

We have

xi(gf)(x)=jgijfxj(gTx)

and so

xixi(gf)(x)=j,kgijgikxkxj(gTx).

We sum over i, for fixed j and k:

igijgik=δjk

since g is orthogonal. We therefore obtain

i2xi2(gf)(x)=j2xj2(f)(gTx),

and therefore

Δ(gf)=gΔ(f).

An element f𝒫 is harmonic if Δf=0. Since dim(𝒫)>dim(𝒫-2), harmonic polynomials must exist. We write 𝒫 for the space of harmonic polynomials.

Lemma 3.41.

On 𝒫, we have

r2Δ=Jx2+Jy2+Jz2+2+.
Proof.

Left as an exercise. ∎

It follows that r2Δ preserves each irreducible subrepresentation of 𝒫 (since Jx,Jy,Jz do). Furthermore, by Schur’s lemma it must act on each irreducible subrepresentation as a scalar. We determine that scalar.

Lemma 3.42.

Suppose that V𝒫 is an irreducible subrepresentation with highest weight ik. Then

(r2Δ)(f)=(2+-k2-k)f=(-k)(+k+1)f.
Proof.

Since r2Δ is a 𝔰𝔬3-homomorphism and V is irreducible, by Schur’s lemma it acts as a scalar on V. It therefore suffices to compute the action on a highest weight vector vV. So

Jzv=ikv,(Jx-iJy)v=0.

It follows that Jz2v=-k2v and, as

Jx2+Jy2 =(Jx+iJy)(Jx-iJy)+i[Jx,Jy]
=(Jx+iJy)(Jx-iJy)+iJz,

we have (Jx2+Jy2)v=0v-kv.

Applying the previous lemma gives the result. ∎

Theorem 3.43.

For every 0,

𝒫=r2𝒫-2.

The space is the irreducible highest weight representation of SO(3) of dimension 2+1, and the space 𝒫, as an SO(3)-representation, is the direct sum of representations of weights i,i(-2),, each occurring with multiplicity one.

Proof.

We use induction on . The case =0 is clear (we just have the trivial representation). Suppose true for -1 with 1.

By the previous lemma, the space is the sum of all the copies inside 𝒫 of the irreducible representation with highest weight i. Since 𝒫-2 does not contain this irreducible representation, by the inductive hypothesis, we have

r2𝒫-2={0}.

Since, 0 as already discussed, its dimension is a positive multiple of 2+1. However, its dimension is at most

dim𝒫-dim𝒫-2=(+22)-(2)=2+1.

It follows that is irreducible, and that we have

r2𝒫-2=𝒫.

The statement about the decomposition into irreducibles follows. ∎

The proof of this theorem shows that

𝒫=r2-2r4𝒫-4

so that every polynomial has a unique decomposition as a sum of harmonic polynomials multiplied by powers of r2.

We can go further and give nice bases for the by taking weight vectors for Jz. First, we have

Lemma 3.44.

The function (x-iy)𝒫 is a highest weight vector of weight i.

Proof.

Exercise! ∎

We then obtain a weight basis by repeatedly applying the lowering operator

Jx+iJy=(ix-y)z+z(y-ix).

The functions thus obtained are known as ’spherical harmonics’ (at least, up to normalization), and give a particularly nice basis for the space of functions on the sphere S23. The decomposition of a function into spherical harmonics is analogous to the Fourier decomposition of a function on the unit circle.

Example 3.45.

If =1, then 𝒫=, and the weight vectors are

x+iy,z,x-iy.

If =2, a basis of made up of weight vectors is

(x-iy)2,z(x-iy),x2+y2-2z2,z(x+iy),(x+iy)2.