Representation Theory IV

2 Representations of Lie groups and Lie algebras - generalities

2.1 Basics

Definition 2.1.

A finite-dimensional (complex) representation (ρ,V) of a Lie group G is a Lie group homomorphism

ρ:GGL(V),

where V is a finite-dimensional complex vector space.

Remark 2.2.

Infinite-dimensional representations are important, but subtle. In general one must equip V with some topology and add topological conditions to all notions which follow. For instance, one might take V to be a Hilbert space.

An example of such a representation arises naturally if you attempt to generalise the regular representation! One must take V to be something like the space of square-integrable functions on the group, rather than the space of arbitrary functions, to get a pleasant theory.

If ρ is a finite-dimensional representation of G as above, then we can take its derivative:

Dρ:𝔤𝔤𝔩(V)=End(V),

mapping from the Lie algebra 𝔤 of G to the space of endomorphisms of V. Note that

End(V)

is a Lie algebra with bracket

[S,T]=ST-TS.

The map Dρ is a Lie algebra homomorphism. Often we write, abusively, ρ instead of Dρ.

Note that choosing an isomorphism Vn induces isomorphisms GL(V)GLn() and 𝔤𝔩(V)𝔤𝔩n,.

Definition 2.3.

A (complex) representation (ρ,V) of a Lie algebra 𝔤 is a Lie algebra homomorphism

ρ:𝔤𝔤𝔩(V),

where V is a complex vector space. That is,

  • ρ is -linear;

  • ρ([X,Y])=[ρ(X),ρ(Y)].

Note that by Theorem 1.37 the differential of a Lie group representation is a Lie algebra representation.

Remark 2.4.

Warning! It is not the case that, if ρ is a Lie algebra representation, then

ρ(XY)=ρ(X)ρ(Y).

Indeed, in general XY need not be an element of the Lie algebra at all, and even if it is the displayed equation will not usually hold.

The notions of G-homomorphism (or 𝔤-homomorphism, or intertwiner), isomorphism, subrepresentation, and irreducible representation stay the same as for finite groups. For example, a 𝔤-homomorphism from (ρ,V) to (ρ,V) is a linear map ϕ:VV such that

ϕ(ρ(X)v)=ρ(X)ϕ(v)

for all vV and X𝔤.

Definition 2.5.

A -linear representation of 𝔤 is a complex representation ρ of 𝔤 such that

ρ(λX)=λρ(X)

for all λ,X𝔤.

If G is a complex Lie group, then a holomorphic representation of G is a complex representation whose derivative is -linear; equivalently, the map GGL(V) is holomorphic.

Theorem 2.6.

Let G be a Lie group, 𝔤 be a Lie algebra.

  1. 1.

    If V1 and V2 are irreducible finite-dimensional representations of G or 𝔤, then

    dimHomG/𝔤(V1,V2)={1if V1V20𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒.

    If V1=V2, then any G- or 𝔤-homomorphism T:VV is scalar.

  2. 2.

    Any irreducible finite-dimensional representation of an abelian Lie group or Lie algebra is one-dimensional.

  3. 3.

    If (ρ,V) is an irreducible finite-dimensional representation of G (or 𝔤) and Z (or 𝔷) is the center of G (or 𝔤) then there is a homomorphism χ:Z× (or χ:Z) such that

    ρ(z)v=χ(z)v

    for all zZ (or 𝔷) and vV. We call this the central character.

Proof.

The proofs are all the same as in the finite group case! ∎

Proposition 2.7.

Let (ρ,V) be a finite-dimensional representation of a Lie group G. Let Dρ be its derivative.

  1. 1.

    If WV is invariant under ρ(G), then W is invariant under Dρ(𝔤).

  2. 2.

    If Dρ is irreducible, then ρ is irreducible.

  3. 3.

    If ρ is unitary, that is, there is a basis for V such that ρ(g)U(n) for all gG, then Dρ is skew-Hermitian, that is, Dρ(X)𝔲(n) for all X𝔤 (using the same basis for V).

  4. 4.

    Let (ρ,V) be another finite-dimensional representation of G. If ρρ, then DρDρ.

If G is connected, then the converses to these statements hold.

So, for connected Lie groups, we can test irreducibility and isomorphism at the level of Lie algebras.

Proof.

For (1), we know that ρ(exp(tX))(w)W for any X𝔤 and wW. Taking the derivative at t=0, it follows that Dρ(X)(w)W as required. Part (2) follows from (1).

For (3), if ρ is unitary, then after choosing a basis appropriately it is a Lie group homomorphism ρ:GU(n). The derived homomorphism therefore lands in the Lie algebra 𝔲(n) of U(n).

For part (4), let T be a G-isomorphism, so that in particular,

Tρ1(exp(tX))T-1=ρ2(exp(tX))

for all X𝔤 and t. Taking the derivative at t=0 gives

TDρ1(X)T-1=Dρ2(X)

so that T is a 𝔤-isomorphism as required.

If G is connected, then G is generated by exp(𝔤). Hence all proofs above can be reversed. For example, for (1), suppose that W is preserved by Dρ(𝔤). If wW and X𝔤, then

ρ(exp(X))w=exp(Dρ(X))w=n=0(Dρ(X))nn!wW

as W is preserved by Dρ(X) and also closed. Since every gG can be written as a finite product of exp(Xi) for Xi𝔤, we see that W is preserved by ρ(G) as required. ∎

2.2 Standard constructions for representations

We give a list of various constructions with representations of Lie groups, and the analogous constructions for their derivatives.

The standard representation of a linear Lie group GGLn() comes from its action on n:

ρ(g) =g
Dρ(X) =X.

The direct sum of representations (ρ1,V1), (ρ2,V2) is (ρ1ρ2,V1V2) with derivative

D(ρ1ρ2)=Dρ1Dρ2.

The determinant representation of GGLn() is det:G* which sends g to det(g). We have

Ddet(X)=tr(X),

which follows from detexp(tX)=ettr(X).

If (ρ,V) is a representation of G, the dual representation (ρ*,V*) of (ρ,V) is defined by

(ρ*(g)(λ))(v)=λ(ρ(g-1)(v)),

for λV* a linear functional on V and gG. It has derivative

Dρ*(X)(λ)(v)=-λ(Dρ(X)v).

Given a basis of V, then the matrix of ρ* with respect to the dual basis is

ρ*(g)=ρ(g)-T,

which differentiates to

(Dρ*)(X)=-Dρ(X)T.

If (ρ1,V1) and (ρ2,V2) are representations of G, then as before the tensor product representation ρ1ρ2 is the representation on V1V2 defined by

(ρ1ρ2)(g)=ρ1(g)ρ2(g).

Then using the product rule one sees

D(ρ1ρ2)=Dρ1idV2+idV1Dρ2.

The symmetric square and alternating square are also as for finite groups. If (ρ,V) is a representation of G then Sym2(V) has a representation Sym2ρ:

Sym2ρ(g)(v1v2)=(ρ(g)(v1))(ρ(g)(v2)),

and

DSym2ρ(X)(v1v2)=Dρ(X)(v1)v2+v1Dρ(X)(v2).

Similarly we have a representation Λ2ρ on Λ2(V):

Λ2ρ(g)(v1v2)=ρ(g)(v1)ρ(g)(v2),
DΛ2ρ(X)(v1v2)=Dρ(X)(v1)v2+v1Dρ(X)(v2).

We can take tensor/symmetric/alternating products of more than one factor. Suppose (ρi,Vi) are representations of G.

  • We form the tensor product

    V1V2Vl.

    It is generated by symbols v1vl subject to the multilinear relations, that is, linearity in each slot:

    v1(avi+bvi)vl=a(v1vivl)+b(v1vivl).

    One has

    dim(V1V2Vl)=i=1ldimVi.

    The action of G is as before: for gG, viVi,

    g(v1vl)=(gv1)(gvl).

    The derivative is, for X𝔤 and viVi,

    X(v1vl)= (Xv1)v2vl
    +v1(Xv2)vl
    +
    +v1v2(Xvl).

    We also write

    Vl=VV.
  • The lth symmetric power is the space Syml(V) generated by symbols v1vl with linearity in each slot and any permutation of the vectors giving the same element. We have

    dimSyml(V)=(n+l-1l)=(n+l-1n-1)

    where dimV=n. Indeed, if e1,,en is a basis for V then a basis for Syml(V) is

    {ei1eil:1i1i2iln}

    from which finding the dimension is a simple counting problem.

    As for higher tensor powers, the actions of G and 𝔤 are

    g(v1vl)=(gv1)(gvl),

    and

    X(v1vl)=(Xv1)v2vl++v1vl-1(Xvl).
  • The lth alternating power is the space Λl(V) generated by symbols v1vl with linearity in each slot and having the alternating property: for any permutation σSl, we have

    vσ(1)vσ(l)=ϵ(σ)(v1vl).

    In particular, switching the places of two components reverses the sign, while v1vl=0 if two of the vectors coincide (more generally, if they are linearly dependent).

    We have

    dimΛlV=(nl)

    where dimV=n. Indeed, if e1,,en are a basis for V then a basis for Λl(V) is

    {ei1eil:1i1<i2<<iln}

    from which finding the dimension is a simple counting problem. In particular, Λnn is one-dimensional generated by e1en.

    The representation on Λl(V) is given again as above: for gG,

    g(v1vl)=(gv1)(gvl),

    while for X𝔤

    X(v1vl)=(Xv1)v2vl++v1vl-1(Xvl).

To give an example of how to justify the claims about derivatives, we do the case of tensor products. Suppose V, W are vector spaces acted on by G. Let X𝔤, vV, wW. We must compute

ddtexp(tX)vexp(tX)w|t=0.

Expanding:

exp(tX)vexp(tX)w =(v+tXv+O(t2))(w+tXw+O(t2))
=vw+t(Xvw+vXw)+O(t2).

This gives

X(vw) =ddtexp(tX)vexp(tX)w|t=0
=Xvw+vXw

as required.

Remark 2.8.

We defined tensor products (and so on) of representations of Lie groups and then differentiated them. We could also directly make these definitions with Lie algebras. For instance, if 𝔤 is a Lie algebra and V is a representation of 𝔤, we define the symmetric square representation on Sym2(V) by

X(vw)=(Xv)w+v(Xw).

2.2.1 Functional constructions

We can construct representations as vector spaces of functions on topological spaces with actions of G. If G acts on a set X, then it also acts on the vector space of functions X by (gf)(x)=f(g-1x). Usually this will be infinite dimensional, and so out of the scope of our course, but sometimes we can impose conditions allowing us to handle it. For example, GLn() acts on n, and hence on the space of polynomial functions in n variables. Imposing a further restriction — to homogeneous polynomials of some fixed degree — gives a finite-dimensional representation. The derivative must be calculated on a case-by-case basis. We will see examples of this on the problem sheet.

2.3 The adjoint representation

Let G be a linear Lie group and 𝔤 be its Lie algebra. Then the adjoint representation Ad of G is the action on 𝔤 by conjugation. We usually write Adg instead of Ad(g), so that

Adg(Y)=gYg-1

for gG and Y𝔤. By Proposition 1.24 (2), gYg-1 is indeed in 𝔤. Thus the map gAdg is a Lie group homomorphism:

Ad:GGL(𝔤).
Remark 2.9.

An alternative definition, that works for general Lie groups, is to consider the conjugation by g map hghg-1 and then take the derivative:

Adg(Y)=ddtgexp(tY)g-1|t=0.

The derivative of Ad, denoted by ad, is called the adjoint representation of the Lie algebra 𝔤. Thus

ad=DAd.

Again, we write adX for ad(X), so we have

ad:X𝔤adXEnd(𝔤)=𝔤𝔩(𝔤).

By Theorem 1.37 we have the formula

Adexp(tX)=exp(tadX)

as elements of GL(𝔤).

Theorem 2.10.

Let G and 𝔤 be as above and let X,Y𝔤. Then

  1. 1.
    adX(Y)=[X,Y]=XY-YX.
  2. 2.

    The map ad is a Lie algebra homomorphism, so that

    ad[X,Y]=[adX,adY].

    Thus, for all Z𝔤,

    ad[X,Y](Z)=[adX,adY](Z)=adX(adY(Z))-adY(adX(Z)).
Proof.
  1. 1.

    Since Adexp(tX)(Y)=exp(tX)Yexp(-tX), taking the differential at t=0, we get

    adX(Y)=XY-YX=[X,Y].
  2. 2.

    The map ad is a Lie algebra homomorphism because it is the differential of a Lie group homomorphism.∎

Remark 2.11.

This explains the origin of the Jacobi identity:

[adX,adY](Z) =adX(adY(Z))-adY(adX(Z))
=[X,[Y,Z]]-[Y,[X,Z]]
=[X,[Y,Z]]+[Y,[Z,X]],

while

ad[X,Y](Z) =[[X,Y],Z]
=-[Z,[X,Y]].

Equating these gives the Jacobi identity.

Remark 2.12.

The first formula, adX(Y)=[X,Y], could have been used to define the adjoint representation for any Lie algebra, without reference to Lie groups.

Remark 2.13.

Warning! It is very easy to misinterpret some of the formulas concerning the adjoint representation. For example, ad[X,Y]=[adX,adY] does not mean that

ad[X,Y](Z)=[adX(Z),adY(Z)],

but (as already noted and proved) that

ad[X,Y](Z)=[adX,adY](Z)=adX(adY(Z))-adY(adX(Z)).

Similarly, Adexp(tX)=exp(tadX) does not mean that exp(tX)Yexp(-tX) is equal to exp(t[X,Y]) but rather is the identity

exp(tX)Yexp(-tX) =exp(tadX)(Y)
=k=0tk(adX)kk!(Y)
=k=0[X,[X,,[X,Y]]]k!tk.
Proposition 2.14.

If G is abelian, so is 𝔤. If, moreover, G is connected, then the converse holds.

Proof.

Suppose that G is abelian. Then, for all gG and Y𝔤,

gexp(tY)g-1=exp(tY).

Taking the derivative at t=0 we see that gYg-1=Y. Thus Ad is trivial. Differentiating, we see ad is trivial, so adX=0 for all X. Thus [X,Y]=0 for all X,Y𝔤 as required.

Conversely, suppose G is connected and 𝔤 is abelian. Then ad is trivial and, since G is connected, Ad is trivial. Thus gYg-1=Y for all gG,Y𝔤. Thus gexp(Y)g-1=exp(Y) for all gG and all Y𝔤. Since exp(𝔤) generates G, we see that G is commutative. ∎

2.4 Representations of U(1) and Maschke’s theorem

We now discuss the representation theory of the unitary group U(1)={eit:t}, which is isomorphic to SO(2), the circle group. Its Lie algebra is i with trivial Lie bracket, which is isomorphic to .

Proposition 2.15.

All irreducible finite-dimensional representations of U(1)={z;|z|=1} are one-dimensional. They are given by

z=eiteint=zn

for n.

Proof.

It follows from Schur’s lemma that all irreducible finite-dimensional representations of U(1) are one-dimensional, so are homomorphisms U(1)×. Since U(1) is connected, such a homomorphism is determined by the derivative 𝔲1𝔤𝔩1()=, which has the form itλt for some λ. As in Example 1.52, this exponentiates to a map U(1)× if and only if λ=in for some n, giving the homomorphism zzn. ∎

Theorem 2.16.
  1. 1.

    All representations of U(1) are unitary.

  2. 2.

    All representations of U(1) are completely reducible, that is, decompose into irreducible representations.

Proof.
  1. 1.

    Exercise (see Maschke’s theorem below).

  2. 2.

    This follows from (1) as in the proof of Maschke’s theorem for finite groups. We sketch another proof. So take (ρ,V), a finite dimensional representation of U(1). Consider its differential Dρ:𝔲(1)=i𝔤𝔩n,. Let A=Dρ(2πi). We may write A=D+N for some strictly upper triangular matrix N and diagonal matrix D such that D and N commute. Then

    ρ(exp(A))=ρ(e2πi)=ρ(1)=I.

    On the other hand,

    exp(A)=exp(D+N)=exp(D)exp(N),

    since D and N commute. It follows that exp(D)=I and exp(N)=I (why?), whence N=0 (why?).

Remark 2.17.

Complete irreducibility does not hold for representations of a general Lie group. For example, the standard representation of

N={(1x01):x}

does not decompose into a direct sum of two one-dimensional invariant subspaces (otherwise we would diagonalize (1x01), which is impossible). Furthermore, it is not unitary (as unitary matrices are diagonalizable).

Some of this actually generalizes substantially:

Theorem 2.18.

(Maschke’s theorem for compact Lie groups) Let G be a compact Lie group.

  1. 1.

    Any finite-dimensional representation of G is unitarizable.

  2. 2.

    (complete reducibility) Any finite-dimensional representation of G is a direct sum of irreducible representations.

Proof.

The second part is proved exactly as for finite groups using the averaging technique. Let (ρ,V) be a finite-dimensional representation and let W be a subrepresentation. Let , be the G-invariant Hermitian inner product on V guaranteed by the first part. Then the orthogonal complement W is also a subrepresentation, and V=WW. Iterating, we obtain that V is a direct sum of irreducible representations.

The proof of the first part also uses the same idea as for finite groups. Take (,) to be any Hermitian inner product on V. Then define

v,w=gG(gv,gw)𝑑g.

This is also a Hermitian inner product, and

hv,hw =gG(ghv,ghw)𝑑g
=kG(kv,kw)d(kh-1) (putting k=gh)
=kG(kv,kw)𝑑k (since dk=d(kh-1))
=v,w.

The challenge here is to show that there is an appropriate notion of gGf(g)𝑑g for which the step “dk=d(kh-1)” is valid — this goes by the name of ‘existence of Haar measure’. For U(1) you can do it by hand, see problem 18. ∎