Representations of Compact Groups (Part 1)

I've written a previous post on representation theory for finite groups. The representation theory of finite groups is very nice, but many of the groups whose representations we care about are not finite. For example, representations of SU(2)SU(2) are important for understanding the behavior of particles with nonzero spin. So we want to extend representation theory to more general groups. A nice family of groups to consider are compact groups. In many ways, compactness is a generalization of finiteness. To use an example from that link, every real-valued function on a finite set is bounded and attains its maximum. This is untrue for real-valued functions on infinite sets: consider the functions f(x)=tanxf(x) = \tan x and f(x)=xf(x) = x respectively on the interval (0,π2)(0, \frac\pi 2). However, continuous real-valued functions on a compact interval must be bounded and must attain their maxima. Similarly, compact groups generalize finite groups, and many of the nice features of the representation theory of finite groups extend to the representation theory of compact groups. This post will mostly follow the notes about representations of compact groups available here

Compact Topological Groups

First, we will start with some nice properties of compact topological groups. Recall that a topological group is a group endowed with a topology so that multiplication and the inverse map are continuous. When studying the representation theory of finite groups, it was often convenient to sum over the elements of the group (e.g. to define our inner product on the space of characters). Clearly we cannot always sum over the elements of an infinite group. But for compact groups, we have a nice theorem that tells us that we can integrate over the group instead, which is just as good.

(Haar measure) Let GG be a locally compact topological group. There exists a non-zero left-GG-invariant measure on GG. This measure is nonvanishing and unique up to positive scalar multiplication

This theorem is tricky to prove for locally compact topological groups. But for Lie groups, it is fairly easy. So we will just show a version of the theorem for Lie groups.

Let GG be a Lie group. Then GG has a left-GG-invariant measure

First, we will show existence. Let nn denote the dimension of GG. Recall that as long as GG is oriented, an nn-form on GG induces a measure. Furthermore, we recall that if we can find a nonvanishing nn-form on GG, then GG must be orientable. So it is sufficient to find a left-invariant nonvanishing nn-form on GG. Let ΛnTeG\Lambda^nT_eG denote the space of kk-covectors on the tangent space to the identity of GG. Pick any nonzero ωeΛnTeG\omega_e \in \Lambda^nT_eG. Now, we can extend ωe\omega_e to a differential form on GG. Let LgL_g denote the automorphism of gg given by left-multiplication by gg. This is continuous. Lg1L_{g^{-1}} sends gg to ee, so we can pull ωe\omega_e back along this map to define ωg=(Lg1)ΛnTgG\omega_g = (L_{g^{-1}})^* \in \Lambda^n T_gG. This defines a differential nn-form ωΩn(G)\omega \in \Omega^n(G) on all of GG. ω\omega is left-invariant by construction. ((Lh)ω)g=(Lh)ωhg=(Lh)(L(hg)1)ωe=(L(hg1)h)ωe=(Lg1)ωe=ωg((L_h)^* \omega)_g = (L_h)^* \omega_{hg} = (L_h)^* (L_{(hg)^{-1}})^* \omega_e = (L_{(hg^{-1})h})^* \omega_e =(L_{g^{-1}})^* \omega_e = \omega_g Clearly this differential form is nonvanishing. And by negating ω\omega if necessary, we see that ω\omega is positive with respect to GG's orientation, so it defines a left-invariant measure on GG.

Now, you might be wondering why it is important that GG is compact, because the above theorems don't require compactness. The nice thing about compactness is that measures only let us integrate functions with compact support - but if GG is compact, then every function has compact support. So we can integrate any real- (or complex-) valued functions on GG. We will write the integral of ff with respect to the Haar measure as Gf(g)dg\int_G f(g)\;dg.

In particular, we can integrate the constant function f(g)=1f(g) = 1 over compact groups. It is convenient to normalize our Haar measure so that G1dg=1\int_G 1 \;dg = 1. I will assume that all Haar measures are normalized in this way.

From now on, I'll assume that all groups are compact Lie groups unless I explicitly state otherwise.

Basic Definitions

We'll start with a whole bunch of definitions. They're essentially the same as the analogous definitions for finite groups, except we require that our maps are continuous. To do so, we have to put topologies on the vector spaces involved.

A topological vector space is a vector space endowed with a Hausdorff topology such that addition and scalar multiplication are continuous.
It turns out that a finite dimensional vector space has a unique topology which turns it into a topological vector space. Rn\R^n naturally has a product topology from the standard topology on Rn\R^n, and any linear isomorphism VRnV \cong \R^n lets us transfer this topology to VV.
Suppose VV has an inner product ,\inrp \cdot \cdot. This gives us a norm v=v,v\|v\| = \sqrt{\inrp vv}, which in turn gives us a metric d(v,w)=vwd(v,w) = \|v-w\|, and thus a topology. This topology makes VV a topological vector space. If VV is complete with respect to this metric, we call VV a Hilbert space.
Given a topological vector space VV, the automorphism group of VV, denoted by Aut(V)\Aut(V) (or GL(V)GL(V) if we are working with a basis) is the group of continuous linear maps VVV \to V with continuous inverses.
A representation of a topological group GG is a pair (V,ϕ)(V, \phi) where VV is a vector space and ϕ\phi is a continuous homomorphism ϕ:GAut(V)\phi:G \to \Aut(V). Frequently, we will write ϕ(g)(v)\phi(g)(v) as gvg \cdot v. Also, we will frequently refer to the representation as VV, leaving the group action implicit. All of the representations I write about today will be assumed to be finite-dimensional unless specified otherwise.
A morphism or GG-linear map between representations VV and WW is a continuous linear map A:VWA:V \to W which commutes with the group action on VV and WW (i.e. g(Av)=A(gv)g(Av) = A(gv)). We denote the set of GG-linear maps between VV and WW by HomG(V,W)\Hom_G(V,W). We will sometimes write EndG(V)\End_G(V) for HomG(V,V)\Hom_G(V,V). The set of finite-dimensional representations of GG together with the GG-linear morphisms form a category.
An isomorphism is a GG-linear map with a GG-linear inverse.
Unless specified otherwise, all functions will be assumed to be continuous.

Useful Constructions

A subrepresentation of a representation (V,ϕ)(V, \phi) is a linear subspace WVW \subseteq V which is invariant under the action of GG. This defines a representaiton (W,ϕW)(W, \phi|_W).

There are several simple subrepresentations we can consider.

  1. For any representation VV, {0}V\{0\} \subseteq V is a subrepresentation because g0=0g \cdot 0 = 0.
  2. Similarly, VV is a subrepresentation of itself.
  3. We also have a subrepresentation VG={vVgv=v}V^G = \{v \in V\;|\; g\cdot v = v\}, the subspace of GG-invariants. Note that the action of GG on VGV^G is trivial.
  4. Given any GG-linear map AHomG(V,W)A \in \Hom_G(V,W), the kernel is a subrepresentation of AA and the image is a subrepresentation of WW.

Given two representations (V,ϕ)(V, \phi) and (W,ψ)(W, \psi), there are several ways we can build new representations out of them.

  1. We can define a representation of GG on the dual space V=Hom(V,k)V^* = \Hom(V, k) (where kk is the base field) by setting g(A)(v)=A(g1v)g(A)(v) = A(g^{-1}v) for AHom(V,k)A \in \Hom(V,k).
  2. We can define a representation of GG on the conjugate space V\overline V. We define V\overline V as follows: it is the same topological abelian group as VV, but the scalar multiplication is changed. Let vv denote an element of VV and v\overline v denote the corresponding element of V\overline V. Then we set λv=λv\lambda \overline v = \overline{\overline \lambda v}. That is to say, we scalar multiply by the conjugate of λ\lambda instead of by λ\lambda itself. The action of GG on V\overline V is the same as the action of GG on VV.
  3. We can define a representation of GG on VWV \oplus W by setting g(v,w)=(gv,gw)g(v,w) = (gv, gw).
  4. We can define a representation of GG on VWV \otimes W by setting g(vw)=(gv)(gw)g(v \otimes w) = (gv) \otimes (gw).
  5. We can define a representation of GG on Hom(V,W)\Hom(V,W) by using the isomorphism Hom(V,W)WV\Hom(V,W) \cong W \otimes V^* for finite-dimensional W,VW,V and using our constructions for taking tensor products and duals of representations.
We can write down an explicit formula for the action of GG on Hom(V,W)\Hom(V,W). Let {ei}\{e_i\} be a basis for WW and {fj}\{f_j\} be a basis for VV. Let {fj}\{f^j\} be the corresponding dual basis of VV^*. Using the definition of the dual representation, GG acts on VV^* by the formula g(fj)(v)=fj(g1v)g\cdot(f^j)(v) = f^j(g^{-1} \cdot v). Therefore, (g(eifj))(v)=(gei)(fj(g1v))(g \cdot (e_i \otimes f^j))(v) = (g\cdot e_i) \otimes (f^j(g^{-1} \cdot v))

Note that fj(g1v)f^j(g^{-1}v) is a scalar and geige_i is a vector in WW. So this is just fj(g1v)(gei)f^j(g^{-1}v)(ge_i). Since gg acts by a linear map, we can factor out the gg to obtain (g(eifj))(v)=g(fj(g1v)ei)=g((eifj)(g1v))(g \cdot (e_i \otimes f^j))(v) = g \cdot (f^j(g^{-1}v) e_i) = g \cdot ((e_i \otimes f^j)(g^{-1}v)) So given any AHom(V,W)A \in \Hom(V,W), we have (gA)(v)=gA(g1v)(g \cdot A)(v) = g\cdot A(g^{-1}v).


HomG(V,W)=Hom(V,W)G\Hom_G(V,W) = \Hom(V,W)^G where the left hand side is the space of GG-linear maps, and the right hand side is the subspace of invariants of the representation Hom(V,W)\Hom(V,W) as defined above.

First, suppose that AHom(V,W)GA \in \Hom(V,W)^G. Then gA=Ag \cdot A = A, so in particular we have (gA)v=Av(g \cdot A)v = Av for any vVv \in V. Using the formula for gAg \cdot A, we see that gA(g1v)=Avg \cdot A(g^{-1} v) = Av for all gG,vVg \in G, v \in V. Multiplying both sides by g1g^{-1}, we find that A(g1v)=g1AvA(g^{-1}v) = g^{-1} Av. Since this is true for all gGg \in G, we conclude that AA is GG-linear. So AHomG(V,W)A \in \Hom_G(V,W).

Conversely, suppose that AHomG(V,W)A \in \Hom_G(V,W). Then A(gv)=g(Av)A(gv) = g(Av) for all gG,vVg \in G, v \in V. So g1A(gv)=Avg^{-1}A(gv) = Av. Letting h=g1h = g^{-1}, we see that gA(h1v)=Avg A (h^{-1}v) = Av for all hG,vVh \in G, v \in V. So AA is in the subspace of invariants Hom(V,W)G\Hom(V,W)^G.

Complete Reducibility and Schur's Lemma

An irreducible representation, or an irrep is a representation VV whose only two subrepresentations are {0}\{0\} and VV itself.
A representation is called completely reducible if it is a direct sum of irreps.
Some representations are neither irreducible nor completely reducible. Consider the set of upper triangular 2×22 \times 2 matrices G={(1x01)aR}G = \left\{\left.\begin{pmatrix} 1 & x \\ 0 & 1\end{pmatrix}\;\right|\; a\in \R\right\}

These all have determinant one, and are thus invertible. Furthermore, the product of two upper-triangular matrices is an upper-triangular matrix, so this is a group. This group has a natural action on R2\R^2 given by the usual matrix-vector product. This defines a representation of GG on R2\R^2.

Note that this representation fixes the subspace VR2V \subseteq \R^2 given by

V={(λ0)λR}V = \left\{\left. \begin{pmatrix}\lambda\\0\end{pmatrix}\;\right|\;\lambda\in\R\right\}

But it doesn't fix any other nontrivial subspaces. So R2\R^2 is neither an irreducible representation nor a completely reducible representation of GG.

It's kind of frustrating that not all representations are completely reducible. One of the nice features of finite groups is that all representations of finite groups are completely reducible. We will show that compact groups are nice in this way as well- all representations of compact groups are completely reducible as well.

(Schur's Lemma)
  1. Let V,WV, W be irreps of GG. Let AHomG(V,W)A \in \Hom_G(V,W). Then AA is either 0 or an isomorphism.
  2. Let VV be a complex irrep of GG. Then EndG(V)=CIdV\End_G(V) = \C \cdot \Id_V (i.e. any GG-linear endomorphism of VV is a scalar multiple of the identity)
  1. Since AA is GG-linear, we know that kerA,imA\ker A, \im A are subrepresentations. Since V,WV,W are irreps, this implies that kerA\ker A is either 00 or all of VV, and imA\im A is either 00 or all of WW. Thus, the only way for AA to be nonzero is if kerA=0\ker A = 0 and imA=W\im A = W. This means that if AA is nonzero, it must be an isomorphism.
  2. Since AA is a complex matrix, it has an eigenvalue λ\lambda. Clearly λId\lambda \Id is a GG-linear endomorphism of VV. Thus, AλIdHomG(V,W)A - \lambda \Id \in \Hom_G(V,W). But AλIdA-\lambda \Id cannot be an isomorphism. So it must be 00. Thus, A=λIdA = \lambda \Id.
Let (V,ϕ)(V, \phi) be a representation. There exists a unique projection AvGEndG(V)Av_G \in \End_G(V) onto VGV^G.

First, we will construct one such projection. Explicitly, we define AvG(v):=GgvdgAv_G(v) := \int_G g \cdot v \;dg This operation averages over the group action, which is why we named the projection AvGAv_G. To show that AvGAv_G is a projection, we have to show that it restricts to the identity on its image. First, we note that the image of AvGAv_G is simply VGV^G. We see that imAvG\im Av_G is contained in VGV^G. Let vv be any vector in VV. Then for any hGh \in G, we have hAvG(v)=hGgvdg=G(hg)vdg=G(hg)vd(hg)h \cdot Av_G(v) = h \cdot \int_G g \cdot v \;dg = \int_G (hg) \cdot v\;dg = \int_G (hg)\cdot v \;d(hg) the last equality follows from the left-invariance of our measure. So we see that hAvG(v)=AvG(v)h \cdot Av_G(v) = Av_G(v), which implies that imAvGVG\im Av_G \subseteq V^G.

Furthermore, any vector of VGV^G is itself fixed by AvGAv_G. If vVGv \in V^G, then AvG(v)=Ggvdg=Gvdg=vAv_G(v) = \int_G g \cdot v\;dg = \int_G v\;dg = v. So in particular, vimAvGv \in \im Av_G. Thus, we see that imAvG=VG\im Av_G = V^G, and AvGAv_G acts as the identity on its image. So it is a projection.

Now, we will show uniqueness. First, note that AvGAv_G commutes with any other TEndG(V)T \in \End_G(V). AvGT(v)=Gg(Tv)dg=GT(gv)dg=TGgvdg=TAvG(v)Av_G \circ T(v) = \int_G g \cdot (Tv)\;dg = \int_G T (g\cdot v) \;dg = T \int_G g \cdot v\;dg = T \circ Av_G (v) Suppose that PP is another projection onto VGV^G. In particular, it is an element of EndG(V)\End_G(V), so it commutes with AvGAv_G. Thus, P=AvGP=PAvG=AvGP = Av_G \circ P = P \circ Av_G = Av_G

Let VV be a representation of GG. Then there exists a GG-invariant inner product on VV.

Let ,\inrp \cdot \cdot be any hermitian inner product on VV. We can view ,\inrp \cdot \cdot as an element of Hom(VV,C)Hom(V,V)\Hom(\overline V \otimes V, \C) \cong \Hom(\overline V, V^*). We can think of the GG-invariant inner products as elements of Hom(V,V)G\Hom(\overline V, V^*)^G. So AvG,Av_G \inrp \cdot \cdot gives us a GG-invariant inner product.

Explicitly, this just means that we can define a GG-invariant inner product ,G\inrp \cdot \cdot _G by the formula v,wG:=Ggv,gwdg\inrp v w_G := \int_G \inrp {gv}{gw}\;dg

If VV is endowed with a GG-invariant inner product, then we call the representation unitary, since for every gGg \in G, ϕ(g)\phi(g) is a unitary operator (or orthogonal if VV is a real vector space). So the above lemma says that for any representation VV, there is an inner product on VV such that our representation is unitary. This perspective will be useful later.
(Maschke) Let VV be a representation of GG and let WVW \subseteq V be a subrepresentation. Then there exists a subrepresentation UVU \subseteq V such that V=WUV = W \oplus U.

Let ,\inrp \cdot \cdot be a GG-invariant inner product on VV. Let U=WU = W^\perp.

We note that UU is a subrepresentation of VV. Let uUu \in U. By definition, u,w=0\inrp u w = 0 for all wWw \in W. Since the inner product is GG-invariant, gu,w=u,g1w\inrp {gu} {w} = \inrp u {g^{-1}w}. Since WW is a subreresentation, g1wWg^{-1}w \in W, so u,g1w=0\inrp u {g^{-1}w} = 0. Thus, gu,w=0\inrp {gu} w = 0 for all wWw \in W, so we conclude that guUgu \in U.

Therefore, V=WUV = W \oplus U.

Any representation of a compact group is completely reducible.

Just reply Maschke's theorem repeatedly. Since our vector space is finite-dimensional, this process must terminate.

Let VV be a representation of GG. Then we can decompose VV into irreducible representations as VE1d1EkdkV \cong E_1^{\oplus d_1} \oplus \cdots \oplus E_k^{\oplus d_k} where the EiE_i are nonisomorphic irreps, and we have di=dimHomG(V,Ei)=dimHomG(Ei,V)d_i = \dim \Hom_G(V, E_i) = \dim \Hom_G(E_i, V) We call did_i the multiplicity of EiE_i in VV, and sometimes denote it [V:Ei][V:E_i].

This follows from the above corollary and Shur's lemma.

Characters

Let (V,ϕ)(V, \phi) be a representation of GG. The character is the function χV:GC\chi_V:G \to \C defined by χV(g)=trϕ(g)\chi_V(g) = \tr \phi(g). Sometimes, we will denote the character by χϕ\chi_\phi
If (V,ϕ)(V, \phi) is a trivial representation of GG (i.e. ϕ\phi sends every gGg \in G to the identity), then χϕ\chi_\phi is the constant function dimV\dim V.
Let (V,ϕ)(V, \phi) and (W,ψ)(W, \psi) be isomorphic representations. Then χV=χW\chi_V = \chi_W.

Let A:VWA:V \to W be a (GG-linear) isomorphism. Then ψ(g)Av=Aϕ(g)v\psi(g)Av = A \phi(g) v. So ϕ(g)=A1ψ(g)A\phi(g) = A^{-1}\psi(g) A. By the cyclic property of the trace, tr(A1ψ(g)A)=tr(ψ(g))\tr(A^{-1} \psi(g) A) = \tr(\psi(g)). Thus, χV(g)=tr(ϕ(g))=tr(A1ψ(g)A)=tr(ψ(g))=χW(g)\chi_V(g) = \tr(\phi(g)) = tr(A^{-1}\psi(g)A) = \tr(\psi(g)) = \chi_W(g)

Our operations on representations define the following operations on the characers
  1. χV=χV\chi_{V^*} = \chi_V^* where χV(g)=χV(g1)\chi_V^*(g) = \chi_V(g^{-1})
  2. χV=χV\chi_{\overline V} = \overline{\chi_V} where χV(g)=χV(g)\overline{\chi_V}(g) = \overline{\chi_V(g)}
  3. χVW=χV+χW\chi_{V \oplus W} = \chi_V + \chi_W
  4. χVW=χVχW\chi_{V \otimes W} = \chi_V \cdot \chi_W
  5. χHom(V,W)=χVχW\chi_{\Hom(V,W)} = \chi_V^* \cdot \chi_W
  6. χVG=av(χV)\chi_{V^G} = av(\chi_V) where av(χV)=GχV(g)dgav(\chi_V) = \int_G \chi_V(g)\;dg considered as a constant function
  1. Since gg acts on VV^* by ϕ(g1)T\phi(g^{-1})^T, and transposing does not change the trace, we see that χV(g)=χV(g1)\chi_{V^*}(g) = \chi_V(g^{-1}).
  2. Since scalar multiplication on V\overline V is conjugated, we have to take the complex conjugate of the entries in the matrix ϕ(g)\phi(g) to get the matrix which acts on V\overline V. Thus, χV=χV\chi_{\overline V} = \overline{\chi_V}.
  3. χVW(g)=tr(ϕ(g)ψ(g))=tr(ϕ(g))+tr(ψ(g))=χV(g)+χW(g)\chi_{V \oplus W}(g) = \tr (\phi(g) \oplus \psi(g)) = \tr(\phi(g)) + \tr(\psi(g)) = \chi_V(g) + \chi_W(g).
  4. χVW(g)=tr(ϕ(g)ψ(g))=tr(ϕV(g))tr(ψW(g))=χV(g)χW(g)\chi_{V \otimes W}(g) = \tr (\phi(g) \otimes \psi(g)) = \tr(\phi_V(g)) \cdot \tr(\psi_W(g)) = \chi_V(g)\chi_W(g).
  5. χHom(V,W)=χWV=χVχW\chi_{\Hom(V,W)} = \chi_{W \otimes V^*} = \chi_V^* \cdot \chi_W.
  6. This one is more complicated. We need to compute χVG\chi_{V^G}. To do so, we use a trick involving the averaging projection.

    Note that the averaging projection AvG:VVGAv_G:V \to V^G acts as the identity on VGV^G and acts as 00 on the orthogonal complement to VGV^G. Thus, ϕ(g)AvG\phi(g) \circ Av_G acts as ϕ(g)\phi(g) on VGV^G and acts as 00 on the orthogonal complement to VGV^G. So trVGϕ(g)=trV(ϕ(g)AvG)\tr_{V^G} \phi(g) = \tr_V (\phi(g) \circ Av_G). (Here trVG\tr_{V^G} denotes the trace over VGV^G and trV\tr_V denotes the trace over VV)

    Therefore, χVG(g)=trVGϕ(g)=trV(ϕ(g)AvG)=trV(ϕ(g)Gϕ(h)dh)=trVGϕ(gh)dh\chi_{V^G}(g) = \tr_{V^G} \phi(g) = \tr_V (\phi(g) \circ Av_G) = \tr_V \left(\phi(g)\int_G \phi(h)dh\right) = \tr_V\int_G \phi(gh)dh Since our measure is left-invariant, this is just χVG(g)=trVGϕ(g)dg=GtrVϕ(g)dg=GχV(g)dg=av(χV)\chi_{V^G}(g) = \tr_V \int_G \phi(g)dg = \int_G \tr_V \phi(g)dg = \int_G \chi_V(g)dg = av(\chi_V)

It turns out that χV=χV\chi_V^* = \overline{\chi_V}. Note that a GG-invariant inner product gives an isomorphism VVV^* \cong \overline V as GG-representations. Since we proved the existence of GG-invariant inner products, it follows that the representations V,V\overline V, V^* are isomorphic, so they have the same characters.
We can put an inner product on the space of complex-valued functions on GG. Given f1,f2:GCf_1,f_2:G \to \C, we define f1,f2(g)=Gf1(g)f2(g)dg=av(f1f2)\inrp{f_1}{f_2}(g) = \int_G f_1(g) \overline{f_2(g)}dg = av(f_1 \cdot \overline {f_2})
For representations V,WV, W we have dimHomG(V,W)=χW,χV\dim \Hom_G(V,W) = \inrp{\chi_W}{\chi_V}
The proof is surprisingly simple using all of the operations we have defined on representations and characters. dimHomG(V,W)=dimHom(V,W)G=χHom(V,W)G=av(χHom(V,W))=av(χWV)=av(χWχV)=av(χWχV)=χW,χV\begin{aligned} \dim \Hom_G(V,W) &= \dim \Hom(V,W)^G\\ &= \chi_{\Hom(V,W)^G}\\ &= av(\chi_{\Hom(V,W)})\\ &= av(\chi_{W \otimes V^*})\\ &= av(\chi_W \cdot \chi_V^*)\\ &= av(\chi_W \cdot \overline{\chi_V})\\ &= \inrp {\chi_W}{\chi_V} \end{aligned}
A representation VV is irreducible if and only if χV,χV=1\inrp {\chi_V}{\chi_V} = 1.
By the above proposition, χV,χV=dimHomG(V,V)\inrp {\chi_V}{\chi_V} = \dim \Hom_G(V,V). We know that we can decompose VV into a direct sum of irreducibles V=iEidiV = \bigoplus_i E_i^{d_i} where dimhomG(Ei,Ej)=δij\dim \hom_G(E_i, E_j) = \delta_{ij}. So dimHomG(iEidi,iEidi)=1\dim \Hom_G(\bigoplus_i E_i^{d_i}, \bigoplus_i E_i^{d_i}) = 1 if and only if V=EiV = E_i for some ii. Thus, χV,χV=1\inrp {\chi_V}{\chi_V} = 1 if and only if VV is irreducible.
χV=χW\chi_V = \chi_W if and only if VV is isomorphic to WW.
We saw earlier that if VWV \cong W, then χV=χW\chi_V = \chi_W. So now, we just have to show that if χV=χW\chi_V = \chi_W, then VWV \cong W. Recall that ViEidiV \cong \bigoplus_i E_i^{d_i} where di=dimHomG(V,Ei)d_i = \dim \Hom_G(V,E_i). Since dimHomG(V,Ei)=χEi,χV\dim \Hom_G(V,E_i) = \inrp {\chi_{E_i}}{\chi_V}, we conclude that if VV and WW have the same characters, then they must be isomorphic.
(Orthogonality of characters) Let E,FE, F be irreps of GG. Then χE,χF\inrp{\chi_E}{\chi_F} is 11 if EE and FF are isomorphic and 00 otherwise.
By Schur's lemma, dimHomG(E,F)\dim \Hom_G(E,F) is 11 if EE and FF are isomorphic and 00 otherwise.

The previous propositions tell us that characters of GG are a decategorification of the category of finite-dimensional representations of GG. Decategorification is the process of taking a category, identifying isomorphic objects and forgetting all other morphisms. This eliminates a lot of useful information, but often makes the category easier to work with. For example, if we decategorify the category of finite sets, we identify all sets with the same cardinality, and forget about all other functions. This just leaves us with the natural numbers, because sets are classified by their cardinality.

Frequently, there are nice structures in the category that still make sense after decategorification. For example, decategorifying disjoint unions of finite sets gives us addition of natural numbers, and decategorifying cartesian products gives us multiplication of natural numbers.

Above, we saw that characters are a decategorification of finite-dimensional GG-representations. Two characters are equal if the corresponding representations are isomorpic, and the direct sums, tensor products, etc. of GG-representations translate nicely into operations on characters. One of the most interesting aspects of this decategorification is that Hom\Homs turn into inner products.

Recall that a pair of adjoint functors are functors F:CD,G:DCF:\mathcal{C} \to \mathcal{D}, G:\mathcal{D} \to \mathcal{C} such that HomD(F(X),Y)HomC(X,G(Y))\Hom_D(F(X), Y) \cong \Hom_C(X, G(Y)) for all XOb(C),YOb(D)X \in \Ob(C), Y \in \Ob(D). Adjoint functors are so named in analogy with adjoint linear operators (Recall that two operators T,UT,U on a Hilbert space are adjoint if Tx,y=x,Uy\inrp {Tx} y = \inrp x {Uy} for all vectors x,yx,y.) This connection between inner products and Hom sets can be formalized to give a categorification of Hilbert spaces.

For an introduction to (de)categorification, you can look here (for a simpler introduction) or here (for a more complicated introduction).

Application: Irreducible Representations of SU(2)SU(2)

Before proceeding, let's use some of this machinery we have built up so far to find all irreducible representations of SU(2)SU(2).

SU(2)SU(2) is the special unitary group of degree 22. It is the group of 2×22 \times 2 complex matrices which are unitary and have determinant 11.

It will be helpful to use another characterization of SU(2)SU(2) as well.

SU(2)S3SU(2) \cong S^3, where S3S^3 is given a multiplicative structure by identifying it with the group of unit quaternions H×\H^\times.

Recall that the quaternions are defined by H={a+jba,bC}\H = \{a + jb\;|\; a,b \in \C\} where jb=bjjb = \overline b j. Then S3={a+jbHa2+b2=1}S^3 = \{a + jb \in \H\;|\; |a|^2 + |b|^2 = 1\}. We have a natural action of S3S^3 on H\H by left-multiplication. This gives us a two-dimensional complex representation of S3S^3. Writing it out explicitly, we see that (a+jb):1a+jb(a+jb):jb+ja\begin{aligned} (a+jb) : 1 &\mapsto a+jb\\ (a+jb) : j &\mapsto - \overline b + j \overline a\\ \end{aligned} Thus, our representation is given by (a+jb)(abba) (a+jb) \mapsto \begin{pmatrix} a & -\overline b\\b & \overline a \end{pmatrix} This matrix is unitary, and has determinant a2+b2=1|a|^2 + |b|^2 = 1. So this is clearly a continuous bijection from S3S^3 to SU(2)SU(2). You can check that this bijection is a group isomorphism.

Since SU(2)SU(2) acts on C2\C^2, we also get an action of SU(2)SU(2) on C[z1,z2]\C[z_1, z_2], the space of complex polynomials in 2 variables. Given ASU(2),pC[z1,z2]A \in SU(2), p \in \C[z_1, z_2], we define (Ap)((z1z2))=p(A1(z1z2))(A \cdot p)\left(\vvec{z_1}{z_2}\right) = p \left(A^{-1} \vvec {z_1}{z_2}\right) We note that this action does not change the degree of monomials. Thus, the space of homogeneous polynomials of degree kk is invariant under this action. So it is a subrepresentation. Let VkC[z1,z2]V_k \subseteq \C[z_1, z_2] denote the space of homogeneous polynomials of degree kk. We will show that {Vk}\{V_k\} are nonisomorphic irreducible representations, and every irreducible representation of SU(2)SU(2) is isomorphic to some VkV_k. First, we'll start with a lemma about the structure of S3S^3.

  1. Every element of S3HS^3 \subseteq \H can be written geiθg1ge^{i\theta}g^{-1}
  2. For fixed θ\theta, {geiθg1}\{ge^{i\theta}g^{-1}\} is a 2D sphere with radius sinθ\sin \theta intersecting C\C at eiθ,eiθe^{i\theta}, e^{-i\theta}
  1. Using our identification of S3S^3 with SU(2)SU(2), we can think of points on the sphere as special unitary matrices. Unitary matrices are unitarily-diagonalizable. Clearly we can rescale these matrices so that the matrices we diagonalize with are in SU(2)SU(2). Finally, note that a diagonal matrix in SU(2)SU(2) must have the form (a00a)\begin{pmatrix} a & 0 \\ 0 & \overline a\end{pmatrix} for aCa \in \C, a2=1|a|^2 = 1. Thus, diagonal matrices in SU(2)SU(2) correspond to points eiθe^{i\theta} on the sphere.
  2. Since the quaternion norm is multiplicative, and elements of S3S^3 have norm 1, we see that geiθg1=geiθg1=1|g e^{i\theta}g^{-1}| = |g||e^{i\theta}||g|^{-1} = 1. Furthermore, geiθg1=g(cosθ+isinθ)g1=cosθ+sinθgig1\begin{aligned} ge^{i\theta}g^{-1} &= g (\cos \theta + i \sin \theta)g^{-1}\\ &= \cos \theta + \sin \theta \; g i g^{-1} \end{aligned} Now, we will consider the map π:ggig1\pi:g \mapsto g i g^{-1}. Note that for unit quaternions, g1=gg^{-1} = \overline g, the conjugate of gg. So we can also write this map π:ggig\pi:g \mapsto g i \overline g. Note that gig=gig=gig\overline{g i \overline g} = g \overline i \overline g = -g i \overline g So giggi\overline g is purely imaginary. Furthermore, since g,ig, i and g\overline g are all unit quaternions, so is their product. Thus, we can think of π\pi as a map π:S3S2\pi : S^3 \to S^2, where we view S3S^3 as the unit quaternions and S2S^2 as the unit imaginary quaternions.

    Furthermore, π\pi is surjective. If we represent vectors in R3\R^3 as imaginary quaternions, then vgvgv \mapsto g v \overline g is a representation of SU(2)SU(2) on R3\R^3 which acts by rotations. Since we can write all rotations in this form, and the rotation group SO(3)SO(3) acts transitively on the two-sphere, we see that {gigˉ}gSU(2)\{gi \bar g\}_{g \in SU(2)} covers all of S2S^2. So π\pi is surjective.

    So since geiθg1=cosθ+sinθπ(g)g e^{i\theta} g^{-1} = \cos \theta + \sin \theta \pi(g), we see that {geiθg1}\{g e^{i\theta} g^{-1}\} is a sphere with radius sinθ\sin \theta. Now, we just have to check that the intersection of this sphere with C\C is e±iθe^{\pm i \theta}. Note that imπ\im \pi is the imaginary unit quaterions, and the only imaginary unit quaternions that lie in C\C are ±i\pm i. Thus, the intersection of {geiθg1}\{g e^{i\theta} g^{-1}\} with C\C is cosθ±isinθ\cos \theta \pm i \sin \theta.

The homogeneous subspaces {Vk}\{V_k\} are nonisomorphic irreducible representations of SU(e)SU(e) and every irreducible representation of SU(2)SU(2) is isomorphic to some VkV_k.

To prove this, we will use characters. For convenience, let us write χVk\chi_{V_k} as χk\chi_k. Recall that the image of eiθS3e^{i\theta} \in S^3 in SU(2)SU(2) is the matrix (eiθ00eiθ)\begin{pmatrix} e^{i\theta}& 0 \\ 0 & e^{-i\theta}\end{pmatrix} Note that the eigenspaces of this operator on VkV_k are {Cz1z2k}\{\C z_1^\ell z_2^{k-\ell}\}_\ell with eigenvalues {e(2k)iθ}\{e^{(2 \ell - k)i \theta}\}_\ell. Therefore, χk(eiθ)==0ke(2k)iθ=e(k+1)iθe(k+1)iθeiθeiθ=sin[(k+1)θ]sinθ\begin{aligned} \chi_k(e^{i\theta}) &= \sum_{\ell=0}^k e^{(2 \ell - k) i \theta}\\ &= \frac{e^{(k+1)i\theta} - e^{-(k+1)i\theta}}{e^{i\theta} - e^{-i\theta}}\\ &= \frac{\sin[(k+1)\theta]}{\sin\theta} \end{aligned} Note that all of these characters are different. This shows that all of the representations VkV_k are distinct. Now, we will show that the characters are orthonormal.

Recall that the inner product on characters is given by χk,χ=S3χk(g)χ(g)dg\inrp{\chi_k}{\chi_\ell} = \int_{S^3} \chi_k(g) \overline{\chi_\ell(g)}\;dg Since the volume of S3S^3 is 2π22\pi^2, we can write dg=12π2dσdg = \frac 1 {2\pi^2}d\sigma where dσd\sigma is the standard volume element on S3S^3. So we want to compute χk,χ=12π2S3χk(g)χ(g)dσ\inrp{\chi_k}{\chi_\ell} = \frac 1 {2\pi^2} \int_{S^3} \chi_k(g) \overline{\chi_\ell(g)}\;d\sigma Recall that characters are constant on conjugacy classes. Since every element of SU(2)SU(2) is conjugate to exactly two unit complex numbers, we have χk,χ=12π20πχk(eiθ)χ(eiθ)vol(orbit)dθ\inrp{\chi_k}{\chi_\ell} = \frac 1 {2\pi^2} \int_0^\pi \chi_k(e^{i\theta}) \overline{\chi_\ell(e^{i\theta})}\vol(\text{orbit})\;d\theta Above, we showed that these orbits are spheres with radius sinθ\sin \theta. Therefore, the volume of an orbit is 4πsin2θ4 \pi \sin^2 \theta. Substituting this and our expressions for the characters, we see that our inner product is χk,χ=2π0πsin[(k+1)θ)sin[(+1)θ]dθ\inrp{\chi_k}{\chi_\ell} = \frac 2 {\pi} \int_0^\pi \sin[(k+1)\theta)\sin[(\ell+1)\theta]\;d\theta Because sines with different frequencies are orthogonal, we conclude that χk,χ=δk\inrp{\chi_k}{\chi_\ell} = \delta_{k\ell} So our characters are orthonormal.

Finally, we will show that these are all of the irreducible representations. Suppose that WW was another irreducible representation. Then 0=χW,χk=GχW(g)χk(g)dg0 = \inrp{\chi_W}{\chi_k} = \int_G \chi_W(g) \overline{\chi_k(g)}\;dg Using the same computational tricks, we see that 0=2π0πχW(eiθ)sin[(k+1)θ]sinθdθ0 = \frac 2 \pi \int_0^\pi \chi_W(e^{i\theta}) \sin[(k+1)\theta]\sin\theta\;d\theta Since sinces form an orthonormal basis for the set of square-integrable functions on the circle, we see that χW(eiθ)=0\chi_W(e^{i\theta}) = 0, which is impossible. Thus, every irreducible representation must be isomorphic to some VkV_k.

Characters made it fairly easy to classify all of the irreducible representations of SU(2)SU(2). Later on, we will generalize some of the computational techniques we used here to find the Weyl character formula and Weyl Integration Formula, which will be very useful for understanding representations. But that will have to be another post, since this one is already much longer than I realized it would be.

Leave a Comment

No comments:

Powered by Blogger.