Real representations and the Frobenius-Schur indicator

Jeffery MensahJuly 6, 2025~5000 words

In this post, we discuss an invariant of a complex representation of a finite group, and show how this yields a connection between irreducible real and complex representations. The story begins with trying to determine whether or not an irreducible complex representation admits an equivariant bilinear form; in trying to solve this problem, we arrive at an invariant of a complex representation known as the Frobenius-Schur indicator. Surprisingly, however, this invariant reveals a lot more about the structure of an irreducible complex representation, and ultimately allows us to classify both real and complex irreducible representations.

Throughout this post, let GG be a finite group.

Equivariant bilinear forms

Let VV be a complex representation of GG. An equivariant bilinear form on VV is a form b ⁣:V×VCb \colon V \times V \to \mathbb{C} such that b(gv,gw)=b(v,w)b(gv, gw) = b(v, w) for all gGg \in G and v,wVv, w \in V. For each bilinear form bb, we may define the left and right kernels

kerL(b)={vb(v,w) for all wV},kerR(b)={wb(v,w) for all vV},\begin{align*} \ker_{\rm L} (b) &= \{v \mid b(v, w) \text{ for all } w \in V \}, \\ \ker_{\rm R} (b) &= \{w \mid b(v, w) \text{ for all } v \in V \}, \end{align*}

which are easily seen to both be GG-invariant subspaces of VV. In particular, they are the kernels of the partial evaluation maps bL,bR ⁣:VVb_{\mathrm{L}}, b_{\mathrm{R}} \colon V \to V^\vee defined by bL(v)=[wb(v,w)]b_{\mathrm{L}}(v) = [w \mapsto b(v, w)] and bR(w)=[vb(v,w)]b_{\mathrm{R}}(w) = [v \mapsto b(v, w)]. Recall that there is an "evaulation" isomorphism eval ⁣:VV\mathrm{eval} \colon V \to V^{\vee\vee}, which takes a vector vv and produces an evaluator evalv ⁣:VC\mathrm{eval}_v \colon V^{\vee} \to \mathbb{C} defined by evalv(ϕ)=ϕ(v)\mathrm{eval}_v(\phi) = \phi(v). Using this isomorphism, we obtain

bL(w)=evalwbL=[vbL(v)(w)]=[vb(v,w)]=bR(w).b_L^\vee (w) = \mathrm{eval}_w \circ b_{\mathrm{L}} = [v \mapsto b_{\mathrm{L}}(v)(w)] = [v \mapsto b(v, w)] = b_R(w).

It follows that both the left and right evaluation maps are dual, and thus have kernels of the same dimension. Thus, if VV is irreducible, either both the left and right kernels are the entire space, in which case b=0b = 0, or zero, in which case we say that bb is nondegenerate.

It turns out that if VV is irreducible, there is at most one (up to scaling) equivariant bilinear form on VV. To see, suppose that a,b ⁣:V×VCa, b \colon V \times V \to \mathbb{C} are two nonzero equivariant bilinear forms. Then bL1aL1 ⁣:VVb_{\mathrm{L}}^{-1} \circ a_{\mathrm{L}}^{\vphantom{-1}} \colon V \to V is endomorphism over C\mathbb{C} and thus has an eigenvector vv with eigenvalue λ\lambda. Unpacking, this means that for all wVw \in V, we have a(v,w)=λb(v,w)a(v, w) = \lambda b(v, w), so aλba - \lambda b is an equivariant bilinear form with a nonzero left kernel, implying a=λba = \lambda b.

For a complex representation VV, the Frobenius-Schur indicator determines whether or not such an equivariant bilinear form exists and if it does, what type of bilinear form it is. It is instructive to compare this situation with that of Hermitian forms on VV.

Equivariant Hermitian forms

Every complex representation admits an equivariant nondegenerate Hermitian form through the use of the "averaging trick": take any positive-definite Hermitian form h0 ⁣:V×VCh_0 \colon V \times V \to \mathbb{C} and define a new map h ⁣:V×VCh \colon V \times V \to \mathbb{C} by

h(v,w)=1GgGh0(gv,gw).h(v, w) = \frac{1}{\abs{G}} \sum_{g \in G} h_0(gv, gw).

By construction, this is an equivariant Hermitian form; moreover for v0v \neq 0 we have h(v,v)=1GgGh0(gv,gv)>0h(v, v) = \frac{1}{\abs{G}} \sum_{g \in G} h_0(gv, gv) > 0, so hh is positive-definite. In particular, hh is nondegenerate. This type of argument does not work for bilinear forms VV, as there are no positive-definite bilinear forms b ⁣:V×VCb \colon V \times V \to \mathbb{C}: if b(v,v)>0b(v, v) > 0, then b(iv,iv)=b(v,v)<0b(iv, iv) = -b(v, v) < 0.

Symmetric and alternating bilinear forms

For a complex vector space VV, denote by T2(V)\mathrm{T}^{2}(V^\vee) the space of bilinear maps V×VCV \times V \to \mathbb{C}. Special types of bilinear forms which are of interest to us are the symmetric forms Sym2(V)\mathrm{Sym}^2(V^\vee), which satisfy b(v,w)=b(w,v)b(v, w) = b(w, v), and the alternating forms Alt2(V)\mathrm{Alt}^{2}(V^\vee), which satisfy b(v,w)=b(w,v)b(v, w) = -b(w, v). Basic examples of forms can be constructed from linear forms ϕ,ψV\phi, \psi \in V^\vee in the following ways:

  1. the outer (or tensor) product ϕψ\phi \otimes \psi, defined by (ϕψ)(v,w)=ϕ(v)ψ(w)(\phi \otimes \psi)(v, w) = \phi(v)\psi(w),
  2. the symmetric product ϕψ\phi \odot \psi, defined by (ϕψ)(v,w)=12(ϕ(v)ψ(w)+ϕ(w)ψ(v))(\phi \odot \psi)(v, w) = \frac{1}{2} (\phi(v)\psi(w) + \phi(w)\psi(v)), and
  3. the antisymmetric (or alternating or wedge) product ϕψ\phi \wedge \psi, defined by 12(ϕ(v)ψ(w)ϕ(w)ψ(v))\frac{1}{2} (\phi(v)\psi(w) - \phi(w)\psi(v)).

One may easily verify that these produce bilinear forms and, as the names suggest, that ϕψ\phi \odot \psi is symmetric and ϕψ\phi \wedge \psi is alternating. Given a basis e1,,ene_1, \ldots, e_n of VV with dual basis e1,,ene_1^\vee, \ldots, e_n^\vee, the products eieje_i^\vee \otimes e_j^\vee form a basis of T2(V)\mathrm{T}^{2}(V^\vee), since a bilinear form bb can be written as

b(v,w)=b(i=1nviei,j=1nwjej)=i,jb(ei,ej)viwj=i,jb(ei,ej)(eiej)(v,w),b(v, w) = b \left(\sum_{i=1}^{n} v^i e_i,\, \sum_{j=1}^{n} w^j e_j\right)= \sum_{i, j} b(e_i, e_j)v^iw^j = \sum_{i, j} b(e_i, e_j) \cdot (e_i^\vee \otimes e_j^\vee) (v, w),

so the outer products are at least spanning, and a linear combination i,jbij(eiej)\sum_{i,j} b_{ij} \cdot (e_i^\vee \otimes e_j^\vee) vanishes only if for each k,k, \ell we have

i,jbij(eiej)(ek,e)=i,jbijei(ek)ej(e)=bk=0,\sum_{i,j} b_{ij} \cdot (e_i^\vee \otimes e_j^\vee) (e_k, e_\ell) = \sum_{i,j} b_{ij} \cdot e_i^\vee (e_k) \cdot e_j^\vee (e_\ell) = b_{k\ell} = 0,

so the outer products are also independent. It follows that dimT2(V)=(dimV)2\dim \mathrm{T}^{2}(V^\vee) = (\dim V) ^2.

One may also determine bases for the spaces of symmetric and alternating forms through a similar argument, although the situation is better illuminated by studying how each subspace sits inside T2(V)\mathrm{T}^{2}(V^\vee). To do this, we define two operators:

  1. a symmetrization operator sym ⁣:T2(V)Sym2(V)\mathrm{sym} \colon \mathrm{T}^{2}(V^\vee) \to \mathrm{Sym}^2(V^\vee) given by (symb)(v,w)=12(b(v,w)+b(w,v))(\operatorname{sym} b)(v, w) = \frac{1}{2} (b(v, w) + b(w, v)), and
  2. an antisymmetrization operator alt ⁣:T2(V)Alt2(V)\mathrm{alt} \colon \mathrm{T}^{2}(V^\vee) \to \mathrm{Alt}^2(V^\vee) given by (altb)(v,w)=12(b(v,w)b(w,v))(\operatorname {alt} b)(v, w) = \frac{1}{2}(b(v,w) - b(w,v)).

Note that both operators are projectors, since

(symsymb)(v,w)=12(12(b(v,w)+b(w,v))+12(b(w,v)+b(v,w)))=12(b(v,w)+b(w,v))=(symb)(v,w)(\operatorname{sym} \operatorname{sym} b)(v, w) = \frac{1}{2} \Big(\tfrac{1}{2} ( b(v,w) + b(w, v)) + \tfrac{1}{2} ( b(w, v) + b(v, w)) \Big) = \frac{1}{2} (b(v, w) + b(w, v)) = (\operatorname{sym} b)(v,w)

and

(altaltb)(v,w)=12(12(b(v,w)b(w,v))12(b(w,v)b(v,w)))=12(b(v,w)b(w,v))=(altb)(v,w).(\operatorname{alt} \operatorname{alt} b)(v, w) = \frac{1}{2} \Big(\tfrac{1}{2} ( b(v,w) - b(w, v)) - \tfrac{1}{2} ( b(w, v) - b(v, w)) \Big) = \frac{1}{2} (b(v, w) - b(w, v)) = (\operatorname{alt} b)(v,w).

Moreover, every bilinear form can be written as the sum of its symmetric and alternating projections, since

(symb)(v,w)+(altb)(v,w)=12(b(v,w)+b(w,v))+12(b(v,w)b(w,v))=b(v,w).(\operatorname{sym} b)(v, w) + (\operatorname{alt} b)(v, w) = \tfrac{1}{2} (b(v,w) + b(w,v)) + \tfrac{1}{2}(b(v,w) - b(w,v)) = b(v, w).

Thus, the space of bilinear forms splits the a direct sum T2(V)=Sym2(V)Alt2(V)\mathrm{T}^{2}(V^\vee) = \mathrm{Sym}^{2}(V^\vee) \oplus \mathrm{Alt}^{2}(V^\vee). The fact that each operator is a projector implies that the products eiej=sym(eiej)e_i^\vee \odot e_j^\vee = \operatorname{sym} (e_i^\vee \otimes e_j^\vee) and eiej=alt(eiej)e_i^\vee \wedge e_j^\vee = \operatorname{alt} (e_i^\vee \otimes e_j^\vee) span Sym2(V)\mathrm{Sym}^{2}(V^\vee) and Alt2(V)\mathrm{Alt}^{2}(V^\vee), respecitively.

Simple combinatorics yield that dimSym2(V)(n+12)\dim \mathrm{Sym}^2(V^\vee) \leq \binom{n + 1}{2} and dimAlt2(V)(n2)\dim \mathrm{Alt}^2(V^\vee) \leq \binom{n}{2}, since there are at most (n+12)\binom{n + 1}{2} distinct nonzero symmetric products and (n2)\binom{n}{2} distinct nonzero antisymmetric products. However, the direct sum composition also implies that dimSym2(V)+dimAlt2(V)=n2\dim \mathrm{Sym}^2(V^\vee) + \dim \mathrm{Alt}^{2}(V^\vee) = n^2, so we must have dimSym2(V)=(n+12)\dim \mathrm{Sym}^2(V^\vee) = \binom{n + 1}{2} and dimAlt2(V)=(n2)\dim \mathrm{Alt}^2(V^\vee) = \binom{n}{2}. To summarize, the dimension and corresponding basis of each space is given below.

SpaceDimensionBasis
T2(V)\mathrm{T}^{2}(V^\vee)n2n^2eieje_i^\vee \otimes e_j^\vee for 1i,jn1 \leq i, j \leq n
Sym2(V)\mathrm{Sym}^{2}(V^\vee)12n(n+1)\tfrac{1}{2} n(n+1)eieje_i^\vee \odot e_j^\vee for 1ijn1 \leq i \leq j \leq n
Alt2(V)\mathrm{Alt}^{2}(V^\vee)12n(n1)\frac{1}{2} n(n-1)e1eje_1^\vee \wedge e_j^\vee for 1i<jn1 \leq i < j \leq n

The Frobenius-Schur indicator

If GG has a linear representation on a complex vector space VV, then we also obtain representations on T2(V)\mathrm{T}^{2}(V^\vee) and the spaces of symmetric and alternating bilinear forms given by (gb)(v,w)=b(g1v,g1w)(gb) (v, w) = b(g^{-1}v, g^{-1}w). Note that the set GG-equivariant bilinear forms is simply the set of invariants of T2(V)\mathrm{T}^{2}(V^\vee), so

{G-equivariant forms V×VC}=T2(V)G=Sym2(V)GAlt2(V)G.\{\text{$G$-equivariant forms $V \times V \to \mathbb{C}$}\} = \mathrm{T}^{2}(V^\vee)^G = \mathrm{Sym}^{2}(V^\vee)^G \oplus \mathrm{Alt}^{2}(V^\vee)^G.

If VV is irreducible, then T2(V)G\mathrm{T}^{2}(V^\vee)^G is either zero or one-dimensional, and thus lies entirely in one of the two summands. This yields three cases:

  1. dimSym2(V)G=0\dim \mathrm{Sym}^{2}(V^\vee)^G = 0 and dimAlt2(V)G=0\dim \mathrm{Alt}^{2}(V^\vee)^G = 0,
  2. dimSym2(V)G=1\dim \mathrm{Sym}^{2}(V^\vee)^G = 1 and dimAlt2(V)G=0\dim \mathrm{Alt}^{2}(V^\vee)^G = 0,
  3. dimSym2(V)G=0\dim \mathrm{Sym}^{2}(V^\vee)^G = 0 and dimAlt2(V)G=1\dim \mathrm{Alt}^{2}(V^\vee)^G = 1.

To distinguish between these cases, we construct the quantity dimSym2(V)GdimAlt2(V)G\dim \mathrm{Sym}^{2}(V^\vee)^G - \dim \mathrm{Alt}^{2}(V^\vee)^G, which takes on the values 1,0,+1-1, 0, +1. Recall that for a representation WW, dimension of the set of invariants WGW^G is given by the inner product the character χW\chi_W with the trivial character χtriv\chi_{\mathrm{triv}}; this simply reads off how many copies of the trivial representation there are in WW. Thus,

χtriv,χSym2(V)χtriv,χAlt2(V)={+1if there exists a nondegenerate symmetric bilinear form on V1if there exists a nondegenerate alternating bilinear form on V+0if there does not exist a nondegenerate bilinear form on V.\inner{\chi_{\mathrm{triv}}, \chi_{\mathrm{Sym}^{2}(V^\vee)}} - \inner{ \chi_{\mathrm{triv}}, \chi_{\mathrm{Alt}^{2}(V^\vee)}} = \begin{cases} +1 &\text{if there exists a nondegenerate \textit{symmetric} bilinear form on $V$} \\ -1 &\text{if there exists a nondegenerate \textit{alternating} bilinear form on $V$} \\ \hphantom{+}0 &\text{if there does not exist a nondegenerate bilinear form on $V$} \\ \end{cases}.

To compute this, we determine the characters of Sym2(V)\mathrm{Sym}^{2}(V^\vee) and Alt2(V)\mathrm{Alt}^{2}(V^\vee). Since the action of each element gGg \in G is always diagonalizable, we may assign to each gg a eigenbasis e1(g),,en(g)e_1(g), \ldots, e_n(g) of VV with corresponding eigenvalues λ1(g),,λn(g)\lambda_1(g), \ldots, \lambda_n(g). Then one may easily check that the bases listed above are also eigenbases for gg on the each space of bilinear forms. Specifically, we have

[g(ei(g)ej(g))](v,w)=ei(g)(g1v)ej(g)(g1w)=(λiλj)1(ei(g)ej(g)).[g (e_i(g)^\vee \otimes e_j(g)^\vee)](v,w) = e_i(g)^\vee(g^{-1}v) \cdot e_j(g)^\vee(g^{-1}w) = (\lambda_{i} \lambda_{j})^{-1} (e_i(g)^\vee \otimes e_j(g)^\vee).

Since each eigenvalue has unit norm, λi1(g)=λi(g)\lambda_i^{-1}(g) = \overline{\lambda_i(g)}, so the character of the representation on T2(V)\mathrm{T}^{2}(V^\vee) is

χT2(V)(g)=i,jλi(g)λj(g)=(i=1nλi(g))2=χV(g)2,\chi_{\mathrm{T}^{2}(V^\vee)}(g) = \sum_{i, j} \overline{\lambda_i(g) \lambda_j(g)} = \left( \sum_{i=1}^{n} \overline{\lambda_i(g)} \right)^2 = \overline{\chi_V(g)^2},

and the characters of the representations on the Sym2(V)\mathrm{Sym}^{2}(V^\vee) and Alt2(V)\mathrm{Alt}^{2}(V^\vee) are

χSym2(V)(g)=ijλi(g)λj(g)=12[(i=1nλi(g))2+i=1nλi(g)2]=12χV(g)2+χV(g2),χAlt2(V)(g)=i<jλi(g)λj(g)=12[(i=1nλi(g))2i=1nλi(g)2]=12χV(g)2χV(g2).\begin{align*} \chi_{\mathrm{Sym}^{2}(V^\vee)}(g) = \sum_{i \leq j} \overline{\lambda_i(g) \lambda_j(g)} = \frac{1}{2}\left[\left( \sum_{i=1}^{n} \overline{\lambda_i(g)} \right)^2 + \sum_{i=1}^{n} \overline{\lambda_i(g)^2} \right] = \frac{1}{2} \cdot \overline{\chi_V(g)^2 + \chi_V(g^2)}, \\ \chi_{\mathrm{Alt}^{2}(V^\vee)}(g) = \sum_{i < j} \overline{\lambda_i(g) \lambda_j(g)} = \frac{1}{2}\left[\left( \sum_{i=1}^{n} \overline{\lambda_i(g)} \right)^2 - \sum_{i=1}^{n} \overline{\lambda_i(g)^2} \right] = \frac{1}{2} \cdot \overline{\chi_V(g)^2 - \chi_V(g^2)}. \\ \end{align*}

Finally, we compute

χtriv,χSym2(V)χtriv,χAlt2(V)=1GgGχV(g)2+χV(g2)21GgGχV(g)2χV(g2)2=1GgGχV(g2).\inner{\chi_{\mathrm{triv}}, \chi_{\mathrm{Sym}^{2}(V^\vee)}}\hspace{-0.3pt} - \hspace{-0.3pt}\inner{ \chi_{\mathrm{triv}}, \chi_{\mathrm{Alt}^{2}(V^\vee)}} = \frac{1}{\abs{G}}\sum_{g \in G} \frac{\chi_V(g)^2 + \chi_V(g^2)}{2} - \frac{1}{\abs{G}}\sum_{g \in G} \frac{\chi_V(g)^2 - \chi_V(g^2)}{2} = \frac{1}{|G|}\sum_{g \in G} \chi_V(g^2).

The last expression is what we refer to as the Frobenius-Schur indicator, and will be denoted here by FS(χV)\mathrm{FS}(\chi_V).

Real representations

The Frobenius-Schur indicator can also be used to understand the real representations of a finite group GG. When working with real representations, results which depend on the base field being algebraically closed no longer directly apply. The most basic of these is Schur's Lemma, which states that all endomorphisms of an irreducible representation over an algebraically closed field are scalar multiples of the identity. For real representations, we state a more general version of the result.

Theorem (Schur's Lemma). Let GG be a finite group, kk be a field, and WW be an irreducible kGkG-representation. Then EndkG(V)\mathrm{End}_{kG}(V) is a division algebra over kk.

Proof. Since EndkG(W)\mathrm{End}_{kG}(W) is already a unital ring, it suffices to show that inverses exist for nonzero elements. If ϕ ⁣:WW\phi \colon W \to W is a nonzero kGkG-equivariant endomorphism, then kerϕ\ker \phi and imϕ\operatorname{im} \phi are subrepresentations of WW. Moreover, since ϕ0\phi \neq 0, the kernel is not the entire space, and the image is not just {0}\{0\}. Therefore, by irreducibility, we must have kerϕ={0}\ker \phi = \{0\} and imϕ=W\operatorname{im} \phi = W. It follows that ϕ\phi is invertible.

\blacksquare

A theorem of Frobenius classifies all the possible finite-dimensional division algebras over R\mathbb{R}; we will only state the theorem here for brevity.

Theorem (Frobenius). Every finite-dimensional division algebra over R\mathbb{R} is isomorphic to either the real numbers, the complex numbers, or the quaternions.

Thus, if WW is an irreducible real representation, then EndRG(W)\mathrm{End}_{\mathbb{R}G}(W) is either R\mathbb{R}, C\mathbb{C}, or H\mathbb{H}.

  • If EndRG(W)R\mathrm{End}_{\mathbb{R}G}(W) \cong \mathbb{R}, then WW just has the structure of a R\mathbb{R}-vector space, so we say WW is real type.

  • If EndRG(W)C\mathrm{End}_{\mathbb{R}G}(W) \cong \mathbb{C}, then we may give WW the structure of a C\mathbb{C}-vector space, so we say that WW is complex type.

  • If EndRG(W)H\mathrm{End}_{\mathbb{R}G}(W) \cong \mathbb{H}, then we may give WW the structure of a left H\mathbb{H}-vector space, so we say that WW is quaternionic type.

Real and quaternionic structures

In the previous section, we showed how a real representation WW can "ascend" into either a complex or quaternionic representation by means of a isomorphism CEndRG(W)\mathbb{C} \to \mathrm{End}_{\mathbb{R}G}(W) or HEndRG(W)\mathbb{H} \to \mathrm{End}_{\mathbb{R}G}(W).

Similarly, a complex representation VV can "ascend" to a quaternionic representation or "descend" to a real representation under certain circumstances. We will call the necessary ascent data a quaternionic structure and the necessary descent data a real structure on VV. To make the situation clearer, we first work in a bit more generality.

Let kkk \subseteq k' be a (possibly skew) field extension. A representation on a kk-vector space WW can be transformed into a representation on the kk'-vector space Wk=WkkW^{k'} = W \otimes_k k' via a process called extension of scalars. Specifically, define the kk'-action by β(wα)=wαβ\beta \cdot (w \otimes \alpha) = w \otimes \alpha\beta and define the GG-action by g(wα)=gwαg(w \otimes \alpha) = gw \otimes \alpha.

The "inverse" of this process transforms a representation on a kk'-vector space VV into a representation on a kk-vector space VkV_{k} with the same underlying set, but with a kk-action kEnd(V)k \to \mathrm{End}(V) given by restricting to kk'-action kEnd(V)k' \to \mathrm{End}(V) to kk. This process is called restriction of scalars.

When working with the extension RC\mathbb{R} \subset \mathbb{C}, extension of scalars is called complexification, and restriction of scalars is called realification.

With these two operations defined, we can rephrase the ascent and descent problems as follows:

  • Ascent. When is a kGkG-representation WW the restriction of scalars of some kGk'G representation VV?
  • Descent. When is a kGk'G-representation VV the extension of scalars of some kGkG-representation WW?

Suitable descent data is described by the following proposition.

Proposition. Let kkk \subseteq k' be a Galois field extension. If VV is a kGk'G-representation, define a kk-structure on VV to be a equivariant semilinear action of Gal(k/k)\mathrm{Gal}(k'/k) on VV; that is, an equivariant action satisfying σ(αv)=σ(α)σ(v)\sigma(\alpha v) = \sigma(\alpha)\sigma(v) for all σGal(k/k)\sigma \in \mathrm{Gal}(k'/k), αk\alpha \in k, and vVv \in V. Then the following categories are equivalent:

{kG-representations}{kG-representations with a k-structure}.\{ \text{$kG$-representations} \} \cong \{\text{$k'G$-representations with a $k$-structure}\}.

To descend from a complex representation VV to a real representation, the data needed is an equivariant semilinear action of Gal(R/C)\mathrm{Gal}(\mathbb{R}/\mathbb{C}), which has group presentation σσ2=1\inner{\sigma \mid \sigma^2 = 1}. Therefore, the data of a real structure on VV is equivalent to the existence of an equivariant map σ ⁣:VV\sigma \colon V \to V such that σ2=id\sigma^2 = \mathrm{id} and σ(zv)=zσ(v)\sigma(zv) = \overline{z}\sigma(v) for all zCz \in \mathbb{C} and vVv \in V (this is known as C\mathbb{C}-antilinearity), so we also call such a map a real structure. If such a map exists, then VV is isomorphic to the complexification of some real representation, and we say that VV (or its character χV\chi_V) is defined over R\mathbb{R} and is a real type representation.

To ascend from a complex representation VV to a quaternionic representation, one needs to equip VV with a left H\mathbb{H}-action which extends the C\mathbb{C}-action, while remaining compatible with the group action. Denote the ring of GG-equivariant endomorphisms of VV (as an abelian group) by EndG(V)\mathrm{End}_G(V). Then one seeks a ring homomorphism

HCj/(j2+1,jzzj (for all zC))EndG(V)\mathbb{H} \cong \mathbb{C}\inner{j}/\Big(j^2 + 1, jz - \overline{z}j \text{ (for all $z \in \mathbb{C}$)}\Big) \to \mathrm{End}_G(V)

which extends the C\mathbb{C}-action CEndG(V)\mathbb{C} \to \mathrm{End}_G(V). Since this map is entirely determined by where jj is sent, it follows that the existence of such a map is equivalent to the existence of an additive endomorphism j ⁣:VVj \colon V \to V such that j2=idj^2 = -\mathrm{id} and j(zv)=zj(v)j(zv) = \overline{z}j(v) for all zCz \in \mathbb{C} and vVv \in V; we call such a map a quaternionic structure. If such a map exists, we say that VV is a quaternionic type representation.

If VV is neither real type or quaternionic type, we say that VV is a complex type representation.

Real-valued characters

If a complex representation VWRCV \cong W \otimes_{\mathbb{R}} \mathbb{C} is defined over R\mathbb{R}, then the matrix elements of each element gGg \in G are all real, since

ρV(g)(ej1)=ρW(g)ej1=i=1n[ρW(g)]ijei1,\rho_V(g)(e_j \otimes 1) = \rho_W(g)e_j \otimes 1 = \sum_{i = 1}^{n} [\rho_W(g)]_{ij} e_i \otimes 1,

and the matrix elements of ρW(g) ⁣:WW\rho_W(g) \colon W \to W are all real. In particular, the character of a complex representation defined over R\mathbb{R} will be real-valued. However, it is not the case that every real-valued character is defined over R\mathbb{R}. As we will shortly see, complex representations with a quaternionic structure also have real-valued characters despite not being defined over R\mathbb{R}; for this reason we also call such representations pseudoreal. To distinguish between these types of representations, we can use the Frobenius-Schur indicator to determine when a complex representation VV has a real-valued character, and if it arises from a real or quaternionic structure on VV.

Theorem. If VV be an irreducible complex representation. A nondegenerate equivariant bilinear form exists on VV if and only if χV\chi_V is real-valued.

Proof. First, suppose that that a nondegenerate equivariant bilinear form bb exists on VV. For each gGg \in G, let e1(g),,en(g)e_1(g), \ldots, e_n(g) be an eigenbasis of VV for gg with eigenvalues λ1(g),,λn(g)\lambda_1(g), \ldots, \lambda_n(g). By nondegeneracy, for each each eie_i, there must exist some eje_j such that b(ei,ej)0b(e_i, e_j) \neq 0. Since bb must either be symmetric or alternating, this also implies b(ej,ei)0b(e_j, e_i) \neq 0. Thus, we may partition the set of eigenvectors into "anisotropic" singletons {ei}\{e_i\} with b(ei,ei)0b(e_i, e_i) \neq 0 and "isotropic" pairs {ei,ej}\{e_i, e_j\} with b(ei,ej)0b(e_i, e_j) \neq 0 and b(ej,ei)0b(e_j, e_i) \neq 0. By equivariance, each anisotropic eigenvector eie_i satsifies

b(ei,ei)=b(g1ei,g1ei)=λi2b(ei,ei),b(e_i, e_i) = b(g^{-1}e_i, g^{-1}e_i) = \lambda_i^{-2} \cdot b(e_i, e_i),

which implies that λi=±1\lambda_i = \pm 1. Similarly, by equivariance, each isotropic pair {ei,ej}\{e_i, e_j\} satisfies

b(ei,ej)=b(g1ei,g1ej)=λi1λj1b(ei,ej).b(e_i, e_j) = b(g^{-1}e_i, g^{-1}e_j) = \lambda_{i}^{-1} \lambda_{j}^{-1} \cdot b(e_i, e_j).

which implies that λi=λj1\lambda_i = \lambda_j^{-1}. Since each eigenvalue has unit norm, this implies that λi=λj\lambda_i = \overline{\lambda_j}, so λi+λj\lambda_i + \lambda_j is real. Thus,

χV(g)=i=1nλi(g)=anisotropicsingletons±1  +isotropicpairsλi+λi,\chi_V(g) = \sum_{i=1}^{n} \lambda_i(g) = \sum_{\substack{\text{anisotropic} \\ \text{singletons}}} \pm 1 \; + \sum_{\substack{\text{isotropic}\\\text{pairs}}} \lambda_i + \overline{\lambda_{i}},

which is real. Conversely, suppose that χV\chi_V is real-valued. The dimension of the set of invariants T2(V)G\mathrm{T}^2(V^\vee)^G is given by the inner product of χT2(V)\chi_{\mathrm{T}^2(V^\vee)} with the trivial character χtriv\chi_{\mathrm{triv}}, so

dimT2(V)G=χtriv,χT2(V)=1GgGχV(g)2=1G[χV(1G)2+g1Gχ(g)2]>0,\dim \mathrm{T}^2(V^\vee)^G = \inner{\chi_{\mathrm{triv}}, \chi_{\mathrm{T}^2(V^\vee)}} = \frac{1}{\abs{G}}\sum_{g \in G} \chi_V(g)^2 = \frac{1}{\abs{G}} \left[\chi_V(1_G)^2 + \sum_{g \neq 1_G} \chi(g)^2 \right] > 0,

since χV(1G)=dimV>0\chi_V(1_G) = \dim V > 0 and χV(g)20\chi_V(g)^2 \geq 0. Thus, there must exist some equivariant bilinear form on VV.

\blacksquare

Thus, if FS(χV)=0\mathrm{FS}(\chi_V) = 0, the representation VV does not have a real-valued character and is consequently not defined over R\mathbb{R}. The two remaining possible values of the indicator correspond to the existence real and quaternionic structures on VV, as we now show.

Theorem. Let VV be an irreducible complex representation. A nondegenerate equivariant symmetric bilinear form exists if and only if VV admits a real structure, and a nondegenerate equivariant alternating bilinear form exists if and only if VV admits a quaternionic structure.

Proof. Suppose there exists a nondegenerate equivariant bilinear form b ⁣:V×VCb \colon V \times V \to \mathbb{C}. Since bb must either be symmetric or alternating, we have b(v,w)=ϵb(w,v)b(v, w) = \epsilon b(w,v), where ϵ=±1\epsilon = \pm 1. Now fix an equivariant positive-definite Hermitian form h ⁣:V×VCh \colon V \times V \to \mathbb{C}. Then the partial evaluation map hR ⁣:VVh_{\mathrm{R}} \colon V \to V^\vee is C\mathbb{C}-antilinear while bR ⁣:VVb_{\mathrm{R}} \colon V \to V^\vee is C\mathbb{C}-linear, which implies that the composition σ=bR1hR ⁣:VV\sigma = b_{\mathrm{R}}^{-1} \circ h_{\mathrm{R}} \colon V \to V is an equivariant C\mathbb{C}-antilinear isomorphism. Thus, σ2 ⁣:VV\sigma^2 \colon V \to V is an C\mathbb{C}-linear isomorphism, which means it is equal to a scalar λEndCG(V)C\lambda \in \mathrm{End}_{\mathbb{C}G}(V) \cong \mathbb{C}. By the previous construction, we have b(v,σw)=h(v,w)b(v, \sigma w) = h(v, w), so

h(σv,σw)=b(σv,σ2w)=b(σv,λv)=λϵb(v,σv)=λϵh(v,v).h(\sigma v, \sigma w) = b(\sigma v, \sigma^2w) = b(\sigma v, \lambda v) = \lambda\epsilon \cdot b(v, \sigma v) = \lambda\epsilon \cdot h(v, v).

Since hh is positive-definite, this implies λϵ>0\lambda\epsilon > 0. We now analyze two cases.

  • If bb is symmetric, ϵ=1\epsilon = 1, which implies that λ>0\lambda > 0. It follows that λ1/2σ ⁣:VV\lambda^{1/2}\sigma \colon V \to V is a real structure on VV.
  • If bb is alternating, ϵ=1\epsilon = -1, which implies that λ<0\lambda < 0. It follows that λ1/2σ ⁣:VV\lambda^{1/2}\sigma \colon V \to V is a quaternionic structure on VV.

Since these cases are mutually exclusive and exhaust all possibilites for bb and structures on VV, the reverse implications also hold.

\blacksquare

We summarize this consequences of this theorem in the following table.

FS(χV)\mathrm{FS}(\chi_V)Representation typeWhat nonzero equivariant bilinear forms exist on VV?
+1+1realnondegenerate symmetric forms unique up to scaling
1-1quaternionicnondegenerate alternating forms unique up to scaling
+0\hphantom{+}0complexnone

Classifying irreducible representations

By using complexification and realification, one can create a correspondence between irreducible real representations with irreducible complex representations with the help of the Frobenius-Schur indicator. The following sections illustrate this procedure.

Complexifying real representations

Given all the irreducible real representations of a finite group GG, it is natural to wonder if we can somehow determine all the irreducible complex representations of GG, and if there is relationship between their characters. To do this, we

  1. show that every irreducible complex representation occurs in some complexified irreducible real representation, and
  2. complexify each irreducible real representation and write it as sum of irreducible complex representations.

Before continuing, recall that the complex conjugate representation V\overline{V} has the same underlying set and group action, but has a C\mathbb{C}-action given by zv=zvz \ast v = \overline{z}v. It is irreducible if and only if VV is, and its character satisfies χV=χV\chi_{\overline{V}} = \overline{\chi_V}. Now for the first step, observe that the map

(VR)C=VRRCVV;vz(zv)(zv) (V_{\mathbb{R}})^{\mathbb{C}} = V_{\mathbb{R}} \otimes_{\mathbb{R}} \mathbb{C} \overset{\cong}{\longrightarrow} V \oplus \overline{V}; \quad \quad v \otimes z \mapsto (z \cdot v) \oplus (\overline{z} \cdot v)

is a isomorphism of complex representations with an inverse given by vw12(v+w)1+12i(vw)iv \oplus w \mapsto \tfrac{1}{2} (v + w) \otimes 1 + \tfrac{1}{2i}(v - w) \otimes i. Let VRU1UnV_{\mathbb{R}} \cong U_1 \oplus \cdots \oplus U_n be the decomposition of the realifcation into irreducible real represenations. Then VV occurs as a subrepresentation of (VR)C(U1)C(Un)C(V_{\mathbb{R}})^{\mathbb{C}} \cong (U_1)^{\mathbb{C}} \oplus \cdots \oplus (U_n)^{\mathbb{C}}. By irreducibility of VV, it belongs entirely to one summand (Ui)C(U_i)^{\mathbb{C}}, so VV occurs in the complexification of some irreducible real representation.

For the second step, let WW be an irreducible real representation and let VWCV \subseteq W^{\mathbb{C}} be an irreducible complex subrepresenation. There exists an equivariant complex conjugation map σEndCG(WC)\sigma \in \mathrm{End}_{\mathbb{C}G} (W^\mathbb{C}) given by σ(wz)=wz\sigma(w \otimes z) = w \otimes \overline{z}; the existence of this just says the complexification is defined over R\mathbb{R}. Since this map is an automorphism, it follows that σ(V)\sigma(V) is also an irreducible subrepresentation of WCW^\mathbb{C}, so Vσ(V)V \cap \sigma(V) is either VV or {0}\{0\}. Note that σ(V)\sigma(V) is isomorphic to the conjugate representation V\overline{V}, via the map σ(v)v\sigma(v) \mapsto v. Indeed, zσ(v)=σ(zv)zvz\sigma(v) = \sigma(\overline{z}v) \mapsto \overline{z}v. We now analyze both cases:

  1. Suppose Vσ(V)=VV \cap \sigma(V) = V. Then σ\sigma corestricts to an map σEndG(V)\sigma \in \mathrm{End}_G(V) such that σ2=id\sigma^2 = \mathrm{id} and σ(zv)=zσ(v)\sigma(zv) = \overline{z}\sigma(v) for all zCz \in \mathbb{C} and vVv \in V. It follows that VV is defined over R\mathbb{R} via the same real structure that defines WCW^\mathbb{C} over R\mathbb{R}. In particular, there are isomorphisms

    Φ ⁣:VσRCV;Ψ ⁣:(WC)σRCWC\Phi \colon V^{\sigma} \otimes_{\mathbb{R}} \mathbb{C} \overset{\cong}{\longrightarrow} V; \quad \Psi \colon (W^\mathbb{C})^{\sigma} \otimes_{\mathbb{R}} \mathbb{C} \overset{\cong}{\longrightarrow} W^\mathbb{C}

    such that Φ\Phi is the restriction of Ψ\Psi, viewing VσRCV^{\sigma} \otimes_{\mathbb{R}} \mathbb{C} as a subrepresentation of (WC)σRC(W^\mathbb{C})^{\sigma} \otimes_{\mathbb{R}} \mathbb{C}. However, (WC)σW(W^\mathbb{C})^{\sigma} \cong W is irreducible and VσV^\sigma is a nonzero real subrepresentation, so Vσ=(WC)σV^\sigma = (W^\mathbb{C})^{\sigma}. It follows that WC=VW^{\mathbb{C}} = V, which means VV is a real type representation. Furthermore, since VV is irreducible, we have

    CEndCG(V)EndCG(WC)EndRG(W)RC,\mathbb{C} \cong \mathrm{End}_{\mathbb{C}G}(V) \cong \mathrm{End}_{\mathbb{C}G}(W^\mathbb{C}) \cong \mathrm{End}_{\mathbb{R}G}(W) \otimes_\mathbb{R} \mathbb{C},

    which implies EndRG(W)R\mathrm{End}_{\mathbb{R}G}(W) \cong \mathbb{R}.

  2. Suppose that Vσ(V)={0}V \cap \sigma(V) = \{0\}. Then σ\sigma restricts to a complex conjugation map σEndG(Vσ(V))\sigma \in \mathrm{End}_G(V \oplus \sigma(V)) as before, so Vσ(V)V \oplus \sigma(V) is defined over R\mathbb{R} via the same real structure that defines WCW^\mathbb{C} over R\mathbb{R}. In particular, there are isomorphisms

    Φ ⁣:(Vσ(V))σRCVσ(V);Ψ ⁣:(WC)σRCWC\Phi \colon (V \oplus \sigma(V))^{\sigma} \otimes_{\mathbb{R}} \mathbb{C} \overset{\cong}{\longrightarrow} V \oplus \sigma(V); \quad \Psi \colon (W^\mathbb{C})^{\sigma} \otimes_{\mathbb{R}} \mathbb{C} \overset{\cong}{\longrightarrow} W^\mathbb{C}

    such that Ψ\Psi restricts to Φ\Phi. But (WC)σW(W^\mathbb{C})^{\sigma} \cong W is irreducible and (Vσ(V))σ(V \oplus \sigma(V))^\sigma is a nonzero subrepresentation, so (Vσ(V))σ=(WC)σ(V \oplus \sigma(V))^\sigma = (W^\mathbb{C})^{\sigma}. Thus WC=Vσ(V)W^{\mathbb{C}} = V \oplus \sigma(V), which is reducible. In particular, we have

    dimCEndCG(WC)=dimCEndCG(VV)=χV+χV,χV+χV={2if V≇V4if VV.\dim_{\mathbb{C}} \mathrm{End}_{\mathbb{C}G} (W^\mathbb{C}) = \dim_{\mathbb{C}} \mathrm{End}_{\mathbb{C}G} (V \oplus \overline{V}) = \inner{\chi_V + \chi_{\overline{V}}, \chi_V + \chi_{\overline{V}}} = \begin{cases} 2 &\text{if } V \not\cong \overline{V} \\ 4 &\text{if } V \cong \overline{V} \\ \end{cases}.

    Also, since WCW^{\mathbb{C}} is reducible, we have EndRG(W)RCEndCG(WC)≇C\mathrm{End}_{\mathbb{R}G}(W) \otimes_{\mathbb{R}} \mathbb{C} \cong \mathrm{End}_{\mathbb{C}G}(W^{\mathbb{C}}) \not\cong \mathbb{C}. Thus, EndRG(W)\mathrm{End}_{\mathbb{R}G}(W) must be C\mathbb{C} or H\mathbb{H}, which yields two subcases.

    • If EndRG(W)C\mathrm{End}_{\mathbb{R}G}(W) \cong \mathbb{C}, then EndCG(WC)CRC\mathrm{End}_{\mathbb{C}G}(W^{\mathbb{C}}) \cong \mathbb{C} \otimes_{\mathbb{R}} \mathbb{C}, which is a 22-dimensional C\mathbb{C}-vector space. Thus V≇VV \not\cong \overline{V}, which means that

      EndCG(WC)EndCG(VV)EndCG(V)×EndCG(V)C×C.\mathrm{End}_{\mathbb{C}G}(W^{\mathbb{C}}) \cong \mathrm{End}_{\mathbb{C}G}(V \oplus \overline{V}) \cong \mathrm{End}_{\mathbb{C}G}(V) \times \mathrm{End}_{\mathbb{C}G}(\overline{V}) \cong \mathbb{C} \times \mathbb{C}.

      Furthermore, since V≇VV \not\cong \overline{V}, the character χV\chi_V is not real-valued (or else χV\chi_V would equal χV\chi_{\overline{V}}). Thus VV is a complex type representation.

    • If EndRG(W)H\mathrm{End}_{\mathbb{R}G}(W) \cong \mathbb{H}, then EndCG(WC)HRC\mathrm{End}_{\mathbb{C}G}(W^{\mathbb{C}}) \cong \mathbb{H} \otimes_{\mathbb{R}} \mathbb{C}, which is a 44-dimensional C\mathbb{C}-vector space. Thus VVV \cong \overline{V}, which means that

      EndCG(WC)EndCG(VV)Mat2×2(C).\mathrm{End}_{\mathbb{C}G}(W^{\mathbb{C}}) \cong \mathrm{End}_{\mathbb{C}G}(V \oplus V) \cong \mathrm{Mat}_{2 \times 2} (\mathbb{C}).

      Furthermore, since VVV \cong \overline{V}, the character χV\chi_V is real-valued (since χV\chi_V equals χV\chi_{\overline{V}}). Thus VV is a quaternionic type representation.

We now relate the character of WW to the character of VV. For a real type representation, VWCV \cong W^\mathbb{C}, so the character χW\chi_W is equal to the character on the complexification χV\chi_V. On the other hand, for complex and quaternionic type representations, χWCχV+χV=2χV\chi_{W^\mathbb{C}} \cong \chi_{V} + \chi_{\overline{V}} = 2\Re \chi_V (in the quaternionic case, this is just 2χV2\chi_V, since χV\chi_V is real-valued). To summarize, we may build a dictionary between real and complex representations, where WW is an irreducible real representation and VV is any irreducible real subrepresentation of WCW^{\mathbb{C}}:

EndRG(W)\mathrm{End}_{\mathbb{R}G}(W)EndCG(V)\mathrm{End}_{\mathbb{C}G}(V)WCW^\mathbb{C}χW\chi_WIs χV\chi_V real?Is VV defined over R\mathbb{R}?FS(χV)\mathrm{FS}(\chi_V)
R\mathbb{R}C\mathbb{C}VVχV\chi_Vyesyes+1+1
C\mathbb{C}C×C\mathbb{C} \times \mathbb{C}VVV \oplus \overline{V}2χV2\Re \chi_Vnono+0\hphantom{+}0
H\mathbb{H}Mat2×2(C)\mathrm{Mat}_{2 \times 2} (\mathbb{C})VVV \oplus V2χV2\chi_Vyesno1-1

Realifying complex representations

We can also solve the inverse problem: given all the irreducible complex representations of a finite group GG, how can we determine all the irreducible real representations of GG, and what is the relationship between their characters? To do this, we

  1. show that every irreducible real representation occurs in some realifed irreducible complex representation, and
  2. realify each irreducible complex representation and write it as sum of irreducible real representations.

Again, the first step is simple: first, note that for any irreducible real representation WW, the complexification WCW^{\mathbb{C}} when treated as a real representation decomposes as

(WC)R=(WRC)R=WR(RiR)=(WRR)(WRiR).(W^{\mathbb{C}})_{\mathbb{R}} = (W \otimes_\mathbb{R} \mathbb{C})_{\mathbb{R}} = W \otimes_\mathbb{R} (\mathbb{R} \oplus i\mathbb{R}) = (W \otimes_\mathbb{R} \mathbb{R}) \oplus (W \otimes_{\mathbb{R}} i\mathbb{R}).

Let WCU1UnW^{\mathbb{C}} \cong U_1 \oplus \cdots \oplus U_n be the decomposition of WCW^{\mathbb{C}} into irreducible complex representations. Then WWRRW \cong W \otimes_\mathbb{R} \mathbb{R} occurs as a subrepresentation of (WC)R(U1)R(Un)R(W^\mathbb{C})_{\mathbb{R}} \cong (U_1)_{\mathbb{R}} \oplus \cdots \oplus (U_n)_{\mathbb{R}}. By irreducibility of WW, it must belong entirely to one summand (Ui)R(U_i)_{\mathbb{R}}, so WW occurs in the realification of some irreducible complex representation.

For the second step, let VV be an irreducible complex representation and let WVRW \subseteq V_{\mathbb{R}} be an irreducible real subrepresentation. The imaginary unit iCi \in \mathbb{C} acts on VRV_{\mathbb{R}} by "remembering" the complex action on VV, yielding an equivariant automorphism i ⁣:VRVRi \colon V_{\mathbb{R}} \to V_{\mathbb{R}}. It follows that iWiW is also irreducible, so WiWW \cap iW is either WW or {0}\{0\}. We analyze both cases.

  1. Suppose WiW=WW \cap iW = W. Then ii corestricts to a map iEndG(W)i \in \mathrm{End}_G(W) which satisfies i2=idi^2 = -\mathrm{id}, so WW can be given the structure of a nonzero complex subrepresentation of VV. By irreducibility, VV must equal WW. It also follows that EndRG(W)\mathrm{End}_{\mathbb{R}G}(W) must be equal to C\mathbb{C} or H\mathbb{H}, which yields two subcases.

    • An isomorphism HEndRG(W)EndRG(V)\mathbb{H} \cong \mathrm{End}_{\mathbb{R}G}(W) \cong \mathrm{End}_{\mathbb{R}G}(V) is the same thing as a quaternionic structure HEndG(V)\mathbb{H} \to \mathrm{End}_{G}(V) on VV, so the latter case occurs if and only if VV is of quaternionic type.

    • An isomorphism CEndRG(W)EndRG(V)\mathbb{C} \cong \mathrm{End}_{\mathbb{R}G}(W) \cong \mathrm{End}_{\mathbb{R}G}(V) is the same thing as saying VV has a complex structure but not a quaternionic structure, which means VV is of complex type.

  2. Suppose WiW={0}W \cap iW = \{0\}. Then ii corestricts to a map iEndG(WiW)i \in \mathrm{End}_G(W \oplus iW) which satisfies i2=idi^2 = -\mathrm{id}, so WiWW \oplus iW can be given the structure of a nonzero complex subrepresenation of VV. By irreducibility, VV must equal WiWW \oplus iW. Thus, we have the isomorphism of complex representations

    VWiW(WRR)(WRiR)WC.V \cong W \oplus iW \cong (W \otimes_{\mathbb{R}} \mathbb{R}) \oplus (W \otimes_{\mathbb{R}} i\mathbb{R}) \cong W^{\mathbb{C}}.

    Note that the individual summands WW and iWiW are not complex subrepresentations, but the direct sum (as real representations) is a complex representation, as the action of ii moves between summands. Since VV is irreducible, we have

    CEndCG(V)EndCG(WC)EndRG(W)RC,\mathbb{C} \cong \mathrm{End}_{\mathbb{C}G}(V) \cong \mathrm{End}_{\mathbb{C}G}(W^\mathbb{C}) \cong \mathrm{End}_{\mathbb{R}G}(W) \otimes_\mathbb{R} \mathbb{C},

    which implies EndCG(W)=R\mathrm{End}_{\mathbb{C}G}(W) = \mathbb{R}. This is not isomorphic to EndRG(V)\mathrm{End}_{\mathbb{R}G}(V) as in the last case; instead, we have

    EndRG(V)EndRG(WiW)EndRG(WW)Mat2×2(R).\mathrm{End}_{\mathbb{R}G}(V) \cong \mathrm{End}_{\mathbb{R}G}(W \oplus iW) \cong \mathrm{End}_{\mathbb{R}G}(W \oplus W) \cong \mathrm{Mat}_{2 \times 2}(\mathbb{R}).

    Finally, since VWCV \cong W^{\mathbb{C}} is the complexification of a real representation, it is of real type.

We now relate the character of WW to the character of VV. For a real type representation, VWCV \cong W^\mathbb{C}, so the character χW\chi_W is equal to the character on the complexification χV\chi_V. On the other hand, for complex and quaternionic type representations, VWV \cong W, so assign an eigenbasis e1(g),,en(g)Ve_1(g), \ldots, e_n(g) \in V to each gGg \in G and let gek=(xk+iyk)ekg e_k = (x_k + iy_k)e_k. Since VRV_{\mathbb{R}} has basis e1(g),ie1(g),,en(g),ien(g)e_1(g), ie_1(g), \ldots, e_n(g), ie_n(g), the matrix elements for gg on WW is

[x1y1y1+x1xnynyn+xn].\begin{bmatrix} x_1 & -y_1 & & & \\ y_1 & \hphantom{+}x_1 & & & \\ & & \ddots & & \\ & & & x_n & -y_n \\ & & & y_n & \hphantom{+}x_n \end{bmatrix}.

It follows that the real character is χW=2χV\chi_W = 2 \Re \chi_V (in the quaternionic case, this is just 2χV2\chi_V, since χV\chi_V is real-valued). To summarize, we may build another dictionary between complex and real representations, where VV is an irreducible complex representation and WW is any irreducible real subrepresentation of VRV_{\mathbb{R}}:

EndRG(W)\mathrm{End}_{\mathbb{R}G}(W)EndRG(V)\mathrm{End}_{\mathbb{R}G}(V)VRV_\mathbb{R}χW\chi_WIs χV\chi_V real?Is VV defined over R\mathbb{R}?FS(χV)\mathrm{FS}(\chi_V)
R\mathbb{R}Mat2×2(R)\mathrm{Mat}_{2 \times 2}(\mathbb{R})WiWW \oplus iWχV\chi_Vyesyes+1+1
C\mathbb{C}C\mathbb{C}WW2χV2\Re \chi_Vnono+0\hphantom{+}0
H\mathbb{H}H\mathbb{H}WW2χV2\chi_Vyesno1-1