is an isomorphism, and V = U W. Proof. Let u 1,..., u m be a basis of U, and add linearly independent

Similar documents
Representation Theory

a(b + c) = ab + ac a, b, c k

INTRODUCTION TO REPRESENTATION THEORY AND CHARACTERS

REPRESENTATION THEORY NOTES FOR MATH 4108 SPRING 2012

Representations. 1 Basic definitions

M3/4/5P12 GROUP REPRESENTATION THEORY

MAT 445/ INTRODUCTION TO REPRESENTATION THEORY

Notes on nilpotent orbits Computational Theory of Real Reductive Groups Workshop. Eric Sommers

Real representations

Math 594. Solutions 5

Groups and Representations

PMATH 745: REPRESENTATION THEORY OF FINITE GROUPS

ALGEBRA QUALIFYING EXAM PROBLEMS

Since G is a compact Lie group, we can apply Schur orthogonality to see that G χ π (g) 2 dg =

REPRESENTATION THEORY. WEEK 4

Exercises on chapter 4

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

REPRESENTATION THEORY OF S n

Representation Theory. Ricky Roy Math 434 University of Puget Sound

Math 250: Higher Algebra Representations of finite groups

Representation Theory

Additional notes on group representations Hjalmar Rosengren, 30 September 2015

L(C G (x) 0 ) c g (x). Proof. Recall C G (x) = {g G xgx 1 = g} and c g (x) = {X g Ad xx = X}. In general, it is obvious that

CHARACTERS OF FINITE GROUPS.

18.702: Quiz 1 Solutions

Representations of algebraic groups and their Lie algebras Jens Carsten Jantzen Lecture III

Algebra Exam Topics. Updated August 2017

SOME ELEMENTARY RESULTS IN REPRESENTATION THEORY

Group Representation Theory

GRE Subject test preparation Spring 2016 Topic: Abstract Algebra, Linear Algebra, Number Theory.

REPRESENTATIONS AND CHARACTERS OF FINITE GROUPS

LINEAR ALGEBRA BOOT CAMP WEEK 4: THE SPECTRAL THEOREM

REPRESENTATION THEORY, LECTURE 0. BASICS

(d) Since we can think of isometries of a regular 2n-gon as invertible linear operators on R 2, we get a 2-dimensional representation of G for

A PRIMER ON SESQUILINEAR FORMS

0.2 Vector spaces. J.A.Beachy 1

Topics in Representation Theory: Fourier Analysis and the Peter Weyl Theorem

A PROOF OF BURNSIDE S p a q b THEOREM

Linear algebra 2. Yoav Zemel. March 1, 2012

Math 121 Homework 5: Notes on Selected Problems

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

REPRESENTATION THEORY FOR FINITE GROUPS

REPRESENTATION THEORY WEEK 5. B : V V k

The Cartan Decomposition of a Complex Semisimple Lie Algebra

A Little Beyond: Linear Algebra

Notation. For any Lie group G, we set G 0 to be the connected component of the identity.

TENSOR PRODUCTS, RESTRICTION AND INDUCTION.

INTRODUCTION TO LIE ALGEBRAS. LECTURE 2.

Topics in linear algebra

Weeks 6 and 7. November 24, 2013

Elementary linear algebra

INTRODUCTION TO LIE ALGEBRAS. LECTURE 7.

1.1 Definition. A monoid is a set M together with a map. 1.3 Definition. A monoid is commutative if x y = y x for all x, y M.

NOTES ON POLYNOMIAL REPRESENTATIONS OF GENERAL LINEAR GROUPS

Algebra Exam Syllabus

Finite Group Representations. for the. Pure Mathematician

Character tables for some small groups

Math Linear Algebra

Some notes on linear algebra

LECTURE 3: TENSORING WITH FINITE DIMENSIONAL MODULES IN CATEGORY O

3 Representations of finite groups: basic results

Answers in blue. If you have questions or spot an error, let me know. 1. Find all matrices that commute with A =. 4 3

MATH SOLUTIONS TO PRACTICE MIDTERM LECTURE 1, SUMMER Given vector spaces V and W, V W is the vector space given by

x y B =. v u Note that the determinant of B is xu + yv = 1. Thus B is invertible, with inverse u y v x On the other hand, d BA = va + ub 2

Notes on Representations of Finite Groups. Ambar N. Sengupta

First we introduce the sets that are going to serve as the generalizations of the scalars.

The Gelfand-Tsetlin Basis (Too Many Direct Sums, and Also a Graph)

ABSTRACT ALGEBRA 1, LECTURES NOTES 5: SUBGROUPS, CONJUGACY, NORMALITY, QUOTIENT GROUPS, AND EXTENSIONS.

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

Final Review Sheet. B = (1, 1 + 3x, 1 + x 2 ) then 2 + 3x + 6x 2

FFTs in Graphics and Vision. Groups and Representations

Test 1 Review Problems Spring 2015

Linear Algebra. Min Yan

Honors Algebra 4, MATH 371 Winter 2010 Assignment 4 Due Wednesday, February 17 at 08:35

UC Berkeley Summer Undergraduate Research Program 2015 July 9 Lecture

MATH 101: ALGEBRA I WORKSHEET, DAY #3. Fill in the blanks as we finish our first pass on prerequisites of group theory.

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

Representations of algebraic groups and their Lie algebras Jens Carsten Jantzen Lecture I

LECTURE 3: REPRESENTATION THEORY OF SL 2 (C) AND sl 2 (C)

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

Fall /29/18 Time Limit: 75 Minutes

LINEAR ALGEBRA BOOT CAMP WEEK 2: LINEAR OPERATORS

Chapter 2 Linear Transformations

1 Invariant subspaces

Linear Algebra Final Exam Solutions, December 13, 2008

Yale University Department of Mathematics Math 350 Introduction to Abstract Algebra Fall Midterm Exam Review Solutions

Review of Linear Algebra

IDEAL CLASSES AND RELATIVE INTEGERS

Supplementary Notes March 23, The subgroup Ω for orthogonal groups

MAT2342 : Introduction to Applied Linear Algebra Mike Newman, fall Projections. introduction

YOUNG-JUCYS-MURPHY ELEMENTS

LINEAR ALGEBRA REVIEW

LECTURE 25-26: CARTAN S THEOREM OF MAXIMAL TORI. 1. Maximal Tori

ELEMENTARY SUBALGEBRAS OF RESTRICTED LIE ALGEBRAS

Exercises Lie groups

CHARACTERS AS CENTRAL IDEMPOTENTS

Two subgroups and semi-direct products

Categories and Quantum Informatics: Hilbert spaces

ALGEBRA QUALIFYING EXAM, FALL 2017: SOLUTIONS

Math 321: Linear Algebra

Transcription:

Lecture 4. G-Modules PCMI Summer 2015 Undergraduate Lectures on Flag Varieties Lecture 4. The categories of G-modules, mostly for finite groups, and a recipe for finding every irreducible G-module of a finite group G. Let s start by thinking about direct sums of vector spaces: Definition 4.1. A vector space V is the direct sum: V = U 1 U m of subspaces U 1,..., U m V if each v V has a unique expression: v = u 1 +... + u m as a sum of vectors u i U i from each subspace. Example 4.1. (a) If v 1,..., v n V are a basis, then V is the direct sum of the one-dimensional subspaces (i.e. lines) L i := {c v i c k}. (b) If U, W V are a pair of subspaces that satisfy: U W = 0 and dim(u) + dim(w ) = dim(v ), then V = U W. (Exercise 4.1) (c) Three distinct lines L 1, L 2, L 3 in the xy-plane inside R 3 satisfy: L i L j = 0 and dim(l 1 ) + dim(l 2 ) + dim(l 3 ) = 3 but R 3 is not a direct sum of L 1, L 2, L 3. Remark. Since every finite dimensional vector space has a basis, it follows that every vector space is the direct sum of lines. For this reason, we say that all vector spaces other than lines and 0 are reducible. Proposition 4.1. Suppose V has finite dimension and U V is a subspace. Then there is a complementary subspace W V such that: q : W V/U is an isomorphism, and V = U W. Proof. Let u 1,..., u m be a basis of U, and add linearly independent vectors (in V ) until a basis: u 1,..., u m, w 1,..., w n m of V is obtained. Then W = {c 1 w 1 + + c n m w n m } has the desired properties. Remark. The analogue of this Proposition is false for abelian groups. Consider, for example, the subgroup of even integers: 2Z Z with coset group Z/2Z Evidently there is no subgroup A Z that maps isomorphically to Z/2Z, because such a subgroup would produce an integer a Z with the property that 2a = 0. The question of whether a group of cosets G/H of a normal subgroup lifts to a subgroup of G is subtle. 1

2 Let G be a group. Definition 4.2. (a) A G-module is a pair consisting of a vector space V (over a field k) together with a representation ρ : G GL(V ). (b) A subspace U V fixed by ρ(g) for all g G is a G-submodule. (c) A linear map f : V W of G-modules is G-linear if: f(g v) = gf( v) for all g G and all v V Remark. Although a G-module is a pair, it is often just written as V, with ρ understood. Similarly, unless we need to be careful, we ll write g v when we mean ρ(g)( v) (as we did in (c) above). Exercise 4.2. (a) The identity map on G-modules is G-linear. (b) A composition of G-linear maps is G-linear. (c) The inverse of an invertible G-linear map is G-linear. (d) The kernel and image of a G-linear map are G-submodules. (e) The space of cosets of a G-linear map is a G-module in such a way that the the map q : V V/U is G-linear. Moment of Zen. The category GMod (fixed G) consists of: (a) the collection of all G-modules, and (b) all G-linear maps Example 4.2. (a) The trivial representation (ρ(g) = id V for all g) makes any vector space into a trivial G-module. Every subspace of a trivial G-module is a G-submodule. (b) Let G = Perm(n) and consider the G-module k n with the natural representation τ e i = e τ(i). Then the addition of coordinates: a : k n k; a( v) = v 1 +... + v n is a G-linear map to the trivial G-module k since: n τ v = (v τ 1 (1),..., v τ 1 (n)) and a(τ v) = v τ 1 (i) = a( v) = τa( v) The G-submodule ker(a) is the standard representation of Perm(n). (c) Consider the two-dimensional standard representation of Perm(3). In terms of the basis e 1 e 2, e 2 e 3, matrices for the action include: [ ] [ ] 1 1 1 0 ρ(1 2) = and ρ(2 3) = 0 1 1 1 i=1 The other matrices are left as an exercise.

Definition 4.3. A G-module V with representation ρ is irreducible if the only G-submodules of V are { 0} and V itself. Example 4.3. (a) Any one-dimensional G-module is irreducible. These include the trivial one-dimensional G-modules, but also any character: χ : G k ; χ(gh) = χ(g) χ(h) E.g. The characters of C n from Example 3.7(a) are χ m (x) = e 2πim n. Consider next the action of D 2n from Example 3.7(b), consisting of the rotation action of C n augmented by the reflections (b) The rotation action of C n D 2n on R 2 is already irreducible when n > 2 because no line is left invariant under the rotations. But: (c) The same rotation action of C n on C 2 is reducible, with two invariant lines, the eigenspaces of x (and all powers of x), spanned by: (1, i) and (1, i), respectively (d) Since the reflection given by y switches the two invariant lines for x, it follows that C 2 with the action of D 2n is irreducible. Proposition 4.2. The only irreducible G-modules over C for an abelian group G are the one-dimensional characters of G. In other words, any set of commuting matrices has a common eigenvector. Proof. If V is a vector space over C, then every element of GL(V ) has an eigenvector. Suppose A and B are commuting elements of GL(V ), and v is an eigenvector of A with eigenvalue λ. Then: A(B v) = B(A v) = B(λ v) = λ(b v) so B v is another eigenvector of A with the same eigenvalue λ. If A 1,..., A n are elements of GL(V ) with a common eigenvector and A n+1 commutes with each of them, then by the same argument, A n+1 acts on the common eigenspaces and therefore shares an eigenvector with A 1,..., A n. This proves that every finite set of commuting matrices shares an eigenvector. If an infinite set of commuting matrices failed to share an eigenvector, then some finite subset would also fail to share an eigenvector (by induction on the dimensions of the shared eigenspaces), so the result applies to infinite abelian groups as well. Example 4.4. Consider the two-dimensional representation of the abelian group C (with addition): [ ] 1 z ρ(z) = 0 1 These matrices share a single one-dimensional eigenspace, the x-axis. 3

4 Definition 4.4. A G-module V is the direct sum: V = U 1 U m of G-submodules U 1,..., U m V if each v V has a unique expression: v = u 1 +... + u m as a sum of vectors u i U i from each submodule. Remark. This is identical to Definition 4.1, but Proposition 4.1 for G- modules is more subtle. The problem is that complementary subspaces to a G-submodule U V are not usually G-submodules. In Example 4.4 above, the y-axis is complementary to the x-axis, but it is not invariant under the action of C, and in fact, in that Example, there is no complementary subspace of C 2 that is a C-submodule! Consider again the dot product on R n from Lecture 2. This is bilinear (linear on each vector) and positive definite: n v v = vi 2 > 0 unless v = 0 i=1 Similarly, there is a standard Hermitian product on C n, namely, n v, w = v j w j C j=1 that is linear in the first vector and conjugate linear in the second, i.e. v 1 + v 2, w = v 1, w + v 2, w and v, w 1 + w 2 = v, w 1 + v, w 2 and c v, w = c v, w but v, c w = c v, w. The Hermitian product is conjugate symmetric, in the sense that v, w = w, v. Remark. We have avoided complex conjugation thus far, but no longer. If c = a + ib, then c = a ib, and cc = a 2 + b 2 = c 2. It follows that the Hermitian dot product is also positive definite, in the sense that: n v, v = v i 2 is real and positive unless v = 0 i=1 Let U V be a subspace of a vector space V defined over R or C. Definition 4.5. The orthogonal complement U V is the set of vectors in V such that u v = 0 (if over R) or u, v = 0 (if over C). Exercise 4.3. When V is a vector space over R or C, then U V is a complementary subspace to U V in the sense of Proposition 4.1. In particular, V = U U and dim(u) + dim(u ) = dim(v )

This doesn t (yet) help with our conundrum. If U V is a G- submodule, there is no reason to expect that U is also a G-submodule. The key is to recognize that there are many positive definite products and to find one that is G-invariant. Definition 4.6. (a) Let V be a vector space over R. A map b : V V R is a positive definite inner product if it is symmetric and bilinear and b( v, v) > 0 for all v 0 (b) Let V be a vector space over C. A map h : V V C is a positive definite Hermitian product if it is conjugate symmetric, and linear in the first vector (and conjugate linear in the second), and h( v, v) > 0 for all v 0 Note. h( v, v) = h( v, v) is automatically a real number. Matrix Notation. An inner product is represented by the matrix: B = (b(e i, e j )) with B = B T and b( v, w) = ( v) T B w where v and w are regarded as one-column matrices. Notice that the ordinary dot product is represented by the identity matrix. Similarly, a Hermitian product is represented by the matrix: H = (h(e i, e j )) with H = H T and h( v, w) = ( v) T H w and the standard Hermitian product is represented by the identity. The orthogonal complement W of a subspace W V can now be taken relative to any positive definite inner (or Hermitian) product and the conclusions from Exercise 4.3 hold, namely that V = W W. Let G be a finite group. Definition 4.7. If V is a G-module over R (or C), then the G-invariant inner product (or Hermitian product) on V is: b( v, w) = 1 g v g w or h( v, w) = 1 g v, g w G G g G Proposition 4.3. The orthogonal complement W V of a G- submodule W V with respect to the G-invariant inner (or Hermitian) product on V is another G-submodule of V. g G 5

6 Proof. The G-invariance of the inner product gives: b( v, w) = 1 g v g w = 1 (gh) v (gh) w = b(h v, h w) G G g G for any h G. Suppose v W, i.e. b( v, w) = 0 for all w W. Then b(h v, w) = b(h v, hh 1 w) = 0 for all w W, because h 1 w W. Corollary 4.1. Suppose V is a vector space over C and A GL(V ) has finite order, i.e. A n = id for some n. Then A is semi-simple. Proof. Let G = C n and let V be a G-module via ρ(x) = A. We know that the matrix A has (at least) one eigenvector v V with eigenvalue λ. The line L spanned by v is a G-submodule of V, whose orthogonal complement under the G-invariant Hermitian product is a G-submodule. The matrix A acts on the orthogonal complement L with an eigenvector in L, and then A acts on the G-invariant orthogonal complement to that eigenvector in L, etc. g G Example 4.5. Consider the matrix from Example 4.2(c): [ ] 1 1 A = ; A 0 1 2 = id Then R 2 is a C 2 -space, and the eigenvector (1, 0) spans an invariant subspace. The C 2 -invariant inner product is: b( v, w) = 1 ( v w + A v A w) 2 which is represented by the matrix: [ ] 1 B = 1 2 Then b((1, 2), (1, 0)) = 0 so (1, 2) is the other eigenvector of A. 1 2 Corollary 4.2. When G is a finite group, any finite-dimensional G- module V over R or C is completely reducible, i.e. 3 2 V = U 1 U 2 U n for some set of irreducible G-submodules U 1,..., U n V. Proof. Choose a non-zero vector v V and consider the G-submodule spanned by the set of vectors W = {g v g G}. This is evidently irreducible. With respect to the invariant inner product (or Hermitian inner product), the orthogonal complement W is also a G-submodule, and V = W W. By induction on the dimension, W is a direct sum of irreducibles, and then so is V.

Since every G-module is a direct sum of finitely many irreducible G-modules, it makes sense to try to classify the irreducible G-modules. We will do this over the complex numbers, where, for example, the irreducible G-modules for an abelian group were seen to be the onedimensional characters. As usual, let G be a finite group. Definition 4.8. The (left) regular G-module over C is: C[G] = e g g G with the action ρ(h)(e g ) = e hg i.e. the vector space with one basis vector e g for each g G, made into a G-module by permuting basis vectors by left multiplication. This is similar to the natural representation of the permutation group, in which the vectors corresponded to elements of the set [n] (rather than elements of the group Perm(n)). Example 4.6. Let G = C n with the ordering {id, x,..., x n 1 }. Then: ρ(x)(e x i) = e x i+1 cyclically permutes the basis vectors, so the matrix for the action is: 0 0 1 1 0 0 A = ρ(x) = 0 1 0.... 0 0 0 Let ζ = e 2πi n. Then: { v ζ m = e 1 + ζ m e 2 + + ζ m(n 1) e n 0 m < n} is a basis of eigenvectors for A (and all powers of A), hence: C[C n ] = C v ζ C v ζ 2 C v ζ n Notice that C v ζ m = χ m as a one-dimensional character of C n, since: A v ζ m = e 2 + ζ m e 3 + + ζ m(n 1) e 1 = ζ m v ζ m and in particular, each summand in the direct sum is a distinct onedimensional character, with the last being the trivial character. Theorem 5.1. Let C[G] = n i=1 U i be a decomposition of the regular representation into irreducible G-submodules over C. Then: (a) Every irreducible G-module U is isomorphic to a summand. (b) If dim(u) = d, then U is isomorphic to d of the summands. 7

8 Corollary 4.3. There are only finitely many irreducible G-modules (up to isomorphism), and their dimensions d 1,..., d m satisfy: G = d 2 1 + d 2 2 + + d 2 m Proof. Immediate from the Theorem, since n i=1 dim U i = G. Before we get to the proof of this, consider some examples: (a) We ve seen C[C n ] as a direct sum of its n characters. (b) Since the only irreducible representations of an abelian group G are one-dimensional characters, it follows that there are G of them, and that C[G] = χ is the direct sum of the distinct characters. (c) We ve seen three irreducible complex representations of Perm(3): trivial, sign and standard (two-dimensional) and since 1 + 1 + 2 2 = 6 = Perm(3), there are no other irreducibles. In particular, since Perm(3) = D 6, the rotation/reflection irreducible representation must be isomorphic to the standard representation. The key ingredient in this proof is Schur s Lemma, which requires us to (finally) think a bit more about the structure of the sets of linear maps between two vector spaces (or G-modules). Observation. Let V and W be vector spaces over k. Then the set of linear maps (morphisms) from V to W is denoted by Hom k (V, W ) and it is itself a vector space, via: (f + g)( v) = f( v) + g( v) and (cf)( v) = f(c v) Remark. We ve already seen one example, namely V = Hom k (V, k). If dim(v ) = m and dim(w ) = n, then dim(hom k (V, W )) = mn. Indeed: Hom k (k m, k n ) = {m n matrices} and the addition is the componentwise addition of matrices. If V and W are G-modules, the set Hom G (V, W ) of G-linear maps is also a vector space. It is a subspace: Hom G (V, W ) Hom k (V, W ) since G-linear maps sum (and scalar multiply) to G-linear maps. Schur s Lemma. (a) If V, W are irreducible G-modules, then: Hom G (V, W ) = { 0} unless V and W are isomorphic (b) If V, W are isomorphic G-modules over C, then: dim(hom G (V, W )) = 1 In particular, End G (V ) = C id U for an irreducible G-module over C.

Proof. The only G-submodules of an irreducible V are 0 and V. If f Hom G (V, W ) consider the G-submodule kernel and image of f. Either ker(f) = { 0} (f is injective), or ker(f) = V (f is zero). Either im(f) = { 0}, (f is zero) or im(f) = W (f is surjective). Thus either f is zero or else f is a bijection, i.e. an isomorphism. This gives (a). For (b), suppose f, g Hom G (V, W ) are an arbitrary pair of nonzero G-linear maps. By (a) they are isomorphisms, so we may consider: id W = g g 1 and h = f g 1 Hom G (W, W ) Then for each scalar λ C, we have: h λ id W Hom G (W, W ) and when λ is an eigenvalue of h (which happens for at least one λ), this G-linear map is not an isomorphism. For such a λ, then, h λ id W = 0, hence λ id W = h and f = h g = λ g, so f is a multiple of g. This proves that Hom G (V, W ) is one-dimensional. Remark. Part (b) of Schur s Lemma doesn t hold if the field is R. For example, if A GL(2, R) is the rotation by 2π/n, then A End Cn (R 2 ) for the irreducible rotation action. Evidently A λ id for any λ. Proof of Theorem 5.1. Let U be an irreducible G-module, and let u U be a nonzero vector. Then the linear map: f u : C[G] U defined by f u (e g ) = g u is G-linear, because f u (he g ) = f u (e hg ) = (hg) u = hf u (e g ) for all h G. This map is not the zero map, since f u (e id ) = u. But if: C[G] = U 1 U 2 U n is the decomposition of C[G], then on each summand, f u : U i U is either zero or an isomorphism, by Schur s Lemma (a), so (at least) one of them is a isomorphism! This gives (a) of the Theorem. For (b) of the Theorem, consider the linear evaluation map: φ : Hom G (C[G], U) U; φ(f) = f(e id ) If φ(f) = u, then f(e g ) = gf(e id ) = g u, so f = f u defined above! And if φ(f) = 0, then f(e g ) = 0 for all g G, and f is the zero map. So φ is an isomorphism. This means in particular that dim(hom G (C[G], U)) = dim(u). But by Schur Lemma (b), the dimension of Hom G (U i, U) = 1 for each irreducible summand of C[G] that is isomorphic to U, and it follows that there are dim(u) of them. 9

10 Exercises. 4.1. Prove that V = U W for subspaces U, W V if and only if: (i) U W = 0 and (ii) dim(u) + dim(w ) = dim(v ). 4.2. (a) Show that G-modules with G-linear maps are a category. (b) Show that the kernel and image of a G-linear map are G-modules. (c) Show that the space of cosets of a G-submodule is a G-module. 4.3. Prove that the orthogonal complement of a subspace U V of a vector spaces over R or C satisfies: V = U U One cheap way to get new irreducible representations from old ones is to twist by a character, i.e. given representations ρ : G GL(U) and χ : G C, consider ρ χ GL(U) defined by g u = χ( u) ρ( u). 4.5. What conditions on ρ and χ guarantee that the G-module on U for ρ χ is not isomorphic to the G-module for ρ? 4.6. Find all the irreducible representations of: (a) D 8 and D 10 (do you see a pattern?) (b) Alt(4) (use the map Alt(4) C 3 = Alt(4)/K 4.) (c) Perm(4) (use the map Perm(4) Perm(3) = Perm(4)/K 4 ) (d) Alt(5) The trace tr(ρ(g)) of an irreducible representation ρ evaluated at an element g G is the same for all elements of the same conjugacy class. This is not surprising. It is a consequence of Exercise 2.9. What is surprising is the following: The number of irreducible representations matches the number of conjugacy classes. There is a Hermitian inner product on the complex vector space V = c C e c (one e c for each conjugacy class c in the group G) such that the vectors v ρ = c trρ(g c)e c are an orthonormal basis of V (length one and orthogonal to each other). 4.7. Find the Hermitian inner product (Hint: It involves the sizes of each conjugacy class) and play with this. Compute the orthonormal basis (a.k.a. the character table) of cyclic groups and the groups whose representations you found in 4.6.