You are on page 1of 111

Group theory for physicists

Literature: I.V. Schensted, A course on the applications of group theory to quantum mechanics W.-K. Tung, Group theory in physics H. Boerner, Darstellungen von Gruppen R. Gilmore, Lie groups, Physics, and Geometry H.F. Jones, Groups, representations and physics J.F. Cornwell, Group theory in physics S. Sternberg, Group theory and physics R. Gilmore, Lie groups, Lie algebras, and some of their applications D.B. elobenko, Compact Lie groups and their representations

Contents
1. Introduction 1.1. Why group theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2. Basic denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3. Examples and general properties . . . . . . . . . . . . . . . . . . . 1.4. Invariant subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5. The symmetric (or permutation) group . . . . . . . . . . . . . . . . 1.6. The action of a group on a set . . . . . . . . . . . . . . . . . . . . 1.7. Equivalence classes and invariant subgroups . . . . . . . . . . . . . 1.8. Cosets and factor groups . . . . . . . . . . . . . . . . . . . . . . . . 1.9. Direct product of two groups . . . . . . . . . . . . . . . . . . . . . 1.10. Homomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.11. Example: Homomorphism between S (2, C) and the Lorentz group 2. Matrix representations 2.1. Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Equivalent representations . . . . . . . . . . . . . . . . . . . 2.3. Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4. Reduction of representations . . . . . . . . . . . . . . . . . . 2.4.1. Reducibility of invariant spaces and representations 2.4.2. OA operators using the example of D3 . . . . . . . . 2.4.3. Four theorems and their consequences . . . . . . . . 2.4.4. Characters of representations . . . . . . . . . . . . . 2.5. The regular representation . . . . . . . . . . . . . . . . . . . 2.6. Product representations and Clebsch-Gordan coecients . . 2.7. Subduced and induced representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 5 6 7 9 11 12 13 15 17 18 20 23 23 24 25 26 26 28 31 32 34 36 40 41 41 42 44 48 48 48 51 52 57 57 57 58 60 61

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

3. Applications in quantum mechanics 3.1. Selection rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. The symmetry group of the Hamiltonian and degeneracies . . . . . . . . . . . . 3.3. Perturbation theory and lifting of degeneracies . . . . . . . . . . . . . . . . . . 4. Expansion in irreducible basis vectors 4.1. Irreducible basis vectors . . . . . . . . . . . . . . . . . 4.2. Projection operators on irreducible bases . . . . . . . . 4.3. Irreducible operators and the Wigner-Eckart theorem 4.4. Left ideals and idempotents . . . . . . . . . . . . . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

5. Representation of the symmetric group and Young diagrams 5.1. Why Sn is important . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. 1-dimensional and associated representations of Sn . . . . . . . . 5.3. Young diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4. Symmetrizers and antisymmetrizers of Young tableaux . . . . . . 5.5. Irreducible representations of Sn . . . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

Contents

5.6. More applications of Young tableaux . . . . . . . . . . . . . . . . . . . . . . . . 6. Lie 6.1. 6.2. 6.3. 6.4. 6.5. 6.6. 6.7. groups Introduction . . . . . . . . . . . . . . . . . . . . . Examples of Lie groups . . . . . . . . . . . . . . Invariant integration . . . . . . . . . . . . . . . . Properties of compact Lie groups . . . . . . . . . Generators of Lie groups . . . . . . . . . . . . . . SO(2) . . . . . . . . . . . . . . . . . . . . . . . . SO(3) . . . . . . . . . . . . . . . . . . . . . . . . 6.7.1. Angle-and-axis parametrization . . . . . . 6.7.2. Euler angles . . . . . . . . . . . . . . . . . 6.7.3. Generators . . . . . . . . . . . . . . . . . 6.7.4. Irreps of SO(3) . . . . . . . . . . . . . . . 6.8. SU(2) . . . . . . . . . . . . . . . . . . . . . . . . 6.8.1. Parametrization . . . . . . . . . . . . . . 6.8.2. Relationship between SO(3) and SU(2) . 6.8.3. Inavriant integration . . . . . . . . . . . . 6.8.4. Orthogonality and completeness relations

62 65 65 66 69 69 70 73 75 75 76 76 77 79 79 80 81 82 83 83 83 84 85 85 87 88 88 89 89 89

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

7. Lorentz and Poincar group 7.1. Relativistic kinematics . . . . . . . . . . . . . . . . . . . 7.2. Generators and Lie algebra . . . . . . . . . . . . . . . . 7.3. Finite-dimensional representations of the Lorentz group 7.4. Unitary irreps of the Poincar group . . . . . . . . . . . 7.4.1. One-particle states and Casimir operators . . . . 7.4.2. Little group and induced representations . . . . . 7.4.3. Massive particles . . . . . . . . . . . . . . . . . . 7.4.4. Massless particles . . . . . . . . . . . . . . . . . . 7.4.5. Tachyons . . . . . . . . . . . . . . . . . . . . . . 7.4.6. Vacuum . . . . . . . . . . . . . . . . . . . . . . . 7.5. Parity and time reversal . . . . . . . . . . . . . . . . . . 8. The 8.1. 8.2. 8.3.

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

8.4. 8.5. 8.6. 8.7.

tensor method to construct irreps of Gl(m) and its subgroups Tensors and tensor spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . Action of the symmetric group on the tensor space . . . . . . . . . . . . . Decomposition of the tensor space into irreducible subspaces under Sn and 8.3.1. Symmetry classes in tensor space . . . . . . . . . . . . . . . . . . . 8.3.2. Totally symmetric and totally antisymmetric tensors . . . . . . . . 8.3.3. Tensors with mixed symmetry . . . . . . . . . . . . . . . . . . . . 8.3.4. Complete reduction of the tensor space . . . . . . . . . . . . . . . Irreps of U(m ) and SU(m ) . . . . . . . . . . . . . . . . . . . . . . . . . . . Complex conjugate representation . . . . . . . . . . . . . . . . . . . . . . Reduction of the product of two irreps . . . . . . . . . . . . . . . . . . . . Applications in (particle) physics . . . . . . . . . . . . . . . . . . . . . . . 8.7.1. The Goldstone theorem . . . . . . . . . . . . . . . . . . . . . . . . 8.7.2. SU(2) isospin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7.3. SU(2) avor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7.4. SU(3) avor and the quark model . . . . . . . . . . . . . . . . . .

90 . . . 90 . . . 91 Gl(m) 92 . . . 92 . . . 93 . . . 94 . . . 95 . . . 96 . . . 97 . . . 98 . . . 100 . . . 100 . . . 100 . . . 101 . . . 104

Contents

A. Appendix 108 A.1. Proof of Cayleys theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 A.2. Proofs of the four theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 A.3. The Kronecker product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

1. Introduction
1.1. Why group theory
group theory = description of symmetries and their consequences (symmetry = invariance under a particular transformation) important in many elds of physics one can learn a lot about a system by analyzing its symmetries (without knowing the dynamics) examples of symmetries in physics: 1) continuous space-time symmetries: a) homogeneity of space = invariance under spacial translations xx+a conservation of momentum b) homogeneity of time = invariance under temporal translations t t + t0 conservation of energy c) isotropy of space = invariance under rotations x Ax convervation of angular momentum d) space-time symmetry (special relativity) = invariance under Poincar transformations (x, t) (x, t) + a see Sec. 7 2) discrete space-time symmetries: a) parity transformation: x x b) time reversal: t t c) discrete translations on a lattice d) discrete rotational symmetries on a lattice 3) permutation symmetries (systems of several identical particles) 4) gauge invariance in electrodynamics (classical and QED) conservation of charge 5) internal symmetries in nuclear and particle physics (spin, isospin, color, etc.) particle spectrum (e.g., degeneracies in the spectrum)

1. Introduction

1.2. Basic denitions


group: a group G is a set of elements for which we can dene a "multiplication" and which staises the following four properties: 1) if A and B are group elements, then AB is also a group element (closure of the group) 2) the multiplication is associative: A(BC ) = (AB )C 3) the group contains one and only one identity element I : for all A G we have AI = IA = A 4) for each A G there is one and only one inverse element A1 G such that AA1 = A1 A = I remarks: the group is dened by its elements and the multiplication rule, i.e., there can be two dierent groups with the same elements but with a dierent multiplication rule in general the multiplication is not commutative, i.e., AB = BA; if the multiplication is commutative, the group is called Abelian if the number of group elements is nite: nite group; then order of the group = number of group elements otherwise: innite group which can be discrete (e.g., translations on an innite lattice) or continuous (e.g., continuous rotations) subgroup: if a subset H G satises the four group properties (with the multiplication rule for G) it is called a subgroup of G each group has two trivial subgroups: {I } and G all other subgroups are called non-trivial the order of G is divisible by the order of H (proof later) for a nte group (with n elements) the structure of the group is completely specied by the multiplication table (with n2 elements): I I A B C . . . A A A2 BA CA . . . B B AB B2 CB . . . C C AC BC C2 . . . ... ... ... ... ... .. .

I A B C . . .

theorem: All elements in a row (or column) of the multiclication table must be distinct (proof in exercices) this leads to the rearrangement lemma: if we multiply all elements of the group {I, A, B, C, . . . } by one of the elements we again obtain all the elements of the group, but in a dierent order isomorphism: two groups G and G are isomorphic, if there exists a map f from G to G which is one-to-one and onto and which preserves the group multiplication, i.e., for A, B G and A , B G we have: if A = f (A) and B = f (B ), then A B = f (AB )

1.3. Examples and general properties

if two groups are isomorphic, they have the same multiplication table (and thus are identical up to the names of the group elements) if the map f between the elements of the two groups is onto but not one-to-one, the two groups are homomorphic see Sec. 1.10

1.3. Examples and general properties


a group structure {I, a, a2 , a3 , . . . , an1 , an = I } is called cyclic group Cn . The smallest non-cyclic group is of order 4. The smallest non-Abelian group is of order 6. group with two elements: Z2 = {I, A} construction of the multiplication table (22 = 4 elements): I 2 = I, IA = A, AI = A, A2 =? (A or I ?)

use the theorem from Sec. 1.2 A2 = I , and thus the multiplication table for Z2 is: I A I I A A A I

this is the only possibility for the multiplication table all groups of order 2 are isomorphic to Z2 examples for groups isomorphic to Z2 : 1) consider the following two transformations from R3 to R3 I: P : (x, y, z ) (x, y, z ) (x, y, z ) (x, y, z ) (parity)

multiplication is dened as consecutive transformations isomorphic to Z2 2) instead of the two spacial transformation we now consider operators acting upon (real or complex) functions of x = (x, y, z ): OI f (x) = f (x) OP f (x) = f (x)
2 OI

I 2 = I,

IP = P,

P I = P,

P2 = I

= OI ,

OI OP = OP ,

OP OI = OP ,

2 OP = OI

isomorphic to Z2 the operators OI and OP are linear, i.e., O(f + g ) = Of + Og 3) permutation of two objects A and B : P1 (AB ) = AB, isomorphic to Z2 P2 (AB ) = BA

1. Introduction

4) consider operators acting on the complex wave function of two particles (which depends on the spatial and spin coordinates): OE (x1 , 1 ; x2 , 2 ) = (x1 , 1 ; x2 , 2 ) OS (x1 , 1 ; x2 , 2 ) = (x2 , 2 ; x1 , 1 ) OE and OS form a group isomorphic to Z2 at rst sight Z2 looks trivial, but we can learn many of the concepts of group theory from Z2 now consider example 2 and two functions fe and fo that satisfy OP fe (x) fe (x) = fe (x) OP fo (x) fo (x) = fo (x) (fe has even parity) (fo has odd parity)

(e.g., fe = x2 yz , fo = xyz ) fe and fo have special properties under the group {OI , OP }: fe is invariant under OP fo only changes sign under OP one of the aims of group theory is the classication of objects that have special symmetry properties under a certain group (in general this is much more dicult than in this simple example) assume that we know the structure of a certain group G in general there are many groups isomorphic to G (e.g., G ) G contains operators that act on certain objects using the structure of G we can immediately classify these objects with respect to their symmetry properties under G the relation
dx dy dzfe fo = 0

is an example of an "orthogonality relation" between objects with special symmetry properties (a.k.a. "selection rule" in QM) an arbitrary function can be written as a sum of an even and an odd function: f = fe + fo with 1 fe = [f (x) + f (x)] 2 1 fo = [f (x) f (x)] 2

this is an example of an "expansion theorem": objects without special symmetry properties can be expressed as linear combinations of (i.e., "expanded in") objects with special symmetry properties

1.4. Invariant subspaces

1.4. Invariant subspaces


rst two denition we need later on: let V be a vector space of dimension m with basis {v1 , . . . , vm } and W be a vector space of dimension n with basis {w1 , . . . , wn }; furthermore assume V W = {0} the direct sum V W is the space with basis {v1 , . . . , vm , w1 , . . . , wn } we have dim(V W ) = m + n example: R2 = R1 R1 the tensor product V W is the space with basis vectors {vi wj } (i = 1, . . . , m and j = 1, . . . , n) the notation vi wj is purely abstract (the basis vectors "consist" of one basis vector each from V and W ) we have dim(V W ) = mn consider a function f1 without special symmetry properties under {OI , OP } and dene f2 (x) = OP f1 (x) = f1 (x) OP f2 (x) = f1 (x) f2 is called the image of f1 under the operation OP (and vice versa) the pair f1 , f2 is called an image pair under OP examples: (x2 + y + z ), (x2 y z ) or (xy 2 + z 2 ), (xy 2 + z 2 ) let S be the set of all functions 1 f1 + 2 f2 with arbitrary coecients 1 , 2 this set is a two-dim. linear function space, which is spanned by f1 and f2 (proof of "S = space" in exerices) f1 and f2 form a basis of S the space S is invariant under {OI , OP }, i.e., if one of the operators acts on an f S the result is still in S proof : OI f = f S OP f = OP (1 f1 + 2 f2 ) = 1 OP f1 + 2 OP f2 = 1 f2 + 2 f1 S however, the space S has smaller invariant subspaces to see this dene new basis functions 1 f 1 = (f1 + f2 ), 2 these also span S since f1 = f f2 = f 1 + f2 , 1 f2 f = 1 f1 + 2 f2 = 1 (f 1 + f2 ) + 2 (f1 f2 ) = (1 + 2 )f1 + (1 2 )f2 1 f 2 = (f1 f2 ) 2

10

1. Introduction

f 1 and f2 have well-dened parity: 1 OP f 1 = OP f1 + 2 1 OP f 2 = OP f1 2 1 1 OP f2 = f2 + 2 2 1 1 OP f2 = f2 2 2 1 f1 = f 1 2 1 f1 = f 2 2

1 and S 2 that are spanned by f the one-dim. spaces S 1 and f2 are separately invariant under {OI , OP }: OP (f 1 ) = f1 S1 OP (f 2 ) = f2 S2 we say that the space S is reducible w.r.t. the action of {OI , OP }, i.e., it can be split into invariant subspaces of smaller dim. 1 S 2 symbolically: S = S S1 and S2 are irreducible, i.e., they cannot be split into smaller invariant subspaces example: consider the two image pairs h1 = x2 + y + z g1 = exyz h2 = x2 y z g2 = exyz

the four products h1 g1 , h1 g2 , h2 g1 , h2 g2 span a 4-dim. space Sh Sg , i.e., f = ah1 g1 + bh1 g2 + ch2 g1 + dh2 g2 this space is invariant under {OI , OP }: OP (h1 g1 ) = (OP h1 )(OP g1 ) = h2 g2 OP (h1 g2 ) = (OP h1 )(OP g2 ) = h2 g1 OP (h2 g1 ) = (OP h2 )(OP g1 ) = h1 g2 OP (h2 g2 ) = (OP h2 )(OP g2 ) = h1 g1 OP f = ah2 g2 + bh2 g1 + ch1 g2 + dh1 g1 Sh Sg h1 g1 and h2 g2 are an image pair they span a 2-dim. invariant subspace S h1 g2 and h2 g1 are an image pair they also span a 2-dim. invariant subspace S every function in Sh Sg can be written in a unique way as a sum of two functions in S and S : f = f + f with f = ah1 g1 + dh2 g2 S f = bh1 g2 + ch2 g1 S symbolically: Sh Sg = S S

1.5. The symmetric (or permutation) group

11

the spaces S and S can be further reduced if we introduce basis functions with well-dened parity: 1 1 f1 = (h1 g1 + h2 g2 ) f2 = (h1 g1 h2 g2 ) 2 2 1 1 f3 = (h1 g2 + h2 g1 ) f4 = (h1 g2 h2 g1 ) 2 2 f = (a + d)f1 + (a d)f2 + (c + b)f3 + (c b)f4 the subspaces Si spanned by the fi are invariant under {OI , OP } (proof as above) Sh Sg = S1 S2 S3 S4 this notation means the following: the space on the LHS and the subspaces on the RHS are separately invariant under the group every function in the space on the LHS can be expressed uniquely as a sum of functions, each of which lies in one of the subspaces on the RHS in this example all irreducible invariant subspaces are 1-dim., but in general they are of higher dimension (for order(G) 6)

1.5. The symmetric (or permutation) group


symmetric group Sn = group of permutation of n objects there are n! permutations order(Sn ) = n! a typical group element is p= 1 2 ... p1 p2 . . . n pn

this notation means: "move the rst object to position p1 , the second object to position p2 , etc." e.g., for n = 6: 1 2 3 4 5 6 6 4 1 2 5 3 applied to gives [a, b, c, d, e, f ] [c, d, f, b, e, a]

a permutation can be split into disjoint cycles (which have no elements in common), e.g., 1 2 3 4 5 6 6 4 1 2 5 3

= (163)(24)(5)

3-cycle 2-cycle 1-cycle

the 3-cycle means "1 goes to 6, 6 goes to 3, 3 goes to 1" within a cycle the numbers can be shifted cyclically: (163) = (613) = (316) but = (136)

12

1. Introduction

the order of the disjoint cycles does not matter the 1-cycles are usually not written explicitly every -cycle ( > 2) can be written as a product of 2-cycles (also called transpositions), e.g., (163) = (13)(16) convention: the right-most cycle is applied rst note: (163) = (16)(13), but this is not in contradiction to the above statement about the order, which referred to disjoint cycles Cayleys theorem: Every group of order n is isomorphic to a subgroup of Sn . example: S3 group elements: element applied to [a,b,c] I [a,b,c] (12) [b,a,c] (13) [c,b,a] (23) (23) (123) (321) I (12) (13) (23) [a,c,b] (123) (123) (23) (12) (13) (321) I (123) [c,a,b] (321) (321) (13) (23) (12) I (123) (321) [b,c,a]

multiplication table (exercises): I (12) (13) (23) (123) (321) subgroups: {I } and S3 (trivial) {I, (12)}, {I, (13)}, {I, (23)} (isomorphic to Z2 ) {I, (123), (321)} (isomorphic to C3 ) I I (12) (13) (23) (123) (321) (12) (12) I (123) (321) (13) (23) (13) (13) (321) I (123) (23) (12)

S3 is a non-Abelian group (true for all Sn with n > 2)

1.6. The action of a group on a set


let G be a group and M be a set the action of G on M is given by a map GM M : that satises: g1 (g2 m) = (g1 g2 )m Im = m (associativity) m M (g, m) gm for g G and m M

this implies that the action is one-to-one and onto, i.e., M is mapped in a one-to-one fashion onto itself (m = gm m = g 1 m ) example: M = R3 , G = rotation group, action = rotation of vectors

1.7. Equivalence classes and invariant subgroups

13

the orbit of a point m M under the action of G on M is G m {gm | g G} The orbit of "typical" points has n = order(G) elements. The orbit of "special" points has less than n elements. Example: G = D3 = symmetry group of an equilateral triangle (isomorphic to S3 = permutation of the corners of the triangle) identity 2 rotations (of 120 and 240 ) 3 reections (axes through the center and one of the corners) M = all points in R2 with origin = center of the triangle

special point (orbit has 1 element)

typical points (orbit has 6 elements)

special points (orbit has 3 elements)

the set of group elements that map an m M to itself, i.e., Gm = {g G | gm = m} is called isotropy group (or stabilizer or stationary subgroup or little group) of m. this set is a group (proof in exercises) in our example: the isotropy group of is {I } the isotropy group of is D3 the isotropy group of is {I, R} through . Z2 where R is the reection about the axis

in all these cases we have (number of elements of orbit) order(isotropy group) = order(group) this is true in general for all nite groups (proof in exercises)

1.7. Equivalence classes and invariant subgroups


a group element y is called conjugate (or equivalent) to a group element x if there exists a group element g such that y = gxg 1 notation: y x examples:

14

1. Introduction

1) for S3 : (13) (12) since (23)(12)(23)1 = (13) 2) rotation group in 3 dimensions: Rn and an angle () = rotation about an axis n
1 = R () with n for arbitrary R we have RRn = Rn ()R n rotations about the same angle, but dierent axes, are equivalent.

physical interpretation: equivalent group elements correspond to the same operation, but in a dierent basis an equivalence class (or simply class) is a set of group elements that are equivalent to each other. Properties (proofs in exercises): the identity forms a class by itself in general the elements of a class dont form a group if the group is Abelian, each elements forms a class by itself every element of G belongs to one and only one class the order of the group is divisible by the number of elements in a class the number of classes is equal to the number of non-equivalent irreducible representations of the group (later) for S3 : the rst class is {I } now conjugate (12) with all elements of S3 I (12)I = (12) (12)(12)(12) = (12) (13)(12)(13) = (13)(321) = (23) (23)(12)(23) = (23)(123) = (13) (123)(12)(321) = (123)(13) = (23) (321)(12)(123) = (321)(23) = (13) (12), (13) and (23) form a class for the remaining two elements we have I (123)I = (123) (12)(123)(12) = (12)(13) = (321) (123) and (321) are equivalent and thus in the same class, i.e., there are three classes: c1 = {I } , c2 = {(12), (13), (23)} , c3 = {(123), (321)}

two elements of S3 are equivalent if they have the same cycle structure (this is true in general for Sn ) a subgroup K G is called conjugate to a subgroup H G if there exists a g G such that K = gHg 1 {ghg 1 | h H } e.g., the subgroup K = {I, (13)} is conjugate to the subgroup H = {I, (12)} since (23)I (23) = I and (23)(12)(23) = (13)

1.8. Cosets and factor groups

15

if ghg 1 H for all h H and all g G, then it is an invariant (or normal or selfconjugate) subgroup of G every group has two trivial invariant subgroups: {I } and G a group is called simple if it does not have any non-trivial invariant subgroups a group is called semi-simple if it does not have any non-trivial Abelian invariant subgroups for S3 the only nontrivial invariant subgroup is H2 = {I, (123), (321)} S3 is neither simple nor semi-simple since H2 is Abelian

1.8. Cosets and factor groups


consider a subgroup H G with order(H ) = nH and elements hi and a group element gG the set gH {ghi | i = 1, . . . , nH } is called left coset of H w.r.t. the element g similarly the set Hg {hi g | i = 1, . . . , nH } is called right coset of H w.r.t. g hence a coset is a subset of G if g H , then gH = Hg = H (rearrangement lemma); therefore some authors require g / H when dening cosets, but we allow for g H (and thus H is also a coset) in the following we mainly consider left cosets (right cosets analogous) properties:
1 two cosets g1 H and g2 H are either identical (if g1 g2 H ) or disjoint, i.e., they 1 have no elements in common (if g1 g2 / H) proof:

suppose they have an element in common: g1 h1 = g2 h2 = g


1 1 1 then g2 = g1 h1 h g2 H = g1 h1 h H and 2 2 H = g1 H , since h1 h2 hH = H for h H (rearrangement lemma) 1 1 in this case we have g1 g2 = h1 h 2 H

converse analogous number of coset elements = order(H ) the group G is split into disjoint cosets of H , with order(H ) elements each: G = H g1 H g2 H order(G) is divisible by order(H ) (see Sec. 1.2) every group element g belongs to one and only one of the dierent cosets of H

16

1. Introduction

for S3 : let H1 = {I, (12)} (not invariant), and H2 = {I, (123), (321)} (invariant). The left and right cosets of H1 are: IH1 = {I, (12)} (12)H1 = {(12), I } (13)H1 = {(13), (123)} (23)H1 = {(23), (321)} (123)H1 = {(123), (13)} (321)H1 = {(321), (23)} H1 I = {I, (12)} H1 (12) = {(12), I } H1 (13) = {(13), (321)} H1 (23) = {(23), (123)} H1 (123) = {(123), (23)} H1 (321) = {(321), (13)}

for H1 the left and right cosets are dierent, and (e.g.) S3 = H1 (13)H1 (23)H1 The cosets of H2 are IH2 = {I, (123), (321)} (12)H2 = {(12), (23), (13)} (13)H2 = {(13), (12), (23)} (23)H2 = {(23), (13), (12)} (123)H2 = {(123), (321), I } (321)H2 = {(321), I, (123)} H2 I = {I, (123), (321)} H2 (12) = {(12), (23), (13)} H2 (13) = {(13), (12), (23)} H2 (23) = {(23), (13), (12)} H2 (123) = {(123), (321), I } H2 (321) = {(321), I, (123)}

for H2 the left and right cosets are identical, and (e.g.) S3 = H2 (12)H2 in general: if H is an invariant subgroup, then left and right cosets are identical, since gHg 1 = H {ghi g 1 } = {hi } with i = 1, . . . , order(H )

(in general the elements on the LHS and RHS are in a dierent order) {ghi } = {hi g } gH = Hg

in this case the partitioning of G into cosets is unique there is a natural "factorization" of G based on this partitioning. if H is invariant, we can view its cosets as elements of a new group: multiplication is dened by (g1 H ) (g2 H ) {g1 hi g2 hj | hi , hj H } = {g1 g2 hk } = (g1 g2 )H closure property is satised
1 with hk = (g2 hi g2 ) hj H H since H is invariant

1.9. Direct product of two groups

17

other group properties: associativity: g1 H (g2 H g3 H ) = g1 H (g2 g3 H ) = (g1 g2 g3 )H = = (g1 g2 )H g3 H = (g1 H g2 H ) g3 H identity element = H , since gH H = gH = H gH inverse element of gH = g 1 H , since gH g 1 H = H = g 1 H gH hence we have a theorem: if H is invariant, then the set of all cosets, i.e., {gH | g G}, is a new group with multiplication law (g1 H ) (g2 H ) = (g1 g2 )H this group is called G/H = factor group, with order(G/H ) =
order(G) order(H )

if H is not invariant the construction above doesnt work anymore, and the set of all (left or right) cosets doesnt form a group (convince yourself using the example of H1 ) example: for S3 the subgroup H2 = {I, (123), (321)} is invariant, and the factor group S3 /H2 has two elements: {I, (123), (321)} and this is isomorphic to Z2 (exercises) {(12), (13), (23)}

1.9. Direct product of two groups


consider two groups A and B with elements a A and b B . The direct product A B consists of all pairs (a, b) with the multiplication law (a1 , b1 ) (a2 , b2 ) = (a1 a2 , b1 b2 ) for nite groups we have order(A B ) = order(A)order(B ) typical examples: suppose A acts on a vector space U and B acts on a vector space V then A B acts on the product space W = U V , and it does so as follows (with u U and v V ): (a, b)(u v ) = (au) (bv ) another example: A and B act on the same vector space (or on the same objects) suppose ab = ba for all a A and b B (but A and B could be non-Abelian) A and B have only the identity in common then G = A B is the set of products g = ab = ba, and for every g G this representation is unique

18

1. Introduction

let G = A B then A and B are invariant subgroups of G proof: gak g 1 = g (ak , I )g 1 = (ai , bj )(ak , I ) (ai , bj )1
(a1 ,b1 ) 1 1 = (ai ak a a A i , bj Ibj ) = (a , I ) =

and analogous for B A is isomorphic to G/B (and B to G/A): G/B = {(ai , bj )B } = {(ai , B )} rearrangement lemma

and this group has the same multiplication law as A (see Sec. 1.8) however, the converse is not true: if H is an invariant subgroup of G, then in general G = H (G/H ) since in general G/H is not an invariant subgroup. example: S3 has the subgroups H1 = {I, (12)} and H2 = {I, (123), (321)}. H2 is invariant S3 /H2 Z2 H1 , but S3 = H1 H2 since H1 is not an invariant subgroup (and since the elements of H1 and H2 dont commute)

1.10. Homomorphisms
a homomorphism from a group G to another group G is a map f from G to G that preserves the group multiplication, i.e., f (g1 g2 ) = f (g1 ) f (g2 ) g1 , g2 G

if this map is one-to-one and onto it is called isomorphism example: the following map from S3 to Z2 is a homomorphism
subset H (123) (321) I (12) (13) subset M (23) I

f
A

(check yourself using the multiplication table)

G = S3

G = Z2

the image of the homomorphism f : G G is the image set of G under f : image(f ) = f (G) = {f (g ) | g G} in our example: image(f ) = {I , A} the kernel (or center) of the homomorphism is the preimage of the identity of G , i.e., the set of all elements of G that are mapped to the identity of G : ker(f ) = f 1 (I ) = {g G | f (g ) = I } in our example: ker(f ) = {I, (123), (321)}

1.10. Homomorphisms

19

from the denition of the homomorphism it follows that I is mapped to I : f (I ) = I inverses are mapped to each other: f (g 1 ) = f (g )1 g G

if G has an invariant subgroup H , there is a natural homomorphism from G to G/H : G g gH G/H

turning this around yields a theorem: Let f be a homomorphism from G to G and onto. Let K be the kernel of f . Then K is an invariant subgroup of G. Furthermore, the factor group G/K is isomorphic to G . Schematically:

K G

a b I

f
homomorphism G G

G q I p G

qK K pK G/K proof: group properties:


isomorphism G/K G

if a, b K , then f (ab) = f (a)f (b) = I I = I and thus ab K associativity is clear f (I ) = I (see above) implies I K for a K and arbitrary g G we have f (gag 1 ) = g I (g 1 ) = g I (g )1 = I gag 1 K , i.e., K is invariant. the elements of the factor group G/K are the cosets gK . Dene the map : G/K G , gK g = f (g )

this map is unique, since g1 K = g2 K implies f (g1 ) = f (g2 ): g1 K = g2 K


1 g1 g2 = k K

g2 = g1 k

f (g2 ) = f (g1 k ) = f (g1 ) f (k ) = f (g1 )


I

we still have to show that is an isomorphism the group multiplication is preserved: (pK )(qK ) = p q = f (p)f (q ) = f (pq ) = (pq ) = (pqK )

20

1. Introduction

the map is one-to-one : if (pK ) = (qK ), then q 1 pK = q 1 K pK = q 1 K (pK ) = [(qK )]1 (pK ) = I q 1 pK = K pK = qK

since f is onto, is also onto

1.11. Example: Homomorphism between S (2, C) and the Lorentz group


let M be Minkowski space, i.e., M = R4 with Lorentz metric x
2 2 2 2 = x2 0 x1 x2 x3

, where x = (x0 , x1 , x2 , x3 )

is called a four-vector. a homogeneous Lorentz transformation (LT) is a linear map from M to M that preserves the Lorentz metric, i.e., x
2

= x

x M

the Lorentz group L = O(3, 1) is the group of all homogeneous LTs now identify each point in M with a Hermitian 22 matrix: X= x0 + x3 x1 ix2 x1 + i x2 x0 x3
2

2 2 2 det(X ) = x2 0 x1 x2 x3 = x

the set of all Hermitian 22 matrices is a four-dimensional real vector space (but not a group). A basis of this space is given by I and the three Pauli matrices: I= 1 0 0 1 , 1 = 0 1 1 0 , 2 = 0 i i 0 , 3 = 1 0 0 1

X = x0 I + x1 1 + x2 2 + x3 3 now let A be an arbitrary complex 22 matrix. Dene the action of A on X by X AXA this induces an action of A on the four-vector x: x (A)x we have (AXA ) = AXA , i.e., AXA is Hermitian, and thus (A)x is again a real four-vector. Furthermore, det AXA = | det A|2 det X

1.11. Example: Homomorphism between S (2, C) and the Lorentz group

21

S (2, C) is the group of complex 22 matrices with determinant 1. If A S (2, C), then det(AXA ) = det X and thus (A)x i.e., (A) represents a LT we also have (AB )X (AB ) = ABXB A = A(BXB )A (AB )x = (A)(B )x
2

= x

i.e., is a homomorphism between S (2, C) and the Lorentz group however, is not an isomorphism since (A) = (A), i.e., the matrices A and -A represent the same LT examples (see exercises): 1) for the matrix U = ei 0 0 ei

(U ) is a rotation about the x3 -axis by an angle of 2 2) for the matrix V = cos() sin() sin() cos()

(V ) is a rotation about the x2 -axis by an angle of 2 3) for the matrix Mr = r 0 0 1/r

(Mr ) is a Lorentz boost in the direction x3 with parameter 2 ln r the homomorphism from S (2, C) to L is not onto: all A S (2, C) are continuously connected to I there are dierent types of LTs: proper LTs: det = +1 improper LTs: det = 1 orthochronous LTs: 00 1 non-orthochronous LTs: 00 1 only the proper and orthochronous LTs are continuously connected to I : "continuous LTs", these form a group L0 the homomorphism only maps S (2, C) to the continuous LTs the image of the homomorphism is image() = (S (2, C)) = L0

22

1. Introduction

homomorphism between SU(2) and O(3): SU(2) S (2, C) is the group of unitary 22 matrices with determinant 1, i.e., AA = I and det A = 1 now let A SU(2) and e0 = (1, 0, 0, 0), i.e., E0 = I E0 AE0 A = AIA = I = E0 i.e., (A)e0 = e0

O(3) is the group of real and orthogonal 33 matrices, i.e., RRT = I for LT of the form = we have e0 = e0 () 1 0 0 R with R O(3) ()

conversely one can show that all LTs satisfying () have the form () O(3) is the subgroup of L for which e0 = e0 is a homomorphism between SU(2) and O(3). Again the map is two-to-one, since (A) = (A) similarly to the discussion above, SU(2) is only mapped to elements of O(3) that are continuously connected to I , i.e., elements with determinant 1 image() = (SU(2)) = SO(3)

2. Matrix representations
2.1. Denitions
consider a -dimensional space S that is spanned by linearly independent functions f1 , . . . , f . Suppose S is invariant under a group G of linear operators A1 , . . . , An , i.e., Aj fi is again in S :

Aj fi =
k=1

(Aj )ki fk

i = 1, . . . , j = 1, . . . , n

()

the (Aj )ki are constant coecients we can view (Aj ) as a matrix (k = row index, i = column index) (Aj ) denes the action of the operator Aj on the basis functions the i-th column of (Aj ) contains the components of the function Aj fi in the basis of the functions f1 , . . . , f for every operator Aj there is one matrix (Aj ), hence altogether we have n such matrices every has dimension the set of these matrices is called (G) we now show that the s satisfy the multiplication law of the group, i.e., (Am )(Aj ) = (Am Aj ) proof: Am Aj fi = Am
k

(Aj )ki fk =
k

(Aj )ki Am fk =

=
k

(Aj )ki
l

(Am )lk fl =
l

(Am )lk (Aj )ki fl =


[(Am )(Aj )]li

=
l

(Am Aj )li fl

(Am Aj ) = (Am )(Aj ) a set of matrices satisfying the group multiplication law is called matrix representation of the group terminology options for (): the functions fi form a basis of the representation (G) the fi transform under G in the representation (G) the fi furnish the representation (G)

23

24

2. Matrix representations

S is called carrier space of the representation the dimension of the representation = dimension of the matrices (I ) = 1 dim. of the representation = Tr (I )
1 1 from the denition of the inverse we have (A j ) = (Aj )

a representation is called faithful if the homomorphism between the group elements and the representation matrices is one-to-one, i.e., dierent group elements are represented by dierent matrices every group has a trivial representation, in which (Aj ) = 1 this representation is not faithful an important theorem (proof in exercises): 1) If a group G has a nontrivial invariant subgroup H , then a representation of the factor group G/H is also a representation of G. This representation is not faithful. 2) Converse: If (G) is an unfaithful representation of G, then G has at least one invariant subgroup H such that denes a faithful representation of the factor group G/H j

2.2. Equivalent representations


consider the same space S as in Sec. 2.1, but with another basis f 1 , . . . , f . The two bases are related by a transformation S

f i =
k=1

Ski fk
1 Sik fi i=1

(i = 1 . . . , ) (k = 1, . . . , )

or

fk =

the matrix S must be invertible, i.e., det S = 0 the functions f i furnish another representation (A1 ), . . . , (An ) of the group G the relation between these two representations is: Aj f i = Aj
k

Ski fk =
k

Ski Aj fk =
k 1 Sml fm = m

Ski
l

(Aj )lk fl =

1 Sml (Aj )lk Ski f m = lk [S 1 (Aj )S ]mi

=
k

Ski
l

(Aj )lk

= Aj ) = S (

(Aj )mi f m
m 1

(Aj )S

same S

G) are called equivalent the representations (G) and ( equivalent representations derscribe the same action of the group elements in dierent basis systems (e.g., rotated coordinate systems)

2.3. Examples

25

2.3. Examples
consider again example 2 from Sec. 1.3 the space spanned by an even function fe (e.g., fe = x2 yz ) has dimension 1 (see Sec. 1.4). From OI fe = fe and OP fe = fe we have (1) (I ) = 1 , (1) (P ) = 1

this is a representation of Z2 (and all groups isomorphic to Z2 ) every function with even parity transforms under {OI , OP } in this representation if we consider another group isomorphic to Z2 , then all functions that are invariant under the two operators of the group also transform in this representation. Example 4 from Sec. 1.3: the function x1 x2 + y1 y2 transforms under {OI , OP } in the representation (1) . the space spanned by an odd function fo (e.g., fo = xyz ) also has dimension 1. From OI fo = fo and OP fo = fo we have (2) (I ) = 1 , (2) (P ) = 1

this is another representation of Z2 (and all groups isomorphic to Z2 ) all odd functions transform under {OI , OP } in this representation isomorphic groups: functions that change sign under the operator corresponding to A also transform in this representation. Example 4 from Sec. 1.3: the function x1 y1 x2 y2 transforms under {OE , OS } in the representation (2) consider two functions f1 and f2 that form an image pair under {OI , OP }: OI f1 = f1 OP f1 = f2 OI f2 = f2 OP f2 = f1

(f1 and f2 span a two-dimensional space) From these equations we immediately obtain the representation of {OI , OP } in the basis of f1 and f2 : (3) (I ) = 1 0 0 1 , (3) (P ) = 0 1 1 0

under a group isomorphic to Z2 all image pairs transform in this representation. 2 Example 4 from Sec. 1.3: the functions x2 1 y2 and x2 y1 transform under {OE , OS } in the representation (3) basis transformation (see Sec 1.4) 1 1 f f 1 = (f1 + f2 ) , 2 = (f1 f2 ) 2 2 = f , O f = f , OP f OI f 2 1 = f1 1 1 I 2

OP f 2 = f2

from this we immediately obtain the representation of {OI , OP } in this basis (4) (I ) = 1 0 0 1 , (4) (P ) = 1 0 0 1

26

2. Matrix representations

(4) is equivalent to (3) (4) = S 1 (3) S with S= 1 1 1 2 1 1

now consider a 4-dim. space spanned by the functions f1 = cosh(xyz ) + x2 + y + z f3 = sinh(xyz ) + x2 y z action of OP : OP f1 = f1 f2 + f4 OP f3 = f2 f3 + f4 the space is invariant under {OI , OP }. The representation of {OI , OP } in this basis is 1 0 (5) (I ) = 0 0

f2 = x2 + y + z f 4 = x2 y z

OP f2 = f4 OP f4 = f2

0 1 0 0

0 0 1 0

0 0 0 1

1 1 (5) (P ) = 0 1

0 0 0 1 0 1 1 1

0 1 0 0

2 = I) for all representations we have 2 (P ) = 1, as expected (since OP

2.4. Reduction of representations


2.4.1. Reducibility of invariant spaces and representations
some representations are more "fundamental" than others. The functions that furnish these representations have special symmetry properties under the group conversely, the symmetry properties of a function can be deduced from the representation of the group in which the function transforms a -dim. space S that is invariant under G is called reducible if there exists a set of linearly independent functions f 1 , . . . , f spanning S , and if this set can be decomposed into two subsets f1 , . . . , fa and fa +1 , . . . , f such that each of these subsets spans a space that is again invariant under G. Let Sa = span f 1 , . . . , fa Sb = span f a +1 , . . . , f , then every function f = fa =
i=1 if i i=1 a

S can be written as the sum of two functions

if i Sa

and

fb =
i=a +1

if i Sb

symbolically: S = Sa Sb

2.4. Reduction of representations

27

G) furnished by the functions what does this imply for the matrix representation ( f1 , . . . , f ? Aj f i for i a doesnt contain components in Sb Aj f i for i > a doesnt contain components in Sa Aj ) are block diagonal: the matrices ( Aj ) = ( a (Aj ) 0 b 0 (Aj ) j = 1, . . . , n ()

with dim a (Aj ) = a and dim b (Aj ) = a both of the sets a (G) and b (G) are matrix representations of the group G a matrix representation of form () and every representation equivalent to it is called reducible if all matrices are already block diagonal we say that the matrix representation is already in the reduced form if representation (G) is not in reduced form we have to check if there exists a Aj ) = S 1 (Aj )S is block diagonal for all j = 1, . . . , n transformation S such that ( (same S for all j !). This is an important problem in representation theory. example: consider the representations (3) and (4) of Z2 from Sec. 2.3 (3) (I ) = (4) (I ) = 1 0 0 1 1 0 0 1 , , (3) (P ) = (4) (P ) = 0 1 1 0 1 0 0 1

(4) is already in reduced form and contains the representations (1) and (2) on the diagonal (3) is equivalent to (4) since (4) (Aj ) = S 1 (3) (Aj )S with S= 1 1 1 2 1 1

a space is called irreducible ("irrep") if it is not reducible the action of the group elements on an irreducible space is given by an irreducible matrix representation (in arbitrary basis) the representations a (G) and b (G) in () could again be reducible, in which case they can be made block diagonal by a suitable basis transformation. This process is continued until only irreps remain, i.e., in general a representation consists of several irreps, and one can nd a transformation S such that a (Aj ) 0 0 b 0 (Aj ) 0 c (A ) S 1 (Aj )S = 0 0 j . . . . . . . . .

.. .

()

with all i (Aj ) irreducible the irreps are the "building blocks" from which all representations are made

28

2. Matrix representations

a representation of the form (), in which all diagonal blocks are irreducible is called fully reduced the same irrep can occur in () more than once Symbolically: (G) = 1 (G) 1 (G) 2 (G) 2 (G) =
a1 -times a2 -times i

ai i (G)

i.e., the representation contains the irrep i ai times from () we see that the invariant space S decomposes into irreducible invariant subspaces. The action of group elements on the basis vectors of the subspaces is given by i

2.4.2. OA operators using the example of D3


D3 = symmetry group of an equilateral triangle (isomorphic to S3 ) p1 =(x1 , y1 ) L3
y x

L2

p3 =(x3 , y3 ) group elements I = identity

L1

p2 =(x2 , y2 )

C = clockwise rotation by 120 = (123) = counterclockwise rotation by 120 = C (321) 1 , 2 , 3 = reections w.r.t. L1 , L2 , L3 = (23), (13), (12) multiplication table as in Sec. 1.5 now consider (one-to-one ) maps A of the xy -plane onto itself (the 6 elements of D3 are examples of such maps A) let p = (x, y ) be a point in the xy -plane the point that is mapped by A to the point p will be called A1 p (i.e., A(A1 p) = p) with every map A we associate an operator OA that acts on an arbitrary function f (x, y ) = f (p) dene the action of the operator OA on f by (OA f ) (p) = f A1 p important: the LHS means: "compute the function OA f at point p", i.e., OA acts on the function f not on the value f (p), i.e., the new function OA f at the point p has the same value as the old function at point A1 p (which is mapped to p by A)

2.4. Reduction of representations

29

3 isomorphic to the 6 operators OA associated with the elements of D3 form a group D D3 since (OA OB )f (p) = OA (OB f ) (p) = (OB f ) (A1 p) = = f B 1 A1 p = f (AB )1 p = (OAB f ) (p) we now consider the action of these operators on certain functions of (x, y ) = p and thus generate representations of D3 S3 consider the function 1 (x, y ) = e(xx1 ) what is OC 1 2 (x, y ) = (OC 1 )(p) = 1 C 1 p = e = e analogously we get
3 (x, y ) = (OC 1 )(x, y ) = e pp3
2 2 (y y )2 1

= e

pp1

C 1 pp1
2

C 1 (pCp1 )

= e

C 1 (pp2 )

= e

pp2

for the reections we have


1 O1 1 (p) = 1 1 p = e
1 1 pp1 2 2

=
2

= e O2 1 (p) = e O3 1 (p) = e

1 1 (p1 p1 ) 1 2 (p2 p1 ) 1 (p3 p1 ) 3

= e = e = e

1 1 (pp1 ) 1 2 (pp3 ) 1 (pp2 ) 3

= e = e = e

pp1 pp3 pp2

= 1 (p) = 3 (p) = 2 (p)

and similarly for 2 and 3 thus we get the following table 3 3 1 2 2 1 3 3 the space S = span(1 , 2 , 3 ) is invariant under D OI OC OC O1 O2 O3 the functions 1 , 2 , 3 furnish a 3-dimensional representation of the group 1 0 0 (5) (I ) = 0 1 0 0 0 1 1 0 0 (5) (1 ) = 0 0 1 0 1 0

1 1 2 3 1 3 2

2 2 3 1 3 2 1

0 0 1 (5) (C ) = 1 0 0 0 1 0 0 0 1 (5) (2 ) = 0 1 0 1 0 0

0 1 0 ) = (5) (C 0 0 1 1 0 0 0 1 0 (5) (3 ) = 1 0 0 0 0 1

30

2. Matrix representations

Is this representation reducible? Yes, since S is reducible, i.e., there is a basis transformation by which S can be decomposed into smaller, invariant subspaces 1 = 1 + 2 + 3 , 2 = 3(2 3 ) , 3 = 21 2 3 3 since the operators OA only interchange the terms in the sum 1 is invariant under D 1 spans a 1-dim. invariant subspace (which is irreducible) and furnishes a 1-dim. irrep of the group ) = (1) (1 ) = (1) (2 ) = (1) (3 ) = 1 (1) (I ) = (1) (C ) = (1) (C = trivial representation of D3 (every group has it) for 2 , 3 we obtain 2 2 3 1 2 2 3 2 3 1 + 2 2 2 3 2 3 1 2 3 2 2 3 1 2 + 2 3 2 3 3 3 1 2 3 2 2 1 23 2 2 3 3 3 1 2 3 2 2 3 1 2 2 3 2

OI OC OC O1 O2 O3

3 , the functions the space span( 2 , 3 ) is invariant under D 2 , 3 furnish a 2-dim. representation of the group (3) (I ) = 1 0 0 1 1 0 0 1 , 1 2 (3) (C ) = 23

3 2 1 2 23 1 2

) = (3) (C

1 2 3 2

23 1 2
3 2 1 2

(3)

(1 ) =

(3)

(2 ) =

1 2

3 2

(3)

(3 ) =

1 2 3 2

the representation of the group in the space span( 1 , 2 , 3 ) is thus given by 1 0 0 (4) (I ) = 0 1 0 0 0 1 1 0 0 (4) (1 ) = 0 1 0 0 0 1

1 0 (4) 1 (C ) = 0 2 0 23 1 0 (4) 1 (2 ) = 0 2 0 23

3 2 1 2

1 0 (4) 1 (C ) = 0 2 0 23 1 (4) (3 ) = 0 0

0 23 1 2
3 2 1 2

0 23 1 2

0
1 2 3 2

since (4) is a representation on the same space as (5) , only in a dierent basis, (4) and (5) are equivalent 1 0 2 3 1 S = 1 1 3 1

(4) (Aj ) = S 1 (5) (Aj )S

with

Aj D 3

2.4. Reduction of representations

31

(4) is already in the reduced form, but (5) is not notation: (4) = (1) (3) (5) = (1) (3) remaining question: is the 2-dim. representation (3) reducible?

2.4.3. Four theorems and their consequences


the following theorems are due to Frobenius and Schur, and the proofs are given in the appendix A.2 Theorem 1: Every representation of a group G by matrices with det = 0 can be converted to a unitary representation (i.e., the matrices of this represetation are unitary). Theorem 2 (Schurs lemma 1): A matrix which commutes with all matrices of an irrep is proportional to the identity matrix. Theorem 3 (Schurs lemma 2): Suppose we have two irreps of a group G with elements A1 , . . . , An : 1 (G) with dimension 1 and 2 (G) with dimension 2 . If there exists a 2 1 matrix M such that M 1 (Aj ) = 2 (Aj )M j = 1, . . . , n

then if 1 = 2 we have M = 0, and if 1 = 2 we have either M = 0 or det M = 0. If det M = 0 the two representations are equivalent since 1 (G) = M 1 2 (G)M . Theorem 4 (orthogonality relations): Let i (G) and k (G) be two non-equivalent, unitary irreps of G with order(G) = n. Then we have
n

i (Aj )
j =1

k (Aj ) = 0

for all , , ,

For the matrix elements of a single unitary irrep with dimension i we have
n

i (Aj )
j =1

i (Aj ) =

n i

Combined:
n

i (Aj )
j =1

k (Aj ) =

n ik i

consequence of Theorem 2: All irreps of an Abelian group have dimension 1. (proof in exercises) consequence of Theorem 4: for xed i, , we collect the n numbers i (A1 ) , . . . , i (An ) in a vector v (i ) with n components

32

2. Matrix representations

for every representation i there are 2 i such vectors (since , = 1, . . . , i ) Theorem 4 says that such a vector is orthogonal to all vectors V (k ) if i = k or = or = however, in n dimensions there are at most n vectors that are mutually orthogonal 2 2 1 + 2 + n. In Sec 2.5 we will show that in fact 2 i =n
i

The sum is over the dierent irreps i, i.e., for a nite group there is a nite number of non-equivalent irreps each of which has nite dimension

2.4.4. Characters of representations


the trace Aj =
k

Aj

kk

of the matrix representing Aj is called character of Aj in the representation (G) from Sec. 2.2: the matrices of a representation transform under a basis transformation S as S S 1 . The trace of a matrix is invariant under such a transformation: Tr S 1 S = Tr SS 1 = Tr () i.e., in two equivalent representations every group element has the same character the characters are basis-independent while the representing matrices depend on the choice of basis characters contain the essential information on the structure of the irrep (the matrices contain a lot of "irrelevant" information) furthermore, all group elements in the same class have the same character, since for two mutually conjugate group elements p and gpg 1 we have Tr gpg 1 = Tr (g )(p)(g 1 ) = Tr (g 1 )(g )(p) = Tr (p) now set = and = in Theorem 4 and sum over and :
n

i Aj
j =1 n

k Aj

n ik i

j =1

i Aj

k Aj = nik

this is an orthogonality relation for characters since the characters depend only on the class we can also write this as nc i c
c

k c = nik

c labels the class, and nc the number of group elements in class c

2.4. Reduction of representations

33

let m be the number of dierent classes of G


i for xed i we can collect the m numbers i 1 . . . , m in a vector with m components

in m dimensions there are at most m orthogonal vectors number of irreps number of classes in the exercises we show that in fact the equal sign holds number of irreps = number of classes the m m matrix i c (with i, c = 1, . . . , m) is called character table of the group for a reducible representation of the form Aj =
i

ai i Aj

we have Aj =
i

ai i Aj

here i (Aj ) is the character of Aj in the irrep i


n 2 n

j =1

Aj

=
ik

ai ak
j =1

i Aj
nik

k Aj = n
i

a2 i

if (G) is irreducible, then one of the ai = 1 and all the others = 0, and thus
n 2

Aj
j =1

=n

if (G) is reducible, then one of the ai > 1 or several ai = 0 and thus


n 2

Aj
j =1

>n

looking at the characters one can conclude whether a given representation is reducible for (3) from Sec. 2.4.2: (3) (I ) + 2 (3) (C ) + 3 (3) (1 )
2 2 2

3) = 4 + 2 + 0 = 6 = order(D

3 (3) is irreducible, and the space span( 1 , 2 ) is invariant and irreducible under D so for S3 we have found a 1-dim. and a 2-dim. irrep (1 = 1 and 3 = 2)
2 from 2 1 + 2 + = n we have

1 + 2 2+4=6 we conclude that there is one more irrep with dimension 2 = 1 (and no other irreps)

34

2. Matrix representations

this remaining irrep can easily be constructed from the multiplication table: ) = 1 (2) (I ) = (2) (C ) = (2) (C (2) (1 ) = (2) (2 ) = (2) (3 ) = 1 }, from this we can compute the character table of S3 : The classes are {I }, {C, C {1 , 2 , 3 }. The irreps are (1) Aj = 1 (2) Aj (3) Aj character table: (1) (2) (3) {I } 1 1 2 } {C, C 1 1 -1 {1 , 2 , 3 } 1 -1 0 for all j see above see Sec. 2.4.2

if we know the characters of all irreps of a group, then for a given representation (which in general is reducible) we can compute how many times the various irreps are contained in it: Aj =
i n

ai i Aj
n

j = 1, . . . , n

k Aj
j =1

Aj =
i

ai
j =1

k Aj
nik

i Aj = nak

ak =

1 k Aj n j =1

Aj =

1 n

n c k c
c

for the reducible representation (5) of D3 we have: 1 a1 = (1 1 3 + 2 1 0 + 3 1 1) = 1 6 1 a2 = (1 1 3 + 2 1 0 + 3 (1) 1) = 0 6 1 a3 = (1 2 3 + 2 (1) 0 + 3 0 1) = 1 6 thus (5) = (1) (3) , as we found earlier

2.5. The regular representation


a group algebra is a set of elements that form a linear vector space, in which both, an addition and a multiplication are dened such that the group properties are satised (with one exception: the zero element of the algebra doesnt have an inverse group element)

2.5. The regular representation

35
n i=1 ci gi

for a group G with elements gi (i = 1, . . . , n) the linear combination coecients ci form an algebra with multiplication law

i=1 n

with

ci gi
j =1

dj gj =
i,j =1

ci dj gi gj

because gi gj G the product is again an element of the algebra the group elements are the basis vectors of the vector space dim(vector space) = order(G) the multiplication law gi gj = gk can also be written as
n

gi gj =
m=1

gm (i )mj

with (i )mj = 1 for m = k and (i )mj = 0 for m = k (and i, j are xed) the n n matrices i (i = 1, . . . , n) form a representation of G, the so-called regular representation (i is the matrix representing gi ) proof: Let ga , gb , gc G with ga gb = gc . Then, ga gb gj =
m

ga gm (b )mj =
m,k

gk (a )km (b )mj

gc gj =
k

gk (c )kj

The LHSs are equal, thus the RHSs are also equal; comparison of the coecients of gk yields: (c )kj =
m

(a )km (b )mj = (a b )kj

c = a b

Theorem: The regular representation contains all irreps of G, and the multiplicity of an irrep k equals its dimension k :
m

k=1

k k (gi )

(m = # of classes)

()

2 .. . 2 .. . m .. . m
2 blocks m blocks

proof:

36

2. Matrix representations

the characters of the regular representation are R (gi ) =


m

(i )mm

for the identity we have


n

Igi =
m=1

gm (I )mj

(I )mj = mj

R (I ) = n

for gi = I : gi gj =
m

gm (i )mj = gj

(i )jj = 0

R (gi ) = 0

with the formula from Sec. 2.4.4 we thus have ak = set gi = I in (): I =
k

1 n

k (gi )
i

R (gi ) =

1 k (I ) n

n = k

k k (I ) 2 k
k

Tr

R (I ) = n =

this is the missing proof of the formula in Sec 2.4.3

2.6. Product representations and Clebsch-Gordan coecients


in physical applications we often deal with vector spaces that are tensor products of smaller vector spaces. Examples: orbital angular momentum and spin of the electron, systems with several identical particles, etc. let U and V be two vector spaces with bases {ui } and {vj }, and let W = U V with basis {wk }, where wk = ui vj with every pair of operators A, B that act1 on U and V , Aui =
i

ui Ai i uj Bj j
j

Bvj =

we can associate a product operator D = A B acting on W : Dwk =


k

wk D k k

with

Dk k Ai i Bj j k = (i, j ), k = (i , j )

often A and B are the same physical operator, but in dierent spaces (e.g., orbital angular momentum of two particles)
1

A acts on U and B on V .

2.6. Product representations and Clebsch-Gordan coefficients

37

if (G) and (G) are two matrix represetations of a group G on the spaces U and V , respectively, then the matrices (g ) = (g ) (g ) with gG

form a representation of G on W = U V , the so-called product representation (proof follows directly from the denition of Dk k above) for the characters we have (g ) = (g ) (g ) we have2
( )(i1)n+k,(j 1)n+ = ij k

with

i, j = 1, . . . , dim( ) k, = 1, . . . , dim( ) = n

in general the product representation is reducible:

with

a n = n n

(n is the dimension of , and the a follow from Sec. 2.4.4) (which are i.e., W can be decomposed into a direct sum of irreducible subspaces W invariant under G), with dim(W ) = n ; the index = 1, . . . , a distinguishes the dierent subspaces that correspond to the same irrep

1 .. . 1 .. . .. .
a1 blocks a blocks

..

there is a basis transformation from the product basis {wk } to a new orthonormal basis } in which the representation matrices are block diagonal { w
= 1, . . . , n labels the basis vectors of W

the basis transformation is written as follows (with k = (i, j ) and in Dirac notation):
w

=
i,j

wij

i, j (, ) , ,
Clebsch-Gordan coe.

()

the CG coecients are the matrix elements of the basis transformation, with (i, j ) = row index (old basis)
2

here we use the Kronecker product of two matrices dened in A.3

38

2. Matrix representations

(, , ) = column index (new basis) (, ) xed for () we write symbolically (sum over repeated indices): w = wU from
aa = w a | w a = wk Uka | wk Uk a = Uka Uk a wk | wk = Uak Uka = U U aa aa

or

w a = wk Uka

we have U U = 1, i.e., the CG coecients form a unitary matrix now dene , , (, )i, j i, j (, ) , , the converse of () wij =
,, w

w = wU 1 = wU

or

wk = w a Uak =w a Uka

, , (, )i, j

()

the CG coecients satisfy orthonormality and completeness relations (this follows from U U = 1 = U U , see exercises): i , j (, ) , ,

, , (, )i, j = i i j j i, j (, ) , , =

, ,
ij

(, )i, j

simplication of the notation: wij


|i, j and w

|, , i, j | , ,

Einsteins summation convention for row/column indices i, j (, ) , ,

let D(g ) be the operator that describes the action of a group element g on W = U V . Then we have D(g ) |i, j = i , j D(g ) |, , = , , (g )i i (g )j j (g )

is carrier space of ) (since U is carrier space of , V is carrier space of and W

D(g ) |, ,

()

= D(g ) |i, j

i, j | , ,

= =

= i ,j
()

(g )i i (g )j j i, j | , , , , i ,j

= , , = , ,

(g )i i (g )j j i, j | , , (see above) ( )

(g ) i ,j

(g )

= , ,

(g )i i (g )j j i, j | , ,

2.6. Product representations and Clebsch-Gordan coefficients

39

()

corresponds to w = wU

() corresponds to w = wU = U U ( ) corresponds to the LHS of ( ) shows that in the new basis the matrices are block diagonal application in QM: coupling of two spins s1 and s2 to a total spin s the 3 components of the spin operator are the generators of the group SU(2) (see Sec. 6.1)
1 spin s1 = 2 2 states with z -components m1 = 1 2 . The two wave functions corresponding to these states

span a 2-dimensional space that is invariant under SU(2) furnish an irrep of SU(2) with dimension 2 analogously for s2 this means we have a product space of dimension 4 with basis states |s1 , m1 |s2 , m2 the two spins can couple to a total spin s = 0 (with m = 0) or s = 1 (with m = 1, 0, 1) the product space can be decomposed into two subspaces that are invariant and irreducible under SU(2). Symbolically: 1 2
dim=2

1 2
dim=2

1
dim=3

0
dim=1

the CG coecients describe the basis transformation from the product basis to the total spin basis the wave functions in the total spin basis furnish a 1- dimensional and a 3- dimensional irrep of SU(2) which basis we should use depends on the problem, e.g.: two spins interacting with an external magnetic eld B = B z : H = (1 s1 + s2 ) B = (1 s1z + 2 s2z )B product basis two spins interacting with each other: H = as1 s2 = total spin basis in Sec. 4 we will learn how to construct the CG coecients, or more generally, how to construct the basis transformation from a reducible representation to a sum of irreps a 2 2 s s2 1 s2 2

40

2. Matrix representations

2.7. Subduced and induced representations


let H (with order nH ) be a subgroup of G (with order nG ) a representation of G is also a representation of H if we restrict it to the (h) with hH this representation is called subduced representation even if (G) is irreducible, in general (H ) is reducible conversely, from an irrep of H one can construct (induce ) a representation of G: let f1 , . . . , fm be the basis states of (H ) if all operators of G act on this basis, in general we end up with a larger space with basis f1 , . . . , fn (n m) this space is invariant under G by construction the basis states furnish a representation ind of G (which in general is reducible) we have (without proof): dim ind (G) = Frobenius reciprocity theorem (without proof): the irrep (G) subduces a representation of H (in general reducible). Suppose this (H ) a times subduced representation contains the irrep (H ) induces a representation of G (in general reducible). Suppose this the irrep induced representation contains the irrep (G) b times then we have
a = b

nG dim (H ) nH

for all ,

3. Applications in quantum mechanics


3.1. Selection rules
in the following we show that the orthogonality relations for irreps lead to selection rules in QM scalar products in QM typically have the form f|g = dx1 . . . wf g

integrals over continuous variables (e.g., position or momentum) sums over discrete variables (e.g., spin) f and g are complex functions (e.g., wave functions) w is a real function (e.g., Jacobi determinant of variable transformation) an operator A is called unitary if it preserves the scalar product, i.e., Af | Ag = f | g

Theorem: Let G be a group of linear, unitary operators A1 , . . . , An . Suppose the , . . . , f transform in the unitary irrep (G) with dim( ) = n , e.g., functions f1 n
n Aj f = =1 (Aj ) f

()

, . . . , g Analogously, suppose the functions g1 n transform in the unitary irrep (G). (I.e., the f and g have special symmetry properties w.r.t. the group. If = , f and g have dierent symmetry properties.) We then have

f g

= v

()

where v is independent of . This implies that two functions with dierent symmetry properties are orthogonal.

41

42

3. Applications in quantum mechanics

Proof: Since the Aj are unitary we can write


f g

=
()

1 n Aj f Aj g n j =1
n

=
n

1 n = n j =1 = 1 n

=1

(Aj ) f =1

(Aj ) g

(Aj ) (Aj ) j

g f

= 1 n

n n

(see Sec.

2.4.3)

f g

=: v

(independent of )

furthermore, an arbitrary function f to which we can apply the operators Aj can be written as a linear combination of the functions with special symmetry properties (= basis functions). This expansion theorem, together with (), leads to selection rules of QM: Proof: an invariant space can be viewed as a direct sum of irreducible invariant spaces a function in an invariant space can be written as a linear combination of the basis functions that span these irreducible spaces it remains to be shown that f can be embedded in an invariant space consider the space R that is spanned by f and all images of f under the operators of G, i.e., (with A1 = I ) R = span(f, A2 f, . . . , An f ) R contains f and is invariant under G; hence the embedding has been shown. (We now have to decompose R into irreducible subspaces. These are carrier spaces of irreps of G, and these irreps determine what basis functions are needed for the expansion of f . This is f -dependent.)

3.2. The symmetry group of the Hamiltonian and degeneracies


let H be the Hamiltonian of a q.m. system and Aj be a unitary operator acting on the same space as H . We then dene
1 H t Aj HA j

"the transform of H under the operation Aj " typically Aj acts on a state | such that Aj | is the state in the new coordinate system. We then have (Aj is unitary):
1 t | H = Aj Aj H = Aj Aj HA j Aj = Aj H Aj

i.e., observers in the two systems measure the same energy

3.2. The symmetry group of the Hamiltonian and degeneracies

43

if Aj H = HAj , we have H t = H , i.e., Aj leaves H invariant and thus the Hamiltonian has the same form in both systems the set of all unitary operators {A1 , . . . , An } that commute with H form a group G, the symmetry group of H (proof is simple) now let Aj G and let | be an eigenstate of H with energy E H | = E | H Aj | = Aj H | = E Aj | ()

i.e., Aj | is an eigenstate of H with energy E if E is non-degenerate we have Aj | | . If E is m-fold degenerate, Aj | is a linear combination of the states |1 , . . . , |m with energy E . In both cases the space S = span(|1 , . . . , |m ) is invariant under the action of the symmetry group of H . degenerate states furnish a representation of G:
m

Aj |i =
k=1

(Aj )ki |i

i = 1, . . . , m j = 1, . . . , n

()

this representation could be reducible or irreducible, but in the typical case it is irreducible: all states that transform in the same irrep of G must have the same energy: H |i = Ei |
k () ()

H Aj |i

= Ei Aj |i (Aj )ki Ei |k
k

(Aj )ki H |k =
Ek |k

(with
irreducible) (no summation)

(Aj )ki Ek = (Aj )ki Ei

now dene a diagonal m m matrix E = diag(E1 , . . . , Em ) (Aj )E E (Aj )


ki ki

= (Aj )k E i = (Aj )ki Ei = Ek (Aj ) i = (Aj )ki Ek for all j

(Aj )E = E (Aj )

according to Schurs lemma 1, E 1m all Ei are the same if is reducible and |i and |k transform in dierent irreps of G, () 0 0 () i k

we have (Aj )ki = 0 for all (Aj ), and Schurs lemma 1 is not applicable we cannot conclude Ek = Ei , i.e., there is no reason why |i and |k should be degenerate

44

3. Applications in quantum mechanics

if states have the same energy even tough they transform in dierent irreps, we speak of "accidental degeneracies". This can have two reasons: 1) ne-tuning of a parameter in H (very unlikely) 2) we have not yet found the full symmetry group (i.e., we have not yet fully understood the problem) from this we learn: the degenerate sates with the same energy transform in an irrep of the symmetry group of H , and therefore can be classied using these irreps number of degenerate states = dimension of the irrep example: Hydrogen atom (without spin-orbit coupling)
2

H=

2m

e2 r

the eigenstates are labeled by the quantum numbers n = 1, 2, . . . (principal quantum number), = 0, 1, . . . , n 1 (orbital angular momentum) and m = , . . . , (z component of angular momentum) (r) = Rn (r)Y
m (, )

the Hamiltonian of a central force problem V (r) = V (r) in three dimensions is invariant under O(3) because of this symmetry the energy doesnt depend on m (2 + 1)-fold degeneracy for xed the Y
m

furnish a (2 + 1)-dimensional irrep of O(3) (see Sec. 6.7.4)

however, the energy doesnt depend on either ("accidental degeneracy") 1 (2 + 1) = n2 n2 -fold degeneracy since n=0 the reason for this is that the symmetry group is larger than O(3): For V (r) 1/r (Coulomb potential) H is invariant under O(4) (H commutes with the Runge-Lenz vector) energy is independent of n2 -fold degeneracy (corresponding to the dimensions of the irreps of O(4))

3.3. Perturbation theory and lifting of degeneracies


typical problem: H= H0

"solvable"

"small perturbation"

let G be the symmetry group of H0 , then there are two cases1 1) H is invariant under G 2) H is invariant under B G
1

There is actually a third case, namely that the perturbation could lead to a symmetry group C of H , such that G C , but this will not be covered in these lectures.

3.3. Perturbation theory and lifting of degeneracies

45

in case 1. the perturbation H does not lead to a splitting of the degenerate states of H0 in case 2. (some of) the degeneracies will be lifted the exact states of H transform in irreps of B the degenerate states of H0 transform in irreps of G for these irreps of G, the matrices corresponding to the elements of B form a (subduced) representation sub (B ) of B . In general this representation is reducible, i.e.,
r

sub (B ) =
i=1

ai i (B )

with

dim(i ) = i

for every irrep of B contained in sub (B ) there is a new energy level i ai new energy levels a1 of them are 1 -fold degenerate, a2 of them are 2 -fold degenerate, etc. example 1: H atom as in Sec. 3.2. If we add a small central potential V (r) (but not 1/r), the O(4) symmetry is broken to O(3), and every energy level is split into n levels with dierent values for . 5 3 1 3 1 1 =2 =1 =0 =1 =0 =0

9 4 1

n=3 n=2 n=1

example 2: central force problem V (r) = 1/r, + small perturbation invariant under the OA operators of D3 i.e., G = O(3), B = D3 S3 = 1 and with the consider in particular the 3-fold degenerate energy level with unperturbed eigenfunctions R(r)Y1,1 (, ) , R(r)Y1,0 (, ) ,

R(r)Y1,1 (, )

these 3 functions furnish a 3-dimensional irrep (G) of O(3) and a 3-dimensional subduced representation sub (B ) of D3 D3 has two 1-dimensional and one 2-dimensional irreps sub (B ) must be reducible the characters of the elements of D3 in the representation sub (B ) are (I ) = 3 , derivation: sub (I ) = 13 (I ) = 3 ) = 0 , (C ) = (C (1 ) = (2 ) = (3 ) = 1

46

3. Applications in quantum mechanics

action of a rotation by an angle about the z -axis: rr , , +

OA f (p) = f A1 p OR Y
m

=Y

m (,

) = f (cos )eim() = eim Y

m (, )

sub =1 OR

= 0
1

ei 0 0 1 0 0 0 ei eim
2 3

) = ( C
m=1

= 1 + 2 cos(2/3) = 0

1 is a reection w.r.t. the yz -plane: x x in spherical coordinates: x = r sin cos i.e., rr , , cos cos , sin sin , y = r sin sin , z = r cos , yy , zz

O1 Y1,1 = O1 f (cos ) (cos + i sin ) = f (cos ) ( cos + i sin ) = Y1,1 O1 Y1,0 = O1 f (cos ) = Y1,0 O1 Y1,1 = O1 f (cos )(cos i sin ) = f (cos )( cos + i sin ) = Y1,1 sub =1 O1 0 0 1 = 0 1 0 1 0 0

(1 ) = 1

character table of D3 from Sec. 2.4.4: 1 2 3 with the formula ak = a1 = a2 = a3 = 1 6 1 6 1 6


1 n

{I } 1 1 2

} {C, C 1 1 -1

{1 , 2 , 3 } 1 -1 0

k c nc (c ) c

from Sec. 2.4.4 we have 210 210 + 311 =1 =0 =1

113 + 113 +

+ 3 (1) 1 301

1 2 3 + 2 (1) 0 +

i.e., the originally 3-fold degenerate level splits into two new levels, one of which is 2-fold degenerate

3.3. Perturbation theory and lifting of degeneracies

47

O(3) 3

D3 1 2

C3 1 1 1

=1

} is a subgroup of D3 the group C3 = {I, C, C we could add to H an even smaller perturbation that is invariant under the OA operators of C3 the two degenerate states furnish a 2-dimensional representation of C3 since C3 only has 1-dimensional irreps, the two states are split by the new perturbation

4. Expansion in irreducible basis vectors


4.1. Irreducible basis vectors
consider an operator representation U (G) of a group G on a space V an invariant, irreducible subspace V under G with dimension n an orthonormal basis { e i } (i = 1, . . . , n ) of V then we have g G
U (g ) e i = ej D (g )ji

(sum over j = 1, . . . , n )

where D (G) is an irrep of G the e i are called irreducible basis vectors of the irrep for two such irreducible bases we have (see Sec. 3.1)
u i vj = ij

1 u v n k k

4.2. Projection operators on irreducible bases


we now learn how an arbitrary vector |x V can be expanded in irreducible basis vectors. The idea is simple: project |x on the basis vector |ei using the projection operator Ei = |ei ei |: Ei |x = |ei ei | x = |ei xi however, the details are more complicated. using the denitions of Sec. 4.1 we rst dene generalized projection operators:
Pji

n n

D (g )1
g G

ji

U (g )

with n = order(G) and i, j = 1, . . . , n


Theorem: For xed |x V and xed j the n vectors {Pji |x } (i = 1, . . . , n ) are either zero, or they transform in the irrep .

48

4.2. Projection operators on irreducible bases

49

Proof:
U (g )Pji |x =

n n n = n n = n

U (g )U (g ) |x D (g )1
g

ji

U (gg ) |x D (g )1
g

ji

= =

U (g ) |x D (g 1 g )1
g

ji

U (g ) |x D (g )1
g

jk

D (g )ki =

= Pjk |x D (g )ki

(no sum over )

this means that, starting from a vector |x V , we can construct an irreducible subspace of V corresponding to the irrep , with basis {Pji |x }. The basis vectors are orthogonal, but not automatically normalized. by varying , j, and |x we can nd all irreducible subspaces (of course we only need those irreps that are contained in the representation we want to reduce) in the following we will assume that the U and D are unitary D1 = D , U 1 = U example 1: reduction of the space S =span(1 , 2 , 3 ) of Sec. 2.4.2 (invariant under D3 S3 ) S3 has two 1-dimensional irreps and one 2-dimensional irrep (1 , 2 , 3 ) the generalized projection operators are
1 P11 = 2 P11 3 P11 3 P12

1 6 1 = 6 1 = 3 1 = 3 1 3 1 = 3

OI + OC + OC + O1 + O2 + O3 OI + OC + OC O1 O2 O3 1 1 1 1 OI OC OC O1 + O2 + O3 2 2 2 2 3 3 3 3 OC + OC O2 + O3 2 2 2 2 3 1 3 3 OC OC O2 + O3 2 2 2 2 1 1 1 1 OI OC OC + O1 O2 O3 2 2 2 2

3 P21 = 3 P22

applied to a vector in S , e.g., 1 : (see Sec. 2.4.2 for the action of the operators on 1 ) = 1:
1 P11 1 =

1 1 1 + 2 + 3 + 1 + 3 + 2 = 1 + 2 + 3 6 3

this function is invariant under D3 and transforms in the trivial irrep 1 .

50

4. Expansion in irreducible basis vectors

= 2:
2 P11 1 =

1 1 + 2 + 3 1 3 2 = 0 6

this had to be zero since 2 is not contained in the 3-dimensional representation (5) . = 3: rst j = 1: 1 1 1 1 2 3 1 + 3 2 2 3 3 P12 1 = 2 + 3 3 + 2 6
3 P11 1 =

1 1 3 + 2 = 0 2 2 =0

now j = 2: 3 2 3 3 + 2 6 1 1 1 1 1 3 P22 1 = 1 2 3 + 1 3 2 3 2 2 2 2
3 P21 1

2 3 21 2 3

the last two functions transform according to 3 . Thus we have the basis transformation of Sec. 2.4.2. example 2: reduction of a product representation let D be a product representation of G on V V . In general we have D a D . How do we nd the irreducible subspaces of V V ? start with the product basis |k,
to it. operators Pji |k, for xed , j, k, the n vectors Pji span an irreducible subspace and apply the generalized projection = e k e

(i = 1, . . . , n ) are either all zero, or they

by varying , j, k, we can nd all irreducible subspaces exercises: reduction of D33 (S3 ) the action of the generalized projection operators on an irreducible basis is
Pji ek =

n n

U (g ) e k D (g )ji = g

n e n

D (g ) k D (g ) ji =
g
n =n i jk

e i

jk

()

from this we can derive more identities (proofs in exercises):


Pji P k = jk P i

(property of projection operators) (inverse of the original denition) (operator form of the Theorem)

U (g ) =
ij

Pji D (g )ij P i D (g )ik i

U (g )P k =

4.3. Irreducible operators and the Wigner-Eckart theorem

51

it follows from these identities and from () that


Pi Pii

is a projection operator on e i is a projection operator on V

P =

i Pi

proof: (no summation over repeated indices)


Pi Pk = Pii Pkk = Pii ik = Pi ik

P P =
ik

Pi Pk = ik

Pi ik =
i

Pi = P

the projection operators are complete, i.e., P = 1

proof: P e k =
Pii ek = i e k () i e i ik = ek

e k

, k

P = 1

summary: decompose the space V into irreducible subspaces: V = labels the irrep counts how often the irrep occurs denote the basis of V by |, , k (i = 1, . . . , n ). We then have P |, , k = |, , k Pi |, , k = |, , i ik
|, , k = |, , j jk Pji , V

4.3. Irreducible operators and the Wigner-Eckart theorem


consider a set of operators {Oi } (i = 1, . . . , n ) acting on a space V . If they transform under a group G as U (g )Oi U (g )1 = Oj D (g )ji

they are called irreducible operators (or irreducible tensors) corresponding to the irrep
consider now a set of irreducible operators {Oi } and a set of irreducible vectors { e j }. How do the vectors Oi ej transform? U (g )Oi ej = U (g )Oi U (g )1 U (g ) e j = Ok e D (g )ki D (g ) j

i.e., they transform in the product representation D

52

4. Expansion in irreducible basis vectors

this product representation can be reduced (see Sec. 2.6), and the vectors Oi ej can be expanded in the irreducible basis vectors w : Oi ej = (we assume here that the operators Oi are normalized in such a way that the vectors Oi ej are normalized, otherwise the CG coecients would not form a unitary matrix.) This leads to the Wigner-Eckart theorem: e Oi ej = m e wm w

, , (, )i, j

, , m(, )i, j 1 n
e k wk k

= m e
Oi

(see Sec. 4.1)

e j

, , (, )i, j

with the so-called reduced matrix element O which does not depend on i, j, or . very important theorem, since the (many) matrix elements on the LHS can be expressed in terms of (only a few) reduced matrix elements on the RHS (the CG coecients are tabulated). in practice the reduced matrix elements are computed as follows: compute as many matrix elements on the LHS as there are reduced matrix elements view the Wigner-Eckart theorem as a system of linear equations for the reduced matrix elements and solve for the latter example: electromagnetic transitions in atoms and nuclei (O(3) symmetry) the states are |j, m (denite angular momentum) the operators are "multipole transition opertors" (i.e., the quantized electromagnetic vector potential A(x, t) expanded in ladder operators for denite angular momentum, see Tung Sec. 8.7)

1 n

e k wk k

4.4. Left ideals and idempotents


so far we have learned how to decompose a reducible representation into irreducible components to do this we need to know the irreps question for this section: How to construct the irreps ? consider the regular representation (see Sec. 2.5), since it contains all irreps D (with multiplicity n = dim(D )) the carrier space of the regular representation is the space of the group algebra (or = span(g1 , . . . , gn ) Frobenius algebra ), G

4.4. Left ideals and idempotents

53

is a vector in G (then we call it |p ), but also an operator on G every element p G (then we call it p): r |q = (ri gi ) qj gj = gi gj ri qj = |gk ri (i )kj qj is the reason for the special properties of the regular this dual role of the elements of G representation can be decomposed into invariant, irreducible subspaces corresponding to the irreps G of G that is invariant under left multiplication is called left ideal: a subspace L of G L = {r} with p |r L p G gi gj = gk (i )kj

a left ideal always contains the zero element of G if L is an irreducible subspace, it is called minimal or irreducible left ideal if we have found the irreducible left ideals, we can construct the irreps of G from them (by acting with the group elements on the basis vectors of the left ideal) task: nd the irreducible left ideals
denote the projection operator on an irreducible left ideal L by P . This operator must have the following properties: , in short: P G = L 1) P |r L r G

2) if |q 3)
r P

. This element can be expanded as |s = proof: consider an arbitrary s G with s L


rP |s = rP P r |s s = r s = rs s = P rs = rs L L

|q L , then P r G = rP

= |

P r

P = P 4) P

furthermore we dene L =

= G

and rst construct the projection operator P on L there is a unique decomposition e = let e be the identity of G. Since e G suitable e L . Then P is given by right multiplication with e , i.e., P |r re proof: r G
e

with

54

4. Expansion in irreducible basis vectors

P is a linear operator (proof simple) we have on the one hand for r G r=

with

r L

and on the other hand r = re = r


e =

re

re L (since L is a left ideal) with

r = re = P r

this proves properties 1. and 2. we have for arbitrary q, r G P r |q = P |rq = rqe rP |q = r qe = rqe P r = rP
(property 3.)

the decomposition of e in invariant components is on the one hand e = 0 + + e + + 0 and on the other hand e = e e = e

e = e e1 + + e e + e e +1 +
L1 L L +1

since the decomposition is unique we have e e = e P P = P


(property 4.)

this derivation works in the same way for projectors on irreducible left ideals, dened by
P |r re

elements of the group algebra satisfying e e = e are called idempotents. If there is an additional normalization factor on the RHS, they are called essentially idempotent. the idempotent e generates the left ideal L : } L = {re | r G an idempotent is called irreducible or primitive if it generates an irreducible left ideal
L = {re | r G}

otherwise it is called reducible and can be written as the sum e1 + e2 of two idempotents (with e1 e2 = 0).

4.4. Left ideals and idempotents

55

How can we nd out if an idempotent is primitive? Theorem: An idempotent e is primitive if and only if ere = r e Proof: 1) Suppose e is a primitive idempotent } is an irreducible left ideal. L = {re | r G an operator R on G : Associate with r G R |q |qere L R is a projection operator on L. Rs |q = |sqere = sR |q s G r G (r is a number)

by Schurs lemma 1, R is proportional to the identity in L. ere = r e and e = e1 + e2 with two idempotents e1 = 0 and e2 = 0 2) Suppose ere = r e r G (and with e1 e2 = 0) ee1 = (e1 + e2 )e1 = e1 e1 + 0 = e1 ee1 e = = e1 = e (by assumption ere = e) 2 2 e1 = e1 = ee = e = 2 =0 or =1 e1 = 0 or e1 = e , i.e., e2 = 0

which contradicts the assumption e1 = 0 and e2 = 0. We also need a criterion to decide whether two primitive idempotents generate equivalent or non-equivalent irreps. Theorem: Two primitive idempotents e1 and e2 generate equivalent irreps if and only if . e1 re2 = 0 for an r G Proof: Let L1 and L2 be the two irreducible left ideals generated by e1 and e2 , and D1 and D2 the corresponding irreps. . Consider the linear map S : 1) Suppose e1 re2 = s = 0 for an r G L1 we have For all p G Sp |q1 = S |pq1 = |pq1 s = p |q1 s = pS |q1 Sp = pS as an operator on L1 operator on L2 since S projects on L2 q1
S

q2 = q1 s L2 .

SD1 (p) = D2 (p)S

by Schurs lemma 2, D1 and D2 are equivalent.

56

4. Expansion in irreducible basis vectors

2) If D1 and D2 are equivalent, there is a transformation S such that SD1 (p)S 1 = , or Sp = pS as a linear map from L1 to L2 D2 (p) p G |s S |e1 L2 s = se2 = e1 s |s = |se2 e1 se2 = s

|s = S |e1 = S |e1 e1 = Se1 |e1 = e1 S |e1 = e1 |s = |e1 s

The trivial representation is generated by the primitive idempotent e1 = 1 n


n

gi
i=1

1 from Sec. 4.2) (this is the generalized projection operator P11 Proof: } with 1) The left ideal generated by e1 is L1 = {re1 | r G

re1 =
j

rj gj

1 n

gi =
i

1 n

rj
j i

gj gi
k

gk

= ce1

where c =
i

ri

is a number

L1 is 1-dimensional and thus irreducible. 2) The elements of L1 are invariant under the group: g ce1 = c n ggi =
i

c n

gk = ce1
k

trivial representation of the group. Summary: The group algebra can be decomposed into left ideals L ( corresponds to the non-equivalent irreps of the group) The L are generated by right multiplication with the idempotents e , with e e = e and e = e. Every L can be decomposed into n irreducible left ideals L ( = 1, . . . , n ).
The L are generated by right multiplication with the primitive idempotents e , with e re = r e

r G

If we have found all primitive idempotents we can easily construct all irreps of the group from them. the task is to nd all non-equivalent primitive idempotents. exercises: reduction of the regular representation of C3 .

5. Representation of the symmetric group and Young diagrams


5.1. Why Sn is important
The representation theory of Sn is the basis for the study of many other groups: nite groups of order n are isomorphic to subgroups of Sn the primitive idempotents of Sn are used to construct the irreps of the classical continuous groups, e.g., U(m), O(m), etc. for systems of identical particles, the symmetry group of the Hamiltonian H always contains Sn as a factor. the eigenstates of H transform in irreps of Sn .

5.2. 1-dimensional and associated representations of Sn


An ("alternating group") = group of the even permutations (i.e., even number of transpositions) An is an invariant subgroup of Sn the factor group Sn /An is isomorphic to Z2 Sn has two irreps given by the irreps of Z2 (see Sec. 1.10) Ds (p) = 1 p Sn (trivial representation)

Da (p) = (1)p

1 for even p 1 for odd p

(1)p is called parity of the permutation p. The 1-dimensional irreps also follow from this theorem: The symmetrizer s p p and the antisymmetrizer a tially idempotent and primitive (proof in exercises). for all p Sn we have spa = sa =
rearrangement lemma

p p (1) p

of Sn are essen-

(1)r qr =
qr q

(1)q
r

(1)q+r qr = a
q

(1)q = 0
=0

=a
rearrangement lemma

s and a generate non-equivalent irreps of Sn with basis vectors |ps and |pa . for all p Sn we have ps = s and pa = (1)p a both irreps are 1-dimensional, with matrix elements 1 or (1)p . = Suppose we have a representation D of Sn with dimension n . Then D and D a D D are called associated representations

57

58

5. Representation of the symmetric group and Young diagrams

also has dimension n . D (p) = (1)p D (p) D


p

(p)

=
p

(p)

2 ?

= n!

is irreducible if and only if D is irreducible. D are equivalent (because in this case all char If (p) = 0 for all odd p, D and D acters are the same, see Sec. 2.4.4), and D is called self-associated. Otherwise they are non-equivalent. the characters of the irreps of Sn are real. Proof: p1 is in the same class as p (p) = (p1 ) = (p)
representation is unitary

this yields a theorem that is important for applications to systems with bosons and fermions. Theorem: Let D and D be two irreps of Sn : 1) D D contains Ds exactly once (not at all) if D and D are equivalent (nonequivalent). 2) D D contains Da exactly once (not at all) if D and D are associated (not associated). (proof in exercises)

5.3. Young diagrams


a partition (1 , . . . , r ) of an integer n is a sequence of positive integers such that
r

i = n
i=1

with

i i+1

two partitions and are equal if i = i for all i. > ( < ) if the rst nonzero term in the sequence i i is positive (negative). a partition is represented graphically by a Young diagram: n boxes arranged in r rows. The i-th row contains i boxes. examples: for n = 3 there are 3 distinct partitions: (3) (2,1) (1,1,1)

for n = 4 there are 5 distinct partitions:

5.3. Young diagrams

59

(4)

(3,1)

(2,2)

(2,1,1)

(1,1,1,1)

every partition of n corresponds to a class of Sn and vice versa: Every class corresponds to a certain cycle structure. The i-th row of the diagram can be interpreted as a i cycle. Each of the numbers 1, . . . , n occurs in one and only one of the cycles i i = n. The number of Young diagrams for n equals the number of classes of Sn , and therefore the number of irreps of Sn . Example: For S3 we have: {e}: {(12), (13), (23)}: {(123), (321)}: three 1-cycles one 2-cycle, one 1-cycle one 3-cycle (1,1,1) (2,1) (3)

a Young tableau is a Young diagram containing the numbers 1, . . . , n in the boxes (each number appearing once). Example: 3 4 1 2 or 2 4 3 1

in a normal Young tableau the numbers 1, . . . , n appear in ascending order, rst from left to right, then from top to bottom. Example: 1 2 3 4 or 1 2 3 4

For every Young diagram there is one and only one normal Young tableau. In a standard Young tableau the numbers increase (but not necessarily in strict order) in the rows and columns. Example: 1 2 4 3 or 1 4 2 3

The notation in the literature is sometimes dierent, e.g., Young graph, Young pattern, Young frame, etc. The normal Young tableau of the partition will be called . An arbitrary tableau is obtained from by a permutation p of the n numbers in the p qp boxes: p p . We have q = .

60

5. Representation of the symmetric group and Young diagrams

5.4. Symmetrizers and antisymmetrizers of Young tableaux


We shall see that for every Young tableau we can dene a primitive idempotent that n . generates an irrep of Sn on the group algebra S Given a Young tableau p we dene
p horizontal permutations hp which only permute the numbers in the rows of . p vertical permutations v which only permute the numbers in the columns of p .

For a Young tableau p we dene symmetrizer : antisymmetrizer : irreducible symmetrizer :


(or Young operator)

sp =
h

hp
p (1)v v v p p p (1)v hp v = s a h,v

ap ep

= =

ep is very important since it is a primitive idempotent. example: Standard tableaux of S3 : 1 = 1 2 3 : all p are h s1 = p p = s (symmetrizer of S3 ) only e is a v a1 = e e 1 = s 1 a1 = s 1 2 : the h are e and (12) s2 = e + (12) 3 the v are e and (13) a2 = e (13) e2 = s2 a2 = e + (12) (13) (321) 1 2 : only e is a h s3 = e 3 all p are v a3 = p (1)p p = a (antisymmetrizer of S3 ) e3 = s3 a3 = a = 1 3 (23) : the h are e and (13) s2 = e + (13) 2 (23) the v are e and (12) a2 = e (12) (23) (23) (23) e2 = s2 a2 = e (12) + (13) (123)

2 =

3 =

(23)

We can learn a lot from this example: (In the following we omit the superscript p to simplify the notation.) 1) For every tableau , the horizontal and vertical permutations each form a subgroup of Sn . 2) The s and a are (total) symmetrizers and antisymmetrizers of the corresponding subgroup. s h = h s = s a v = v a = (1)v a
r

s s = n s a a = n a

with

n =
i=1

i !

5.5. Irreducible representations of Sn

61

s and a are essentially idempotent, but in general not primitive. 3) The e are primitive idempotents (exercises). 4) e1 = s and e3 = a generate the two 1-dimensional irreps of S3 (see Sec. 5.2). e2 3 (by right multiplication with e2 ): generates a 2-dimensional left ideal L2 of S ee2 = e2 (12)e2 = (12) + e (321) (13) = e2 (23)e2 = (23) + (321) (123) (12) r2 (13)e2 = (13) + (123) e (23) = e2 r2 (123)e2 = (123) + (13) (23) e = e2 r2 (321)e2 = (321) + (23) (12) (123) = r2

i.e., L 2 = span(e2 , r2 ). Since e2 is primitive, L 2 is irreducible. The Young operators of the normal Young tableaux generate all irreps of the group. 5) e2
(23)

also generates a 2-dimensional irrep:

This irrep must be equivalent to the irrep generated by e2 (since there is only one 2-dimensional irrep). The irreducible left ideal generated by e2 L2
(23) (23) (23)

is = (123) (13) + (23) (321).

= span e2

(23)

, r2

(23)

with

r2

This is orthogonal to the other left ideals L1 = span(e1 ), L3 = span(e3 ) and L2 . 3 is the direct sum of the four irreducible left ideals. The decomposition of the 6) S identity is 1 1 1 (23) 1 e = e1 + e2 + e2 + e3 6 3 3 6 the regular representation of S3 is reduced by the Young operators of the standard Young tableaux.

5.5. Irreducible representations of Sn


we now generalize what we learned about S3 in Sec. 5.4 to Sn (some proofs in exercises, some omitted) again we omit the superscript p for simplicity; everything is valid for arbitrary Young tableaux. Theorem 1: The symmetrizers of the Young tableau satisfy s ra = r e e2 = e e is essentially idempotent. , n r S with r , numbers, and = 0

62

5. Representation of the symmetric group and Young diagrams

Theorem 2: The Young operator e is primitive idempotent and therefore generates an n . irrep of Sn on S Theorem 3: The irreps generated by e and ep (with p Sn ) are equivalent. Theorem 4: The Young operators e and e generate inequivalent irreps if the corresponding Young diagrams are dierent (i.e., if = viewed as partitions). Theorem 5: The Young operators of the normal Young tableaux generate all inequivalent irreps of Sn . (I.e., we can identify the dierent irreps with the dierent Young diagrams.) Theorem 6: a) The irreducible left ideals generated by the Young operators of the standard tableaux are linearly independent. n . b) The direct sum of these left ideals generates S

5.6. More applications of Young tableaux


proofs omitted (too lengthy) the dimension of the irrep D corresponding to the Young diagram equals the number of standard tableaux and is given by n = n! with
i<j ( i i

i!

j)

n! i,k hik

n! = order of Sn i, j = 1, . . . , r (number of rows of the diagram) k = 1, . . . , i (number of boxes in row i)


i

= i + r i = number of boxes along a hook from i, k to the right and downwards example: h23 = 7

hik = "hook length" of the box i, k

this implies that Sn has only two 1-dimensional irreps (Ds and Da from Sec. 5.2):

Ds =

n boxes

Da =

n boxes . . .

5.6. More applications of Young tableaux

63

associated to an irrep D is obtained by transposing the corre the representation D sponding Young tabeau , i.e., exchanging rows and columns:

recursive calculation of the characters of the irreps of Sn : the boundary [] of a Young diagram = "south-east boundary" = set of boxes whose right edge or lower edge or right lower vertex is part of the border of the diagram. (Sometimes [] is called staircase.) 1 3 2 Example: 6 5 4 7 in general: [] = all boxes (i, j ) so that (i + 1, j + 1) / . a skew hook = connected part of the boundary that can be removed to leave a proper Young diagram. In the example above: 1-2, 1-4, 1-5, 1-7, 2, 2-4, 2-5, 2-7, 4, 4-5, 4-7, 7 a box at the end of a row is the beginning of a skew hook a box at the end of a column is the end of a skew hook every hook corresponds to a skew hook and vice versa1 . The hook length then equals the length of the skew hook. 1 3 2 Example: The skew hook 1-5 corresponds to the hook 6 5 4 7 the leg length of a skew hook is the number of vertical steps (=number of rows covered by the skew hook - 1) let c be a class of Sn with cycle structure c = (a1 , a2 , . . . , aq ) (in any order). What is the character c of this class in irrep D ? select an arbitrary cycle of c with length ai . denote by c the class of Snai obtained by removing the cycle ai from class c start with the Young diagram of D , determine all skew hooks with length ai , and denote the Young diagram of Snai obtained by removing such a hook by we then have c =

+ sign for skew hooks with even leg length - sign for skew hooks with odd leg length
1

This should rather be taken as an additional condition for skew hooks.

64

5. Representation of the symmetric group and Young diagrams

use the same algorithm to calculate the c recursion


=0 1 (and include the sign depending on the leg if nothing remains, () length)

if there is no skew hook with length ai , then c =0 to make this method ecient, the order in which the ai are removed should be chosen to minimize the total number of skew hooks to be removed example: S13 , c = (7, 4, 2), = (6, 3, 3, 1) = there is only one hook with length 7. This hook corresponds to the skew hook (with leg length 2) (7,4,2) = +(4,2)
(6,3,3,1) (2,2,1,1)

now there is only one hook with length 4. This hook corresponds to the skew hook (with leg length 2) (7,4,2) = +(2) = 1
(6,3,3,1) (2)

there is also a non-recursive (but less ecient) method: determine all possibilities to "disassemble" the Young diagram completely by successfully removing skew hooks with lengths a1 , . . . , aq denote by k the sum of the leg lengths of the skew hooks occuring in possibility then c =

(1)k

6. Lie groups
6.1. Introduction
so far we considered nite discrete groups. We now proceed to consider continuous groups in this case the group elements are continuous functions of nitely many real parameters: G g = g (1 , . . . , n ) g () "order n" or "dimension n" or "n parameters"

i.e., a group element is specied by a point on a n-dimensional manifold the group elements g () and g ( ) are "close to each other" if their "distance" ( i )2 )1/2 in parameter space is small
i (i

the parameterization in not unique. We could also parameterize the group elements by n linearly independent well-behaved functions i () (i = 1, . . . , n) the multiplication law is now 1 . . . = f1 (1 , . . . , n ; 1 , . . . , n )

g ()g ( ) = g ( )

with

n = fn (1 , . . . , n ; 1 , . . . , n ) the structure of the group is determined completely by the functions f1 , . . . , fn (similar to the multiplication table for nite groups) the group must have an identity element I . I corresponds to a certain point 0 = 0 , . . . , 0 ) in parameter space. (1 n we have for all i and
0 0 0 0 fi (1 , . . . , n ; 1 , . . . , n ) = fi (1 , . . . , n ; 1 , . . . , n ) = i

for every g G there must be an inverse element g 1 . This corresponds to the point = ( 1 , . . . , n ) in parameter space, and from gg 1 = g 1 g = I we have
0 fi (1 , . . . , n ; 1, . . . , n ) = fi ( 1, . . . , n ; 1 , . . . , n ) = i

We only consider groups for which we can invert these equations (at least in principle):
0 0 i = hi (1 , . . . , n ; 1 , . . . , n )

the group is called Lie group if the functions fi and hi are analytic for all i (i.e., all derivatives are continuous) a Lie group is called compact if the intervals of the parameters i are nite and closed; otherwise it is called non-compact

65

66

6. Lie groups

6.2. Examples of Lie groups


1) SO(2) rotations about a xed axis (e.g., z ) 1 parameter: rotation angle multiplication law: R(1 )R(2 ) = R(1 + 2 ), i.e., f = 1 + 2 abelian group identity: = 0 inverse element: R()1 = R() 2) Translation group: translation (for example) in xy -plane 2 parameter group (i.e., order 2) multiplication law: T (x1 , y1 )T (x2 , y2 ) = T (x1 + x2 , y1 + y2 ) identity: x = y = 0 inverse: T (x, y )1 = T (x, y ) (Note that if we had parameterized the coordinate system dierently, e.g., in polar coordinates (r, ), the multiplication law would not be quite as straightforward.) 3) General linear group Gl(N, C): group of non-singular linear homogeneous transformations in an N -dimensional complex space the dening or fundamental representation is given by complex N N matrices A such that det A = 0. The vectors of the space transform as
N

zi =
j =1

Aij zj

we have 2N 2 (real) parameters the multiplication law is A = AA or Aij (A, A ) =


k

Aik Akj

2N 2 equations (for real and imaginary parts of Aij ) identity:

inverse of A is A1 All the following examples are subgroups of Gl(N, C). 4) Unitary group U(N ): group of linear unitary transformations in an N -dimensional complex space unitary means that the quantity |zi |2 = |z1 |2 + + |zN |2
i

is left invariant

6.2. Examples of Lie groups

67

this imposes some restrictions on the matrices of U(N ): |zi |2 =


i ijk

(Uij zj ) Uik zk =
ijk

Uij Uik zj zk = k

|zk |2

i.e.,
Uij Uik = jk i

U U = 1N = U U

this implies: k=j:


i Uik Uik = i Uij Uik = 0 i i Uij Uik = 0

|Uik |2 = 1

N equations N (N 1) equations

k<j:

the total number of parameters is thus 2N 2 (N + N 2 N ) = N 2 also we have 1 = det U U = | det U |2 5) Special unitary group SU(N ): subgroup of U(N ) restriction to unitary matrices with det U = 1 N 2 1 parameters 6) Orthogonal group O(N ): group of orthogonal transformations in an N -dimensional real space orthogonal means that x2 i
i

det U = ei

(pure phase)

is preserved the xi transform as xi =


j

Oij xj

thus we have (xi )2 = =


i ijk T Oki Oij xj xk = j !

x2 j

T Oki Oij

= jk

O O = 1N = OOT
T

so that k=j:
i 2 Oik =1

N equations 1 N (N 1) equations 2

k<j:
i

Oij Oik = 0

68

6. Lie groups

the number of parameters is 1 1 1 N 2 N + N 2 N = N (N 1) 2 2 2 here we have 1 = det OOT = (det O)2 det O = 1

the parameter space of O(N ) decomposes into two disconnected pieces: a) transformations with det O = 1 continuously connected to the identity SO(N ) which is a subgroup of O(N ) b) transformations with det O = 1 not continuously connected to the identity they are products of rotations and reections the most important case is O(3) 3 parameters homomorphic to SU(2) (see Sec. 1.11) often parameterized by Euler angles 7) non-compact unitary group U(n, m): Transformations in an (n + m)-dimensional complex space that leave |z1 |2 + + |zn |2 |zn+1 |2 |zn+m |2 invariant i.e., the metric is diag(1, . . . , 1, 1, . . . , 1)
n m

8) non-compact orthogonal group O(n, m): transformations in an (n + m)-dimensional real space which leave
2 2 2 x2 1 + + xn xn+1 xn+m

invariant (metric is the same as in 7.) 9) Lorentz group O(3,1) (see Sec. 1.11): inhomogeneous Lorentz group = Poincar group = O(3,1) + 4-dimensional translations x = x + a most denitions (subgroups, factor groups etc.) are analogous to those for nite groups, e.g., U(n) = (SU(n) U(1))/Zn , O(n)/SO(n) = Z2

6.3. Invariant integration

69

6.3. Invariant integration


in many proofs we used the rearrangement lemma in the form
n n n

f (Ai ) =
i=1 i=1

f ( Ai B ) =
i=1

f (BAi )

()

with n = order(G) and Ai , B G. instead of the sum over i we now have integrals over the parameters 1 , . . . , n of the group G. In the following we write A = g (1 , . . . , n ) G. To obtain an analogue of () we need an integration measure d(A) = d(1 , . . . , n ) = (1 , . . . , n ) d1 . . . dn
density function

such that d(A)f (A) =


G G

d(A)f (AB ) =
G

d(A)f (BA)

for arbitrary B G. the integration is over the entire parameter space of G:


G

d(A)f (AB ) =

A =AB G

d(A B 1 )f (A ) =
G

d(A )f (A )

we want this to hold for arbitrary functions f and therefore have to require d(A) = d(AB ) = d(BA) B G

such a d(A) is called invariant integration measure or Haar measure if we nd such a measure, many of the results obtained for nite groups can be carried over to continuous groups

6.4. Properties of compact Lie groups


out of the examples in Sec. 6.2, U(N ), O(N ), and their subgroups are compact Theorem: For a compact Lie group there exists an invariant integration measure. (For concrete calculations we often do not need to know it explicitly.) a compact Lie group has a countably innite number of irreps, all of nite dimension. every matrix representation is equivalent to a unitary representation every representation can be decomposed into a sum of irreps the matrix elements of the irreps are now functions of the group parameters and satify orthogonality and completeness relations:
d(A)D (A) ij D (A)k = G

ik j n

d ( A)
G vol(G)

n D (A)ij D (A
ij

) ij

= (A A )

70

6. Lie groups

n = dim(G) vol(G) = volume of the group (normalization arbitrary). Typically one normalizes such that vol(G) = 1, which we will do in the following orthogonality relation for characters: d(A) (A) (A) = vol(G)
G

criterion for the reducibility of the representation : d(A)| (A)|2 = vol(G)


G

for irreducible basis functions we have as in Sec. 4.1:


u i vj = ij

1 n

u k vk k

with the following denition of the scalar product: f|g =


G

d(A)f (A)g (A)

Peter-Weyl Theorem: Every well-behaved function f (1 , . . . , n ) on the parameter space of G can be expanded in the irreducible representation functions of G: (vol(G) = 1)
n c ij D (1 , . . . , n )ij =1 i,j =1

f (1 , . . . , n ) = with
c ij = n Dij f = n G

d(A)D (A) ij f (A)

6.5. Generators of Lie groups


many properties of Lie groups follow from the behavior of the group elements near the identity (innitesimal transformations). Finite transformations are then obtained by many successive innitesimal transformations.
0 = 0 for all i, i.e., g (0, . . . , 0) = I in the following we choose the parameters such that i

because of e0 = 1 it is sensible to write an innitesimal group element as g ( 1 , . . . , n ) = e


i
n S j =1 j j

the operators Si are called generators of the group (one generator per parameter) for the special groups with det(g ) = 1 the generators are traceless: det(g ) = eTr ln g = e
i
j

j Tr Sj !

=1

Tr Sj = 0

6.5. Generators of Lie groups

71

for unitary groups with gg = 1 the generators are Hermitian: g = g 1 e


i
j j Sj

=e

j Sj

Sj = Sj

the generators are formally dened through derivatives: iS1 = lim i.e., g (, 0, . . . , 0) g (0, . . . , 0) 0 g Si = i i =0 etc.

here we have to distinguish two cases: 1) the group elements A are maps of a coordinate system on itself 2) the group elements are OA operators acting on functions (see Sec. 2.4.2) example: SO(3), i.e., rotations in 3 dimensions. Here we use the following (widely used) conventions: (a) = for a mathematically positive rotation angle (counterclockwise) (b) xi = Aij xj , i.e., x = Ax (c) the OA operators are dened by OA f (x) = f (A1 x), as in Sec. 2.4.2 case 1: rotation by an angle about the z -axis: x = Ax with cos sin 0 Az () = sin cos 0 0 0 1 0 0 0 analogously : Sx = 0 0 i 0 i 0

Sz = i lim

Az () 1 0

0 i 0 = i 0 0 0 0 0

0 0 i Sy = 0 0 0 i 0 0

case 2: for an innitesimal rotation about the z -axis we have x + y x 1 0 1 y x y 1 0 = A ( ) x = z z z 0 0 1


1 OAz ()f (x) = f (A z x) = f (x + y, y x, z )

= f (x, y, z ) + (yx xy )f (x, y, z ) + = (1 iSz )f (x, y, z ) + Sz = i(yx xy ) = Lz L is the angular momentum operator: 1 L=rp=r i Sx = Lx (with , = 1)

analogously :

Sy = Ly

physical meaning: the generators are operators corresponding to measurable quantities the commutator of two generators is a linear combination of all generators. To see this, consider innitesimal transformations with one generator each:

72
gi p gij
1 gj

6. Lie groups

gj
1 gi

gi = eii Si = 1 + ii Si +
1 1 gj gi gj gi =

2 i 2 S + 2 i

(no sum over i) 2 i 2 S 2 i

1 ij Sj +
+ = =

2 j 2 S 2 j

1 ii Si +

1 + ij Sj +

2 i 2 S 2 j

1 + ii Si +

2 i 2 S 2 i

= gij =
!

1 + i j [Si , Sj ] + 1 + i j
k

ck ij Sk +

since gij must be a group element whose innitesimal parameters are proportional to i j .
n

[Si , Sj ] =
k=1

ck ij Sk

the coecients ck ij are called structure constants of the Lie group the commutation relations of the generators correspond to the multiplication law of the group near the identity, i.e., they determine the local structure of the group, but not the global (topological) properties
k for the structure constants we have ck ij = cji and the Jacobi identity n m n m n cm ij cmk + cjk cmi + cki cmj = 0 m

the generators form a vector space of dimension n, the Lie algebra L(G) of the group G the law of addition of the generators follows from the multiplication law of the group furthermore a multiplication can be dened: Si Sj [Si , Sj ] the Lie algebra is closed under this multiplication a subalgebra is a set of generators that is closed under multiplication if H is a subgroup of G, L(H ) is a subalgebra of L(G) a subalgebra L(H ) of L(G) is called invariant if [Sg , Sh ] L(H ) for all Sg L(G) and Sh L(H ) a Lie group is called simple if it does not have any nontrivial abelian invariant Lie subgroups (discrete subgroups are allowed). It is called semi-simple if it does not have any nontrivial abelian invariant Lie subgroups

6.6. SO(2)

73

the structure constants form a representation of the Lie algebra, the adjoint representation: Si ck ij

these are n matrices (i = 1, . . . , n) with k = row index and j = column index. Proof: These n matrices satisfy the same commutation relations as the generators. if the Lie group is simple, the adjoint representation is irreducible; otherwise it is reducible. the Killing form is a scalar product of A, B L(G) dened as: (A, B ) Tr Dad (A)Dad (B ) Tr ad (AB ) applied to the generators we obtain the Cartan metric
n

gij Tr

ad

(Si Sj ) =
k, =1

ck i cjk

(i, j = 1, . . . , n)

det gij = 0 is a necessary and sucient condition for G to be semi-simple the number of mutually commuting generators is called rank of G: [Si , Sj ] = 0 for i, j = 1, . . . ,

(every group has at least rank 1) The commuting generators span the Cartan subalgebra and can be diagonalized simultaneously for a group of rank there are Casimir operators (polynomials in the generators) that commute with all generators. Example: Quadratic Casimir operator C2 = ij gij Si Sj (with gij the Cartan metric)

6.6. SO(2)
SO(2) = group of rotations in the plane about a xed point 1 parameter; natural choice: rotation angle with 0 < 2 . "dening" representation: Action of SO(2) on a 2-dimensional vector xi Rij xj with R() = cos sin sin cos

abelian group: R(1 )R(2 ) = R(1 + 2 ) innitesimal rotation with generator J : R(d) = eidJ = 1 idJ + R( + d) = R()R(d) = R() idR()J dR() = R() + d + d dR() = iR()J d

this is a dierential equation with boundary condition R(0) = 1 solution: R() = eiJ (this is true for nite rotations, not just innitesimal ones)

74

6. Lie groups

similarly to Sec. 6.5 we obtain J= this gives (with J 2 = 1): R() = eiJ = 1 iJ 1 2 3 + iJ 2! 3! cos sin = 1 cos iJ sin = sin cos 0 i i 0

now consider a vector space V and a representation of SO(2) by unitary operators U () on V : in analogy to the calculation above we nd U () = eiJ with a Hermitian operator J (so that U is unitary) since SO(2) is abelian, all irreps are 1-dimensional. Thus we have for a vector | in an irreducible subspace J | = | U () | = ei | with R (so that U is unitary) because of R(2 ) = R(0) we have e2i = 1 = m Z, i.e., we have irreps U m () = eim that are characterized by integers m: (a) m = 0: R() U 0 () = 1 (trivial representation) (b) m = 1: R() U 1 () = ei ; this is an isomorphism between SO(2) and the complex numbers on the unit circle (c) m = 1: R() U 1 () = ei ; like (b), but in the other direction (d) m = 2: R() U 2 () = e2i ; homomorphism between SO(2) and the unit circle such that the circle is covered twice, etc. for higher m. Only the irreps with m = 1 are faithful the dening representation is reducible and can be reduced by diagonalizing J= 0 i i 0 R= U 1 U 1

it has eigenvalues 1 and eigenvectors e = x i y: Je = e , R()e = ei e , i.e.,

the invariant integration measure is d/2 , since d = d( + ) = d( + ) with = const, and vol(SO(2)) = 1 orthogonality and completeness relations:
2

d m n U () U () = 2 U n ()U n ( ) =

d i(mn) e = mn 2 ein( ) = ( )

0 n=

0 n=

the

U m ()

eim

are the basis functions of the Fourier expansion

6.7. SO(3)

75

multivalued representations and global properties of the group: consider the following map: R() U 1/2 () = ei/2 U 1/2 ( + 2 ) = ei/2i = U 1/2 () U 1/2 ( + 4 ) = ei/22i = U 1/2 () () denes a 1-to-2 map, i.e., with every R() we associate two numbers ei/2 . This is a two-valued representation of SO(2) in the sense that the group multiplication is preseved if both numbers are acceptable (e.g., if only bilinears like the scalar product 1 | 2 are physically relevant, since these are then still invariant under a rotation by 2 ) more generally we can also dene m-valued representations: R() U n/m () = ein/m ()

(n and m must not have a common denominator) in general, continuous groups have multivalued representations only if the parameter space is multiply connected (i.e., it is topologically nontrivial) in case of SO(2) there are closed paths in parameter space (0 < 2m) that wind m times around the circle and therefore cannot be continuously deformed into each other according to our current knowledge: in classical physics only single-valued representations occur in quantum physics only single- and two-valued representations occur

6.7. SO(3)
6.7.1. Angle-and-axis parametrization
SO(3) = group of rotations in 3 dimensions; 3 real parameters we can parametrize a rotation as Rn : ( ) with rotation angle and rotation axis n
z n

sin cos cos

y x

n = sin sin

intervals of the parameters: 0 , Rn ( )

0 < 2,

0 , since Rn (2 ) =

rotations about a xed axis form a subgroup of SO(3), which is isomorphic to SO(2) for an arbitrary rotation R SO(3) we have
1 RRn = RRn ( )R ( )

hence all rotations by the same angle are in the same class

76

6. Lie groups

a rotation can be visualized by a vector = n . The tips of these vectors ll a ball B with radius :
z

n y

the parametrization has a redundancy: Rn ( ) = Rn ( ). Therefore two points that are diametrically opposite on the surface of the ball B can be identied with each other this means that there are two types of closed curves in parameter space: (a) curves that can be shrunk to a point by continuous deformations (b) curves for which this is not possible
a p

curve b is also closed! (in parameter space)

these global properties inuence the possible representations of the group

6.7.2. Euler angles


a rotation can also be parametrized by Euler angles: R = R3 ()R2 ( )R3 ( ) with cos 0 sin cos sin 0 1 0 , R3 ( ) = Rz cos 0 R2 ( ) = Ry ( ) = 0 ( ) = sin sin 0 cos 0 0 1 parameter intervals: 0 , < 2 , 0

relation to angle-and-axis parameters: 1 = ( + ) 2 , tan = tan 2 sin + 2 , cos = 2 cos2 + cos2 1 2 2

6.7.3. Generators
from Sec. 6.5 (with S J ): 0 0 0 J1 = 0 0 i 0 i 0

0 0 i J2 = 0 0 0 i 0 0

0 i 0 J 3 = i 0 0 0 0 0

6.7. SO(3)

77

this can also be written as (Jk )ij = ikij = iijk with the totally antisymmetric tensor the generators transform under SO(3) like vectors, i.e., RJk R1 = J R
K

(proof by matrix multiplication using the Euler angle parametrization) the generator of a rotation about the axis n = nk e k is Jn = nk Jk (proof in exercises)

the rotation in both parametrizations can be expressed in terms of the generators:


iJk nk Rn ( ) = e

R(, , ) = eiJ3 eiJ2 eiJ3 the Lie algebra of SO(3) is [Ji , Jj ] = iijk Jk the rank of SO(3) is 1, i.e., there is only one Casimir operator. The structure constants are ck ij = iijk the Cartan metric is gij ij , and thus
2 2 2 C2 = J 2 = J 1 + J2 + J3

with

[Jk , J 2 ] = 0

for

k = 1, 2, 3

if the Hamiltonian of a quantum mechanical system is invariant under rotations, we have [H, Rn ( )] = 0 [H, Jk ] = 0 for all n and for k = 1, 2, 3

as shown in Sec. 3.2, the states of the system then transform in irreps of SO(3)

6.7.4. Irreps of SO(3)


every representation of a Lie group is also a representation of the Lie algebra a representation of the Lie algebra provides us with a representation of the Lie group if certain global (topological) conditions are met we rst construct irreps of the Lie algebra on a space V (see QM I). To do so we construct irreducible subspaces of V as follows: 1) choose a suitable starting vector 2) generate an irreducible basis by repeated application of the generators on this vector Schurs lemma 1 implies that J 2 1 in an irreducible subspace all states |m in the irreducible subspace have the same eigenvalue of J 2 the irreps of the SO(3) algebra can be labeled by the eigenvalues of J 2

78

6. Lie groups

as a starting vector we choose an eigenstate of the commuting operators J 2 and J3 : J3 |m = m |m dene J = J1 iJ2 a short calculation gives J3 (J |m ) = (m 1)(J |m ) J |m |m 1 or 0 or 0

since the irreducible subspace has nite dimension, the sequence must terminate, say for m = j at the upper end and for m = at the lower end J3 |j = j |j J+ |j = 0 this yields
2 J 2 |j = (J3 + J3 + J+ J ) |j = j (j + 1) |j

, ,

J3 | J |

= | =0

J2 |

2 = (J3 J3 + J+ J ) |

= ( 1) |

since all states in the irreducible subspace have the same eigenvalue of J 2 , we obtain j (j + 1) = ( 1) 2 solutions: = j and = j + 1, but by assuming j we just have = j

since we go from j to

= j in integer steps we obtain j (j ) = 2j N

3 we have irreps Dj of the SO(3) algebra with j = 0, 1 2 , 1, 2 , 2, . . .

the number of basis vectors in irrep j is 2j + 1 for the orthonormal basis vectors |jm we have J 2 |jm = j (j + 1) |jm J3 |jm = m |jm

dim(Dj ) = 2j + 1

J |jm = [j (j + 1) m(m 1)]1/2 |j, m 1 the irreps of the SO(3) algebra yield irreps of the group SO(3): U (, , ) |jm = jm Dj (, , )m m here U is the operator that implements the rotation R(, , ) on V . Multiplying by jn| gives Dj (, , )nm = jn|eiJ3 eiJ2 eiJ3 |jm = ei(n+m) jn|eiJ2 |jm
=:dj ( )nm

6.8. SU(2)

79

3 for j = 0, 1, 2, . . . these irreps are single-valued, and for j = 1 2 , 2 , . . . they are two-valued. Proof (with = = 0 , = 2 ):

Dj [R3 (2 )]nm = ei2m jn | jm = (1)2m nm = (1)2j nm

(1)2(j m) = 1

1 with a suitable R for arbitrary n we have Rn (2 ) = RR3 (2 )R

2j D j [ Rn (2 )] = (1) 1

for all n

the two-valued irreps exist since the parameter space is doubly connected (see Sec. 6.7.1) however, all these irreps are single-valued irreps of SU(2) (see Sec. 6.8.2) relation between representation matrices and spherical harmonics and Legendre functions: Y P
m (, )

2 +1 D (, , 0) m0 4 ( + m)! d ()m0 ( m)!

m (cos )

= (1)m

characters: since all rotations by the same angle are in one class, it is sucient to consider R3 ( ):
j

j ( ) =
m

Dj [R3 ( )]mm =
m=j

eim =

sin (2j +1) 2

sin 2

in particular, we have for the dening or fundamental 3-dimensional representation: 1 ( ) = 1 + 2 cos further properties in Sec. 6.8

6.8. SU(2)
6.8.1. Parametrization
SU(2) = group of unitary 22 matrices with determinant 1; 3 real parameters let A SU (2) with A= a b . c d

The conditions AA = 1 and det A = 1 give 4 equations: |a|2 + |b|2 = 1 |c|2 + |d|2 = 1 ac + bd = 0 ad bc = 1

the solution of these equations yields the standard parametrization of SU(2): A= cos ei sin ei sin ei cos ei with 0 2 0 , < 2

80

6. Lie groups

the matrix A can also be written as A= r0 ir3 r2 ir1 r2 ir1 r0 + ir3 with ri R 2 + r2 + r2 + r2 = 1 and det A = r0 1 2 3

now interpret the ri as Cartesian coordinates in a 4-dimensional Euclidean space the parameter space of SU(2) is the surface of the unit sphere in this space the parameter space is compact and simply connected (as opposed to the parameter space of SO(3))

6.8.2. Relationship between SO(3) and SU(2)


with every 3-dimensional vector x = x1 , x2 , x3 ) we can associate a Hermitian, traceless 22 matrix:
3

X=
i=1

xi i

with the Pauli matrices 1 = then det X = det x3 x1 ix2 x1 + i x2 x3 = |x|2 0 1 1 0 , 2 = 0 i i 0 , 3 = 1 0 0 1

now consider a linear map of X that is induced by an A SU(2) (see Sec. 1.11): X X = AXA ()

X is again Hermitian and traceless and can therefore be associated with a 3-dimensional vector x because of detX = detX we have |x |2 = |x|2 the SU(2) transformation () induces an SO(3) transformation in the 3-dimensional space of the vectors x the two SU(2) matrices A yield the same rotation the map from SU(2) to SO(3) is 2-to-1. In other words: The kernel of the homomorphism, i.e., the preimage of the identity of SO(3), consists of the two SU(2) matrices 1 0 0 1 and 1 0 0 1 SU (2)/Z2

these form a Z2 subgroup of SU(2). Symbolically: SO(3)

in the parametrization of A by the ri we consider r1 , r2 , r3 as independent parameters and


2 + r2 + r2 ) r0 = 1 (r1 2 3

6.8. SU(2)

81

A = 1 corresponds to r1 = r2 = r3 = 0. Near the identity we have A = 1 ii dri the {i } form a basis of the Lie algebra of SU(2) the three matrices 1 2 i satisfy the same commutation relations as the generators of SO(3): k , 2 2 = ik
m

m 2

SO(3) and SU(2) have the same Lie algebra (but not the same global properties) if k we replace Jk 2 a general SU(2) matrix is obtained by exponentiation: A = ei 2 with = n we obtain (exercises) A = 1 cos i n sin 2 2

in Sec. 6.7.4 we have constructed all irreps of the Lie algebra of SO(3) (and thus of the Lie algebra of SU(2)). Since SU(2) is simply connected, all these irreps are single-valued irreps of SU(2)

6.8.3. Inavriant integration


reminder: we require d(A)f (A) =
G G

d(A)f (B 1 A) =
G

d(BA)f (A)
!

to be invariant for an arbitrary function f , therefore: d(A) = d(BA) general method to compute the Haar measure (for an arbitrary group): for a given parametrzation A( ), compute
A i A express the products A1 as linear combinations of the generators: i

A ( ) this denes a matrix A the weight function is then

n A( ) ( )ki ( ) = Sk A i k=1

( ) A ( ) = det A and the Haar measure is: d(A) = A ( ) proof: see Tung Sec. 8.2 for SU(2) we obtain in the standard parametrization of Sec. 6.8.1 (exercises): d(A) = 1 sin(2)d d d 4 2
n i=1 di

82

6. Lie groups

6.8.4. Orthogonality and completeness relations


the irreducible representation functions of SU(2) and SO(3) satisfy the orthogonality relations (with vol(SU (2)) = 1): (2j + 1)
j d(A)Dj (A) mn D (A)m n = jj mm nn

in the parametrization by Euler angles this simplies to 2j + 1 2


1

d(cos )dj ( )mn dj ( )mn = jj


1

no summation over n and m

Peter-Weyl theorem: The irreducible representation functions form a complete basis of the space of square-integrable functions of the group parameters, i.e., f (A) =
jmn j fmn Dj (A)mn

with

j fmn = (2j + 1)

d(A)Dj (A) mn f (A)

more generally, f (A) need not be a scalar function, but can also be a vector or an operator

7. Lorentz and Poincar group


7.1. Relativistic kinematics
in the following = c = 1, g = diag(1,-1,-1,-1) for Minkowski space, t = x0 , x = (x1 , x2 , x3 ), x = (x0 , x) Lorentz transformations (LT) leave
2 x2 = x x = g x x = x2 0 |x|

invariant space-time splits into 3 distinct regions:


x0 light cone time-like (future) space-like space-like x1 time-like (past)

future cone: x2 > 0 and x0 > 0 past cone: x2 > 0 and x0 < 0 space cone: x2 < 0
these 3 regions are separated by the light cone (x2 = 0)

7.2. Generators and Lie algebra


classication of homogeneous LTs and homomorphism with Sl(2, C): see Sec. 1.11 in the following we only consider continuous (i.e., proper and orthochronous) LTs, i.e., the group L0 Poincar group = Lorentz group and 4-dimensional translations x = x + a with 1 = T

the Poincar group is the symmetry group of all relativistic quantum eld theories elementary particles transform in irreps of the Poincar group the Poincar group has 10 generators: 3 for rotations in 3-dimensional space, e.g., Jz = i x y y x

for rotations about the z -axis. (Here and later we choose case 2 of Sec. 6.5, i.e., OA operators; we could have chosen case 1, e.g., maps of the coordinate system to itself)

83

84

7. Lorentz and Poincar group

3 for boosts, e.g., Kx = i t for boosts along the x-direction 4 for translations, e.g., Px = i for translations in the x-direction in covariant notation: P = i x , J = i x x x x J0i = Ji0 = Ki Jij = Jji = ijk Jk x x x t

this leads to the Lie algebra [P , P ] = 0 [P , J ] = i g P g P [J , J ] = i g J g J + g J g J translations form an abelian subgroup rotations form a subgroup pure boosts do not form a subgroup rotations plus boosts form the continuous Lorentz group L0 introducing the denitions: 1 Mi = (Ji + iKi ) 2 we obtain [Mi , Mj ] = iijk Mk [Ni , Nj ] = iijk Nk the Lorentz group has the same algebra as SU(2)SU(2) [Mi , Nj ] = 0 and 1 Ni = (Ji iKi ) 2

7.3. Finite-dimensional representations of the Lorentz group


the irreps of SU(2)SU(2) are products of irreps of SU(2). 1 Notation: (j1 , j2 ) with j1,2 = 0, 2 , 1, 3 2 , 2, . . . These irreps have nite dimension (2j1 + 1)(2j2 + 1) and are also irreps of L0 trivial representation: (0,0) (scalars) two fundamental representations:
1 1) spinor or Weyl representation: (0, 2 ) with dimension 2

the representation matrices are the Sl(2, C) matrices A from Sec. 1.11

7.4. Unitary irreps of the Poincar group

85

1 the 2-dimensional objects that transform in (0, 2 ) are called spinors

2) conjugate spinor representation: ( 1 2 , 0) the representation matrices are A


1 ,0) there is another type of spinor, , that transforms in ( 2

these two spinor types transform into each other under a parity transformation: a Dirac spinor contains both types of spinors, i.e., = , and transforms in the 1 0 ) (1 reducible representation (0, 2 2 , 0) of L ; however, this is an irrep of the full Lorentz group if we include parity, since parity mixes and . The Dirac equation can be derived by applying Lorentz boosts to a Dirac spinor

1 four-vectors transform in the irrep ( 1 2, 2)

antisymmetric tensors of rank 2 (6 components) transform in (1,0)(0,1). Example: Electromagnetic eld strength tensor F symmetric tensors of rank 2 (10 components) can be decomposed into the trace (which transforms in (0,0)) and 9 traceless components (which transform in (1,1)). Example: Energy-momentum tensor T L0 and SU(2)SU(2) have the same Lie algebra, but: L0 corresponds to the exponentiation of {iJk , iKk }, while SU(2)SU(2) corresponds to the exponentiation of {iMk , iNk } the denition of Mk and Nk shows that the sets {Jk , Kk } and {Mk , Nk } cannot be simultaneously Hermitian thus unitary nite-dimensional irreps of SU(2)SU(2) give non-unitary nite-dimensional irreps of L0 physical states have to transform in unitary irreps since symmetry operations are realized by unitary transformations (otherwise transition probabilities between states are not preserved) the unitary irreps have innite dimension since the Lorentz group is non-compact

7.4. Unitary irreps of the Poincar group


7.4.1. One-particle states and Casimir operators
let C be a Casimir operators of the group G, i.e., [C, Si ] = 0 for all generators Si . All basis states of an irrep of G are eigenstates of C with the same eigenvalue (Schurs lemma 1) the irreps of G can be labeled by the eigenvalues of the Casimir operators in the following we consider the action of the Poincar group on the Hilbert space of one-particle states. We will learn: the unitary irreps are characterized by the particle type (mass and spin/helicity) for a given particle type (= irrep) the basis states of the irrep correspond to the possible physical states (four-momentum and spin projection)

86

7. Lorentz and Poincar group

within an irrep the states can be labeled by the eigenvalues of commuting generators because of [P , P ] = 0 we can choose the states in Hilbert space to be eigenstates |p, of four-momentum with P |p, = p |p, here, is a placeholder for all other indices that are necessary to label a state uniquely one Casimir operator of the Poincar group is C1 = P P = P 2 C1 |p, = p2 |p, = (E 2 |p|2 ) |p, = m2 |p,

with m = rest mass of the state (i.e., of the particle) irreps can be labeled by m (among others) a second Casimir operator is C2 = W W , with 1 W = J P 2

the Pauli-Lubanski (pseudo-)vector. (We will look at the eigenstates of C2 later.) the states transform under pure translations as follows: T (a) |p, = eia P |p, = eia p |p, i.e., in unitary irreps of T4 (with dimension 1 since T4 is abelian) pure LTs are implemented on the Hilbert space by unitary operators U (). U () generates a state with four-momentum p = p: P U () |p, = U () U ()P U () |p, = U () P (1 ) |p, = p U () |p, = p U () |p, (see Sec. 4.3)

the state U () |p, is a linear combination of states with four-momentum p : U () |p, =

p ,

D(, p)
unitary representation of the Poincar group

()

the Hilbert space is reducible and can be decomposed into irreducible subspaces w.r.t. the Poincar group (see Sec. 7.4.2), i.e., the matrices D can be brought to block diagonal form the states of a particular particle type can be identied with the basis states of an irrep of the Poincar group

7.4. Unitary irreps of the Poincar group

87

7.4.2. Little group and induced representations


four-momenta p can be divided into six categories: category 1. p2 = m2 > 0, p0 > 0 2. p2 = m2 > 0, p0 < 0 3. p2 = 0, p0 > 0 4. p2 = 0, p0 < 0 3. p2 < 0 3. p 0 standard vector k (m, 0, 0, 0) (m, 0, 0, 0) (k, 0, 0, k ) (k, 0, 0, k ) (0, 0, 0, k ) (0, 0, 0, 0) little group SO(3) SO(3) E2 E2 SO(2,1) L0

massive particles massless particles tachyons vacuum

the cases 2., 4., and 5. are presumably not realized in nature within a category we can, for xed p2 , obtain an arbitrary p by applying a suitable LT to a common "standard vector" k (see table): p = L(p) k or in short p = L(p)k

the states with four-momentum p are then dened by: |p, U L(p) |k, the set of all LTs A that leave a given p invariant, i.e., A p = p is a subgroup of the Poincar group, the so-called little group (or stabilizer or isotropy group) of p little groups for the standard vectors: see table (for E2 see Sec. 7.4.4). (Within a category the little group has the same structure for all p , i.e., it is conjugate to the little group of k ) now consider the action of an arbitrary LT (with p = p) on |p, : U () |p, = U ()U L(p) |k, = U L(p) |k, = U L(p ) U L(p ) U L(p) |k, = U L(p ) U L(p )1 L(p) |k, L(p )1 L(p) leaves k invariant L(p)k = p L(p)k = p = p L(p )1 L(p)k = L(p )p = L(p )1 L(p )k = k A(, p) = L(p )1 L(p) is an element of the little group of k , i.e., U (A) |k, =

k,

D(A)
unitary representation of the little group

88

7. Lorentz and Poincar group

and thus U () |p, = U L(p ) U (A) |p, =

U L(p ) k, p ,

D(A)

D A(, p)

compare with () in Sec. 7.4.1: the representations of the Poincar group are induced by the irreps of the little group (see Sec. 2.7) in general, irreps of a subgroup induce reducible representations of the group. However, in our case the representation of the Poincar group induced by the irreps of the little group are again irreducible. (For proof, see Tung Theorem 10.10 (ii))

7.4.3. Massive particles


case 1 in Sec. 7.4.2
1 the little group is the rotation group SO(3) with irreps labeled by s = 0, 2 , 1, . . .

C2 = W W applied to the standard vector (in the rest frame of the particle) yields C2 |k, = = m2 J 2 |k, = m2 s(s + 1) |k, with s = intrinsic angular momentum ("spin") of the particle spin is indeed a kind of angular momentum the unitary irreps of the Poincar group are labeled by two parameters: mass m and spin s the unitary irreps are innite-dimensional since, for xed p2 = m2 , p can take on innitely many values (= J3 if the the index corresponds to the eigenvalues of the helicity operator J P boost is in the z -direction), see Tung Sec. 8.4.1 the operator commutes with the P and yields the spin component in the direction of motion under a LT the indices are mixed within an irrep

7.4.4. Massless particles


case 3 of Sec. 7.4.2 the generators of the little group are A = J1 + K2 , B = J2 K1 , and J3 . These three generators have the same Lie algebra as the Euclidean group E2 (translations and rotations in a plane) we have [A, B ] = 0 A and B can be diagonalized simultaneously , B |k, a, b = b |k, a, b

A |k, a, b = a |k, a, b

7.5. Parity and time reversal

89

assume that there exists a pair of eigenvalues a, b = 0. This would lead to a continuum of eigenvalues, since after a rotation by an angle we have |k, a, b A |k, a, b B |k, a, b

U 1 [R()] |k, a, b = (a cos b sin ) |k, a, b = (a sin + b cos ) |k, a, b


however, in nature we do not observe massless particles that have such a continuous degree of freedom for physical states we require a = b = 0 the physical states |k, are then distinguished by the eigenvalues of J3 : J3 |k, = |k, as in Sec. 7.4.3 corresponds to the helicity = spin projection on the direction of motion (recall k = (0, 0, k )) can only take on integer and half-integer values (corresponding to single-valued and double-valued irreps, respectively) for massless particles is invariant under LTs states with dierent helicity can in principle be viewed as dierent particles, but: Parity transformations exchange states with opposite helicity if the theory is invariant under parity, the two states correspond to the same particle (with two helicity states). Examples: photon with = 1 graviton with = 2 but: If neutrinos were massless, we would have neutrinos with = 1 2 1 and antineutrinos with = 2 , since the weak interactions violate parity (however, we now know that neutrinos are massive)

7.4.5. Tachyons
see Tung Sec. 10.4.5

7.4.6. Vacuum
see Tung Sec. 10.4.1

7.5. Parity and time reversal


read Tung Sec. 11 and 12 further reading: Weinberg, The quantum theroy of elds, Vol. I, Sec. 2

8. The tensor method to construct irreps of Gl(m) and its subgroups


8.1. Tensors and tensor spaces
Let Vm be an m-dimensional vector space and {g } be the set of invertible linear transfomations on Vm . The transformations g form a group which is isomorphic to Gl(m, C) let {|i , i = 1, . . . , m} be a basis of Vm a matrix representation of Gl(m) is given by g |i = |j gji (sum over j ) , with det(gij ) = 0

this is the so-called dening or fundamental representation of Gl(m) this representation is irreducible, and we simply call it g
n = V V (n factors) is called tensor space the product space Vm m m n is given by a basis of Vm

|i1 . . . in |i1 |in in short: |i


n

with

ik = 1, . . . , m

n can be written as an arbitrary element |x Vm

|x = |i1 . . . in xi1 ...in in short: |x = |i


n x{i} .

The x{i} are called tensor components of |x

n: the elements of Gl(m) induce the following transformation on Vm

g |i

= |j

n D (g ){j }{i}

with

D(g ){j }{i} = gj1 i1 . . . gjn in

the D(g ) form an mn -dimensional representation of Gl(m), the product representation n g g , with carrier space Vm
n transform under Gl(m) as follows: the |x Vm

|xg g |x = g |i

n x{i}

= |j

n D (g ){j }{i} x{i}

= |j

g n x{j }

xg {j } = D (g ){j }{i} x{i}

the objects that transform like |x are called tensors of rank n under Gl(m). In particular, vectors are tensors of rank 1 if we restrict the operators or matrices g to the subgroups U(m), SU(m), or O(m), we can dene tensors under these groups

90

8.2. Action of the symmetric group on the tensor space

91

8.2. Action of the symmetric group on the tensor space


consider a permutation p =
n: mation on Vm

1 ... p1 . . .

n pn

Sn and associate with it a linear transfor-

|xp p |x = |i1 . . . in xp i1 ...in

with

xp i1 ...in xip1 ...ipn


n

here p only acts on the tensor components x{i} , but not on the basis tensors |i

on the other hand, we could also let p act on the basis tensors (but not on the tensor components). Because of |xp = |i1 . . . in xip1 ...ipn = |ip1 . . . ip 1 xi1 ...in n
1

the action of p on the basis tensors is given by


p p |i1 . . . in = |ip1 . . . ip 1 = |i n
1 1

Therefore we have p |i
n

= |j

n D (p){j }{i}

with

1 = jp i . . . jp in D(p){j }{i} = j1 ip1 . . . jn ip n 1 1 n 1

example: m = 2, n = 3, p = (123), and |x = |111 + 3 |112 + 4 |122 + 2 |211 + 5 |212 + 4 |221 i.e., we have x111 = 1 x211 = 2 , , x112 = 3 x212 = 5 , , x121 = 0 x221 = 4
(123)

, ,

x122 = 4 x222 = 0

possibility 1, action on tensor components: xi1 i2 i3 = xi2 i3 i1 , i.e., x111 = 1 x211 = 3 and thus |x(123) = |111 + 2 |121 + 4 |122 + 3 |211 + 4 |212 + 5 |221 possibility 2, action on basis tensors: p1 = (321) p |i1 i2 i3 = |i3 i1 i2 p |111 = |111 p |211 = |121 and thus |x(123) = |111 + 3 |211 + 4 |212 + 2 |121 + 5 |221 + 4 |122 , , p |112 = |211 p |212 = |221 , , p |121 = |112 p |221 = |122 , , p |122 = |212 p |222 = |222
(123) (123)

, ,

x112 = 0 x212 = 4
(123)

(123)

, ,

x121 = 2 x221 = 5
(123)

(123)

, ,

x122 = 4 x222 = 0
(123)

(123)

92

8. The tensor method to construct irreps of Gl(m) and its subgroups

important point #1: The representation matrices D(g ) and D(p) have the following symmetry: For p Sn and {ip } = ip1 . . . ipn we have D{j }{i} = D{j p }{ip } i.e., the matrices are invariant if the same permutation acts simultaneously on the indices of i and j (only the order of the factors of g in D(g ) or of the Kronecker deltas in D(p) changes) important point # 2: the matrices D(g ) (g Gl(m)) and D(p) (p Sn ) commute, i.e., pg |i (proof in exercises)
n are in both the representation D(g ) of Gl(m) and the representation D(p) of Sn on Vm general reducible n

= gp |i

from Sec. 5 we know how to decompose D(p) in irreducible components: by applying n the irreducible symmetrizers of the algebra S in the following we will see that this process also leads to a reduction of D(g )

8.3. Decomposition of the tensor space into irreducible subspaces under Sn and Gl(m)
8.3.1. Symmetry classes in tensor space
let (as in Sec.5): p be a Young tableau ep be the corresponding Young operator n } be the irreducible left ideal generated by e L = {re | r S in the following we will learn: a subspace of the type span(re | )
n and arbitrary for xed | Vm r Sn (i.e., re L )

is invariant and irreducible under Sn a subspace of the type span(ep | )


n for xed ep and arbitrary | Vm

is invariant and irreducible under Gl(m)


n can be decomposed such that the basis tensors have the form the tensor space Vm |, , a with = symmetry class, given by the Young diagram = index for the dierent invariant irreducible subspaces under Sn a = index for the dierent invariant irreducible subspaces under Gl(m) p n for a given Young tableau the {ep | | | Vm } are called tensors of symmetry

8.3. Decomposition of the tensor space into irreducible subspaces under Sn and Gl(m)

93

n for a given Young diagram the {rep | | r Sn , | Vm } are called tensors of symmetry class

n } for xed | , then T () is either empty, consider a subspace T () = {re | | r S or T () is invariant and irreducible under Sn the representation of Sn on T () is given by the irrep that is generated by e on n S proof: let |x T (), then we have by denition |x = re | p |x = pr e | T ()
n r S

n for some r S

T () is invariant under Sn let {ri e } be a basis of L , then {ri e | } is a basis of T (). The action of Sn on L is: p |ri e = |pri e = rj e D (p)ji the action of Sn on T () is pri e | = rj e | D (p)ji T () is irreducible and the reprsentation matrices on T () are the same as n those on S p Sn

8.3.2. Totally symmetric and totally antisymmetric tensors


n boxes
, i.e., es = s the total symmetrizer of Sn . Because of pes = es (for let =s = all p Sn ), Ls is 1-dimensional for given | the irreducible subspace Ts () = span(es | ) is 1-dimensional. Such tensors are totally symmetric (in all indices):

es | =

1 n!

p |i
p

n {i}

= |i

1 n!

{ip }
p

the tensor components are symmetric in all indices all p Sn leave a totally symmetric tensor invariant the representation of Sn on Ts () is the 1-dimensional trivial representation (p 1) example: m = 2, n = 3 es = 1 6 [e + (12) + (13) + (23) + (123) + (321)]. In this case there are 4 dierent totally symmetric tensors: | = |111 | = |112 | = |122 | = |222 es | = |111 1 es | = |112 + |121 + |211 3 1 es | = |122 + |221 + |212 3 es | = |222 |s, 1, 1 |s, 2, 1 |s, 3, 1 |s, 4, 1

94

8. The tensor method to construct irreps of Gl(m) and its subgroups

the space spanned by the tensors of symmetry class s is called Ts totally antisymmetric tensors ( = a) exist for n m, i.e., only up to rank n = m (this is obvious from the fact that for n > m there are at least two of the n indices of the tensor components that are the same and thus all the components vanish) La and Ta () are 1-dimensional the representation of Sn on Ta () is the 1-dimensional irrep p (1)p example: tensors of rank 2 (n = 2) in m dimensions es |ii = |ii 1 es |ij = |ij + |ji 2 i = 1, . . . , m i=j

1 2 2 m+ 1 2 (m m) = 2 (m + m) totally symmetric tensors

es |ii = |ii 1 es |ij = |ij |ji 2


2 1 2 (m m) totally antisymmetric tensors

i = 1, . . . , m i=j

8.3.3. Tensors with mixed symmetry


consider again the example of rank-3 tensors in m = 2 dimensions, and in particular tensors of symmetry = = 1) rst pick | = |112 e | = [e + (12)] |112 |211 = 2 |112 |211 |121 = 2 |121 |211 |112 |, 1, 1 |, 1, 2
1 2 3

with e = [e + (12)][e (13)].

From Sec. 5.4: L = span(e , (23)e )

(23)e | = (23) 2 |112 |211 |121

3 , re | is a linear combination of these two tensors for all r S these two mixed tensors form a basis for a 2-dimensional subspace T (1) that is invariant and irreducible under S3 (see Sec. 5.4) 2) now pick | = |221 e | = 2 |221 |212 |122 (23)e | = 2 |212 |221 |122 |, 2, 1 |, 2, 2

this is a basis for another 2-dimensional invariant irreducible subspace T (2) |, 1, 1 and |, 2, 1 are tensors of symmetry and span the 2-dimensional space T (1) {e | | | V23 } (exercises) T (1) is invariant under Gl(2): ge | = e g | T (1) T (1) is irreducible under Gl(2) (exercises)

8.3. Decomposition of the tensor space into irreducible subspaces under Sn and Gl(m)

95

|, 1, 2 and |, 2, 2 are tensors of symmetry and span the 2-dimensional space (23) T (2) = {e | | | V23 }. T (2) is also invariant and irreducible under Gl(m) the spaces T (a) (a = 1, 2) contain all tensors of symmetry class thus the 8-dimensional tensor space V23 is fully reduced: V23 = Ts T (1) T (2) = Ts T (1) T (2)
invariant under

(23)

S3

invariant under

Gl(2)

we can choose as basis tensors of V23 : the 4 totally symmetric tensors |s, , 1 ( = 1, . . . , 4) from Sec. 8.3.3 the 4 tensors |, , a with = 1, 2 and a = 1, 2

8.3.4. Complete reduction of the tensor space


from the example of Sec. 8.3.3 we can deduce the following general properties (proofs see Tung Sec. 5.5)
n that are invariant and irreducible under S and belong to the same two subspaces of Vm n symmetry class are either identical or disjoint

two invariant irreducible subspaces belonging to dierent symmetry classes are always disjoint the tensor space can be fully decomposed into invariant irreducible subspaces under Sn :
n Vm =

T ()

here only Young diagrams with at most m rows occur the basis tensors of T () are |, , a with a = 1, . . . ,dim T () the basis tensors can be chosen such that the representation matrices of Sn on T () are identical for all belonging to the same symmetry class : p |, , a = |, , b D (p)ba
independent of

p Sn

n w.r.t. the symmetry classes of S automatically leads to a the decomposition of Vm n n into invariant, irreducible subspaces under Gl(m): decomposition of Vm

the subspaces T (a) that are spanned by the |, , a with xed and xed a are invariant and irreducible under Gl(m) the irrep of Gl(m) on T (a) that is furnished by the |, , a is independent of a: g |, , a = |, , a D (g )
independent of a

g Gl(m)

96

8. The tensor method to construct irreps of Gl(m) and its subgroups

there are two formulas for the dimensions of the irreps of Gl(m) constructed in this way:

m1 k=1

det (i m i)mj

dim D

m1 k=1

i,j =1,...,m

of boxes in row i of

= =
ij

(i j i + j )
i<j

m+ji hij

product over all boxes of , i = row index, j = column index

hooklength of box ij (see Sec. 5.6)

back to the example of V23 : dim D 3 1 =2 1 1 2 3 1 = =2 3 1 1 = det , , dim D = det 4 1 =4 0 1 2 3 4 = =4 3 2 1

Ts is a 4-dimensional invariant irreducible subspace under Gl(2) (or: Ts consists of 4 1-dimensional invariant irreducible subspaces under S3 ). = fundamental representation of Symbolic notation for V23 = Ts T (1) T (2): Gl(m) (here m = 2), then =

8.4. Irreps of U(m ) and SU(m )


the irreps of Gl(m) from Sec. 8.3.4 are also (subduced) representations of the subgroups of Gl(m). In general these do not have to be irreps, but for U(m) and SU(m) they are irreps (but not for O(m)) because of U(m) = SU(m) U(1) the irreps of U(m) are given by
k DU (m) (U ) = (det U ) DSU (m) (U )

for

U U(m)

with

kZ

therefore in the following we only need to consider SU(m) for SU(m) the two irreps corresponding to the Young diagrams (1 , . . . , m ) and (1 + k, . . . , m + k ) are equivalent, e.g., for m = 5 and k = 1:

Proof: the two diagrams dier by a factor of (det U )k , and det U = 1 for U SU(m) from the proof one also sees that the Young diagram (m boxes) corresponds to the

trivial representation of SU(m), i.e., U 1 for all U SU(m). Functions that transform under SU(m) in this representation are called SU(m) scalars or SU(m) singlets (but such functions transform in the totally antisymmetric irrep of Sm )

8.5. Complex conjugate representation

97

Irrep of SU(2): fundamental representation: trivial representation: with dimension 2 with dimension 1

m = 2 the Young diagrams have at most 2 rows every irrep corresponding to a diagram with 2 rows is equivalent either to , e.g.,

or to a diagram with 1 row that is obtained by cutting o all columns with 2 boxes, e.g.,

except for , we only need to consider diagrams with 1 row dimensions of the irrep corresponding to a diagram with 1 row and k boxes:
k 2+j1 (k + 1)! m+ji = = =k+1 hij k j + 1 k! j =1

ij

for SU(2), every k N corresponds to one and only one irrep with dimension k+1 irreps of SU(3): fundamental representation: trivial representation: with dimension 3

with dimension 1

m = 3 the Young diagrams have at most 3 rows all irreps are equivalent either to (1 , 2 , 0) with

or to a diagram with at most 2 rows: =

(1 + 2)2 1 + 2 1 1 1 dim( ) = det (2 + 1)2 2 + 1 1 = (1 + 2)(2 + 1)(1 2 + 1) 2 2 0 0 1

8.5. Complex conjugate representation


in the following we only consider SU(m) in the fundamental representation of SU(m) we have m objects, 1 , . . . , m , that transform as i i = j U ji for U SU(m) ()

the fundamental representation has dimension m and is is often denoted by m the matrices U also form a representation of SU(m) since the map U U preserves the group multiplication. This representation is called (complex) conjugate representation and is denoted by m

98

8. The tensor method to construct irreps of Gl(m) and its subgroups

conjugation of () yields
j i i j (U i ) = j (U ) j we now dene an object i that transforms like i (but is not equal to i )

i j (U )i j i.e., the i transform in the conjugate representation. upper (lower) indices are called contravariant (covariant) more generally one can dene tensors that transform in the product representation m m m m:
p factors q factors
q 1 i1 ...ip

j ...j

q k1 U k1i1 . . . U 1 ...kp

...

kp j1 ip (U )

. . . (U )

jq
q

and then reduce them (using the Young diagrams of Sp and Sq ) for SU(2) the representations 2 and 2 are equivalent SU(2) has only real and pseudoreal irreps (the latter are equivalent to their conjugate representations) SU(3) for the irreps with Young diagram = (1 , 2 , 0) dene p1 = 1 2 = # of columns with 1 box p2 = 2 = # of columns with 2 boxes 1 1 dim(D ) = (1 + 2)(2 + 1)(1 2 + 1) = (p1 + 1)(p2 + 1)(p1 + p2 + 2) 2 2

this irrep is also denoted by (p1 , p2 ) the irreps (p1 , p2 ) and (p2 , p1 ) have the same dimension and are complex conjugates of each other, e.g., = (0, 1) (1, 0) = = (0, 3) (3, 0) =

=3 = 10

8.6. Reduction of the product of two irreps


suppose we have two irreps D and D of Gl(m), U(m), or SU(m) with Young diagrams and . Problem: Reduction of the rpoduct representation D Dm u there is a simple graphical rule for this: 1) For 1 i # of rows of , write the number i in all boxes of row i of . 2) Consider all possibilities to add the boxes of to the Young diagram in the order "rst the 1s, then the 2s, etc." , according to the following rules: a) In every step the resulting Young diagram must be allowed and must not have more than m rows.

8.6. Reduction of the product of two irreps

99

b) The same number must not appear more than once in the same column. c) If the numbers are read in the order "rows from top to bottom, every row from right to left", then in this sequence of numbers there must never (meaning at every step of reading) be more is than (i 1)s. 3) If two Young diagrams created in this way have the same shape, they are only counted as dierent if the is are distributed dierently. 4) For SU(m), columns with m boxes can be deleted (except for the trivial irrep) always check the dimensions on both sides! example 1: SU(2) 54= j =2 j = = = =
1 1

3 = 2
1 1 1 1 1


1 1

1 1 1

=
1

1 1

1 1 1

1 1 1

1 1 1

7 2 j= 5 2

j= 3 2 j= 1 2

=8642= j =

example 2: SU(3)

33= or 33= 33=

= = =

=81
1

1 2

1 2

1 2

=81

=63
1

3 3 3 = (6 3) 3 = = 10 8 8 1 88= = = =
1 1 1 1 2

=
1 1

1 1

1 2

1 1 2

1 1 2

1 1 2

1 2

2 1

1 1 2

= 27 10 10 8 8 1

100

8. The tensor method to construct irreps of Gl(m) and its subgroups

8.7. Applications in (particle) physics


8.7.1. The Goldstone theorem
of a system has a global, continuous symmetry group G of suppose the Hamiltonian H order nG with generators Si , i.e., Si ] = 0 [H, for all i = 1, . . . , nG

normally the groundstate |0 of the system is then also symmetric under G, i.e., eii Si |0 = |0 Si |0 = 0 for all i

however, it could happen that the ground state does not have the full symmetry but only a smaller symmetry group H G of the order nH , i.e., Si |0 = 0 for i nH = 0 for i > nH

Terminology: The symmetry is broken spontaneously from G to H . |0 = 0. What is the energy of the state Si |0 ? Let H i |0 = (HS i Si H ) |0 = [H, Si ] |0 = 0 HS the state Si |0 has the same energy as the ground state there are nG nH such states. These states are called Nambu-Goldstone (NG) bosons because of 0 = E = p2 + m2 m for p 0, the NG bosons are massless

example: Ferromagnet in 3D is invariant under G = SO(3) H at low temperature (T < Tc ) the ground state exhibits a spontaneous magnetization: M =0 the ground state is only symmetric under H = SO(2), i.e., under rotations in the plane perpendicular to M the number of NG bosons is nG nH = 3 1 = 2, these are spin waves, i.e., slow variations in M (x). They propagate in the plane perpendicular to M 2 degrees of freedom another example: Chiral symmetry breaking in QCD (NG bosons = pions)

8.7.2. SU(2) isospin


experimental observation: among the observed hadrons there are small groups (multiplets ) with roughly the same mass (= eigenvalue of H ), e.g., proton p and neutron n: the 3 pions, 0 , and + : theoretical explanation: the strong interactions are independent of the electric charge mp mn 940 MeV m0 m+ m 140 MeV

8.7. Applications in (particle) physics

101

the small mass dierences within a multiplet are due to electroweak interactions as usual, degenerate states should transform in irreps of an internal symmetry group nd a group that explains the observed particle spectrum (i.e., degeneracies = dimensions of the irreps) an early attempt (before the discovery of quarks) was isospin symmetry: consider p and n, and dene an object with two components: N= p n

this object exists in a 2-dimensional space (isospin space ) consider SU(2) transformations in this space, with generators I1 , I2 , I3 by denition we take I3 |p =
1 2

|p and I3 |n = 1 2 |n

the strong interactions should be invariant under SU(2)isospin , i.e., the Hamiltonian of the strong interactions should commute with all 3 generators: [H, I ] = 0 N transforms in the 2-dimensional fundamental (or doublet ) representation (I = 1 2) of SU(2)isospin similarly other particles transform in other irreps of SU(2)isospin , e.g., the pions form an isospin triplet (I = 1) with + : I3 = 1 , 0 : I3 = 0 , : I3 = 1

the hypercharge Y is dened via 1 Q = I3 + Y 2 with Q = electric charge

e.g., p and n have Y = 1, the 3 pions have Y = 0 the isospin multiplets are distinguished by other quantum numbers of the strong interactions (B = baryon number, Y , I , J = spin, P = parity). For all particles within a multiplet the quantum numbers are equal, only I3 diers

8.7.3. SU(2) avor


we now know that hadrons are composed of quarks q with spin
3 baryons ( qqq ): spin = 1 2, 2, . . . 1 2

mesons ( qq ): spin = 0, 1, . . . baryon number B :


1 +3 for quarks, 1 3 for antiquarks

+1 for baryons, 1 for antibaryons, 0 for mesons and all other particles the interactions of quarks are described by QCD

102

8. The tensor method to construct irreps of Gl(m) and its subgroups

in nature there are Nf = 6 quark avors : u, d, s, c, b, t. Two of them are very light (u, d), one is light (s), and 3 are heavy (c, b, t) in low-energy experiments we only observe hadrons containing u and d for now, consider only Nf = 2, i.e., a 2-dimensional avor space the reason for the isospin invariance of the hadron masses is that for mu md the QCD Lagrangian is invariant under SU(2)f lavor , i.e., the internal symmetry group is SU(2)f lavor the 2-dimensional fundamental representation of SU(2)f lavor is furnished by q= u d up quark down quark
1 1 (Q = 2 3 , I3 = 2 ; Y = 3 ) 1 1 1 ) (Q = 3 , I3 = 2 , Y = 3

1 i.e., q transforms as a doublet under SU(2)f lavor (I = 2 ,Y = 1 3)

in the quark model the two nucleons have quark content p uud n udd 1 (Q = 1, I3 = ; Y = 1) 2 1 (Q = 0, I3 = ; Y = 1) 2

( means that we ignore permutations of the quarks for now), i.e., we have product states of the form 1 2 (here,
1 2

1 2

1 2
1 2

denotes the 2-dimensional fundamental represnetation with I =

and Y = 1 3)

since particles transform in irreps of the symmetry group, we have to decompose the product representation into irreducible components: 1 2 1 2 1 2 = (1) (0) 1 2 = 3 2 1 2 1 2

or expressed in dimensions: 2 2 2 = (3 1) 2 = 4 2 2 or expressed as Young diagrams: = = =

in Sec. 8.7.4 we will learn: p the doublet corresponds to a linear combination of the two 2-dimensional irreps n (I = 1 2 , Y = 1) in the RHS
3 the 4-dimensional irrep (I = 2 , Y = 1) corresponds to the -baryons

in the quark model, mesons consist of quarks and antiquarks . Antiquarks are obtained by applying the charge conjugation operator C = i2 K (where K = complex conjugation operator): Cu = u Cd = d

8.7. Applications in (particle) physics

103

consider a SU(2) transformation of the quark doublet: g= SU (2) with + = 1

u d u d

=g

u d

u d

= Cg

u d

= gC

u d

i.e.,

= g

u , thus the anti-doublet d

u d

transforms in 2

since SU(2) is pseudo real, 2 is equivalent to 2, i.e., we can also combine u and d in a doublet in such a way that it transforms in 2 to see this consider h = 0 1 1 0 u d SU(2) u d

g = h gh

= h gh

u d

= gh

u d

u d = transforms in 2 just like d u 1 I3 (d) = 1 2 , I3 (u) = 2 i.e., h

u , i.e., as an isospin doublet with d

for the mesons we start with product states of the form 2 2 2 2. The decomposition into irreducible components is: 2 2 =31 u d d u

from this we can read o the triplet and singlet states (see example at the end of Sec. 1 1 (|12 + |21 ), |22 } and singlet = { (|12 |21 )}) 8.3.2: triplet = {|11 , 2 2 the isospin triplet (I = 1, Y = 0) corresponds to the pions: I3 = 1 I3 = 0 I3 = 1 + = ud 1 0 = (uu + dd) 2 + = du
1 (uu 2

all these states are symmetric under u d, d u

the singlet state is

dd). In Sec. 8.7.4 we will see that this is the -meson

the introduction of the doublet

d was not absolutely necessary; we could also have u constructed the generalized projection operators for the decomposition 2 2 = 3 1 to obtain the projections on span(uu dd) and span( + , 0 , ). The choice of an eigenbasis of I3 in the 3-dimensional subspace then gives the quark content of the pions

104

8. The tensor method to construct irreps of Gl(m) and its subgroups

8.7.4. SU(3) avor and the quark model


at higher energies also the strange quark shows up consider Nf = 3, i.e., a 3-dimensional avor space with internal symmetry group SU(3)f lavor additional quantum number: strangeness S , with Y = B + S B u d s
1 3 1 3 1 3

I
1 2 1 2

I3
1 2 1 2

Y
1 3 1 3 2 3

S 0 0 1

Q
2 3 1 3 1 3

QCD processes leave S (and thus Y ) invariant LQCD is only invariant under SU(3)f lavor if mu = md = ms . Because mu md < ms , this symmetry is not exact but broken explicitly to SU(2)I U(1)Y no perfect degeneracy, but small mass dierences within an SU(3) multiplet (GellMannOkubo formula) the 3-dimensional fundamental representation of SU(3)f lavor is furnished by q = (u, d, s)T in the quark model mesons consist of a quark and an antiquark (which transform in 3), i.e., we start with product states of the form 3 3 and decompose them in irreducible components as in Sec. 8.6, yielding 33=81 or =

i.e., we should nd multiplets of roughly degenerate mesons containing 8 particles or 1 particle, respectively experimental observation: The lightest (groundstate) mesons indeed form an octet and a singlet (together also called nonet ), with quantum numbers B = 0 and J P C = 0+ Y pseudoscalar meson octet (scalar beacaue of J = 0, pseudo because of P = 1) pseudoscalar singlet: 1 with I = Y = 0
(su) (ds)

0 K

+ K (us)

I=1 2 , m = 496 MeV

(du)

-1

(uu, dd, ss)

-1 2

8 0
(uu, dd)

(ud)

1 2

I = 1, m = 137 MeV + I 1 3 I = 0, m see below

(sd) -1 K 0

I=1 2 , m = 496 MeV

in reality it is a bit more complicated: consider the three states with I3 = Y = 0: 0 is the state of the isospin triplet, i.e., 0 = 1 is the SU(3) singlet state, i.e., 1 =
1 (uu 3 1 (uu 2

dd)

+ dd + ss)

8 is the SU(3) octet, isospin singlet state. This state must be orthogonal to 1 0 and 1 , i.e., 8 = (uu + dd 2ss) 6

8.7. Applications in (particle) physics

105

1 and 8 have the same quantum numbers (I = 0 and J P C = 0+ ) if SU(3) were exact, 1 and 8 would be physical states (i.e., particles) since they transform in dierent irreps of SU(3) however, SU(3) is broken explicitly states transforming in dierent irreps, but with the same quantum numbers, can mix (548 MeV) = 8 cos 1 sin (958 MeV) = 8 sin + 1 cos the physical particles are and . is called nonet mixing angle (experimental value: = 22.6 ) in addition there are excited qq -states (rotations, vibrations, etc.). The rst excited meson nonet has quantum numbers B = 0 and J P C = 1 in the quark model baryons consist of 3 quarks, i.e., we start with product states of the form 3 3 3. Decomposition into irreducible components as in Sec. 8.6 yields or =

3 3 3 = 10 8 8 1
S MS MA A

totally symmetric tensors under S3 , i.e., permutations of quarks tensors of mixed symmetry: symmetric under exchange of the rst two quarks MA = tensors of mixed symmetry: antisymmetric under exchange of the rst two quarks A = totally antisymmetric tensors i.e., we should nd multiplets of (nearly) degenerate baryons consisting of 10, 8, and 1 particles, respectively with = = Experimental nding: The lightest (i.e., ground state) baryons form an octet and a decouplet: + 3+ baryon octet (B = 1, J P = 1 baryon decouplet (B = 1, J P = 2 ) 2 )
(udd)

S MS

(uud)

Y
I = 1 , m = 939 MeV 2 (ddd)

n
(dds)

(udd)

(uud)

+ ++

(uuu)

3 , m = 1232 MeV I = 2

(uds)

(uus)

+ I3

I = 0, m = 1116 MeV I = 1, m = 1193 MeV


(dds) (dss)

(uds)

uss) ( +

I = 1, m = 1385 MeV

(uus)

I3
1 , m = 1530 MeV I = 2

(dss)

(uss) +

I =

1, 2

(sss)
m = 1318 MeV

I = 0, m = 1672 MeV

Where are the singlet and the second octet? baryons are fermions, whose wave functions must be totally antisymmetric (in space, spin, avor, color)

106

8. The tensor method to construct irreps of Gl(m) and its subgroups

baryons are color singlets, i.e., they transform under SU(3)color in the irrep the spin-avor part must be totally symmetric for the spins of the 3 quarks in the baryon we have (here the Young diagrams are for SU(2)spin ) = = or


MA

222=4 2 2
S MS

i.e., we must combine 10 8 8 1


MS

MA

SU (3)f lavor

and

4 2 2

MS

MA

SU (2)spin

this gives the following possibilities for (SU(3) , SU(2)) multiplets S: (10, 4) MS : (10, 2MS ) MA : (10, 2MA ) A: (1, 4) + + + + (8, 2) (8, 4) + (8, 2) + (1, 2) (8, 4) + (8, 2) + (1, 2) (8, 2)

e.g., the totally symmetric octet corresponds to the linear combination 1 (8, 2)S = (8MS , 2MS ) + (8MA , 2MA ) 2 and similarly for the other combinations only the totally symmetric spin-avor multiplets (10,4) and (8,2) give a totally antisymmetric wave function of the baryon in the ground state there is only the decouplet and one octet, but no singlet and no second octet; however, the latter two exist in excited states another method to derive this: consider the 3 quarks as 6 states with 3 avors and 2 spin states per avor approximate SU(6)spinf lavor symmetry the decomposition into irreducible components under SU(6) is 6 6 6 = 56S 70MS 70MA 20A the 56-dimensional irrep of SU(6) subduces a representation of SU(3)f lavor ; this representation is reducible, and we obtain: 56S = 103/2 81/2
dim.=104 dim.=82 3 2)

this corresponds to the baryon decouplet (spin = =1 2 ), respectively

and the baryon octet (spin

the SU(3)f lavor symmetry is seen not only in the mass spectrum, but also in scattering experiments:

8.7. Applications in (particle) physics

107

e.g., consider scattering of a meson (P ) from the pseudoscalar meson octet and a baryon (B ) from the baryon octet the scattering amplitudes for, e.g., p or Kp or scattering are all related due to SU(3) symmetry using the Wigner-Eckart theorem, all scattering amplitudes for P B scattering can be expressed in 6 complex matrix elements and the CG coecients in the decomposition (see Sec. 8.6) 8 8 = 27 10 10 8 8 1 or =

more examples: see Halzen & Martin, Quarks and Leptons, Sec. 2

Appendix A
A.1 Proof of Cayleys theorem
we consider a group G = {g1 , . . . , gn } of order n. if we look at the multiplication table of the group, we notice that the i-th row is given by (gi g1 , . . . , gi gn ) = p(g1 , . . . , gn ) where p Sn , i.e., every row is just a permutation of the group elements. from this it is obvious that there must be an isomorphism between G and a subgroup of Sn (i.e., the subgroup includes the permutations that appear in the multiplication table).

A.2 Proofs of the four theorems


Theorem 1: Every representation of a group G (of nite order) can be converted to a unitary representation. suppose that the action of g G on a vector space V is represented by D(g ). Then we need to show that S such that SD(g )S 1 = U (g ) dene a matrix M by M=
g

gG

U (g ) : unitary

D(g ) D(g )

this matrix is positive denite, and can thus be written as M=


g

D(g ) D(g ) S S

now we look at the scalar product U (g )x U (g )y = SD(g )S 1 x SD(g )S 1 y = S 1 x D(g ) S SD(g )S 1 y now we use that D(g ) S SD(g ) = D(g )
h

D(h) D(h)D(g ) =
h

D(hg ) D(hg ) = S S

so we have U (g )x U (g )y = S 1 x S SS 1 y = x | y and thus U (g ) is unitary

108

A.2 Proofs of the four theorems

109

Theorem 2 (Schurs lemma 1): A matrix A which commutes with all matrices of an irrep U (G) is proportional to the identity matrix. without loss of generality we can take the matrices U (g ) of the irrep to be unitary, and the matrix A to be Hermitian. Otherwise we can decompose A = (Hermitian + antiHermitian). Then we can choose the basis of the carrier space such that A is diagonal: A |ni = i |ni then, with [A, U (g )] = 0, we have AU (g ) |ni = U (g )A |ni = i U (g ) |ni U (g ) |ni is also an Eigenvector of A with Eigenvalue i therefore the vectors |ni form an invariant subspace (for xed i); but this can only be true if this subspace is already the whole carrier space itself, since U (G) is an irrep. Hence the additional label i is infact superuous, and we can set i = A = 1 Theorem 3 (Schurs lemma 2): Suppose we have two irreps of a group G: 1 (G) with dimension 1 and 2 (G) with dimension 2 . If there exists a 2 1 matrix M such that M 1 (g ) = 2 (g )M gG () species the degeneracy

then if 1 = 2 we have M = 0. If 1 = 2 , then either M = 0 or det M = 0. In the case M = 0 the two representations are equivalent. without loss of generality we can assume 1 2 also we can take the two irreps to be unitary. Then, taking the hermitian conjugate of (), we have 1 (g ) M = M 2 (g ) 1 (g 1 )M = M 2 (g 1 )
()

M 1 (g 1 )M = M M 2 (g 1 ) 2 (g 1 )M M = M M 2 (g 1 )

M M is a square matrix that commutes with all 2 (g ). Therefore by Schurs lemma 1 we have M M = c1 , det M M = c2

if 2 = 1 = , then M is a square matrix with det M = c/2 if c = 0, M is non-singular and 1 and 2 are equivalent if c = 0, we have M M = 0. This implies M = 0, since 0 = MM
kk

=
i

Mki (M )ik =
i

Mki Mki = i

|Mki |2

Mki = 0

if 1 < 2 M is rectangular. Dene a square matrix N = (M 0), i.e., the missing space is lled up with zeros. Then, NN = MM , det N N = |det N |2 = 0 = det M M since M M = c1 it follows that M M = 0 and therefore M = 0 (see above)

110

Appendix A Appendix

Theorem 4 (orthogonality relations): Let i (G) and k (G) be two non-equivalent, unitary irreps of G with order(G) = n. Then we have
n

i (Aj )
j =1

k (Aj ) =

n ik i

rst we dene a matrix M by M=


j 1 k i (A j )X (Aj )

()

where X is an arbitrary i k matrix. multiplying this by k (A ) from the right gives M k (A ) =


j 1 k k i (A j )X (Aj ) (A ) =

= i (A )
j

i (Aj A )1 X k (Aj A ) =

= (A )M where we have used that Aj A also runs over all group elements. now we look at the case i = k . Then i and k are non-equivalent and it follows that M = 0 (froms Schurs lemma 2). writing () in component form yields M =
j 1 k i (A j ) X (Aj )

until now the matrix X was arbitrary; we now choose all the X = 0 except for one particular element X which we set to 1 then, 0=
j 1 k i (A j ) (Aj ) = j k i (Aj ) (Aj ) = j

i (Aj )

k (Aj )

for the case i = k we have M i (Aj ) = i (Aj )M i.e., M commutes with all i (Aj ), and therefore M = c1 again, writing () in component form gives c =
j 1 i i (A j ) X (Aj )

here we can also choose all X = 0 except for one X = 1, so that (c = c ) c =


j 1 i i (A j ) (Aj )

A.3 The Kronecker product

111

to evaluate c we set = and sum over :


i n i 1 i i (A j ) (Aj ) = j =1 =1 j =1 n 1 i (Aj )i (A j ) n

c
=1

=
i

=
j =1 n

(1)

n = i

1 i inserting c into the equation above and using the fact, that i (A j ) = (Aj ) = (i (Aj ) ) , gives the equation from the theorem

A.3 The Kronecker product


the Kronecker product of a n m matrix A and a k following way: A11 B . .. AB = . . . An1 B
A11 B11 . . .

matrix B is dened in the

Anm B

A1m B . . . = A11 B1 . . . A11 Bk . . . An1 B1 . . . An1 Bk .. . .. . .. . A1m B11 . . . A1m Bk1 . . . .. . .. . A1m B1 . . .

A B 11 k1 . . = . An1 B11 . . .

.. . .. . .. .

An1 Bk1

Anm B11 . .. . . . Anm Bk1

Anm Bk

Anm B1 . . .

A1m Bk . . .

with this in mind one can see that the formula, (A B )(i1)k+s,(j 1)
+t

= Aij Bst ,

given in Sec. 2.6 is indeed correct, since to get a matrix element with Aij one has rst to jump down (i 1) B -blocks of (row) size k and jump over (j 1) B -blocks of (column) size to the right. From there one can move to Bst in the usual manner.

You might also like