Sei sulla pagina 1di 300

SYMMETRY IN PHYSICS

VOLUME 2: FURTHER APPLICATIONS


SYMMETRY IN PHYSICS
VOLUME 2:
FURTHER APPLICATIONS

J. P. ELLIOTT and P. G. DAWBER


School of Mathematical and Physical Sciences
Un iversity of Sussex, Brighton

M
MACMILLAN
© J. P. Elliott and P. G . Dawber 1979

All rights reserved . No part of thi s publication may be reproduced or


transmitted, in any form or by any means, without permission

First edition 1979


Reprinted (with correction s) 1984

Publi shed by
Scientific a nd Medi cal Di vision
MACMILLAN PUBLISHERS LTD
London and Basingstoke
Companies and repre sent atives throughout the world

British Library Cataloguing in Publi cat ion Data

Elliott, J P
Symmetry in physics.
Vol. 2: Further applications
I. Symmetry (Physics)
I. Title II. Dawber, P G
530 QC793.3.S9

ISBN 978-0-333-38271-4 ISBN 978-1-349-07637-6 (eBook)


DOI 10.1007/978-1-349-07637-6
ISBN 978-0-333-11820-7 Set of2 vols he
ISBN 978-0-333-38272-1 Set of2 vols pb
Contents of Volume 1

Prefa ce xvii

Introduction 1
1.1 The place of symmetry in physics 1
1.2 Examples of the consequences of symmetry 3
1.2.1 One particle in one dimension (classical) 3
1.2.2 One particle in two dimensions (classical) 3
1.2.3 Two particles connected by springs (classical) 4
1.2.4 One particle in three dimensions using quantum
mechanics-spherical symmetry and degeneracies 5
1.2.5 One particle in one dimension using quantum
mechanics -parity and selection rules 6
1.2.6 The search for symmetry---elementary particle
p~~a 7
1.3 Summary 8
2 Groups and Group Properties 9
2.1 Definition of a group 9
2.2 Examples of groups 11
2.3 Isomorphism 16
2.4 Subgroups 17
2.5 The direct product group 17
v
vi Contents

2.6 Conjugate elements and classes 18


2.7 Examples of classes 19
2.7.1 The rotation group 91 3 19
2.7.2 The finite group of rotations D 3 20
2.7.3 The symmetric group g 3 21
2.8 The class structure of product groups 21
2.9 The group rearrangement theorem 22
Bibliography 22
Problems 22
3 Linear Algebra and Vector Spaces 24
3.1 Linear vector space 25
3.2 Examples of linear vector spaces 27
3.2.1 Displacements in three dimensions 27
3.2.2 Displacement of a set of N particles in three
dimensions 28
3.2.3 Function spaces 28
3.2.4 Function space with finite dimension 29
3.2.5 Wave functions 29
3.3 Linear operators 30
3.4 The multiplication, inverse and transformation of operators 32
3.5 The adjoint of an operator-unitary and Hermitian
operators 34
3.6 The eigenvalue problem 35
3.7 Induced transformation of functions 36
3.8 Examples of linear operators 38
3.8.1 Rotation of vectors in the xy-plane 38
3.8.2 Permutations 39
3.8.3 Multiplication by a function in function space 39
3.8.4 Differentiation in function space 40
3.8.5 Induced transformation of functions 40
3.8.6 Further example of induced transformation of
functions 41
3.8.7 Transformed operator 41
Bibliography 42
Problems 42
4 Group Representations 43
4.1 Definition of a group representation 43
4.2 Matrix representations 44
4.3 Examples of representations 45
4.3.1 The group D 3 45
4.3.2 The group ~2 46
4.3.3 Function spaces 47
4.4 The generation of an invariant subspace 48
4.5 Irreducibility 50
4.6 Equivalent representations 52
Contents vii

4.6.1 Proof of Maschke's theorem 53


4.7 Inequivalent irreducible representations 54
4.8 Orthogonality properties of irreducible representations 54
4.8.1 Proof of Schur's first lemma 58
4.8.2 Proof of Schur's second lemma 60
4.9 Characters of representations 60
4.10 Orthogonality relation for characters of irreducible
representations 61
4.11 Use of group characters in reducing a representation 62
4.12 A criterion for irreducibility 63
4.13 How many inequivalent irreducible representations?-The
regular representation 64
4.14 The second orthogonality relation for group characters 66
4.15 Construction of the character table 67
4.16 Orthogonality of basis functions for irreducible
representations 68
4.17 The direct product of two representations 70
4.18 Reduction ofan irreducible representation on restriction to a
subgroup 73
4.19 Projection operators 74
4.20 Irreducible sets of operators and the Wigner - Eckart
theorem 78
4.21 Representations of direct product groups 81
Bibliography 83
Problems 83
5 Symmetry in Quantum Mechanics 85
5.1 Brief review of the framework of quantum mechanics 85
5.2 Definition of symmetry in a quantum system 89
5.3 Degeneracy and the labelling of energies and eigenfunctions 90
5.4 Selection rules and matrix elements of operators 91
5.5 Conservation laws 92
5.6 Examples 93
5.6.1 Symmetry group C 3 93
5.6.2 Symmetry group D3 95
5.6.3 Symmetry group S 2 96
5.6.4 Symmetry group :Jt 2 96
5.7 Use of group theory in a variational approximation 97
5.8 Symmetry-breaking perturbations 99
5.8.1 Examples 100
5.8.2 Magnitude of the splitting 101
5.9 The indistinguishability of particles 102
5.10 Complex conjugation and time-reversal 103
Bibliography 104
Problems 104
6 Molecular Vibrations 106
viii Contents

6.1 The harmonic approximation 107


6.2 Classical solution 108
6.3 Quantum mechanical solution 109
6.4 Effects of symmetry in molecular vibrations 110
6.5 Classification of the normal modes 113
6.5.1 The water molecule 115
6.5.2 The ammonia molecule 116
6.6 Vibrational energy levels and wave functions 117
6.7 Infrared and Raman absorption spectra of molecules 120
6.7.1 Infrared spectra 120
6.7.2 Raman spectra 121
6.8 Displacement patterns and frequencies of the normal modes 122
Bibliography 124
Problems 124
7 Continuous Groups and their Representations, Including Details
of the Rotation Groups [Jl2 and [Jl3 125
7.1 General remarks 126
7.2 Infinitesimal operators 127
7.3 The group [Jl2 130
7.3.1 Irreducible representations 131
7.3.2 Character 131
7.3.3 Multiplication of representations 132
7.3.4 Examples of basis vectors 132
7.3.5 Infinitesimal operators 133
7.4 The group [Jl 3 134
7.4.1 Infinitesimal operators 135
7.4.2 Irreducible representations 137
7.4.3 Characters 140
7.4.4 Multiplication of representations 141
7.4.5 Examples of basis vectors 143
7.4.6 Irreducible sets of operators and the Wigner-Eckart
theorem 146
7.4.7 Equivalent operators 147
7.5 The Casimir operator 148
7.6 Double-valued representations 150
7.7 The complex conjugate representation 153
Bibliography 153
Problems 154
8 Angular Momentum and the Group fJt 3 with Illustrations from
Atomic Structure 156
8.1 Rotational invariance and its consequences 156
8.2 Orbital angular momentum of a system of particles 158
8.3 Coupling of angular momenta 159
8.4 Intrinsic spin 161
8.5 The hydrogen atom 166
Co ntents ix

8.6 The structure of man y-electron atoms 170


8.6.1 The Ham iltonian 170
8.6.2 The Paul i principle and shell filling 171
8.6.3 Atom s with more than one valence electron- LS
coupling 173
8.6.4 Classificatio n of term s 176
8.6.5 Ordering of terms 179
Bibliography 181
Problems 181
9 Po int Groups with an Application to Cr ysta l Fields 183
9.1 Point -group operations and notation 184
9.2 The stereogram 184
9.3 Enumeration of the point groups 186
9.3.1 Proper groups 186
9.3.2 Improper groups 191
9.4 The class structure of the point groups 192
9.4.1 Proper point groups 193
9.4.2 Improper point groups 193
9.5 The crystallo graphic point groups 196
9.6 Irreducible representations for the point groups 197
9.7 Double-valued repre sentat ion s of the point groups 199
9.8 Time-reversal and magnetic point groups 201
9.9 Crystal field splitting of at omic energy levels 202
9.9.1 Definition of the physical problem 202
9.9.2 Deduction of the manner of splitting from symmetry
con siderations 204
9.9.3 Effect of a magnetic field 209
Bibliography 210
Problems 211
10 Isospin and the G roup S U 2 213
10.1 Isospin in nuclei 214
10.1.1 Isospin labelling and degeneracies 215
10.1.2 Splitting of an isospin multiplet 218
10.1.3 Selection rules 221
10.2 Isospin in elementary particles 222
10.2.1 Collisions of n-mesons with nucleons 223
10.3 Isospin symmetry and charg e-independence 223
Bibliograph y 224
Problems 224
11 The Group SU 3 with Applicat ions to Element ar y Particle s 226
11.1 Compilat ion o f so me relevant da ta 227
11.2 The hypercharge 230
11.3 Baryon number 231
11.4 The group SU 3 232
11.5 Subgroups of S U 3 233
x Contents

11.6 Irreducible representations of SU 3 233


11.6.1 Complex conjugate representations 241
11.6.2 Multiplication of representations 242
11.7 Classification of the hadrons into SU 3 multiplets 243
11.8 The mass-splitting formula 244
11.9 Electromagnetic effects 247
11.10 Casimir operators 248
Bibliography 249
Problems 249
12 Supermultiplets in Nuclei and Elementary Particles-the Groups
SU 4 and SU 6 and Quark Models 251
12.1 Supermultiplets in nuclei 252
12.2 Supermultiplets of elementary particles 255
12.3 The three-quark model 257
12.4 The nine-quark model 260
12.5 Charm 262
Addendum (mid-1978) 262
Addendum (late 1983) 263
Bibliography 264
Problems 264

Appendix I Character Tables for the Irreducible Representations of


the Point Groups 265
Appendix 2 Solutions to Problems in Volume 1 275

Index to Volumes I and 2 (adjacent to p. 280)


Contents of Volume 2

Preface xvii

13 Electron States in Molecules 281


13.1 Linear combinations of atomic orbitals (LCAO) 282
13.2 Examples 284
13.3 Selection rules for electronic excitations in molecules 287
Bibliography 288
Problems 288
14 Symmetry in Crystalline Solids 289
14.1 Translational symmetry in crystals 289
14.2 The translation group 5"(111,112,113) 290
14.3 The Brillouin zone and some examples 293
14.4 Electron states in a periodic potential 294
14.4.1 The nearly-free electron model 295
14.4.2 Metals and insulators 299
14.4.3 The tight-binding method 302
14.5 Lattice vibrations 306
14.5.1 The one-dimensional monatomic lattice 306
14.5.2 Three-dimensional crystals with several atoms per
unit cell 309
14.6 Spin waves in ferro magnets 311
xi
xii Contents

14.7 Excitons in insulators (Frenkel excitons) 313


14.8 Selection rules for scattering 314
14.9 Space groups 315
14.9.1 Irreducible representations of space groups 316
14.9.2 Application to electron states 320
14.9.3 Other excitations 323
Bibliography 323
Problems 324
15 Space and Time 325
15.1 The Euclidean group g 3 326
15.1.1 Translations 326
15.1.2 The group operators 328
15.1.3 The irreducible representations 328
15.1.4 The groupgz 331
15.1 .5 The physical significance of the Euclidean group g 3 331
15.1.6 Scalar products and normalisation of basis vectors 333
15.2 The Lorentz group !i' 334
15.2.1 The Lorentz transformation 335
15.2.2 The regions of space -time 339
15.2.3 Physical interpretation of the Lorentz
transformation 340
15.2.4 Infinitesimal operators 343
15.2.5. The irreducible representations 344
15.3 The Lorentz group with space inversions !i's 347
15.4 Translations and the Poincare group flJ 349
15.4.1 Translations in space-time 349
15.4.2 The Poincare group and its representations 351
15.4.3 Casimir operators 356
15.4.4 Definition of scalar product 359
15.5 The Poincare group with space inversions fYJ s 360
15.6 The Poincare group with time inversion fill 362
15.7 Physical interpretation of the irreducible representations of
the Poincare group 363
15.7.1 Mass 364
15.7.2 Spin 366
15.7.3 Parity 368
15.7.4 Time-reversal 369
15.7.5 Some consequences of time-reversal symmetry 373
15.8 Single-particle wave functions and the wave equations 375
15.8.1 The group 91 3 376
15.8.2 The group g 3 377
15.8.3 The Poincare group with s = O-the
Klein-Gordon equation 379
15.8.4 The Poincare group with s = i -the Dirac
equation 380
Contents xiii

15.8.5 Particles with zero mass and spin Iml = t-the


Weyl equation 387
15.8.6 Particles with zero mass and spin Iml = I-the
Maxwell equations 389
Bibliography 390
Problems 391
16 Particles, Fields and Antiparticles 393
16.1 Classical mechanics of particles 394
16.1.1 Lagrange formalism 394
16.1.2 Hamiltonian formalism 394
16.1.3 Examples from relativistic mechanics 396
16.2 Classical mechanics of fields 398
16.2.1 The transformation of fields 398
16.2.2 The Lagrange equation for fields 399
16.2.3 The electromagnetic field 400
16.3 Quantum fields 401
16.3.1 Second quantisation 402
16.3.2 Field operators 404
16.3.3 The physical role of field operators 405
16.3.4 Causality and the spin-statistics theorem 408
16.3.5 Antiparticles 409
16.3.6 Charge conjugation and the PCT theorem 411
16.3.7 Field for particles with non-zero spin 413
Bibliography 423
Problems 423
17 The Symmetric Group Y' n 425
17.1 Cycles 426
17.2 The parity of a permutation 427
17.3 Classes 428
17.4 The identity and alternating representations-symmetric
and antisymmetric functions 430
17.5 The character table for irreducible representations 431
17.6 Young diagrams 434
17.7 The restriction from Y' n to Y' n - 1 434
17.8 The basis vectors of the irreducible representations 436
17.9 Examples of basis vectors and representation matrices 438
17.10 The direct product of two representations 439
17.11 The outer product of two irreducible representations 441
17.12 Restriction to a subgroup and the outer product 443
17.13 The standard matrices of the irreducible representations 445
17.14 The class operator I T(Pij ) 450
i <j
Bibliography 450
Problems 451
18 The Unitary Group UN 452
xiv Contents

18.1 The irreducible representations of UN 453


18.2 Some examples 456
18.3 The chain of subgroups
UN-+UN-l-+UN-2-+'" -+U 2-+U t 457
18.4 A labelling system for the basis vectors 459
18.5 The direct product of representations of UN 461
18.6 The restriction from UNtO its subgroup SUN 462
18.7 The special cases of SU 2 , SU 3 and SU 4 464
18.8 The infinitesimal operators of UN 466
18.9 The complex conjugate representations of UN and SUN 467
18.10 The use of the group UN in classifying many-particle wave
functions 469
18.10.1 The use of subgroups of UN 471
18.11 Characters 475
18.12 Group integration and orthogonality 476
18.13 The groups SU 2 and e, 478
18.13.1 The parameters of SU 2 478
18.13.2 Infinitesimal operators and irreducible
representations of SU 2 480
18.13.3 Connection between the groups fit 3 and SU 2 480
18.13.4 Explicit formula for the parameters of a product
of rotations 482
18.13.5 Examples of SU 2 basis vectors 482
Bibliography 483
Problems 483
19 Two Familiar 'Accidental' Degeneracies -the Oscillator and
Coulomb Potentials 485
19.1 The three-dimensional harmonic oscillator for one particle 486
19.2 The three-dimensional harmonic oscillator for many
particles 491
19.3 The harmonic oscillator in n dimensions 492
19.4 The symmetry group of the Coulomb potential 492
19.4.1 The groups e, and !e 494
19.4.2 The classification of states of the Coulomb
potential 495
Bibliography 496
Problems 497
20 A Miscellany 498
20.1 Non-invariance groups 498
20.2 The Jahn- Teller effect and spontaneously broken
symmetries 502
20.2.1 The adiabatic approximation 502
20.2.2 The role of symmetry 503
20.2.3 Spontaneous symmetry breaking 505
20.3 Normal subgroups, semi-direct products and little groups 507
Contents xv

2004 The classification of Lie groups 510


2005 The rotation matrices 519
Bibliography 522
Problems 523

Appendix 3 Topics in Representation Theory 524


A.3.1 Symmetrised products of representations 524
A.3.2 Use of a subgroup in reducing product
representations 527
A0303 Class multiplication 529
Appendix 4 Some Results Pertaining to the Group f.Jl 3 531
AA.1 An integral over three spherical harmonics 531
AA.2 The spherical harmonic addition theorem 532
AA.3 Group integration 533
Appendix 5 Techniques in Atomic Structure Calculations 539
A.5.1 Term energies for p2 and p 3 configurations 539
A.502 Recoupling coefficients (6j- and 9j- symbols) 543
A.503 Transition strengths 547
A.5.4 The crystal field potential 549
A.5.5 Use of symmetry to deduce ratios of splittings 550
Problems on appendices 4 and 5 553
Appendix 6 Solutions to Problems in Volume 2 555

Index to Volumes I and 2 ( adjacent to p. 558)


Preface to Volume 2

In volume 1 the fundamental role of symmetry was developed and applied to a


wide range of problems in classical and quantum physics. In th is volume we
mo ve on to some more difficult applicat ions and to a more general study of the
symmetric and unitary groups, particular cases of which were met in volume 1.
These topics would probably only arise at postgraduate level, but would none
the less be of interest to the more inquisitive undergraduate.
This volume begins with a further application of the use of ' point groups'-
to the motion of electrons in a molecule-and then, in chapter 14, moves away
from symmetries with a fixed point to study discrete translations and the ir
applications to crystal structure. The theory of relativity is of profound
importance in the philosophy of physics and, when speeds become comparable
with that of light, it has practical importance. For all the systems discussed in
volume 1 we were able to ignore relativity because the speeds of the particles
involved were sufficiently small. Chapter 15 describes the symmetry in four-
d imensional space-time which is the origin of relativity theory and discusses its
consequences, especially in relation to the classification of elementary
particles. The concepts of momentum, energy, mass and spin are interpreted in
terms of symmetry using the Lorentz and Poincare groups and a natural place
is found in the theory for particles, like the photon, with zero mass. Chapter 16
is concerned with fields, in contrast to the earlier chapters which dealt with
particles or systems of particles. We first describe classical fields, such as the
xvii
xviii Preface

electromagnetic field, using four-dimensional space-time. This is followed by a


brief account of the theory of relativistic quantum fields which provides a
framework for the creation and annihilation of particles and the existence of
antiparticles. Chapters 17 and 18 contain details of two general groups, the
'symmetric' group of all permutations of n objects and the 'unitary' group in N
dimensions, and an intimate relation between these two groups is discussed .
Particular cases of these two groups have been met earlier. Chapter 19
describes some unexpected symmetries in two familiar potentials, the Coul-
omb and the harmonic oscillator potentials, and a number of small,
unconnected, but interesting topics are collected into the last chapter.
The text includes worked examples and a selection of problems with
solutions. A bibliography of references for further reading is given at the end of
each chapter for those who wish either to follow the physical applications into
more detail or to study some of the mathematical questions to a greater depth.
As in volume 1,roman type is used for operators and italic type for numbers.
Vectors are printed in bold face and four-vectors carry a circumflex.

Brighton. Sussex. 1979 J. P. E.


P.G. D.
13
Electron States
in Molecules

In studying the structure of molecules one should strictly consider the motion
both of the nuclei of the constituent atoms and of their electrons. However,
because of the large mass of the nuclei, relative to that of an electron, one may
approximately separate the motions of these two kinds of particle . This
'Born - Oppenheimer' approximation is discussed briefly in section 20.2. For
our present purpose we shall be considering the motion of the electrons
relative to the nuclei which are assumed to be fixed. Our interest will centre on
those molecules for which the arrangement of fixed nuclei is invariant under
some point group of symmetry operations. This is precisely the same
symmetry group which we used in discussing molecular vibrations in chapter
6. The physical difference between that problem and the present one is that in
studying molecular vibrations the electrons were ignored, except in so far as
they were responsible for the potential in which the nuclei moved. From the
experimental viewpoint the molecular vibrations involve energies of the order
of (10)-2 eV, whereas the excitation of the electronic motion, which we are
about to discuss, requires energies of the order of 1 eV.
Even with the assumption of fixed nuclei the calculation of the electronic
wave function is a formidable task -see Eyring et al. (t944), Murrell et al.
281
282 Electron States in Molecules 13 .1

(1970) and Murrell and Harget (1972) in the bibliography-more complex


than the calculation of the electronic wave functions of a single atom which we
described in chapter 8. We therefore restrict the present discussion to the
simple model in which each electron is assumed to move independently in a
fixed field due to the nuclei and the other electrons. This is similar to the central
field approximation in atoms except that the field now has some point group
symmetry instead of being spherical. Having found the single-electron wave
functions we can then construct a Slater determinant for the many-electron
wave function as described in subsection 8.6.2. One usually refers to a single-
particle wave function in the molecule as a 'molecular orbital' to distinguish it
from an atomic orbital which refers to a single atom.

13.1 Linear combinations of atomic orbitals (LeAO)

If we consider an electron moving close to one of the nuclei in the molecule it


will experience a field which must be quite similar to that in an isolated atom,
so that in this region we would expect the wave function to be similar to a free-
atom wave function 4>nlm(r-r,) centred on the nucleus at r,. This leads us to
construct wave functions by taking linear combinations of the least-bound
atomic orbitals 4>nlm (r - r,) centred on the different nuclei r.. Such a wave
function is called an LeAO molecular orbital. In a simple calculation it is
reasonable to assume that only valence electrons are involved in these orbitals,
whilst the electrons in the inner full shells of the atoms remain in their
unperturbed atomic state. A more complete calculation could include some of
these lower energy states and also some of the unoccupied excited states.
The set of atomic orbitals 4>nlm (r - r,), with n and I fixed and with r, running
over the positions of equivalent nuclei, provides the basis for a representation
T of the symmetry group. To see this, we show that the effect of a group
operation T(G a) on the orbital 4>nlm(r-r,) is to turn it into another atomic
orbital centred on one of the nuclei r.. In detail, using the general definition
(3.37) we have
T(G a)4>nlm(r-r,l = 4>nlm(G a- 1r-r,l
= 4>nlm{G a- 1 ( r - Gar,l}
=L D~!m(Gal4>nlm,(r-Gar,l
m'

= L D~!m(Gal4>nlm·(r-r"l (13.1)
m'

where r, = Gar, and OIl} is the familiar (21 + lj-dimensional irreducible


representation of &3' The representation T will have dimension (21 + I)N"
where N, is the number of nuclei equivalent to the one at r..
From general theory we expect an energy eigenstate to transform ir-
reducibly under the symmetry group '§, so that we shall be able to label the
13.1 Electron States in Molecules 283

molecular orbitals by irreducible representation labels of '§. To find which


labels occur we must reduce the representation T and for this we need to know
the character of T which may be deduced from equation (13.1) using an
argument like that in section 6.5. From this equation one sees that diagonal
matrix elements occur only when r, = r", i.e when the nucleus at r, is unmoved
by Ga. The contribution to the character of T from such a nucleus is then
simply the character of D(l) which is given from equation (7.42) as

X(I)(R (¢ ) ) = sin (l + t)¢ /sin t¢


X(/)(S(¢)) = cos (l+t)¢/cost¢ (13.2)

where R(¢) is a proper rotation and S( ¢) = O'hR(¢) = IR(¢ + n) is a mirror


rotation, in the notation of section 9.1. The characters of the (2/ + 1)N,-
dimensional representation T are then obtained by multiplying these ex-
pressions by the number of nuclei left unmoved by the group operation R(</J)
Dr S( </J). Once we have the character of T, its reduction is found by using the
known table of irreducible characters using the general method of section 4.11.
To construct a molecular orbitall/l~~) (r) which transforms according to one
of the representations T(2) of '§ occurring in the above reduction we may use
the projection method explained in section 4.19. From equation (4.51) we can
co nstruct the un-normalised molecular orbital

(13.3)
a

where the sum runs over the group elements G, and any of the possible values
m and r, may be used.
If a representation T(') occurs more than once , or if we wish to consider the
mixing of different atomic orbital labels nl, then we must form a trial wave
function
1/1(2) (r) = I C ~~ ) i I/I~~) i (r) (13.4)
i.n,l

and determine the coefficients C~~ ) i by the variational method, In this sum, the
index i distinguishes molecular orbitals when T(2) occurs more than once in the
reduction ofT. To perform the variational calculation, or even to estimate the
relative energies of the single molecular orbitals (13,3) for different IX, one must
assume some form for the single electron Hamiltonian H(r). Since the
functions I/I~~) i (r) are not orthogonal the matrix formulation of this variational
problem follows the method of section 5,7. The wave functions and energies
for the molecular orbitals with symmetry IX are then given by the roots of the
matrix equation (5.23); in our present notation,
(H(') - ES('))c(') =0 (13.5)
where C (2) is the column vector of coefficients c ~~ )i in equation (13.4). The
284 Electron States in Molecules 13 ,2

matrix elements of and Sla) are given by

ft/J~i)i·
H(a)

Hl~l. j n' " = (r)H t/J~W (r)dr


s!a) . =
mi. jn 'L'
f'I'(')i·(r)'I,(~),i(r)dr
¥'
nl V' nI

Since t/J'.:)i (r) is a known linear combination of atomic orbitals these matrix
elements are all given in terms of matrix elements between atomic orbitals. It is
usually assumed that the matrix elements of H(r) between atomic orbitals
vanish unless the two sites t and t' are identical or nearest neighbours in the
molecule. Further assumptions often made are that the diagonal elements are
just the single-particle energies in the free atom since the atomic orbitals are
well located about a single nucleus and H(r) must approximate the sum of the
atomic potentials. Care must be taken in these energy calculations since the
atomic orbitals on different sites are not orthogonal. In the simplest
approximation, one might assume them to be orthogonal, i.e. S(') is the unit
matrix.

13.2 Examples

To illustrate the techniques developed in the preceding section we will look


at two examples. The first will be a hypothetical H 3 molecule with the three
hydrogen atoms at the corners of an equilateral triangle and the second will be
the water molecule used as an illustration in chapter 6.

H 3 molecule
The H 3 molecule is illustrated in figure 13.1 with the three atoms being
labelled by a, band c. The symmetry group is D 3 h , but since we shall only be
using atomic s-states in our calculation and these have reflection symmetry in
the plane we may ignore the horizontal reflection and use just the group D 3 •

a b
H H

Figure 13.1
13.2 Electron States in Molecules 285

The characters are given in table 13.1 together with the character for the three
Is-orbitals calculated as outlined above.

Table /3 ./

DJ E 2C J 3C z

AI 1 1 1
Az 1 1 -1
E 2 -1 0

X(ls) 3 0 = Al +E

This reducible representation consists of a singlet A I and a doublet E. Thus


instead of the three wave functions 4>15 (a), 4> Is(b) and 4>ls (c) we can construct
one linear combination e (A 1) which transforms like A I and a pair '" ~E) and '" ~E),
which transform according to E. Using the projection (13.3), we find
",(Atl oc {4>l5(a)+4>ls(b)+4>l5(C) }
",~E) oc {24>I.(a) - 4>1.(b) - 4>ls(c) } (13.6)

",~E) oc {4>I.(b) - 4>l5(C)}


In this simple example the irreducible representations occur only once so that
no variational calculation is necessary and the states (13.6) are our best one-
electron molecular orbitals. The energies of these states in the simplest
approximation are then given by
E(A ,) = 1- (3e ls + 6A.) = els + 2A.
E\El = E~E) = !-(2CIs - 2A.) = Cis - A.
where Cis is the energy of the free hydrogen Is-orbital and A. is the integral

A. = f4>~s(a)H(r)4>ls(b)dr
The main contribution to this integral is the energy due to the interaction
between the electron in the state 4>15 (a) and the nucleus at b and vice versa.
Since this is due to an attractive force we would expect A. to be negative and
hence the totally symmetric state ",(A,) to be lowest in energy. Hence, if we put
the three electrons into the states (13.6) in a way consistent with the Pauli
exclusion principle, the ground state of the molecule will have two electrons
with opposite spins in the lowest orbital ",(Ad and one electron in one of the
degenerate states ",(E). The total wave function for this state is the Slater
determinant constructed from these three functions and since each term in the
expansion of the determinant is a product with symmetry A I X A I X E = E it
is an E-orbital doublet, the degeneracy arising from the two possible states for
the third electron .
286 Electron States in Molecules 13.2

H 20 molecule
The symmetry group for the water molecule shown in figure 13.2 is C 2 v ' The
hydrogen atoms have one electron each and the oxygen atom has eight
electrons in a ground sta te configuration Is 2 2s 2 2 p 4 . The atomic orbitals that
y

.---+x

H H
a b
Figure 13.2

we will use are the Is-orbitals on the hydrogen nuclei which we denote <PI s(a)
and cPI s(b), and the 2p-orbitals on the oxygen nucleus , denoted cP2px, cP2p y,
cP2pz ' The atomic Is- and 2s-orbitals in oxygen are more tightly bound so we
neglect them in this simple calculation. The binding of the 2p-orbital in oxygen
and the 1s-orbital in hydrogen are approximately the same so that we must
take account of them both. The character table for the group C 2v is shown in
table 13.2, together with the characters for the spaces generated by the atomic
orbitals. ((Tv denotes a reflection in the plane of the molecule and (T~ a reflection
in the perpendicular plane containing the y-axis.)
Table 13.2

C Zv E Cz (Jv (J v

1 1 1
1 -1 -1
-1 1 -1
-1 -1 1

0 2 0 =A I -tBI
-1 1 1 = Al -tBI -tB2
Since all the irreducible representations are one-dimensional it is easy to
write down the symmetrised orbitals as
cP\~') oc {cPI s(a) + <PIs(b)}
cP\~" oc {cPIs(a) - cPI s(b)}
-I,(A,) _ -I, -1,(8,) _ -I, -1,(8 ,) - -I,
'+'2p - '+'2py ' '+'2p - '+'2px ' '+'2p - '+'2pz
13 .3 Electron States in Molecules 287

Since there are two A 1 -type orbitals we must allow them to mix. If we denote
the free-atom energies of 4>1$ and 4>lP by 6 1 s and 6 1 p then the effect of the
mixing will be to produce a pair of A I -type energies as shown in figure 13.3. In
the same way, the two BI-type orbitals will mix as shown. One may argue
qualitatively that, since the bond angle is greater than 90°, the overlap between
A,. (B .J and A,. (B d is greater than that between A,. (A d and ",(Ad . Hence one expects
'l'l s 'l'2p 'l' l s 2p
that the two B I -type levels will be split more widely than the two A 1 -type
levels, as shown in the figure. In the simplest approximation, the Brlevel has
the free-atom energy 6 1 p ' This ordering ofsingle-particle levels, as shown in the
figure, is consistent with experiment.
8,

Figure 13.3

To construct the wave function for the water molecule we must now place
electrons into these mo lecular orbitals, bu ilding up a Slater determinant. There
are six electrons to be placed, two of them coming from the Is-orbits on the
hydrogen atoms and four from the 2p-orbits in oxygen (recall that the oxygen
atom has a configuration of I S l 2s 1 2 p4 ). To describe the molecular ground
state we therefore place two electrons (spin up and down) in each of the lowest
three levels B A I and B 1 of figure 13.3.
"

13.3 Selection rules for electronic excitations in


molecules

Optical absorption can take place when the molecule undergoes a transition
from the ground state to an excited state. The simplest excited states can be
considered to be single-particle excitations from one occupied molecular
orbital to a higher unoccupied one. The energies involved are of the same order
of magnitude as in the atomic case so that absorption can be in the visible,
infrared or ultraviolet regions of the electromagnetic spectrum. The selection
rules governing the transitions can be obtained in the same way as for the levels
of an atom in a crystal field, see subsection 9.9.2. In the case of H 1 0, because
there are two electrons in each orbital, the ground state has A 1 symmetry (like
288 Electron States in Molecules

a closed shell). Excited states may be constructed by promoting an electron


from the least-bound orbital Bi into either ofthe unoccupied A 1 or B 1 orbitals
to form states of symmetry B 1 or A 1 , respectively, using table 13.2. Only the
first of these will be excited by electric dipole transitions since, from the table,
the vector representation does not contain A 1 .

Bibliography

Eyring, H., Walter, J. and Kimball, G. E. (1944). Quantum Chemistry (Wiley,


New York)

or one of the more specialist books

Murrell, 1. N., Kettle, S. F . A. and Tedder, J. M. (1970). Valence Theory (Wiley ,


New York)
Murrell, J. N. and Harget, A. J. (1972). Semi-empirical Self-consistent Field
Molecular-orbital Theory of Molecules (Wiley , New York)

Problem

13.1 The methane molecule consists of four hydrogen atoms arranged at the vertices
of a regular tetrahedron and a carbon atom at the centre (symmetry group Td ) .
Show that the four Is atomic orbitals on the hydrogen atoms lead to a singlet and
a lriplet molecular orbital. Note that the 2s and 2p atomic orbitals on the carbon
atom also lead to a singlet and a triplet of the same symmetries and hence that
these will mix to give bonding states.
14
Symmetry in
Crystalline Solids

In this chapter we discuss the effects of the translational and rotational


symmetry in crystals . The first three sections are concerned with the
translational symmetry and section 14.4 discusses in detail how the group
theoretical results are used in determining the wave functions for electrons in a
crystal. The next three sections sketch briefly how these results can be taken
over for various types of elementary excitation in crystals (lattice vibrations,
spin waves and excitons), and section 14.8 deals with the selection rules for
scattering processes involving the excitations. In the final section we include
the effects of the point-group symmetries. There is no attempt in this section to
discuss all aspects of space group theory but we give a short description of the
nature of the group elements and irreducible representations of the simpler
space groups followed by an application to electron states .

14.1 Translational symmetry in crystals

A crystal is formed by arranging atoms or ions in a space lattice defined to be


the set of points
(14.1 )
289
290 Symmetry in Crystalline Solids 14.2

with ni integers. The a, are called primitive translation vectors and must not be
coplanar. The parallelepiped defined by ai' a2 and a 3 is called the primitive
cell. The simplest crystals have just a single atom at each lattice point but more
generally an array ofatoms with specific relative orientation is associated with
each lattice point. Figure 14.1 shows a two-dimensional crystal with two atoms
per unit cell.

x denotes atom of type A


o denotes atom of type B

Figure 14.1

As defined by equation (14.1) the space lattice, and consequently the solid, is
infinite in extent. In reality a crystal is finite but contains a very large number of
atoms ~ 102 0 per em:'. Here we will consider only the infinite crystal from
which one expects to calculate what are known as the bulk properties of the
material, such as conductivity and specific heat. These will be relevant in
samples with a small surface to volume ratio, i.e. the number of atoms
influenced by the surfaces of the crystal must be small compared with the total
number.
The infinite crystal will generally have symmetries which involve rotation
but for the moment we consider only the pure translational symmetry
=
described by the group of translation operators P(n) P(n, n 2 n3) which
displace an arbitrary point r to r + n,
(14.2)
where the primitive translation vectors a, are fixed and the ni take integer
values. We denote the group by Y (a I ' a 2 , a 3 ) and observe that it has an
infinite number of elements.

14.2 The translation group :Y (a" a 2 , a 3 )

The multiplication rule for translations is easily derived from the defining
14.2 Symmetry in Crystalline Solids 291

equation (14.2) since


P(m) P(n) r = P(m) (r+n) = r+n+m (14.3)
= r+(n+m) = P(n+m)r

This shows immediately that the group is Abelian and, in particular, the
translations along the directions of the three primitive vectors a, commute with
each other so that the group can be considered as a direct product of three
subgroups of translations along these directions. We now consider just one of
these subgroups :1(a 1 , 0,0). This is an Abelian group and therefore its
irreducible representations are one-dimensional. From equation (14.3) it
follows that the matrix elements T(n 1 00) corresponding to the group
elements P(n 1 00) satisfy

(14.4)

Although we use the expression 'matrix element' all these matrices are one-
dimensional so that there is only one matrix element. We use the notation
T(n) == T(n 1 n2 n 3) rather than the cumbersome T(P(n)) for a representation
of the translation group. Given that the matrix element for the identity
P(O 0 0) is unity, i.e. T(O 0 0) = I and defining the constant o: by a = T( I 00) it
follows immediately from equation (14.4) that T(n 1 00) = rx n , . For T to be a
unitary representation rx must be a complex number of modulus unity and it is
conventional to write a = exp( - 2rrik d, where k 1 is real and serves as a label
for the representation

(14.5)

Since n 1 is always an integer it follows from (14.5) that two representations


labelled by k 1 and k'l = k 1 + p are identical if p is an integer. Thus to describe
all unitary irreducible representations, k 1 needs to take any value over an
interval of length unity . As we shall see later the obvious interval 0 ~ k 1 < 1 is
not usually the most convenient in three-dimensional problems.
The irreducible representations of the complete group of translations
:1 (ai' a 2 , a 3 ) are obtained as direct products of representations of the type
(14.5) and are thus one-dimensional and labelled by three parameters k 1 , k 2
and k 3 with the explicit form

T(k)(n) = T(k,OO)(n l OO)x T(Ok,O)(On 2 O)x T(OOk,)(00n3)

=exp[ -2rri(k 1 n 1 +k2n2+k3n3)] (14.6)

The exponent here has the form of a scalar product but, since the primitive
translation vectors a, are neither orthogonal nor normalised, it is necessary to
292 Symmetry in Crystalline Solids 14.2

introduce what is called a 'reciprocal basis' defined by


2na2/\ a3
bl = , etc. (14.7)
a l ·a 2/\a3

before we can write it as a scalar product. The reciprocal basis vectors have the
property
(14.8)

which may also be taken as a definition oftheb j • We now define a vector k with
components k, in the basis b,
(14.9)

so that
(14.10)

The irreducible representation matrix element may then be written as


T(k)(n) = exp (- ik .n) (14.11)

where n denotes the group element and k labels the irreducible representation.
In one dimension we noted that the representations k I + P with p any integer
were all the same as the representation k 1 and in three dimensions this will be
true for representations obtained by adding integers to any of the k j • This
result is more conveniently expressed if we introduce a 'reciprocal lattice' of
points given by the vectors
(14.12)

where the m j are integers. We now see that the representations k and k +Km
are equivalent,

T(k+X.)(n)=exp[ -i(k+Km).n] =exp(-ik .n)=T(k)(n) (14.13)

We conclude this section by deriving the simple rule for constructing


product representations of the translation group. Since the irreducible
representations of ff are one-dimensional the rule is trivial. The direct product
of the representations labelled by k and k' is the representation k + k' since

T(k)(n) X T(k'Xn) = exp( -ik.n)exp( -ik'.n)

= exp[ -i(k+k').n] = T(k+k'\_) (1·U4)


However, because of the equivalence noted above we may add a reciprcaJ
lattice vector without changing the result so that
T (k)x T(k ') = T(k+k'+X.) (14.15)
14.3 Symmetry in Crystalline Solids 293

14.3 The Brillouin zone and some examples

We have seen that the different representations may be constructed by


allowing the parameters k" k 2 and k 3 to range over a unit interval. In terms of
the vector k this defines a region of space in which k must lie and this region
may be visualised by using the reciprocal basis and equation (14.9). If we were
to choose the intervals 0::;; k, < I then the region would be the parallelepiped
defined by the reciprocal basis vectors b, , b 2 and b 3 • However it is usually more
convenient to choose a region which is as symmetric as possible about the
origin, allowing the k, to be negative. This reg ion is called the 'Brillouin zone'
and its boundaries may be defined in terms of the reciprocal lattice given in
equation (14.12). The Brillouin zone is defined as being bounded by the planes
which bisect normally the vectors joining the origin to all neighbouring
reciprocal lattice points.
Some examples of two-dimensional lattices are illustrated in figure 14.2
together with their reciprocal lattices and Brillouin zones. In these two-
dimensional examples the vectors b i are more quickly deduced from the
relation (14.8) than from the explicit definition (14.7).

Simple Hexagonal Face-centred


rectangular square

• ozL • .t... • • •
• •oLe
°z
Space lattice • 0, • 0, 0,
• • • • • • • • •
• • •
• •
• • •
.
~ ~ • • •~•
,•
Reciprocal
lattice and
first Brillouin • ',I b, b, • •
zone (shaded)
• •
• Ib,l:•21T 110,1• • • • • •
Ib,l: 41TI/310,1 Ib,1:2 /2 1TI 10,1

Figure 14.2

One reason for choosing the Brillouin zone rather than any other unit cell is
that it demonstrates most easily the point group symmetry of the reciprocal
lattice. This is in fact the same as the symmetry of the real space lattice even
though these lattices may not be the same. The proof is as follows . The
294 Symmetry in Crystalline Solids 14.4

reciprocal lattice vectors K m defined by equation (14.12) satisfy the equation


(14.16)
for all lattice points n and it is easy to show the converse, that any vector which
satisfies the condition (14.16) for all n must be a reciprocal lattice vector. I f we
now take a rotation R, from the point group which leaves the space lattice
invariant then R j- 1 n is also a lattice point so that exp(iK", .R j- 1 n) = 1. But
since the exponent is a scalar product this implies that exp(iRjKm .n) = I and
hence RjK m is also a reciprocal lattice point. This shows that all reciprocal
lattice vectors are taken into reciprocal lattice vectors by any R, in the group
and hence that the reciprocal lattice has the same point-group symmetry as the
space lattice .

14.4 Electron states in a periodic potential

The simplest way to treat electrons moving in a crystal is to take an


independent electron model where each electron is assumed to move
independently in a fixed potential V(r), which represents its interaction with
the nuclei and also, in an average sense, its interaction with the other electrons
in the solid . We assume the potential V(r) to have the translational symmetry
of the crystal V(r) = V(r + n) . This periodic potential is to be regarded in the
same sense as the central field used in describing the motion of electrons in an
atom, see subsection 8.6.1.
We can now use the results of the preceding sections to find the
characteristic properties of electrons moving in such a potential. Firstly we
define the effect of the translations on the wave function of an electron,
following the general definition (4.8):
T(n)¢(r) = ¢(p -I(n)r) = ¢(r-n)

This transformation T(n) satisfies the multiplication law of the group


T(m) T(n) ¢(r) = ¢(r- m - n) = T(m + n) ¢(r) (14.18)

as required generally by any representation of a group.


From our general results on labelling and degeneracy in section 5.3 we
would therefore predict that in a crystal with no symmetry other than
translational (i.e. no special relations between the aj) the eigenstates for the
electrons would transform according to the irreducible representations of the
translation group, i.e. they would be labelled by a k-vector and satisfy the
transformation properties

T(n) ¢ (k)(r) = T(k)(n) ¢ (k)(r) = exp ( - ik . n) ¢(k)(r) (14.19)


so that, using equation (14.17),
¢ (kl(r - n) = exp ( - ik . n) ¢ (k)(r) (14.20)
14.4.1 Symmetry in Crystalline Solids 295
Also, since the representation k is one-dimensional the eigenstate ¢ (kl(r)
would be non-degenerate. The important result (14.20) is the famous 'Bloch
theorem' for crystal states. Given the behaviour of the eigenfunction in one cell
of the crystal the behaviour in all other cells is determined by the k-Iabel and
involves a simple phase change. The theorem is often stated in a slightly
different form , by writing the eigenfunction (Bloch state) as

(14.21)
Substitution into equation (14.20) now shows that the function uk(r) defined
by equation (14.21) is a periodic function with the periodicity of the space
lattice
(14.22)
In other words the function uk(r) is an invariant with respect to the translation
group !I.
To find uk(r) we must write down the Schrodinger equation for an electron
moving in a periodic potential. The Hamiltonian is
p2
H = 2M+ V(r) (14.23)

with V(r) = V(r-n) . Thus, using the form (14.21) for the eigenfunctions
¢(k)(r), the Schrodinger equation is

[ ~~2 V2 + V(r) J
uk(r) exp (ik .r ) = e(k)uk(r) exp iik . r) (14.24)

which may be written as

[2:(V + ik)2 + el k ) - V(r) J


uk(r) = 0 (14.25)

Since uk(r) is periodic, this differential equation need be solved only within a
single unit cell using the periodic boundary condition on uk(r) at the edges of
the cell. Group theory has thus enabled us to replace a problem for the
complete crystal by one for a single cell. For each value of k , equation (14.25)
will have a set of solutions en(k ). We can see immediately from equation (14.25)
that as k is slowly varied, the solutions to this differential equation will change
continuously showing that the en(k) must be continuous functions of k . For
each n the energies en(k) are said to form a band since they vary over a limited
range of values as we now illustrate.

14.4.1 The nearly-free electron model

We will now discuss the general features of the el k ) by taking a very simple
model for the crystal. This model regards the electrons as being nearly free,
treating the periodic potential V(r) as a weak perturbation to the free-electron
296 Symmetry in Crystalline Solids 14.4.1

states. Clearly this approximation can only be appropriate for the outermost
atomic electrons (valence electrons) which, particularly in a metal, move fairly
freely through the crystal. We shall see in the next section that this model is
capable of explaining how elements from different sections of the periodic
table form metals or insulators.
If we put V(r) = 0 it is soon verified that equation (14.25) has solutions of
the form
u"k(r) = V;1 exp(iK".r)

Ti 2
e ~ ( k ) = 2M(K,,+k)2 (14.26)

where, because of equation (14.16), the periodic boundary conditions are


satisfied provided that K" is a reciprocal lattice vector (14.12).
For any fixed k there is therefore one solution for each choice of n, i.e, for each
choice of the three integers nt , n2 and n3. The 'bands' are then labelled by n
although it is sometimes the practice simply to number the bands 1, 2, 3 .. .,
starting with the band of lowest energy and progressing upwards con-
secutively. The constant vc- 1 is included to make the wave function u/Ik(r)
normalised over a unit cell of volume Vc = Iat /\ a2· a31 . This is most
convenient since, because of the periodicity, most integrals can be reduced to
integrals over a unit cell. We notice that the full wave function is
ljI(k)(r) = exp(ik .r)u"k(r) = vc- t exp[i(K,,+k).r] (14.27)
"
and equation (14.26) is just the usual expression for the energy of the free-
electron plane wave state (14.27). The slightly unusual form of the wave vector
is simply a consequence of limiting k to the Brillouin zone .
Figure 14.3 shows a plot versus k of e~ (k) given by equation (14.26) for a one-
dimensional crystal, i.e. atoms arranged along a line with separation a. In this
case the single reciprocal basis vector has length b = 2n /a and the Brillouin
zone boundaries are at ± tt]«. The figure shows how sections of the single free-
electron curve e(k) = Ti 2 k 2 / 2 M have been translated by reciprocal lattice
vectors into the Brillouin zone.
We now move away from the free-electron limit V(r) = 0 and consider the
effect of the periodic potential V(r) as a perturbation on the free-electron
states. Since V(r) is invariant under the translation group it will only couple
states belonging to the same irreducible representation, i.e. states with the same
k vector in the Brillouin zone . The perturbation series for the energy is thus

e,,(k) = e~ (k) + v c-
1
f
cell
exp[ - i(k + K II ) · r] V(r)exp[i(k + K,,).r]dr

+ Vc
-2" I
L.
J exp( - iKm·r) V(r)exp(iK".r)dr 1
0 0
2
+ (14.28)
m e,,- em
14.4.1 Symmetry in Crystall ine Sol ids 297

\
I
I \
,, I
I
I
I

\
\
1
I
\
\
,
1
I
I
I
I

\
\
I
,
I \
\
I
I
I
,
I I
I
I

, I
I
I

\ I \ ' I I , I
"v
I I
\'I
\ te, (k)
I :I
(
1\
, ~\
, I
I
I
II
I

, I I I
\ \ I I
,
I

(a)
\
\ I
I , I
I
I
I ,
I

I
\ \ I ' I I
,, \ ~
I ,
'(
1 \ I
I
I I
I

, \
I \ I \
,'
I I
i 2 I
, ;'
\ I
\
\
\
I
I \
, 1 I
Ie (k)=U I
1 / 0 2M I
\ 11 \1 \ I I I
, I I
\.l.
~
I,
I ,, / *,
1 1\
I \ I
1,\
,, I:
le.,( k)

,
I \ I
/ I / I

,,
I
,
/ I

-, ,
I
, 'A "- ,
>(
/
,
1 / /
/ I

" "- ,,>' " , ./ ~-2( k )


" -, .... '
-
I
"-
-" "
:::'c:
I I
':J.-""
"- I .- "- "-
"-
» >............... - _X_ ' -
-3"1. - "1" 0k "I.
__
3"/11

, I
I

\ I I
\ I I
\ I I
\ I
, I
,I
(b) ,
\

\\ I /
'.,"
", />,

"
\
,
' I'"
I

:/
I

I /
/
'

I
\
\ I
/ J..,
I \ /';' , /
I
" ....._ /
/ :',
I , _/
/ '\
<: »
// I
:
,
' - ""
"
,
,
. . -r _/ I ,_...
<;
I
~_ /
"' .... _ ....
, ,,.... - ........... ~ -",.- ... ....... ~ - ,,'
I

,
I

Figure 14.3

where the unperturbed energies e~ (k ) are given by equation (14.26). Since the
integrals are periodic it is sufficient to restr ict the integration to a unit cell, the
same region over which the wave functions were normalised. The first-order
term is j ust a con stant Vo, equ al to the average value of V(r). The matrix
298 Symmetry in Crystalline Solids 14.4.1

elements appearing in the second-order term are

I exp[ - i( K m - K,,) .r] V(rldr = I exp( - .«; _".r) V(r)dr


and we see that they are just the Fourier coefficients V", _" appearing in a
Fourier series expansion of the periodic function VIr)

V(rl = I exp(iKp .r) Vp


p

Vp = "e-II exp(-iKp .r) V(rjdr (14.29)


unit
cell

The perturbation series (14.28) can thus be written in the simple form

£,,(k) = £~(k) + Vo + I 1Vpl:


.0 + ... (14.30)
P C"-I.,, .p

where we have written p = m - n. The second-order term has the usual


'repulsion' effect between states and, because of the energy denominator, has
most effect when two states nand n + p have nearly the same energy.
lfwe take the simple one-dimensional model of figure 14.3(aj we expect the
'repulsion' to produce a spectrum qualitatively like that of figure 14.3(b) .
Notice that the unperturbed bands touch at the zone boundaries and at k = 0
so that at these points the perturbation expansion breaks down and degenerate
perturbation theory must be used . We will now calculate the splitting, between
the lowest two levels. near the point of degeneracy at the zone boundary
k = - ttla and ignoring the coupling to higher bands. Writing the solutions as
linear combinations of the two unperturbed states
Uk(X) = COUOk(X)+CIUlk(X) (14.31 )
the Schrodinger equation (14.25) leads to the matrix equation

(
£Z(k ) + Vo - £(k ) 0 VI )((:0)=0 (14.32)
VI £I(k)+ Vo-£(k) (I

where £(k) is the new energy and Vt is the first Fourier coefficient in the one-
dimensional Fourier expansion corresponding to (14.29),

Vm = ! (" expt - 2rrimx /a)V(x)dx (14.33)


a Jo
The solutions are
£±(kj = Vo + t[£Z(k)+£?(kj] ± H [£Z(kj-£?(kj]2 +4 Vi}! (14.34)
Away from the zone boundary, where VI is small compared with the difference
14 .4.2 Symmetry in Crystalline Solids 299

in energies s8(k) - s?(k ), the solutions approach the unperturbed values. At the
zone boundary itself s8(k ) = f;?(k ) and the roots are

(14.35)

This shows that a 'band gap' appears at the zone boundary with a width 21 V.I.
A similar calculation for the point of contact between bands labelled nand m
would yield a splitting of 21 V(n _ m) Iand the form of these results is illustrated in
figure 14.3(b).
This separation of the continuous range of allowed energies in the
unperturbed system into bands of allowed energies with gaps containing no
electron states is the central result of the theory of electron states in crystals.
We will sho w below how it can give rise to the difference between metals and
insulators. Before doing this however we look at the form of the wave
functions at the zone boundary. If we take VI to be negative, corresponding to
an attractive ionic potential at the lattice sites, the solutions to equation (14.32)
corresponding to the energies (14.35) a re (' ±= =F (' ± and hen ce from equations
o I
(14.31) and (14.27)
I
'P ~ ,,/a(x) = -2 -exp t -inx/a){l-exp(2nix/a)l
( a) l

= - ( ~ Yi sin ( nax ) (14.36)

and (14.37)

The wave function 'P ~ ,,/a ( x ) therefore has a large electron density near
x = ± a/2 and con sequently midway between lattice po ints, wherea s 'P ::: " /O(x )
has a large den sity at the lattice points and would be expected to have the lower
energy which is consistent with equation (14.35).

14.4,2 Metals and insulators

In the last section we found the general behaviour of the s(k ) curve for a
valence electron in a crystal. We now discuss the many-electron ground state
which is obtained by putting electrons into these single-electron states, taking
care of the Pauli exclusion principle. To do this it is convenient to consider a
crystal of finite size so that we can count the number of electrons and the
number of states. We shall therefore consider a very large but finite crystal, and
the boundary conditions we impose at the surfaces will restrict the values of k
which are acceptable for the Bloch states (14.21). The actual choice of
boundary conditions will not affect bulk properties and it is usual to use cyclic
conditions. We take a crystal with edges parallel to the primitive lattice
translations a, having N, cells along the edge parallel to a.. The total number of
300 Symmetry in Crystalline Solids 14.4.2

cells is then N = N I N 2 N 3' The cyclic boundary conditions require the wave
functions on opposite edges of the crystal to have the same value so that, for
instance
"'(000) = ",(N I 00) = "'(0 N 2 0) = "'(00 N 3 ) (14.38)
Using Bloch's theorem (14.21) this means that
exp(2n ik l N d = exp(2nik 2N 2 ) = exp(2nik 3N 3 ) = 1
and hence the possible values k jp of k, are given by kip = pi N j , with p any
integer. The allowed values of k thus form a lattice of points within the
reciprocal lattice having N, points along the unit vector hi and con sequently N
points in each unit cell. We must however restrict k to lie in the Brillou in zone
which has the same volume as a unit cell, and the above result can be expressed
by saying that there are just N allowed values of k uniformly distributed
throughout the Brillouin zone for a crystal containing N cells. Since the N, are
very large numbers there will be no real distinction between continuous and
discrete values of k when we discuss bulk properties except in so far as it
enables us to enumerate the states.
We now return to the one-dimensional crystal offigure 14.3(b) but assume it
to be finite with N cells so that there are just N Bloch states corresponding to
the N values of k in each of the bands n. We can put two electrons with opposite
spin in each of these states so that each band will hold 2N electrons. If the
atoms of our solid have just one valence electron each, like sod ium, then the
crystal with N atoms will have N valence electrons which, in the ground state,
will occupy the lowest N single-particle states and so only half fill the first
band. If, on the other hand, the atoms have two valence electrons the ground
state will have all the single-particle states in the first band occupied. On
applying an electric field a current can only flow if some electrons can be
accelerated and hence move to states with infinitesimally higher energy. For
the one-valence electron atoms, such as sodium, this is possible since there are
unoccupied electron states immediately above the highest occupied state. We
then have a conductor. For atoms with two valence electrons, like magnesium,
the band is full and the next available state is at the bottom of the second band
so that the minimum energy which can be given to an electron is equal to the
band gap 2 VI' We would therefore expect magnesium to be an insulator.
Unfortunately this disagrees with experiment but the reason is that we have
used a one-dimensional model. In a three-dimensional calculation we have
band gaps at all points on the three-dimensional Brillouin zone boundary but
they occur at different energies so that unless the gaps are large there may be no
energy region which is always in a band gap. Thus atoms with an even number
of electrons which, from the one-dimensional arguments, we would expect to
be insulators, are not always so in reality. We can illustrate this effect in two
dimensions by drawing the nearly -free electron energy bands from equation
(14.26)for a rectangular latt ice choosing k to lie first along the x-direction then
along the y-direction. The results are plotted in figure 14.4for the case a 1 = ~a2
14.4.2 Symmetry in Crystalline Solids 301

and demonstrate that for this system there will be no gap in energy between the
first and second bands since the gaps on the x- and y-boundaries of the
Brillouin zone occur at energies Ji2rr;2 j2Mai and Ji 2rr;2 j2Mai , see problem
14.3,and as shown in the figure this ditTerenceis generally large compared with
the gaps . The figure also shows how maxima and minima can now occur at
points away from the zone boundary or centre where the free-electron curves
intersect. Such features are important in considering excitations of electrons
from one band to another but we do not discuss this aspect (see Elliott and
Gibson, 1974, in the bibliography).

I
I
I
\ I
\ I i
\ I I
\
,, ,
I
/
I
\

, I
I I
I

\
, I,
, I
I "

1",,/

,
"
-,
_ _ _ _---"e::-_-::-;L--_-=::~:-~_ ____:;;:~------="------ k along x -a xis
Tr/ o ,

I ; ....
__/ \ I I "-
__ \ I
I '
\ I I
\ I I
I

,I
\ I I
\ I t{k)

I
I "-
/
/
/
,I
I
,
I "-
"-
II " ,
// I
I ,

--
/ I
,
I
I "

Figure 14.4 Nearly-free electron energy band s for a rectangular lattice with "I = 1"2'
The dotted curves are free-electron parabolas

A more complete picture of e(k ) is usually obtained by drawing the contours


of constant energy as a two-dimensional diagram in the first Brillouin zone for
each band. The first two bands are illustrated in figure 14.5 for the same system
as figure 14.4. The contours in figure 14.5 clearly indicate the maxima offigure
14.4 at the points k; = ±~rr; /a2' k, = O.
302 Symmetry in Crystalline Solids 14.4.3

'--t-----'--j-- k, \........f-+-+--1--. k,

(0 ) (b)

Figure 14.5 Surfaces of constant energy for a rectangular lattice. The dotted curves
are free-electron circles: (a) first band ; (b) second band

14.4.3 The tight-binding method

The nearly-free electron method developed in subsection 11.4.1 was


appropriate for the valence electrons in a crystal but if we want to find wave
functions for the electrons in the inner shells or core states a free-electron wave
function would clearly be a poor starting point. For these electrons the wave
functions must look much more like free-atom orbitals, being only slightly
affected by the interaction with neighbouring atoms. This suggests that we
might use a perturbation theory approach starting from the free atomic states
and treating the interactions between atoms as a perturbation. This is of course
degenerate perturbation theory since there will be degenerate atomic orbitals
cPn'm (r - n) corresponding to each atomic state cPn'm (r) and each site n.
Qualitatively the effect of the interactions will be to mix the degenerate states
to form a band of states with a spread of energy which depends on the strength
ofthe interactions. This effect is illustrated in figure 14.6, where we have shown
a very small bandwidth for the deep core states which are strongly screened by
the outer shells from the effects of neighbours, and a greater spread for the
higher energy states, which could eventually coalesce .
We first consider an 's-band' formed from electrons in atomic s-states with
1= m = O. To lowest order in perturbation theory the states will mix to form a
band of states each of which is a linear combination of these orbitals. The new
states must transform according to irreducible representations of the crystal
translation group .Y, i.e. they will be Bloch states labelled by a k vector and can
be found by projection (see section 4.19) from an orbital at n = O.
¢~k) (r) oc P(k)¢s(r)

oc I exp(ik.n)cPs(r-n) (14.39)
n
14.4.3 Symmetry in Crystalline Solids 303

n =2 ------<:

rsotoreo Intero cting


ot orns otorns

Figure 14.6

We notice that projection from an orbital centred on a different site only


changes the overall phase of ¢ ~kY(r) and thus produces the same state.
If the original atomic level is not an s-state there will be additional
degeneracy corresponding to the allowed values of m. In this case we must
construct linear combinations of the Bloch states obtained from each orbital
nlm. We write
¢ (k l (r)
nl
= L c km ¢(knlm) (r ) ( 14.40)
m

where

¢(~ l (r) =
nm
L" exp(ik .n)¢nlm(r-n)
304 Symmetry in Crystalline Solids 14.4.3

is the Bloch state for orbital m. The coefficients c~ are generally found by a
variational calculation as in section 5.7. In subsection 14.9.2 we show how they
may be found from symmetry considerations alone if the crystal has additional
point-group symmetry and the vector k points along a symmetry direction.
We can see that the method is analogous to the LCAO method for molecular
orbitals discussed in chapter 13, with the irreducible representation label k for
the translation group replacing CI. for the point group. The calculation is less
general than that used in chapter 13 in that we have restricted the sum (14.40)
to atomic orbitals with a single value of n and I. The method can therefore be
extended by taking in more atomic orbitals belonging to atomic states of
nearby energy, e.g. states with different values of I but the same principal
quantum number n. In this case the variational calculation yields a matrix
equation exactly equivalent to equation (13.5).
For the simple Bloch states (14.39) the energy is obtained by evaluating the
diagonal matrix element of the one-electron Hamiltonian H(r) in this state,
Hkk
ss(k ) = - (14.41)
I kk

where

H kk =f cP~k~(r)H(r)cP~k)(r)dr

I kk = JcP~k)*(r)cP~k)(r)dr
Using the specific form (14.39) of the Bloch functions these expressions can be
simpl ified

H kk = ~, exp[ik .(n-n')] f-: (r-n')H(r)cPs(r-n)dr

= ~. exp[ ik. (n - n')] f cP: (r' - n' + n) H (r') cPs(r') dr'

where r' = r - n and we have used the fact that H (r) = H (r'). Introducing
p = n' - n as a variable instead of n' gives

H kk = I eXP(-ik.P)fcP;(r-p)H(r)cPs(r)dr
f1,p

=I exp(-ik ,p)Hp (14.42)


f1 ,p

where
14.4.3 Symmetry in Crystalline Solids 305

In the same way we can write

I kk =L exp(-ik .p)I p

f
lI.p

with Ip = 4>;(r-p)4>s(r)dr (14.43)

The energy given by equation (14.41) can therefore be written

L exp(-ik.p)H p
s, (k) = -'-p -- ---- (14.44)
Lexp(-ik.p)Ip
p

To evaluate the integrals Hp and I p we take the Hamiltonian


Jj2
H(r) =- 2M V
2
+ V(r) (14.45)

with V(r) the periodic potential and rewrite it as


Jj2
H(r) =- 2M V2 + Vat(r) + [V(r) - Va,(r)] (14.46)

where Vat (r) is the central-field potential of the free atom. The first two terms
then form the Hamiltonian for a free atom at the origin and the s-state wave
function is an eigenstate of this Hamiltonian with energy t:s ,

(14.47)

Using this equation we can now write the matrix element H pas

Hp = f 4>;(r-p)[t:s + V(r)- Va,(r)]4>s(r)dr (14.48)

and equation (14.44) becomes

Lexp( - ik .p) Vp
t:s(k ) = s, + .!:..p -- - - -- (14.49)
Lexp( -ik .p)Ip
p

with

Vp f
= 4>; (r-p)[ V(r)- Vat(r)] 4>s(r)dr (14.50)

The first term on the right-hand side in equation (14.49) is just the free-atom
energy level and the second term gives the k dependence which spreads the
306 Symmetry in Crystalline Solids 14.5

level into a band. We can see qualitatively how this spreading depends on the
unperturbed state in the following way. If ¢. (r) is a deep core state such as a I s-
state, its wave function is very localised so that I p and Vpeffectively vanish for p
not equal to zero because there will be negligible overlap of the wave functions.
This means that the second term on the right of equation (14.49) reduces to
Vol 10 which is independent of k and the bandwidth is zero . For a less deep
state a crude approximation might be to neglect the overlap in I p but keep the
term in Vp for p a nearest neighbour. For a simple cubic crystal, with one atom
at each lattice point and for which the three lattice vectors a, are of equal length
a and mutually perpendicular, there are six nearest neighbours p . In this case
the energy (14.49) is given by
E.(k) = G. + Vo + 2 VI (cos ak; + cos ak, + cos ak:) (14.51)
where VI is the value of Vp for a nearest-neighbour site (all these will have the
same value in a cubic crystal). For a valence electron, several more terms should
be kept in both Vp and I P' The range 0 f values 0 f G. (k) as k varies is called the
bandwidth and the example given in equation (14.51) is seen to have a
bandwidth of 121 VII. To calculate the value of the bandwidth we must, of
course, take a model for the periodic potential V(r). A simple approximation is
to take V(r) to be the sum of the atomic potentials Va,(r), i.e.
V(r) = I Va,(r - n) (14.52)
II

With this choice , the shift Vo is small and negati ve since the perturbation in
equation (14.50) is very small in the central cell where the wave functions ¢.(r)
are large, and negative everywhere. For the same reason VI is also negative .
In a realistic calculation (see Harrison, 1970 and Zirnan, 1972 in the
bibliography) of the band structure it is necessary to choose V(r) more
carefully, since with the above choice the long range nature of the atomic
potential would mean that the series (14.49) would converge only very slowly.
It is therefore convenient to take V(r) to be a sum of localised potentials which
are similar to those of a free atom or ion in a small region of atomic size and
zero elsewhere . The energies Gnl must then be calculated for these potentials.

14.5 lattice vibrations


The vibrations of a crystal can be analysed by the methods developed in
chapter 6 for molecules . We shall start our discussion with a one-dimensional
monatomic model, i.e. identical atoms equally spaced along a line at intervals a.
Later, we generalise the results to three dimensions with several atoms per unit
cell.

14.5.1 The one-dimensional monatomic lattice

We showed in chapter 6 that the normal modes of vibration must transform


according to irreducible representations of the symmetry group which in the
14.5.1 Symmetry in Crystalline Solids 307

present problem means they must be classified by k, and that they can be
obtained by projection. Although in this simple problem with identical masses
the concept of mass weighting is unnecessary, we none the less use it so that
equations from chapter 6 may be taken over directly. The general displacement
is given by a vector q in a vector space with dimension equal to the number of
atoms. Thus we denote by e(n) a displacement of the nth atom by an amount
M -land q = LqnMle(n), where qn is the magnitude of the displacement of the
n
nth atom and M is the mass .
A displacement U(k) with the symmetry k is then obtained by projection from
e(O), as in equation (14.39):
U(k) = L exp(ikna) T(n)e(O)
n

= L exp(ikna)e(n) (14.53)
n

In this one-dimensional case, k = ±Ikl = 2nk 1 l a in the notation of equation


(14.9). This equation is analogous to equation (6.34) in the molecular vibration
problem.
The displacement U(k) obtained in this way is unique since an identical
displacement (apart from a phase factor) would be obtained by projection
from a displacement on a different site. This means that U(k) must be a normal
displacement and that there is only one normal mode for each k.
The frequency of this normal mode is given by equation (6.35) and follows
immediately if we know the form of the potential energy of the crystal for an
arbitrary displacement of the atoms. If we take this to have the same form as
equation (6.1) allowing qn to be complex,
V =! "L.m e:»:«: (14.54)

where qn is the magnitude of the displacement of the nth atom, then equation
(6.35) yields
L M - 1 Bnmexp( - ikna)exp(ikma)
n.m (14.55)
L exp( - ikna)exp(ikma) bnm
n.m
(Since the displacement U(k) is complex we must use the complex form (3.7) for
the scalar products in equation (6.35).) Because of translational invariance the
constant Bnm can depend only on the difference n - m so that, introducing a
variable p = n - m and writing B p for Bnm , we have

wf = M- 1 L [L Bpexp( -ikpa)]IL 1
p n

= M -I L Bpexp( - ikpa) (14.56)


p
308 Symmetry in Crystalline Solids 14.5 .1

A system of this type is the one-dimensional chain of atoms of mass M


connected to their nearest neighbours by Hooke's law forces with force
constant A.. The potential energy can then be written

v = L 1).(qn -qn- d 2
n

= 11. L (q; - 2qnqn - 1 + q; - d


n

By changing the dummy variable n in some of the terms this can be written as

v = 11. L (2q; -
n
qnqn - 1 - qnqn + d

so that it has the form of equation (14.54) with Bnn = 21., B np = -). for nearest
neighbours, and Bnp = 0 for all other pairs . Equation (14.56) now gives

I
rof = M [21. - Aexp( - ika) - Aex p (i k a ) ]

2A
= M(l-cos ka)
A plot of rof against k is shown in figure 14.7. The Brillouin zone boundaries
are at ± rr./a and the curve shows a great similarity to that of the lowest band of
states in the free-electron model. The degeneracy between states with opposite
values of k arises from the fact that this particular crystal has a point group
with inversion symmetry (see section 14.9).

-----:!:, - - - - - - - ------"'-...-"'----- - - - - - - , = - - _ k
o
Figure 14.7

The striking difference between these results and those ofchapter 6 is that, in
the latter, the spectrum of normal mode frequencies was discrete whereas here
it is continuous. This arises from our use of an infinite chain. On restriction to a
14.5.2 Symmetry in Crystalline Solids 309

chain offinite length with N atoms, the continuous curve of figure 14.7 would
break into N dist inct roots corresponding to N equally spaced values of k. In
real crystals the number N is so large that the distinction between discrete and
continuous spectra is of no consequence.

14.5.2 Three-dimensional crystals with several atoms


per unit cell
If we extend the discussion of the previous section to a three-dimensional
crystal with a number s of atoms per unit cell we can construct 3s orthonormal
basis vectors transforming according to the representation k of the translation
group ;Y by projecting on each of the 3s linearly independent atomic
displacements in a single cell. As in equation (14.53) they have the form

u:~) = Lexp(ik .n)e'i(n) (14.57)


"
where e,i(n) is a unit displacement, in the mass-weighted sense, of the tth atom
in the cell n. The suffix i = x, y or z specifies the direction of the displacement.
The normal modes will be linear combinations of these 3s vectors

U(k) = L c k . u(k) (14.58)


i .. 1; tJ II

To find the normal frequencies and the coefficients c':/I we follow the method of
chapter 6 by diagonalising the 3s x 3s matrix of the potential energy for the
displacement (14.58) in the basis (14.57). As an obv ious generalisation of
equation (14.54) we first write V in terms of the simple coordinates,

V =t L Bli".r'jmql~(n)ql)m) (14.59)
t ,t' ,iJ .n,m

where q,i(n) is the Cartesian component, in the i-direction, of the displacement


of the atom t in the cell n. Then from equations (14.57) and (14.58) we see that
modes labelled by k have components given by
qk(n) = ckexp(ik. n)jMI
u tl (

and substituting in equation (14.59) gives the value of V in a mode k as

Vk = t ttL'i] Ek./I,'J, c k.* c ~rj (L 1)


/I n
where k
E t i ," j=
1"
(M, M " w; .
B /io.r'jp exp( - ik .p ) (14.60)

Here we have written p = n - m and 0 denotes any fixed cell. It then follows as
a generalisation of equation (6.35) that the are eigenvalues of the matrix wi
(14.60).
To show the effect of having more than one atom in the unit cell we consider
a one-dimensional chain with two atoms per unit cell and Hooke's law nearest-
310 Symmetry in Crystalline Solids 14.5.2

.. a

M, Mz A A
. ... 0. 00000 000000.000000 00000.00000000000. 00000 Oll'"
I 2 I 2 I 2 I 2

.. cel l
n -I
••
cell
n ••
cell
n+ 1
... cell
n+2

Figure 14.8

neighbour forces, see figure 14.8. The potential energy is

v = !- I }.{ [q 2 (n) - q d n) ] 2 + [q d n) - q 2 (n - 1)] 2 }


n

= FI [2q i(n) + 2 q~ ( n ) -
n
q2(n)qdn) - qdn)q 2(n)

-qtln)q 2(n -l) -q2(n)ql(n+ 1)]

To get the last equation we have manipulated the dummy variable n. Hence the
numbers Bt o • t • p in the one-dimension al equi valent of eq ua tio n (14.59) are
B IO . 10 = B 20. 20 = 2),

B 20. 10 = B 10. 20 = B 10. 2 - 1 = B 20. 11 = -},


The roots wt of the matrix E ar e thu s found by so lving the equation

2). 2
-M - wk
1
=0
),
-------.-, [I + exp( -
(M I M 2 )1
.
ika)]

i.e.wt = A. [( M_1 +_1)±{(.


M
_I+_
M M
1)2__M 2 (l-coska)}!J _
I 2 I 2 IM 2

Figure 14.9 shows a plot o f wt as a function of k in the Brillouin zone. Again


this curve shows great similarity to the free-electron curves for the lowest two
bands. If more atoms are added to the unit cell, more bands are added so that
there is alway s one band for each atom in the unit cell. As k tends to zero, wt
tends to zero as k 2 and if we quantise the vibration as in chapter 6 the quanta of
excitation are called 'phonon s'.
In three dimension s the matrix E with elements given by (14.60) will ha ve
dimension 3s x 3s so that there are 3s roots for each value of k, This means that
there are 3 bands for each band in figure 14.9. The three lowest bands touch at
k = 0 and correspond at k = 0 to the three translational degrees of freedom,
which are not really vibrations.
In crystals with point group symmetries, degeneracies may occur at special
points in the Brillouin zone , such as along a line with rotational symmetry. At
14 .6 Symmetry in Crystalline Solids 311

o a
-tt

Figure 14.9

such values of k the matrix E can be partialIy diagonalised by symmetry


arguments which we outline in section 14.9.

14.6 Spin waves in ferromagnets

In a simple model of a ferromagnet, the atom at the site n in a crystal has a


magnetic moment J.l." and at sufficiently low temperatures (below the Curie
point) these moments line up in more or less the same direction so that the
crystal has a finite magnetic moment in the absence of any applied field.
Experiment shows that the g-factor is very close to 2 so that we can assume that
it arises from unpaired spins on the atoms and that the orbital contribution is
quenched by the crystal field as described in subsection 9.9.3. Although the
origin of the interaction energy is electrostatic, arising from exchange
interactions between the electrons, the phenomenon can be discussed in terms
of an effective Hamiltonian co ntaining a Heisenberg interaction term

V=-LJ"mS" ,Sm (14.61)


".m
where S"is the spin operator for the atom at site n. It is usual to take J 1111 = 0
and J 11m positive so that the ground state has alI the atomic spins paralIel and
hence alI magnetic moments paralIe\. In this situation every term SII'Sm takes
its maximum value 52, where 5 is the value of the atomic spin , and the ground
state energy is simply

Eo = _52 L J llm (14.62)


II.m
These statements can be justified if we denote the state in which alI spins are
312 Symmetry in Crystalline Solids 14.6

parallel, in a direction which we choose to call the z-direction, by 10 ) and write


the operator in the form S" . Sm = !(SII+Sm_ + S,,_Sm+) + S"zSmz' following
(7.35). Then we have

S"zIO ) = SIO >


and SII+IO ) = ° for all n so that
s n,smIO) = S21° ) (14.63)
The problem we set ourselves is to find the nature and the energy of the low-
lying excited states of the system which will be thermally populated at low
temperatures. The interaction (14.61) is naturally assumed to have trans-
lational invariance so that the constants J lIm depend only on the difference
m - n = q. We shall write J nm = J II ' The eigenfunctions of Vmust therefore be
labelled by k and it is evident that the ground state 1 °>
is translationally
invariant, with k = 0. Starting from the ground state 10 >with all spins parallel,
one may construct many different wave functions with given k. Our object here
is to construct the low-lying eigenfunctions. Bearing in mind the aligned
structure of the ground state one might seek an excited state by turning the
spin on a single site p with the lowering operator Sp_ and then projecting to
produce a state with definite k . Such a trial function would be written

Ik >= L exp(ik.p)Sp_IO > (14.64)


p

and we now show that this is in fact an eigenfunction of V for all k. The
operator S• .S. commutes with S" unless p = n or m, since spin operators on
different sites commute. Thus using equation (14.63)
Sn,Sml k ) = S2/k ) +exp(ik. n)[ (S",Sm)' S,,_] 10>
+ exp(ik. m) [(S",Sm)' Sm-] 10)

and the commutators follow from the familiar commutation relations (7.28)
and (7.30) for the spin operators giving
[(Sn ,Sm),Sn-] = Sm-S"z-S,, _Smz
so long as m # n. (The case of n = m may be ignored since J "" = 0.) Putting
together these two equations with (14.61) gives

+ exp(ik. m)(S,,_ - Sm_)} 10)

= Eolk >+ S L s,{[ 1 - exp(ik .q) ]exp(ik. n)S,,_


",m
+ [1 -exp( - ik . q) ]exp(ik . m)Sm_ } 10 ) where q = m-n
14.7 Symmetry in Crystalline Solids 313

=E o lk ) + S L J q [ l - ex p (ik · q ) ] lk )
q

+ SL J q [ l - exp( - ik . q )] Ik )
q

= Eolk ) + 25 L J q{l - cos k . q ) Ik ) (14.65)


q
This proves that Ik) is an eigenstate with an excitation energy of

,
25 I:J, (1 - cos k.q), where q runs over all sites relative to a fixed site. For each k
we have an excitation of thi s kind, which is called a 'magnon'. The analogy with
the lattice vibration problem of section 14.5 is apparent, with the turning of the
spin at a site corresponding to the displacement at that site. Notice that as
k -> 0 the excitation energy also tends to zero like k 2 , so that excitations may be
produced at infinitesimally low excitation energy. This contrasts with the finite
value which would be found for the expectation value of V in the wave function
Sp -10 ) in which only the spin at site p is turned. The state Ik ) corresponding
to k = 0, although having zero excitation, is not identical with the ground state
since it is given by I: Sp - 10 ) . However the operator I: Sp _ is the total spin of
p p

the system and this corresponds to an infinitesimal rotation of the direction of


alignment and does not represent any internal excitation of the system. This
degeneracy between states with different directions of magnetisation is
inherent in the form of V which is rotationally invariant. In a real crystal,
anisotropy would remove this degeneracy and would show up as additional
terms in V destroying its rotational invariance.
The operator from equ ation (14.64) which is usually denoted

a~ = ./VL exp(ik .p)Sp _


p

where .v' is a normalisation factor, creates a magnon of wave vector k . A two-


magnon state

can then be constructed. This is not quite an exact eigenstate but the error is of
the order 1/ N , where N is the number of atoms in the crystal and is large. To
the same order of accuracy the magnon creation and annihilation operators
obey Bose - Einstein commutation relations (see section 19.1 and subsection
16.3.1) and we can talk about a state in which there are nk magnons with wave
vector k. Each magnon carries an excitation energy, given in equation (14.65).

14.7 Excitons in insulators (Frenkel excitons)

In a simple insulator such as a rare-gas solid, with no valence electrons, the


314 Symmetry in Crystalline Solids 14.8

overlap of atomic wave functions is very small indeed and to a very good
approximation the ground state has individual electrons occupying the atomic
orbitals on individual atoms. The many -electron wave function is just a
product of these orbitals. As in the spin-wave case we might try to construct
excited states by raising an electron on one atom to a higher unoccupied orbital
in that atom, but again we must use projection to construct a state with definite
I..
Let ¢J denote the many-electron ground state of an atom and ¢J' the excited
state formed from ¢J by exciting one electron into a higher orbital. Thus ¢Jpand
¢J~ refer to such atomic wave functions at site p . For the entire crystal we have
the corresponding ground state

'f'0 = n¢J
p p

and for an excitation at site p ,

(14.66)

The projected state with the correct translational symmetry k , called an


exciton, is then given by

'f'(k)=
ex
Lp exp(ik .p)'f'p (14.67)

Note that this differs from the tight-binding model where the states we
constructed were one-electron states. An exciton of wave vector k is thus an
excitation which propagates through the crystal (but of course does not carry
charge) . The energy will depend on the k vector, but only to a small extent, and
will be approximately equal to the-atomic excitation energy.

14.8 Selection rules for scattering

Selection rules for scattering processes in crystals can be derived immediately


from the simple relationship (14.15) for the product of representations. For
instance if we consider the scattering of an electron in a crystal
(electron - phonon scattering) then the probability for scattering from a state
of wave vector k to a state of wave vector k' due to the absorption ofa phonon
of wave vector q will depend on a matrix element which will vanish unless the
product representation T(q)x T(k) contains the representation T(k'). From
equation (14.15) this mean s that
k' = q+k+K", (14.68)
Scattering with K m = 0 is usually called direct and that in which K m f= 0 is
called an Umklapp process. Equation (14.68) is often referred to as the
conservation of 'crystal momentum'. In a scattering process the energy must
14.9 Symmetry in Crystalline Solids 315

also be conserved so that an additional condition is


nw'l + e(k ) = e(k') (14.69)
Similar relationships hold for electron - magnon, magnon - phonon,
electron - photon, phonon - photon and magnon - exciton scattering.

14.9 Space groups

So far in this chapter we have looked at the consequences of the translational


symmetry in crystals. However, the infinite crystal may also be invariant under
one of the finite point groups of rotations (proper and improper) about each
lattice point which we described in chapter 9. This will occur for example when
the primitive lattice vectors a, are symmetrically placed relative to each other.
We now consider the group t§ which is obtained by combining the elements
P(n) of the translation group :Y with the elements R, of a point group of
rotations about a lattice point which we call the origin.
The general group element is written {R j , n} and is defined by its effect on
an arbitrary vector
{Rj , n } r = P(n) Rjr = R jr + n (14.70)
We can easily verify that these elements satisfy the group postulates by
deducing the multiplication law from equation (14.70).
{Rj,m}{ R j,n}r = {Rj,m} (R jr+n) = Rj(Rjr+n)+m
= RjRjr+Rjn+m
= {RjRj,Rjn+m}r
so that {Rj,m} {R jon} = {RjRj,R jn+m} (14.71)
Because the lattice is invariant under the point group, the vector R jn + m gives
a lattice point if nand m are lattice points and hence the product (14.71) is
contained in the set (14.70).
As a special case of (14.71) we see that rotations and translations do not
commute, in fact putting m = 0 and R, = E (the identity) we have
RjP(n) = { Rj, O}{ E,n} = {R j, R jn } = P(Rjn)R j (14.72)

which can also be written as


(14.73)

Thus the group t§ is not a direct product of :Y with the point group. As
another special case of (14.71) we deduce the inverse
{R j,n}-l = {R - 1, -Rj-1n}
j (14.74)

by asking what choice of Rj and m leads to the identity {E, O} on the right-
hand side of (14.71).
316 Symmetry in Crystalline Solids 14.9.1

Although R, refers to rotations about a lattice point at the or igin, the group
t§ includes rotations about other points because the product P(n)RjP( - n)
represents a rotation R, about the lattice point n (see problem 14.6) rather than
the origin. Furthermore, for particular lattices, t§ also includes rotations about
points other than lattice points. For example, in a two-dimensional square
lattice in the xy-plane it is verified that the operation {R z(n I2) , -ad is a
rotation R, (nI2) about the point (t,t, 0) at the centre of a cell. Reflections in
planes bisecting the cells may be generated in the same way.
In some crystals there may be additional symmetry operations which are not
of the form (14.70). These are the so-called screw-axis and glide-plane
transformations which are combinations of translations with rotations in
which neither of the separate transformations is a symmetry element. We shall
not investigate the larger groups obtained by including such operations but
refer the reader to Nussbaum (1966) and Birman (1974}-see the bibliography.
This neglect will not invalidate our results but means that in some crystals
there will, for example, be further degeneracies unexplained by our dis-
cussions. The word 'symorphic' is used to describe the group t§ defined above,
with 'non-syrnorphic' referring to the more general space groups with glide-
planes or screw-axes.
The inclusion of time-reversal will also enlarge the space groups in much the
way that it enlarged the point groups in section 9.8. In particular for magnetic
crystals there will be what are known as magnetic space groups in which time-
reversal occurs only in conjunction with a rotation or translation and is not
itself a symmetry operation. For instance, in many antiferromagnets, trans-
lation by a lattice vector must be accompanied by time-reversal to reverse the
directions of the spins . We shall not discuss these magnetic space groups any
further but a full account may be found in Bradley and Cracknell (1972) in the
bibliography to chapter 9.

14.9.1 Irreducible representations of space groups

If we have a representation T of the space group t§ then the operator


corresponding to the group element {R j , n} is denoted T( R j , n). These
operators must satisfy the same relations as the group elements themselves, so
that from the definition (14.70) we can write
T(R j, n) = T(n) T(R;) (14.75)
and in particular from (14.73)
T(n) T( R;) = T( R;) T( R j-
1
n) (14.76)
In function space the representation operators are defined in the usual way
to satisfy
(14.77)
where we have used the result (14.74). We shall be particularly concerned with
14.9.1 Symmetry in Crystall ine Solids 317

the effect of group transformations on basis vectors e" which have been chosen
to transform irreducibly with respect to the subgroup of translations !!7 and
hence carry a label k, We therefore derive here a few useful results. First, under
a simple translation, we have from equation (14.11), by definition of e",
T(E, n)e" = T(n)e" = T I")(n)e" = exp( - ik . n)e" (14.78)

Under a more general transformation { R j , n } we form the new vector

ei. = T(R j, n)e" = T(n)T(Rde"


= T(RdT(R j-
1
n)e"

using the result (14.76), and hence


e't = T(R j, n)et = exp( - ik . R j - 1
n)T(Rde"

= exp( - iR jk. n)T(Rj)e" (14.79)


We now consider the effect of a pure translation on this new vector e"
T(E,m)e~ = T(E,m) [T(R j, n)ed = T(R m +
j, n )e"
= exp [ -iRjk.(n+m)]T(Rj)e,k
= exp( - iRjk. m) [T ( R n)e,,] (14.80)
j,

= exp( - iRjk . m ) e ~

This last equation shows that the vector e~ transforms irreducibly under the
subgroup of translations and belongs to the representation T lR;"I. (We use the
general method given in section 20.3.)
These general results will now be used in constructing and classifying
(labelling) the irreducible representations of the space group r§. Let L denote
the vector space of an irreducible representation. Since the translations form a
subgroup !!7 of r§ we are free to choose basis vectors which belong to
irreducible representations T I") of !!7. Let e" be such a basis vector. We can
then generate a set of vectors by operating on e"
with all the group elements
T(Rj, n). Before continuing the discussion we must make clear that there are
two different kinds of vectors involved. There are the vectors k in ordinary
three-dimensional space and there are the vectors e k in the vector space L 0 f
the representation. We shall refer to both kinds simply as 'vectors' but it will be
clear from the context and from the notation which kind of vector is meant. It
is convenient to discuss first the case when k is a vector with no special
symmetry, by which we mean that none of the rotations R, leaves k invariant.
In this case, the set of vectors k, = Rjk are all different , since if k, = k j then
R j- 1 Rjk = k and k would have special symmetry. The set of vectors k, is called
the 'sta r' oik , Hence, from equation (14.80), the set of vectors e", = T(RJet in L
all belong to different irreducible representations T(k ,) of !!7 and are therefore
linearly independent. Furthermore, we now show that they form an invariant
space and thus provide a representation of r§ . In other words, the set of vectors
e,,; generated from e" by the action of group rotations T(R j ) only, form a basis
318 Symmetry in Crystalline Solids 14.9.1

for a representation F of the complete group <§. To do this we calculate the


effect of a general transformation {R j , m } on the vector ek . :
T (R j ,m )e k • = T (R j,m)T (R dek

= T (R jR j,m)e k
using equation (14.71). If we call the group element RjR j = R, we can use
eq uatio n (14.79) and write
T (R j ,m )e k , = exp( - i R, k . m)T (R,)e k
= exp( - ik. . m)e k
. (14.81)

where k, = R,k = Rjkj. Th is eq ua tio n gives the general matrix element for a
general group element { Rj , m } o f <§. We note that the dimension of the
representation I" is equal to g, the number of elements in the point group. As in
subsectio n 7.4.2 we can now argue that the set e k, forms a basis for the space L
o f the irreducible representation because, if it did not, th en the space L would
have been reduced. Thus the structure of the irreducible representations of <§
is described by the basis vectors ek with the matrix elements (14.81). The
representation is labelled by a n arbitrar y vector k and denoted by flkl. There is
clearly an equ ivalence between r 1k)and r 1k.) for any k, in the sta r of k and, as
with the translation group, there is equivalence if k is replaced by k + K. , where
K. is an y reciprocal lattice vector.
If k has special symmetry (e.g. lies on an n-fold rotation axis or in a mirror
plan e) some o f the group elements R, will leave k invaria nt. These clements
must form a subgro up o f the po int group a nd it is called the 'litt le group' of k .
If g, denotes th e order o f the little group of k ; and g the order of the point
gro up, then one sees (problem 14.7) that the sta r of k now contains only gig,
vectors, which we call k p ' In general, there is more than one rotation R with the
property Rk = k p , for fixed k and k p , but for each k p we choose just one of
these and denote it by H p . It is easy to show that any rotation may be written
R, = HpG a , where G , belongs to the little group.
The irreducible repr esentati on s of <§ will now have a more complicated
structure since the vectors e k are no longer necessaril y linearly independent.
The subset of the e k for which k, = k, correspond ing to elements in the little
group, will form a subspace L ' of L which is invariant under the little group.
Consequently this subspace provides a representation of the little group. Since
the little group must itself be one of the point groups its irreducible
representations are well known and are denoted by T (2 ). It can be shown that if
L is to give an irreducible representation of the space group, then the
representation of the little group given by the subspace L' mu st be irreducible
with respect to the little group. Thus if S2 denotes the dimension of T(2) then
there will be a set of S2 linearly independent basis vectors eka~, where
JJ = 1,2, ... , Sa' which provide the representation T (2) of the little group and
all behave like T lk)under translations. By taking elements H p which lie outside
the little group we can now generate a corresponding set ofS2new bas is vectors
14 .9.1 Symmetry in Crystalline Solids 319

e liX~ for each ofthe gig, vectors k p in the star. Thus the representation of r§ has
dimension SiXg/g, and is denoted by rid).
To construct the represent ation matrix elements, corresponding to equation
(14.81) for the case of special symmetry, we must first write the general rotation
in the form R j = HpG a , where G, is an element of the little group and H, is one
of a set of rotations, one for each point in the star, with the property Hpk = k p'
Starting from the e" lJ we define the general basis vector by
e'iX ) = T(H )e ('"
kplJ p klJ
then
T(R.)' n)e(iX'
kplJ
= T(n)T(R ).)T(H P )e(ikXu)

= T(n)L, T( ~) (G )T(H )e("',


IJ IJ b q ku
IJ

which gives the matrix elements. In der iving this result we have defined H, and
G, as the factor s in the anal ysis of the rotation RjHp = HqG b since any
rotation can be analy sed in th is way. The vector k q is given by k q = Hqk which
is the same as Rjk p because k is unchanged by G b • The numbers T ( ~) (G b ) are
the standa rd irreducible repre sentation matrices for the little gro~p.
Some example s ofstars' for two-d imensional latt ices a re illustrated in figure
14.10and for a three-dimensional example we take a simple cubic lattice where
the Brillouin zone is also a cube of side Ztt]«. The point group here is 0h and

(0 ) (b) ( c) ( d)

Figure 14.10 Stars of k for a two-dimensional square lattice with symmetry group C 4v :
(a) general k ; (b) k alo ng x-axis: (c) k along sq uare diagon al; (d) k at zone boundary
alo ng x-ax is
for a general k the star o f k will have 48 vecto rs k, The special values of k with
symmetry are shown and labelled using the standard solid-state phys ics
not ation in figure 14.11. For a general point ~ along the kx-axis the little group
is C 4 v with 8 elements and the star has 6 points. For the particular point X on
the face of the zone , with k x = Tr ia, there is additional symmetry since it is
320 Symmetry in Crystalline Solids 14 .9.2

-t----*.

Figure 14.11

equivalent to the k-vector with k x = - ttl a. The little group of X therefore has
a reflection plane perpendicular to the x-axis and is D 4 h with 16 elements. The
star of X has thus only 3 points. At the centre r of the cube the little group
becomes the full cubic group 0h ' In a similar way if we proceed from r to M
along the line r M, then at a general point L the little group is C 2 v with four
elements and the star of L has 12 po ints. At the special point M, however, all
four corners of the cube in the kxkz-plane are equivalent and the resulting
symmetry group is D 4h with 16 elements so that the star of M has only three
points M, M 2 and M 3 in figure 14.11.

14.9.2 Application to electron states

In section 14.4 we saw how the translational symmetry led to a labelling of


the eigenfunctions by k corresponding to the irreducible representation T(k)of
the translation group .:1. However , there was no degeneracy because these
representations are all one-dimensional. The inclusion of the point-group
symmetry has led to representations rlk)ofthe space group C§ which are again
labelled by k but which have dimension g (the order of the point group) for
arbitrary k . This means that the value £(k ) of the energy for some k will be
repeated at all k, in the star of k or in other words that the constant-energy
surfaces when plotted as a function of k will have the full point-group
symmetry.
For a vector k with special symmetry the irreducible representation r lak)
has dimension sagl g/ giving rise to a degeneracy of the same amount. This
corresponds to a degeneracy of order gig/ due to the number of points in the
star and a degeneracy of order Sa at each such point corresponding to the
irreducible representation of the little group.
14.9.2 Symmetry in Crystalline Solids 321

We now consider how we can use these results in a band-structure


calculation for the motion of the electrons in a crystal using the tight-binding
method of subsection 14.4.3. We shall illustrate the procedure by calculating
the form ofe(k) for the bands constructed from atomic p-functions when k lies
along the x-axis in a cubic crystal, with the little group C 4 v of the point .1 in
figure 14.11. The character table for this group is shown in table 14.1 and the
irreducible representations labelled .1 1 to .1 5 , The three atomic p-functions

Table 14.1 Character tables for the group C 4 y at ~, O; at F and D4 h at X, for the
common elements (i.e. for the elements which appear in C4 y ) . The irreducible
representations are here labelled by ~j , F, and Xj, respectively, rather than by the
standard notation of appendix A.l in order to distinguish the representations of the
different groups. The groups O, and D4 h are product groups with an S2 factor so that
their representations carry a ± label. We have only listed the odd parity representations
since we are primarily interested in the p-states. The only even parity states to occur
belong to the trivial identity representations [i and Xi .

C4 y E C2 2C4 2C1 y 2C1d

~1 1 1 1 1 1
~2 1 1 1 -1 -1
~3 1 1 -1 1 -1
~4 1 1 -1 -1 1
~5 2 -2 0 0 0

°h
[i 1 1 1 -1 -1
[- I 1 -1 -1 1
2
[3 2 2 0 -2 0
[,;- 3 -1 1 1 1
[; 3 -1 -1 1 -1

D4 h
Xi 1 1 1 -1 -1
Xi 1 1 1 1 1
X3 1 1 -1 -1 1
X,;- 1 1 -1 1 -1
X;; 2 -2 0 0 0
Xvector 3 -1 1 1 1

transform as the vector representation, see section 5.4, with character given in
the last row of the table. This reduces to a sum of .1 1 and .15 under the little
group. Of the three p-functions, Px belongs to .1 1 and the pair Py and pz belong
to .15 ' (Remember that x is the four-fold axis.) As in equation (14.39) we can
construct three Bloch states

r/J(k) =
q
L exp(ik.n)r/Jq(r-n)
/I
(14.82)

where q = x , y or z and, for the case being considered with k along the x-axis,
322 Symmetry in Crystalline Solids 14 .9.2

they transform in the same way as the p-functions under the little group. If
these states are now used in a variational calculation, as outlined in section
14.4.3, we may therefore conclude that the states </J ~k) and </J~k) will be
degenerate, labelled by the representation f( ti ,k I, while </J If) is labelled by f(ti,k) .
Because they belong to inequivalent representations these states will not mix.
If we add to the variational calculation th e Bloch state </J(k) s
formed from an
atomic s-state, it will o f course have .11 symmetry and therefore mixes with the
state </J(k)
5
. The matrix elements of H in the sa me nearest-neighbour approxi-
mation as equation (14.51) are therefore

H~, = 1:, + V,JO) + 2 V,,(O 0 a)(2 + cos kxa)


Hk sx = 2i V,x(a
. 0 Ojsin kxa
H~x=l:p+ Vu(O)+2[2Vxx(00a)+ Vxx(aOO)coskxa]
H~y = H~z = I: p + Vxx(O)+ 2[ Vxx(O 0 a) + Vxx(a 0 0) + Vx x(O 0 a)cos kxa]
where we have extended the notation of equation (14.51) with, for example,

Vsx(P) = f </Ji(r - p)[ VIr) - Vu,(r)]</Jxlr)dr

and used the fact that


VxxlO 0 a) = Vxx(O a 0) = Vyy(a 0 0) = V}')'(O 0 a)
Vxxla 0 0) = V}')'(O a 0)
Vxx(O) = Vyy(O)
Vsx(O a 0) = V,x(O 0 a) = 0
The 2 x 2 matrix for the mixture of s- and p-states can now be diagonalised and
rather than write out detailed expressions for th e energies we have sketched the
results in figure 14.12. The figure demonstrates that, as the magnitude of k

El k )

rtl....---
';:;------_-:-- 1-- --+ x -axis
r x
Figure 14.12 Tight-binding energy bands for s- and p-like atomic states with k along
the x-axis
14.9.3 Symmetry in Crystalline Solids 323

tends to zero, the energies for ~ 5 and ~ 1 coalesce and the s and p wave
functions are no longer mixed . Thi s additional degeneracy is however to be
expected since, at the point F, the little group becomes the full octahedral
group O; and the three Bloch functions of equation (14.82) transform
according to the irred ucible representation r 4- of 0h' The representations ~ 1
and ~ 5 of C 4 v , which label curves which come together at I", may be obtained
from the representation r 4- by the process of reduction of r .; on restriction
from O; to the subgroup C 4 v , see sectio n 4.18. This connection of ~1 and ~5
with r;; is called a compatibility relation. In this way we can also label the
irreducible representation at the po int X by irreducible representations of the
little group at X which is D4h and from the character table we deduce the labels
sho wn in figure 14.12.

14.9.3 Other excitations

All the symmetry properties of t; (k ) which we have derived in the previous


section for the energy bands of electrons in solids apply to the other excitations
discussed in sections 14.5, 14.6 and 14.7. For instance, in a lattice-vibration
calcula tion the three orthogonal atomic displacements at a site transform in
the same way as the p-function s, so that the arguments used for the tight-
binding bands for p-functions in subsection 14.9.2 can be taken over in
calculating the vibrational spectru m of a simple cubic lattice.

Bibliography

For general background reading o n the physics of this chapter we suggest


one of the following:

Kittel, C. (1971). Introducti on to Solid State Ph ys ics (Wiley, New York)


Elliott, R. J. and Gibson, A. F. (1974). An Introduction to Solid State Physics
and it s Applications (Macmillan, London)
Harrison, W. A. (1970) Solid Stat e Theory (McGraw-Hili, New York)

A more advanced treatment is to be found in

Ziman, J. M. (1972). Principles of the Theor y of Solids (Cambridge University


Press)

and for a specialist review on symmetry and electron states in metals see

Nussbaum, A. (1966). Solid St. Ph ys., 18, 165

For further read ing on space groups see


324 Symmetry in Crystalline Solids

Birman, J. L. (1974). Theory of space groups and infra-red and Raman lattice
processes in insulating crystals, Handb. Phys., XXV (2b); Light and matter
(1b)

Problems
14.1 Show that the vo lume of the Brillouin zone is (2n)3 /v" where Vc is the volu me of
the primitive cell.
14.2 Show that an y vector K m which satisfie s equation (14.16) is a reciprocal latt ice
vector.
14.3 Show that the free-electron parabolas in figure 14.4 are given by

eOn,n, ()k = 2M
~ { ( 2n" 1
at +k x
)2 + (2n"2 )2}
~ + k).

where "I and "2 are integers.


14.4 Draw the Brillouin zone for a plane square lattice in two dimensions with
symmetry group C 4 v and mark on it the six kinds of special point where the little
group of k is non-trivial. Wh at is the little group at each of these po ints ? Draw up
a table of compatibility relati ons for the irreducible representations for k along
the x-di rection of figur e 14.10 as k -00 and as k -s n ]a.
14.5 In subsection 14.4.1 we used deg enerate perturbation theory to construct the
wave functions and hence the energies at the zone boundary for the nearly-free
electron model. The fact that the wave funct ion s were odd or even is an example
of the states transforming irr ed ucibly under the little group at that po int.
Consider the two-dimensional square lattice of the prev iou s problem and do a
similar calculation using the four degenerate plane waves corresponding to
k = ( ± nf a, ± x /a ). These all ha ve the same reduced k vecto r which is one of the
special po ints in the zone. The little group at the po int is the full symmetry group
C 4 v ' By calculating the character of the representation generated by the four
equ ivalent plane waves, show that the degeneracy must be split into a doublet
and two singlets and hence use projection operators to con struct the eigenstates.
Calculate the nearly-free electron energies of these states to first order in V(r ).
14.6 Show that the product P(n ) R, P( - n) represents a rotation through the sa me
angle as R; but about the point n.
14.7 Show that if g, is the order of the little group of k and g is the order of the full
point group, then the star of k contains gig, vectors.
14.8 Show that the three Bloch functions rjJ(k )of equation (14.82) do , in fact, transform
in the same way as atomic p-functions'centred at the origin, under operations of
the little group.
15
Space and Time

Much of this book has been concerned with symmetries in space. So far as we
can tell, space is homogeneous in the sense that an experiment carried out in
one laboratory will give the same result as in another and the result will not
depend on the direction in which the experiment is set up. This implies that the
laws of nature are invariant under translations and rotations and leads to the
conservation of linear momentum and angular momentum and all the
associated phenomena discussed in earlier chapters. (Within a crystal the
positioning of the atoms produces a potential field which lacks these
continuous symmetries and is invariant only with respect to finite rotations
and translations, as discussed in chapter 14.) Time is also believed to be
homogeneous in the sense that an experiment carried out today will give the
same result as one carried out tomorrow. This invariance with respect to time
translations leads to conservation of energy, a feature which is built into the
Schrodinger equation which we have used hitherto as our starting point. The
introduction of the theory of relativity by Einstein rested on the idea that space
and time possess a deeper homogeneity; that the laws of nature are invariant
under a set of transformations (Lorentz) which operate in the four-
dimensional 'space' of space and time . The transformations include rotations
and translations in both space and time as special cases. The concept of the
(rest) mass of a particle has a natural place in this theory and one finds for
325
326 Space and Time 15.1

example that the Schrodinger equation does not satisfy the new invariance. It
is necessary to derive a new formulation for both classical and quantum
mechanics to incorporate the new invariance. However, the corrections
implied by this modification are important only at velocities near that oflight,
about 3(10)8 m/sec. Thus it turns out that for most classical and quantum
mechanical applications, including the structure of atoms, molecules and
nuclei, the relativistic corrections are small. In this chapter we shall study the
groups associated with the Lorentz transformations and see how their
representations are related to physical quantities. The improper rotations
generated by the inclusion of the space inversion may also be incorporated in
the four-dimensional framework along with the time inversion operation.
The concept 0 f four-dimensional space and time is first introduced in section
15.2 where we describe the Lorentz group. Translations in four dimensions are
included in section 15.4 leading to the Poincare group. The improper
operations of space and time inversion are included in smaller sections, 15.3,
15.5 and 15.6. Although some indication of the physical significance of the
transformations is given in section 15.2 we delay a discussion of the
significance of the irreducible representations until the last two sections. The
chapter begins in section 15.1 with a discussion of the Euclidean group which
combines rotations with translations. This group is concerned entirely with the
three space dimensions and in a sense does not belong to the present chapter.
Also the physical interest in the Euclidean group is quite small but we include it
here for two reasons. Firstly it is very similar to the Poincare group but
somewhat simpler. Secondly, a group isomorphic with the two-dimensional
Euclidean group occurs in our study of the Poincare group and we shallbe able
to quote from the results of section 15.1. A reader who is anxious to move into
four dimensions could defer section 15.1 until he meets this quotation in
section 15.4.

15.1 The Euclidean group C 3

We have studied rotations and inversion in some detail in earlier chapters and
in chapter 14 we described the symmetry of crystals with respect to finite
translations and finite rotations-the 'space groups'. We now study the group
of continuous translations and rotations following the procedure used in
section 14.9 for the space groups. We stress that the present section is
concerned entirely with the three space dimensions.

15.1 .1 Translations

First we consider pure translations in one dimension. There is an infinite


number of group elements P (~) denoting translations along the x-axis P( ~)x
=x +~, where -00 < ~ < 00. The multiplication law for the translation group
is obviously P(~d P(~2) = P(~I+~2) and it is an Abelian group. In fact the
multiplication law is the same as that for the group !Yi 2 (see section 7.3). The
15.1.1 Space and Time 327

irreducible representations are therefore one-dimensional of the form


T(k) (¢) = exp( - ik ¢) (15.1)
with any value of k. There is now no restriction that k should be an integer since
there is no periodicity in the variable ¢ but k must be real if the representation
is to be unitary. The transformation of functions is defined to be
l/J'(x) = T( ¢) l/J(x) = l/J (P( - ¢)x) = l/J(x - ¢)

following equation (3.37) so that the function l/J(x) = exp(ikx) transforms like
the irreducible representation T(k). Following section 7.3 the infinitesimal
operator then takes the form P x = - a/ox.
The exten sion to translations in three dimensions is trivial since translations
all commute. Thus the three-dimensional translation group is a direct product
of three one-dimensional translation groups in x, y and z. The general group
element is thus denoted by P(p) with the definition
P(p)r=r+p (15.2)
where r is a general position vector. Writing p = (¢, I}, () there are clearly three
parameters and three infinitesimal operators P x ' P y and P z which in a space of
functions l/J(r) are given by - a/ox, - %y and - %z. The irreducible
representations are obvious generalisations of equation (15.1), being labelled
by three numbers kx ' k; and k, which form a vector k,
Tlkl(p)=exp(- ik.p) (15.3)
If the Hamiltonian of a quantum system is translationally invariant then the
infinitesimal operators P,; P y and P z represent conserved quantities, essen-
tially the three components of the total linear momentum of the system . As
they stand, these operators are skew-Hermitian, being the infinitesimal
operators of a unitary transformation, but they can be made Hermitian by the
inclusion of a factor i. The dimension is also inappropriate for a momentum
but this can be put right by including a factor h. One actually defines the
physical momentum operators Pq by Pq = iliP q • The reason for associating
these operators with momentum is that in classical mechanics also the
translational inva riance of a system leads to the conservation of the total linear
momentum, as explained in section 16.1.
Since the irreducible representations of the translation group are one-
dimensional this invariance leads to no degeneracy, but simply to the labelling
of the eigenfunctions by the representation labels k. From the expression given
above for T(k)(p) one sees that the infinitesimal operator P, has the eigenvalue
- ik q in a state with given k . Thus the momentum is given by hk .The presence
of any potential V(r), depending on position r, will violate translational
symmetry unless V is trivially a constant. Thus for a single particle, translation
invar iance is relevant if the particle is moving freely. For a system with more
than one particle, any internal potential like V(r 1 - r2) is clearly invariant
under translations and thus the total momentum of the system is conserved.
328 Space and Time 15.1 .2

15.1.2 The group operators

The 'Euclidean group' iff3 in three dimensions is formed by the products of


the proper rotations R(a) and the translations P(p). Here a = (a.x> ay, a z )
denotes the three parameters of the rotation R(a) in the notation of section 7.4.
The general group element is thus written as P(p) R(a) and its effect on a
position vector r is given by
P(p)R(a)r = R(a)r+ p (15.4)

using equation (15.2). Notice that rotations and translations do not commute,
which is obvious geometrically, and in fact taking the product (15.4) in reverse
order we have
R(a) P(p)r = R(a) (r + p) = R(a)r+ R(a)p
so that we may conclude that
P(p)R(a) = R(a)P(R-1(a)p) (15.5)
Because of this result it follows that one need only consider group elements of
the form P(p) R(a), since any product in the reverse order may be written in this
way using equation (15.5).
Following our general definition (3.37) of the transformation of functions
we define the operator Ttp, a) in function space by
T(p, a) "'(r) = "'{ [P(p)R(a)] - I r } = "'{ R - I (a)p - 1 (p)r}
= ", {R -1(a)(r-p)} (15.6)
which is similar to the notation used in section 14.9.

15.1 .3 The irreducible representations

Since the rotations and translations do not commute, the group iff3 is not a
direct product group so that the labelling of the irreducible representations of
iff3 is a non-trivial problem. However, the existence of the translations as an
Abelian subgro up enables us to solve this problem following the procedure
used in section 14.9 for the space groups.
We use the notation T(p, a) to denote the representation operator cor-
responding to the group element P(p)R(a) and, by definition, the T(p, a) must
satisfy the same multiplication laws as the group elements, i.e. from equation
(15.5).

T(p, a) = Ttp, O)T (0, a) = T(O, a)T (R - I(a)p , 0) (15.7)

Before constructing the representations of iff3 we need to establish the


following property of rotations, namely that any rotation R may be written
R = RxyRz where R, is a rotation about the z-axis and Rx y is a rotation about
an axis lying in the xy-plane. To justify this factorisation we define k = Rk o,
where k o lies along the z-axis. Now define Rx y to be that unique rotation about
15.1.3 Space and Time 329

an axis lying in the xy-plane which carries k o into k, viz. Rxyk o = k. From these
definitions we see that
R;/ Rk o = R;/k = ko
so that the rotation R;y1 R leaves k o unmoved and must therefore be a rotation
about the z-axis, i.e. R;/ R = R. which gives the desired result,
(15.8)
The irreducible representations E of the Euclidean group tff 3 may now be
generated by a sequence of steps as follows . (See also section 20.3.)

Step (I). Choose basis vectors Ik > which belong to irreducible repre-
sentations T(k) of the subgroup of translations so that
T(p, 0) Ik > = exp( - ik . p)lk > (15.9)
(It will be more convenient in this chapter to use the notation Ik > for a basi s
vector rather than ek which would be more consistent with section 14.9.)
Step (2) It is now shown that the vector T(O,a)lk > transforms like the
representation T(R(a)k)of the subgroup of translations. Using equations (15.7)
and (15.9) we have
T( p, O)(T(O, a)lk » = Ttp, a)lk >
= T(O, a)T(R - I (a) p, 0) Ik >
= T(O, a)exp[ - ik .R - I(a ) p ]Ik >
= exp[ - iR(a)k . p] (T(O, a)lk » (15.10)
which is the desired result. This shows that if the representation E contains a
basis vector Ik >then it must also contain vectors Ik' >for all k' = R(a)k, i.e. for
all k' with Ik'i = Ikl.
Step (3) We now choose a reference direction, usually the z-axis, denoting a
vector in this direction by k o , a basis vector which transforms like T(k o ) under
translations by Iko > and a rotation about k o by R(a o). As a result of step (2)
above, the vector T(O,ao) Ik o > will also transform like Ik o > under translations
because R(ao)k o = ko . Thus the set of basis vectors ofE with the same label k o
will form an invariant space under rotations about the z-axis and may
therefore be labelled also by the irreducible representations of the group 9t 2 of
rotations about a fixed axis. Thus we may introduce the more detailed notation
Ikorn >in place oflk o >where rn labels the :Yi 2 representation (see section 7.3)
and
(15.11)
Step (4) We are now in a position to construct the representation E. It will be
labelled by Ikl = Ikol and rn. Starting with a single basis vector Ikorn >we
generate the set
Ikrn> = T(O,ii(k))lkorn > (15.12)
330 Space and Time 15.1 .3

where k is any vector oflength Ikol and R(ii(k)) is that unique rotation about an
axis a(k) in the xy-plane which carries k o into k, i.e. R(ii(k) )k o = k , We use the
tilde on ato signify that this vector lies in the plane perpendicular to k o , i.e. the
xy-plane. There is therefore an infinite set of basis vectors Ikm ) in E which all
have the same m and the same length of k but with different directions of k. To
prove that this set is invariant under the group Iff3 we consider the effect of a
general operation T'(p, a) on an arbitrary basis vector Ikm ),
T(p, a) Ikm ) = T(p, O)T(O, a)T(O, o(k) )Ikom )
= T(p, O)T(O, a(k') )T(O, c)lkom )
= T(p, O)T(O, a(k'))exp( -imc)lkom )
= exp( - imc)T(p, O)lk 'm )
= exp( - imc)exp ( -ik' . p) Ik 'm > (15.13)
In deducing this result we have used equation (15.8) for the rotation
R(a)R(d(k)) writing
R(a) R(a(k)) = R(o(k')) Rz(c) (15.14)
which defines the rotations R(a(k')) and Rz(c). The vector k' is defined by
k' = R(d(k'))ko or equivalently by k' = R(a)k using equation (15.14). Thus
given aand k one can deduce k'and cso that equation (15.13)actually provides
the representation matrix elements of E.

By an argument like that used for the group :Ji 3 in subsection 7.4.2 it follow s
that, if it is to be irreducible, the representation E can have only one basis
vector with the labels Ikom ) . Thus the irreducible representations of g 3 are
given by equation (15.13). To be definite, we choose a particular axis, the z-axis,
in generating the basis vectors but it can be argued quite simply that
representations generated by taking k o in different directions are equivalent so
long as k o ha s the same length. Thus the direction of k o is irrelevant in labelling
the inequivalent, irreducible representation of rff 3 and we use the notation
Ellkl.mJ for E, where Ikl denotes the length of all vectors k appearing in the
representation.
As an example of these representations we see that the representation Ellkl.OJ
may be generated, using equation (15.12) from the function expUlklz) of
position ofa single particle. The basis vectors are given by IkO) = exp(ik.r),
I.
where k runs over all vectors with fixed length Ik To generate the
representations Elikl.m) one need s to take at least two particles and the function
exp[ilkl(Zt +zz)] {(xt-xz)+i(Yt-Yz)}m would suffice for Ikom ).
The procedure used here in the study of the irreducible representations of
rff 3 is another example of the general method of 'little groups' used in
subsection 14.9.1. Here the little group is the group 9t z which leaves the vector
k o unchanged and the 'sta r' contains an infinite number of points. This method
will be used later in studying the Poincare group and in fact it is largely for that
15 .1.4 Space and Time 331

reason that we have d iscussed f! 3 in detail.


In the particular case k o = 0 the discussion above will be inappropriate,
since no non-zero vectors k can then be formed by step (4). However in this
case the representation space is translation invariant and so the irreducible
representations a re precisely those of the rotation group .~ 3 described in
chapter 7 and we denote them by E(o.j) == DW. Thus, only for k = 0 does the
group f! 3 have finite-dimensional irreducible representations.
Since :Ji3 is a subgroup of iff3 it is in principle possible to label the basis
vectors of an irreducible representation of iff3 by the Jl 3 labelj. However this
is more complicated than the method given above which uses the subgroup of
translations. Notice that since translations and rotations do not commute, one
cannot use both labels k and j. In quantum mechanics this statement is
equivalent to saying that in a system with iff3 symmetry the momentum and
the angular momentum are both con served but since their operators do not
commute they may not be measured simultaneously. The choice of basis made
for Ellkl.m) corresponds to one in which the momentum is diagonalised
together with the angular momentum about the direction of the momentum.
Such a basis is somet imes said to be a 'helicity representation'.

15.1.4 The group s,


In this section we briefly describe the Euclidean group iff2 in two
dimensions, not because it is of interest in its own right but because it will be
used later, in section 15.4, in constructing irreducible representations of the
Poincare group. Following the steps set out for iff3 we choose basis functions
to satisfy the two-dimensional form of equation (15.9). There is then no
element but the identity which leaves the two-dimensional vector k invariant.
Thus the irreducible representations are labelled simply by ElIk!l. An exception
occurs when k = 0, since all states of the representation are then translation
invariant and the group iff2 is effectively restricted to Jl 2 • Thus when k = 0
each representation T(III, of 91 2 (see subsection 7.3.1) provides a representation
of If 2 ' We denote it by E(O, III) to remind us that k = O.

15.1.5 The physical significance of the Euclidean group iff3

If a Hamiltonian is invariant under the full Euclidean group then by the


usual arguments we expect the eigenstates of the system to be labelled by the
possible irreducible representations Ellkl.m) rand E(o. j) of iff3 and we also expect
to find degeneracies. For a physical system to have iff3 symmetry, not only
must the Hamiltonian be independent of position, to satisfy translation
invariance, but it must also retain rotational invariance. For a single particle
this is therefore relevant only for a free particle, with wave function exp (i k . r)
and energy ,,2 k 2 12M. Furthermore this function must correspond to the
representation E<lkl. O) as remarked above. The infinite dimensionality of this
332 Space and Time 15.1.5

representation corresponds to the infinite variety of directions of k with given


Ikl, and physically corresponds to the infinity of possible directions of the
momentum of a free particle with fixed energy.
The other representations, E II k l, m) with m t= 0 and E(o. '', cannot be
accommodated if we restrict to a single classical particle. For example, E(O, j)
would imply a particle at rest but with angular momentumj. For a system of
particles the representation E(o. j) has obvious relevance; their centre of mass is
at rest but they have an internal angular momentumj. Such a representation,
with integer i. would result from a pair of particles interacting through some
potential V(r l - r 2 ).
To appreciate fully the relevance of the representation E(lkl,mj when m t= 0 it
is helpful to study the infinitesimal operators of S 3 and to construct the
Casimir operators (see section 7.5). The six infinitesimal operators are simply
the P x' P, and P, for "translations and the Xx, X, and X, for rotations. The P,
commute among themselves , the commutation relations of the X, were given
in equation (7.25) and those between the Pq and X, are soon deduced from the
multiplication law (15.5), see problem 15.1. The results are
(15.15)

with cyclic rearrangements. The operator p 2 = P~ + P; + P; is clearly an


invariant and will serve as one Casimir operator and one soon verifies from
equation (15.10) that P , X also commutes with all P, and X, and is therefore a
second Casimir operator. (Note that X . X is not translation invariant.) The
Casimir operators must necessarily be proportional to the unit operator within
an irreducible representation and their values in E' lki.m) are seen to be·- k 2 and
- km, respectively . The latter follows quickly by choosing the particular basis
vector Ikom> and using equation (15.11) for small Qo . Physically these two
Casimir operators determine the magnitude of the momentum and the
component of angular momentum in the direction k of motion. If we use the
explicit differential form P, = - clex, etc., with X = r /\ P from equation
(7.21), which would be appropriate for the operators relating to a single
classical particle, then one sees immediately that P .X = 0 as expected,
implying m = O. Now consider a pair of classical particles so that P x = P XI
+P x = - oiox i - clox 2, etc., and X = r l /\ PI +r2/\ P 2 = t(r l +r2)/\ (PI
+P/)+t(r l -r2) /\ (PI - P 2 ) . The latter identity simply expresses X as the
sum of an external and an internal part and, on forming the product P. X, only
the internal part survives , giving P .X = t(P1 +P 2) ·(rl -r2) /\ (P I-P 2 ) ·
Thus the Casimir operator P .X measures the component of internal or relative
angular momentum in the direction of motion of the centre of mass of the pair.
Within the representation Etlkl. mi. all states Ikm ) will have the same magnitude
Iklfor the total momentum of the pair but the direction of k will vary from
state to state, The component of internal angular momentum in the direction
k for the state/knl) is always m, independent of k , For the energy to depend on
m the Hamiltonian must contain terms like P. X which in physical terms is a
scalar product of total momentum and internal angular momentum, Notice
that the magnitude of the internal angular momentum plays no part in the
15 .1.6 Space and Time 333

representation ElI"l.m) only the projection is relevant. If the internal motion


were in a state of internal angular momentum 1then the (21 + 1) projections m
= I, 1- 1, . . ., -I would belong to different representations oft! 3 and could
have different energies. In other words the gp 3 symmetry of internal motion is
not contained withinC 3 ' (An extension to the Lorentz group, which includes
transformations between systems moving with uniform relative velocity, is
necessary before the internal angular momentum gains relevance from the
point of view of symmetry.)
The discussion in th is sectio n has been confined to one or two classical
particles, although the extension from two to any number is straightforward.
One can of course introduce an intrinsic spin s of a particle in the way
described in section 8.4. One does not attempt to define new internal
coordinates but this new degree of freedom is confined to the vector space of
2s + 1 states which transform according to O(s) under internal rotations. When
at rest, such a particle would trivially belong to the representation E(o. s) oft! 3
but in motion the significance of s would be lost. As in the case of the internal
angular momentum of the pair discussed above, only the spin along the
direction of motion has significance in C 3 ' For a particle with spin s the
infinitesimal operator X is the sum of the usual orbital part r /\ P and a spin
part X, which is defined as a matrix operator in the (2s + l j-dimensional spin-
space of the particle, as described in section 8.4.

15.1.6 Scalar products and normalisation of basis vectors

This small section may be omitted since it will not be used directly, but we
feel it necessary to comment on the fact that the irreducible representations of
C 3 have a continuum of basis vectors. Throughout the general discussion of
chapter 4 and in alI previous applications to physical problems the irreducible
representations have had finite dimension and were therefore expressible as
finite matrices. Here, for the first time, we have found irreducible repre-
sentations E(lkl.m) which have a continuous infinity of linearly independent
basis vectors Ikm >corresponding to different directions of k , This situation
will occur again for the Poincare group in section 15.4.
Let us restrict to vectors IljJ >belonging to a particular representation space .
Then, in the case of a finite representation, we write
N
IljJ >= L ljJ;ji > (15.16)
i= 1

where Ii>denotes a set of linearly independent basis vectors and the ljJ j are
numbers. We then define the scalar product of two vectors Ic/J >and IljJ >as
N

<c/J lljJ >= L= 1 c/Jil/J j


j
(15.17)

which implies that


(15.18)
334 Space and Time 15.2

A transformation T is unitary, with respect to this definition of scalar product,


if <T4>!TIjJ > = <4> IIjJ > for all 14> > and IIjJ >. For a representation like
E(lkl,m) with a continuous basis Ikm > we write , in place of (15.16),

IIjJ > = f IjJ(k) Ikm > dn k (15.19)

where the integration runs over all directions of k and IjJ (k) is a function of
direction of k. We then define a scalar product as

<4> IIjJ> = f 4>*(k)IjJ(k)dn k (15,20)

We may now verify that the representation E(lkl,m) given by equation (15.13) is
in fact unitary. From equations (15.13) and (15.19)

ITIjJ > = f IjJ(k)exp(-imc-ik' ·p)lk'm >dn k

f
= IjJ(R -1k')exp(- imc - ik' ·p)lk'm >dn k·

using the fact that, since k ' = R(a)k with fixed a, the surface integral dn k may
be replaced by dn k" i.e. the Jacobian has the value unity. With a change of
notation for the integration var iable this may now be written as

ITIjJ > = f
IjJ(R - I k)exp( - imc- ik . p)lkm > dn k

so that comparing with equation (15.19) and using the definition (15.20) of
scalar product we have

<T 4> IT IjJ> = f4>*(R - 1k)IjJ(R- 1k)dn k

= f 4>* (k)IjJ(k)dn k = <4> IIjJ >

The analogue to equation (15.18) may be deduced only if we make use of the
Dirac b-function of the angles of k' and k. Notice that the definition of scalar
product (15.20) is not quite the same as the integral over all momentum space,
since here we keep the magnitude of k fixed.

15.2 The Lorentz group .!f

Having discussed the way in which rotations and translations come together in
the Euclidean group we now temporarily discard translations and study the
15.2.1 Space and Time 335

extension of the rotation group :Jt 3 into four dimensions when time is
included. We define the four-dimensional space by associating with each
'event', or happening, a 'point' with four coordinates (x , y, z, cr) which denote
the place (x , y, z) and time l of the event , where c is the velocity of light. Thus all
fo ur coordinates have the dimen sion of length. We shall denote the point by a
vector e = (x, y, z, et ) where e is the four- vector from the 'origin' (0, 0, 0, 0) to

°
the point (x, y, z, et ). Remember that an event at the o rigin in four-space
implie s an event at the position x = y = Z = at time t = 0.

15.2.1 The Lorentz transformation

In the rotation group :Jt 3 we are interested in vectors, = (x, y, z) in a three-


dimensional space and the transformations r' = R(a), which preserve distance
and angle, i.e. which preserve the scalar product rv.r, = X 1X 2 + YIY2 +Z IZ2
between any two vectors, 1 and rj . In the Lorentz group Y we are interested in
vectors e = (x, y, z, et ) in a four-dimensional space with real transformations
e' = Le which preserve the scalar products defined as
(15.21)

Such transformations L are called ' Lorentz' transformations and by definition


they satisfy the equation
(15.22)
for all vectors el and e2' The notation e has been used for a four-vector to
distinguish it from a three-vector in ordina ry space which would have been
.e
written as e. We use the simple notation e l 2 for the four-dimen sional scalar
product (15.21). The reason why we are interested in such transformation is,
quite simply, that one finds from experiment that the laws of physics are
invariant with respect to them. The definition (15.21) could ha ve been made
with a different sign throughout without any essential change in the
transformation but the sign difference between the space and time terms in
equation (15.21) is crucial. If the sign of all terms were the same then the group
would be 9i 4 but nature shows no invariance with respect to &4'
The scalar product of a vector with itself e.e = - x 2 - y2 - Z 2 + e 2 t 2 is said
to measure the 'length' of the vector, or the 'interval' between events occurring
at opposite ends of the vector. The interval is defined as (e .e)!. Since from the
definition (15.21) a scalar product may have either sign it follows that an
interval may be real or imaginary. A real interval is called time-like and an
imaginary interval is called space-like for obvious reasons. The interval
between any two events at el and e2 is similarly defined as
(i l - e2 .el - e2)! and from equ ation (15.22) the interval between any two
events is unchanged by a Lorentz transformat ion .
If we define unit vectors ex,e}',ez ' e, as (1, 0, 0, 0), etc., then we note, from the
definition (15.21) of scalar product, that ex.ex = - 1, etc., for the space vectors
while t,.e, = 1. It is con venient to introduce the matrix g with elements defined
336 Space and Time 15 .2.1

by gij = ei .e j , where i = x, y, z or t, so that

( -~ o

~)
0
-1 0
g=
o o -1
o o 0

We define the matrix elements L j i of L by the equation

Lej = L Ljje j
j

so that

Given an arbitrary vector e = L eiei with components e, then L carries e into


j

the new vector

with components ej = L L jie i . Given a fixed vector ewith components e, with


i
e
respect to the basis i then its components with respect to the transformed
basis e~ = Le j are given by

The defining property (15.22) of the Lorentz transformation leads to the


relation

between matrix elements of L which in matrix form is


g = LtgL (15.23)
where t,' denotes the transpose ofL. Since g2 = 1 this implies that geg = L -1
which differs from an orthogonal transformation, see section 3.5, through the
presence of the matrices g. One can soon verify that L - 1 is itself a Lorentz
transformation and that the set of all L form a group, see problem 15.2.
[Because of the unusual form (15.21) of the scalar product, in which the
'metric' is not the unit matrix, these formulae differ from those given in section
3.7 through the presence of factors gii ' It is possible to avoid the explicit
appearance of these factors by defining what are called covariant components
ei for each vector e which are defined as ei = giiei ' The original set e, are called
15.2.1 Space and Time 337

the contravariant components. In this way the factors gjj may be absorbed by
switching from one set to the other. Thus for example a scalar product takes
the form

with the last two expressions involving both covariant and contravariant
components but avoiding the g jj' In this book we prefer not to introduce the
extra sophistication of having both sets of components and therefore we shall
always use the contravariant components e j denoted by a lower index and
transforming according to the equations above.]

Proper and improper transformations


In section 7.4 we distinguished between proper and improper rotations
depending on the sign of the determinant. A similar distinction occurs in the
Lorentz group but there are now four different types of transformation in the
sense that one cannot pass continuously from one type to another. From
equation (15.23) we have
(det L)2 = 1
so that one can distinguish between transformations with det L = ± 1.
Furthermore if we write out equation (15.23) explicitly for the matrix element
in the fourth row and fourth column we have

so that

Since the Lij are real it follows that L;' ~ 1and hence the values of L" can be in
either of two separated regions L" ~ 1 or L" ~ - 1. Thus we have the four
types of transformation L distinguished by

(l) det L = 1, L" ~ I which contains the identity


(2) det L = -1 , L" ~ 1 which contains the space inversion I
(3) det L = - 1, L" ~ - 1 which contains the time inversion I,
(4) det L = 1, L,,~-1 which contains the space - time inversion II,

Transformations of type (I), which are all connected continuously to the


identity, are called proper. They form a group which we shall call the 'Lorentz
group' !fl. Extended Lorentz groups may be formed by includ ing either type
(2) or type (3) with type (1). We refer to these groups as !fl. and !fl"
respectively. Inclusion of all four types gives an even larger group which we
denote by !fl.,. The word 'orthochronous' is sometimes used to describe
transformations with L" ~ I and the word homogeneous to distinguish all
338 Space and Time 15.2.1

these groups from those which include translations. We shall not use this
nomenclature. In thi s section we confine our attention to the group 2 .

Pure Lorentz transformations


Before con sidering the general Lorentz transformation we look at two
spec ial cases,

and

1 0 o
o 1 o (15.24)
Q z(b) = 0 0
cosh b
( o 0 -sinh b

The first of these, written in partitioned form, is simply a rotation , since the
condition (15.23) implies that RtR = 1 and the time coordinate is unchanged.
The second leaves all points in the xy-plane unchanged and is called a pure
Lorentz tran sformation (o r 'boost")in the z-d irection. It cont ains a single real
parameter b and each value of b in the range - 00 < b < 00 corresponds to a
different transformation Q, (b). The transformed vector e is given by

e' = Q z(b)e = (x, y , z cosh b - ct sinh b, ct cosh b - z sinh b)


and the tran sformation ma y look a little more fam iliar if we use the parameter
P= tanh b instead of b, so that
z' = (z - pct)/(1 - P2)1, y' = y, x' = x
ct ' = (ct - pz)/(1 - P2)! (15.25)

(We shall see later that p has the physical significance of a velocity p = vi c
measured relative to the velocity c of light.) We may clearl y define a similar
boost transformation in an arbitrary direction defined by a unit 3-vector u.The
notation Q(b) will be used where b = bu and b defines the magnitude of the
boost, as in equation (15.24). The transformed vector e' = Q(b)e may be
deduced as a generalisation from the result for Q z(b) and is most easily
expressed by giving the space and time components e' and ct' of e',
e' = e + { (cosh b -1)(e .u) - ct sinh b} u
ct' = ct cosh b- (e .u)sinhb (15.26)
15.2.2 Space and Time 339

The 4 x 4 matrix of Q (b) may be read off from these equations. For example
Qxx(b) = 1 + (coshb -1)u;, QyAb) = (coshb -1)uxu y,
Q,x(b) = -uxsinhb, Qu(b) = coshb
There is similarity between equation (15.26) and the corresponding expression
given for rotations at the beginning of section 7.4.
Since Q(b) and R(a) are both Lorentz transformations it follows that their
product is also a Lorentz transformation for any a and b. Further than this we
now show that any Lorentz transformation L may be written as such a
product. In principle this result is very similar to the factorisation (15.8) of an
arbitrary rotation. We first define eo = (0,0,0, 1). Then, for each L we may
e
construct a vector = Leo . From equation (15.26) one sees that there is a
unique boost Q(b) with the property Q(b)eo = e. Thus we have
eo = Q-l(b)e = Q-l(b)Leo

showing that the product Q - 1 (b) L leaves eo invariant and must therefore be a
pure rotation. We may therefore write Q - I (b) L = R(a) so that
L = Q(b)R(a) (15.27)
We shall find it convenient to parameterise L in this way by the six parameters
needed to define the two three-vectors a and b. We note that whereas the
magnitudes of the rotation parameters are limited in the usual way lal ~ 11:, the
boost parameters are unlimited. This is clear from equation (15.24) which
shows that Q z satisfies the defining equations of a Lorentz transformation for
all b.

15.2.2 The regions of space -time

Before discussing the physical meaning of the Lorentz transformations it is


helpful to study their geometrical meaning. It is sufficient to consider a simple
boost Qz(b) in the z-direction and for this we need to draw only the z and ct-
axes as in figure 15.1. Because of the definition (e.e)i oflength', vectors ein the
(z, et)-plane with the same length will not lie on circles but on the hyperbolae
- Z2 + e t = e.e = A, where A is a constant. Since A may take either sign it is
2 2

convenient to distinguish vectors of the following six types

(l)e .e<O called space-like


(2) e . e > 0 with t > 0 called future time-like
(3) e.e 0 > with t < 0 called past time-like
(4) e.e = 0 with e -# 0 and t >0 called future null
(5) e. e = 0 with e -# 0 and t <0 called past null
(6) e = 0 the zero vector

This classification divides space-time into regions as shown in figure 15.1.


Space-like and time-like regions are separated by 45° lines of z = ±et which
340 Space and Time 15.2 .3

Post null-e----'......, Future null

---':2~~-+---;~~--+-f---~~-;---cl

Figure 15.1

contain the null vectors. Notice that a null vector, with e.e = 0, does not have
all its components zero; this property is restricted to the single zero vector
e = 0. Although figure 15.1 is restricted to a single space dimension, the
classification is quite general since it is based on the scalar product and the sign
of t. In the general case the space-like and time-like regions are separated by
what is called the 'light cone', meaning the set of points satisfying
x 2 + y2 + Z2 = c 2 t 2 and so forming a three-dimensional 'surface' in four-space.
Since a Lorentz transformation preserves the length of a vector it carries a
point described by a vector einto another point e' with the same A, i.e. on the
same hyperbola. (One imagines a family of hyperbolae corresponding to
different A.) Thus the space-like, time-like, null and zero regions are invariant
under Lorentz transformations. Furthermore all six regions remain invariant
under the proper Lorentz transformations defined earlier. These do not permit
transformation between future and past regions (see problem 15.4). (NB The
two-dimensional diagram 15.1 suggests that proper Lorentz transformations
cannot connect regions with positive and negative z but this can be achieved in
four dimensions with the help of a rotation about one of the other space axes.)
One often refers to pictures like figure 15.1 as 'Minkowski diagrams'. An
individual at rest with respect to the space axes x, y, z would move parallel to
the time axis in the Minkowski diagram while a speed of v relative to the space
axes results in a path making an angle tan - 1 (v/c) with the time axis. Starting
from the origin , with speeds less than that of light, one can reach points in the
future time-like region only.

15.2.3 Physical interpretation of the lorentz transformation

The physical interpretation of the transformation R(a) is clear and easy to


15 .2.3 Space and Time 341

visualise-it is a rotation. On the other hand we have introduced the boost


Q(b) simply as a mathematical transformation in space- time . We now explain
the physical meaning and significance of the boost. For this purpose it is
sufficient to study the boost (15.25) in the z-d irection and we may use figure
15.1.
Consider an experiment relating to events occurring in a system at rest with
respect to a frame of reference. We denote the points at which these events
occur by vectors 2 = (x, y, z; er) in this frame . Suppose that the experiment is
repeated with the system now moving, relative to the frame, with velocity - v
in the z-direction. Then instinctively, using pre-relativity ideas, one would
expect the events to occur at the same times t but that their z-coordinates
would have shifted by amounts - ot. The later the event, the more it would
have shifted in the intervening time t. In other words we might have expected
the events in this second experiment to occur at points 2 = (x, y, z', et'), where
z' = z - ot, r' = r (15.28)

However when the velocity v is comparable with c, the velocity of light,


experiment disagrees with these instinctive conclusions-see Muirhead (1973)
in the bibliography. The essence of relativity theory, and this agrees with
experiment, is the statement that e' is obtained from e by the Lorentz
transformation (15.25), where f3 = vic, rather than by (15.28). Notice however
that if VI~ c, a condition which is satisfied in our everyday experiences because
of the large value of c = 2.998( 10)8m /sec, then the Lorentz transformation
(15.25) reduces to (15.28). It is not therefore too surprising that instinct has
failed us. There are a number of startling conclusions from relativity theory
which we now bring out by studying three particular cases of the general
experiment described above. (We should remark that we could equally well
have described the events It as the events eas seen by an ob server moving with
velocity v in the z-direction, The two viewpoints, of the moving system or the
moving observer, are both acceptable. We choose to use the former (active)
description since it follows our treatment of rotations in which the system is
rotated rather than the frame of reference. Most relat ivity books use the latter
(pass ive) description and we have made our formulae agree with them by using
- v rather than v for the velocity of the system.)

(1) Time dilation and the decay of a meson


Consider a meson being produced at the origin x = y = z = r = 0 at rest
and decaying a time r later, i.e. at € = (0,0,0, f). Th is defines a two-event
exper iment at rest. Now , in the way described generally above, we consider a
second experiment in which the meson is produced at the origin with velocity
- v. According to relativity theory it will decay at the point i' given from
equation (15.25) by x' = y' = 0, z' = - Vf(1- f32)-!, t' = f(1- f32)-!. In other
words the life of the meson, as measured in the fixed frame, has increased by a
factor (1 - f32)-!. This is referred to as the dilation of time and implies that time
342 Space and Time 15.2.3

is not absolute, as one assumed in pre-relativity days. Were it not for this effect
the mesons produced by collisions in the upper atmosphere would have
decayed long before reaching the surface of the earth. They have speeds close
to e so that the factor (1 - fJ 2) - t is large. Notice however that the 'interval'
between the two events, birth and death of the meson, is unaffected by the
velocity .

(2) Contraction of rods


The argument in thi s problem is simplified if we first rearrange equation
(15.25) by eliminating I to obtain
z' = - Bet' + z(1 - fJ 2)1 (15.29)
Now consider the difference between the space coordinates of two events at the
same time 1': ZB - ZA = (Z8 - ZA)(I - fJ 2)!. Physically, the same I' means taking a
picture of the moving system. Thus if A and B are events marking the ends of a
rod then we have shown that the mov ing rod will have its length contracted by
the factor (I - fJ2 )1, as recorded by the instantaneous picture. We stress that, in
our active picture, mea surements of position z' and time I' are made in the
same , fixed, frame.

(3) Relative velocities and the invariance of the velocity of light


As our th ird illustration of the consequences of the Lorentz transformation,
consider an object moving with uniform velocit y - V so that its path in space
time is given bye = (0,0, - VI, et). Now repeat this experiment but where the
object has velocity - V with respect to the system mo ving with velocity - v
relative to the fixed frame . The new path is given from the equations (15.25) by
z' = -(V+ v)t(1 - fJ 2)- 1
ct' = (e+ vV/e)t(l-fJ 2)- t
so that z'[t' = - (V+ v)/( I + vV/e 2 )

The object therefore mo ves with a constant speed V = (V + v)/( I + v V/e 2)


with respect to the fixed frame . Thus relative velocities are not simply
additive. Furthermore we see that if V = c, then V' = c, which
shows that the velocity of light is the same with respect to all frames of
reference. This remarkable result, observed by Michelson and Morley, was one
of the earliest experiments which led to the development of relativity theory
and is the reason why the velocity of light appears in the Lorentz transfor-
mation. One could have defined a Lorentz transformation with any other
number for e but it would not have agreed with experiment. It follows from the
expression for V' and if Vand v are both less than c, then so also is V'. In other
words one cannot exceed the velocity of light by superimposing velocities, no
matter how close Vand v are to e. When one studies relativistic dynamics one
finds that an infinite amount of energy would be required to accelerate a
15 .2.4 Space and Time 343

particle to precisely the speed of light. Experimental results in high energy


particle accelerators are completely consistent with this conclusion. In fact c is
an upper limit to the speed of a particle and one sees that the Lorentz
transformation has a singularity when v = c. There is nevertheless some
speculation as to whether it is possible to have objects moving with v > c,
which remain on the other side of this singularity. The name 'tachyon' has been
given to such particles and some of their properties have been discussed
theoretically, but so far no evidence has been found for their existence-see
Goldhaber and Smith (1975) in the bibliography to chapter 12.
To conclude this section we stress that the statement of Lorentz invariance
implies that physical laws are the same in frames moving with uniform relative
velocity. The surprise was that the transformation relating motion in two such
frames was not what we expected . The term 'Galilean invariance' is used to
describe the same statement but when the intuitive transformation (15.28) is
used to relate the two frames . Consequently Galilean invar iance is valid only
for v ~ c.

15.2.4 Infinitesimal operators


The Lorentz group !e is a Lie group and we have seen in equation (15.27)
that the general element L is defined by six parameters. This number may also
be deduced from the fact that a real 4 x 4 matrix contains 16 parameters and
the relation (15.23) between two symmetric 4 x 4 matrices imposes 10
constraints. To find the infinitesimal matrices we let a and b become small in
equation (15.27). Taking the infinitesimal form of R(il) from subsection 7.4.1
and that ofQ(b) from equation (15.26), keeping only first order terms in b, we
have L = 1 + A where the infinitesimal matrix A is given by

~x -~.<
Q
y
-b<)
=bOb: = q ;~y.z(QqXq+bqYq) (15.30)
-by -b z

where bq are the three components of the boost vector b, the matrices X, are as
defined in equation (7.24) and the matrices Y q are given by

YX=( -1
~ ~ ~ -~)'YY=(~ ~ ~ -!),
0 0 0 0 -1 0 0

y,~O ~ J-0 (15.31)


344 Space and Time 15 .2.5

These three matrices describe infin itesimal boosts in the three space directions.
The commutation relations of the infinitesimal operators follow from the
matrices (15.31)
[X.~,x.j = x. [x; x.] = X.~ [x., x .j = x,
[Y,, Y y ] = - X, [Y y , Y,'] = - X.~ [Y. , Y,] = -x y

[X Y,'] = 0 [X )" Y y ]=0 [X: , YJ = 0 (15.32)


[X Y,'] = Y, [Xy , Yz ] = v, [X., y~] = v,
[Xj, Y.~] = - Y, [Xz, Y y ] = - Y, [X.. . Y,'] = - Y,
We notice that the linear combinations
Aq = t (X q + iY q ) and B, = t (X q - iY q ) (15.33)
have much simpler commutation relations
[A x, A,'] = Az [A y , A,'] = Ax [A,, A.~] = A y
(15.34)
[B.~ , By] = Bz [By, a .j = -B.~ [B z ' B~] = By

together with [A q , B",] = 0 for all q and q', In other words if we regard the A"
and B, as the infinitesimal operators the y will describe a product group
because all A q commute with all Bq • Furthermore the commutation relations
(15.34) of the A q and B, separately are precisely those of the group 9i 3 , see
equation (7.25). Thus the A q , B, describe the infinitesimal operators of the
product group fJt 3 x 9i 3 • This does not mean that !e and 9i 3 x 9i 3 are
isomorphic because in !e the X, and Y, appear with real coefficients, while in
9i 3 x 9i 3 it is the A q and B, which have real coefficients. However it does
enable us to deduce and label the irreducible representation of!e from those of
9i 3 x fJt 3 because the arguments are based on the algebra of the infinitesimal
operators as in section 7.4.

15.2.5 The irreducible representations

We now move away from the basic four-space of the Lorentz group to study
representations in more general vector spaces. We shall however use the
same notation for the infinitesimal operators, so that whereas in the preceding
section A q and B, have denoted 4 x 4 matrices, in the following they will
denote the corresponding operators in some vector space which provides the
basis of the representation . We know from the general arguments ofsection 7.1
that the commutators of the infinitesimal operators are the same in all
representations and may therefore be taken from equation (15.32).
Since the structure of the irr ed ucible representations O (jJ of fJt 3 was deduced
in subsection 7.4.2 entirely from the commutation relations and since the
irreducible representations of a product group are labelled by a pair of labels,
one for an irreducible representation of each group, it follows that the finite
15 .2.5 Space and Time 345

dimensional irreducible representations of the Lorentz group may be labelled


by a pair ofindicesj andj' which take values 0, t, 1,1, . .. , etc., as for ~3 ' These
representations, for which we use the notation Lu./l have dimension
(2j + 1)(2j' + 1). It may be shown that the representations with j and j' both
integers and those with j and j' both half-integers are single valued, but those
with one index an integer and the other a half-integer are double valued.
The two ind ices may be interpreted in terms of the eigenvalues of the two
Casimir operators (see section 7.5) A 2 = A~ + A; + A; and B2 = B~ + B; + B;.
Using equation (7.36) and remembering the factor i introduced just before
equation (7.26) we have eigenvalues - jU + 1) and - j'U' + 1) for A 2 and B 2 .
Alternatively, and for convenience later, we may use the two combinations
given below
X2 - yZ = 2(A 2 + B2 ) with eigenvalue - 2 UU + 1) + j'V + I)}
X .Y = - i(A 2 - B2 ) with eigenvalue iUU + 1) - j'V + I)} (15.35)
The basis vectors of the representation Lu./l may be denoted by ei,,!~ , where, as
in the group :Jt 3 , the lower indices m=j, U-l) , ... .-j and m'=j',
V' - 1), . . . , - j' denote the eigenvalues of A: and B:. Precisely, quoting from
subsection 7.3.5 we have
(15.36)
The 4 x 4 Lorentz matrices L must themselves form a single-valued four-
dimensional representation which can only be ut
t). To verify this we need
only calculate from (15.31) and (7.24) the matrices

_H) (~! H)
o
x'- y' ~(-~ -2
o = -3
o o 0 000 3
and X .y =0
which agrees with the formulae (15.35) when j = j' = 1. One may even go
further and deduce the matrices of the operators X, and Y q knowing the
matrices (see equations (7.40) and (7.39)) of the infinitesimal operators of ~ 3 '
Thus for example the matrices (15.31) may be deduced from the OH) matrices
by a suitable change of basis (see problem 15.5). In particular it is worth noting
that X, = A z + B, so that from equation (15.36) the basis vector of LU' j')
labelled by m, m' will be an eigenvector of X, with eigenvalue - i(m + m') and
will transform according to the representation TIm + m'l of the group ~ 2 of
physical rotations about the z-axis.
The reduction of products of two irreducible representations of!f' is given
from the corresponding rule for the rotation group. The result is

LU,J.lx LU,.i') = t UJ·J ') (15.37)


s.r
346 Space and Time 15 .2.5

where VI - j 2/ s J s U. +j 2) and V'I - j~1 s J ' ~ VI +j~).


The physical rotation group Yt3 is a subgroup of the Lorentz group !l' and
hence Viol ) must reduce into a sum of irreducible representations D (l ) on
restriction from !l' to Yt3' The rule is

V M )= t D<Jl (15.38)
[;- /1 ~ J ~ (j+n

which may be derived simply in two steps . First we note from equation (15.37)
that

But in the representation LU' 0) we must have B, = 0 so that X, = Aq • Hence


for rotations Vi, 0 )= D(j). In the same way L(O,n = DU'), so far as rotations are
concerned, and this gives the desired result (15.38) using equation (7.44). As an
example of this result we have

which is to be expected since the 4 x 4 Lorentz matrices them selves give the
representation LIB) and the four-space breaks up into the three -space
components, which transform like 0(1), and the time component which is
rotation invariant.
We should also notice that whereas for the group Yt3 the complex conjugate
of a representation D U I is equivalent to DUI with a change of basis
e: = ( - 1)i- m e _m in the notation of subsect ion 7.4.2 (see also section 7.7), the
complex conj ugate of V i ,n is equivalent to Ui' ,i). a distinct representation,
Thus U M ) is equivalent to its complex conjugate only when j = j'. The
justification for this statement comes directly from equation (15.33) but we
must be careful to distinguish the operator Aq in the complex conjugate
representation which we denote by (Aq )* from the complex conjugate A: of the
operator Aq • Now, in the complex conjugate representation, X, and Y, are
replaced by X: and Y: so that

Since the representations D U) of fJi3 are self-conjugate this implies an


interchange of j and j'.
Notice also that the representations U M ) are not unitary. To be unitary, the
infinitesimal operators must be skew-Hermitian, i.e. have pure imaginary
eigenvalues. But we have seen, in discussing 91 3 in section 7.4, that Aq and B,
are skew-Hermitian so that from equation (15.33)Xq is skew-Hermitian but Y,
is Hermitian, Thus no matter what basis we choose, these finite-dimensional
irreducible representations of the Lorentz group are not unitary, The identity
representation U o.O) = 1 is naturally an exception to th is result.
15.3 Space and Time 347

15.3 The Lorentz group with space inversions!£ s

The addition of the space inversion I to the rotation group fJt 3 was a trivial
process. The inversion commutes with rotations, leading to the product group
0 3 = fJt 3 X S 2 (see section 7.4) and the irreducible representations of 0 3 are
labelled by the pair of labels j, n, where n = ± is the parity.
The addition of the space inversion I to the Lorentz group ff is not so simple
because I does not commute with the Lorentz transformations. However from
the separation (15.27) of an arbitrary L into a rotation R(a) and a boost Q(b)
we need only concern ourselves with the products of 1 with boosts. For the
particular boost Qz(b) one sees from the matrix (15.24) and the matrix

I =
(-~ -~ ~ ~)
o
o
0
0
-I
0
0
I
(15.39)

that IQz (b) = Q, ( - b)l, and it is clear that this will be true for any direction in
place of z. Thus lQ(b) = Q( - b)1 or in other words IQ(b)1 = Q( - b) showing
that 1 simply changes the direction of the boost, a result which might have been
expected. In terms of infinitesimals this means that
IXq = Xql, lY q = -Yql (15.40)
so that
Aqi = IB q and Bql = IAq (l5.4la)
which implies that
A2 I = IB 2 and B21 = IA 2 (15.4lb)
The structure and labelling of the irreducible representations of the group ff s
may now be deduced quite simply from these multiplication laws which are of
course valid not only in the 4 x 4 matrices but in all representations. As in the
preceding section, we still use the same notation A q , etc., to denote both the
infinitesimal 4 x 4 matrices and the corresponding operators in a general
representation. However, in line with the notation of section 4.1, we use the
notation T(I) for the transformations induced by the inversion operator I in a
general representation space.
Let us denote the basis vectors of an irreducible representation Lu.n of ff
r.
bye-:t.'m' with m = j,U -I), .. . , -j, m'= U' -I), .. ., -j'. Then using (15.41)

A2[T(l)e~~ ,] = T(l)B2e~~ , = -j'U' + l)[T(I)e!!~ ,]


while
348 Space and Time 15 .3

and similarly, using equations (15.36) and (15.41a)


A:CT(I)e:,{~ ,] = -im '[T(I)e:,{~ ,] (15.42)
while
Bz[T(I)e~~.] = -im[T(I)e~~ ,]
From the set of equations (15.42) one concludes that the set of (2j + 1)(2j' + 1)
vectors T(I)e:,{~, form the basis ofa representation U /)) of!i' and in fact we use
the notation
!~!m = T(I)e:,{~ (15.43)
Furthermore, since 12 = E, the identity, we have
e:,{~ , = T(I)f~!m

'*
Now so long asj' t . the vectors e:,{~, and!~~, must be linearly independent
and so, from the above equations, the set of 2(2j + 1)(2j'.+ 1) vectors e:,{~, and
r: with fixed j and j' is invariant under all the operators A q , B, and I of the
group !f's ' Hence they form an irreducible representation which we shall
denote by L~j.j'), although the order ofj and j' is irrelevant in this case . Clearly,
on restriction to the subgroup !f' we have the reduction
(15.44)
From equation (15.43) we see that the basis vectors e andfdo not have definite
parity. This may be achieved only by taking the combinations
(15.45)
which have even or odd parity depending on the sign ±. These combinations,
although not eigenfunctions of A2 or B 2, are nevertheless eigenfunctions of the
operator X 2 - y2 defined in equation (15.35).
A separate discussion is required for the special case of j = i ' since the
vectors e!.!m' and!!.!m are not then necessarily linearly independent. In fact we
shall show that the representation generated from e!.!m' is not irreducible if they
are linearly independent. This is done by using a reductio ad absurdum
argument. Suppose that !!.!m' and e!.!m' were independent, then the irreducible
representation which contains e!.!m' must also contain the basis vectors!!.!m' and
so have dimension 2(2j + 1)2. But since
T(I)(e:,{m ' +f:,{m ') = n: + e:,{'m
and as in equation (15.42), both e andftransform identically under A q and Bq ,
it follows that the subspace of (2j + 1)2 basis vectors (e:,{m ' +f:,{m ') is inva riant
under all the group operations. Thus the representation will have reduced,
contrary to the hypothesis that it is irreducible. The only escape from this
contradiction is thatf:,{m' = ce:,{m" for some constant c and since 12 = E then
c 2 = 1 and c = ± 1. Thus, for each i. there are two possible (2j + 1)2_
dimensional irreducible representations of !f's which we denote by L(j~ ) and
15.4 Space and Time 349

L(j- ) which are both identical with UM) for proper Lorentz transformation
and have the respective properties
for Ui +) (15.46)
and
Notice that the ± sign on U i± ) does not mean that all the basis vectors have
even or odd parit y. The equations (15.46)show tha t T(I) is not diagonal in the
mm'-basis. For example , the four-dimensional representation of the matrices L
is the Llt- ) representation. The basis y,ex' e e:,e,
in which it was set up actually
has T(I ) in diagonal form with three odd-parity vectors and one ofeven parity.
From equation (15.46) the matrix of T(l) in the mm'-basis would be

J)
- I o o
o -1
T(l) ~ ~ - 1 o
( o o

15.4 Translations and the Poincare group fJ'

The pre vious two sections have been concerned with the (homogeneous)
Lorentz transformations, which leave the origin unmoved. We now consider
the inclusion of space - time translations with the Lorentz group which
generates the Poincare group f!J or, as it is sometimes called, the in-
homogeneous Lorentz group. We begin with a short section on the trans-
lations themselves before moving on to the Poincare group. The inclusion of
the improper elements of space and time inversion is left for later sections, as is
the discussion of the physical significance of the different representations of
the Poincare group.

15.4.1 Translations in space-time

We define translations in space - time by simply extending the arguments of


subsection 15.1.1 to a fourth dimension. The translation operator is written
pee), where e = (Ex . Ey , Ez' s.) is a four-vector parameter, and is defined by the
equation
(15.47)
Notice that the time translation adds a quantity E, to the value of ct,
corresponding to a change E,/C in t. As usual, the infinitesimal operators
P,, Py , P,, P, are defined by writing, for small e,
pet) = 1 +ExPx +EyPy +EzPz -E,P, = 1-~.P (15.48)
The choice of a negative sign in this definition of P, means that it denotes a
350 Space and Time 15.4 .1

negative time translation, whereas the space components denote positive


translations as in section 15.1. The reason for this apparently perverse
definition is to make P x ' P y ' P z and P, the components of a four-vector P. They
then transform like the components of any other four-vector when going from
one frame to another, see problem 15.6, and we are able to use the scalar
product in equation (15.48). (Had we not introduced this negative sign we
would have needed to distinguish between covariant and contravariant
vectors, see the note in subsection 15.2.1.) The four-d imensional translation
group is an Abelian product group just as in three dimensions so that the
irreducible representations are one-dimensional and labelled by four numbers
kx ' k y , k z and k,. By defining the four-vector k = (k x , k y , k z , k,) we may write the
representation as TIkI, and if we denote the basis vector by Ik ) then the
representation is given by
T(f:) II ) = exp(iI. e) II ) (15.49)
For the three-space translations this is consistent with subsection 15.1.1 and is
the defining equation for the parameters I.
Comparing the preceding two equations shows that in T (1 )the infinitesimal
=
operators have the values P -ik. If we consider the explicit form of the
infinitesimal operators on a function l/J (e ) of position e then we have for small
e, following subsection 7.3.5,

T(e)l/J(e) = l/J(e-e) = l/J(e) - ( L>q -a + 1:,-;\ iJ- ) l/J(e) (15.50)


q (ieq u(et )
so iilal " u rn equation (15.48),
D iu
Pz = - - and P = - - (15.51)
2z ' c iJt
As an example of a typical function transforming like T(k ) we choose
l/J(e) = exp( - .t .e) = exp(ik . r - ik,ct) (15.52)
where k and k, denote the three-space part and fourth component of I,
respectively.
In discussing the physical significance of the space - time translations we
need only consider the time translation since we have already seen in
subsection 15.1.1 how translation invariance in space leads to momentum
conservation. Invariance under time translation means that the infinitesimal
operator P, must also represent a conserved quantity, which we call energy.
Again this operator is skew-Hermitian and has inappropriate dimension. The
actual energy operator is written H = ineP, . The invariance of the Hamil-
tonian under time translations means that it does not depend explicitly on the
time. For a Hamiltonian which is invariant under translations in both space and
time one expects to find states labelled by I corresponding to momentum hk ,
as explained in subsection 15.1.1, and energy inc( - ik,) = hck, which is usually
denoted by E. For a system with time translation symmetry only we retain only
15.4.2 Space and Time 351

the label k, or E. In fact the familiar expression (5.3) for a stationary state
simply reflects the fact that under time translations it transforms according to
the representation labelled by k, = E[hc, We note that the Schrodinger
equation (5.2) is obtained by equating the explicit form in otC/ for the energy
operator, deduced from P, to so me assumed form such as T + V for the energy
operator H. Notice that since P is a four-vector operator it follows that the
energy and momentum operators come together into a four-vector operator
(P." P,., PZ' H ie) with eigenvalues n(kx ' k)", k:, k,). In classical mechanics also ,
one sees that the lack of any explicit time dependence in the Hamiltonian leads
to the conservation of energy (see subsection 16.1.2).
In principle, k, may take either sign which would imply both positive and
negative energies. There are difficulties in assigning physical significance to
sta tes with negative energies; for example, why should the positive energy
states not decay into them'? The postulate that all negative energy states are
occupied has been made but was never sat isfactory. The modern interpretation
is to postulate that all free particles have positive energy since this is consistent
with observation. Nevertheless the representations corresponding to negative
k, have a role to play when one con structs a quantum field. We discuss this role
briefly in chapter 16.

15.4.2 The Poincare group and its representations

The Poincare group ;!J' is obtained by including both Lorentz transfor-


mations L and translations P(e) and their products. These two kinds of
transformation do not commute but they have a simple multiplication law. To
find it, consider the two products
P(t)Le = Le+l
and
LP(t)e = L(e+t) = u+ U'
from which one concludes that
LP(f) = PILe) L (15.53)
The most general Poincare transformation may therefore be written
e' = P(e)U = U+ 6 (15.54)

with the Lorentz transformation acting first, followed by the translation. The
group f!I! is very similar to the Euclidean group of section 15.1 except that there
are now four dimensions and the rotations have been replaced by Lorentz
transformations. There are now ten parameters, six from L and four from P{t) .
We shall use the notation T(t, L) to denote the transformation induced in
representation space by the Poincare transformation P(t)L and, as usual , the
operations T(t, L) will satisfy the same multiplication rules as the correspond-
ing group operations.
352 Space and Time 15 .4.2

We now construct the irreducible representations of the group 9 following


closely the four steps described in subsection 15.1.3 for the Euclidean group.

Step (1). We choose the basis vectors Ik) to belong to irreducible


representations of the translation subgroup and thus to satisfy equation
(15.49).
Step (2). Next we argue that iin irreduci~le rePlesentation containing Ik)
must also contain basis vectors Ik' ) , where k ' = Lk for all L. In other words k '
has the same 'length' as k and lies in the same region of space - time in the sense
defined in subsections 15.2.1 and 15.2.2. To justify this statement consider

T(~) T(L)lk ) = T(L) T(L - I ~)Ik)


= T(L)exp(ik . L -1~)lk >
= exp(iLk . ~) T( L) l k > (15.55)

using equations (15.53), (15.49) and (15.22). Thus if Ik >is a basis vector then
>,
T(L)lk which must also belong to the representation, transforms according
to the representation of the translation group labelled by k' = Lk, and the
assertion above is proved.
The irreducible representations of f!} may therefore be first classified
according to the region of four -space in which the four-vector k lies and as we
saw in subsection 15.2.2 there are six regions . The structure of the repre-
sentation is quite different for each region so that the remaining two steps must
be discussed separately for each region.

Region (6), k= 0
°
This case may be quickly disposed of since with k = all basis vectors must
be invariant under translations and the group is effectivelyrestricted from f!} to
Y . The irreducible representations are then simply the Uj.j') described in
subsection 15.2.5.

Regions (2) and (3), time -like k (the representations plk.s»)


Step (3). To construct the irreducible representations in this case we choose,
for convenience, the simple time-like vector ko= (0,0,0, k) of length k. To be
definite, we take only k > 0, the future time-like case, corresponding to region
(2).Identical results follow for the past time-like case. Since k o is unchanged by
any rotation R(a) it follows from equation (15.55) that the vector T(R(a))lko >
transforms like Iko ) under translations. Hence the set of all basis vectors
which transform like Iko >in a representation forms an invariant space under
Yi 3' so that they may be simultaneously labelled by ko and by an irreducible
representation label of the group 913 , We use the symbol s( = 0,1, 1,1, .. .) for
the Yi 3 representation label and ms = s, s - 1, . . . , - s to label the Yi 3 basis
vector (see subsection 7.4.2), so that we henceforth write Ikosms in place of >
15 .4.2 Space and Time 353

Ik o ) . In the notation of subsection 7.4.2 there will then be a set of(2s + 1) basis
vectors with the property
T(R(a))lkosm s) = L. Db)mJm, (a) Ikosm~ ) (15.56)
m,
as well as the property (15.49).
Step (4). Starting from these (2s + 1) basis vectors we now argue that the
representation is generated by constructing all vectors of the kind
Ik sms ) = T(Q[h(k)] )Ikosms> (15.57)
where k isany four-vector in the same region as ko and having the same 'length'
and h(k) describes that unique pure Lorentz transformation which carries k o
into k, i.e.
(15.58)
One sees from equation (15.26) that the three-vector h(k) has magnitude
cosh - 1 (k,/k) and direction - k, where k denotes the three-vector part of k and
k, its fourth component.
To show that the set (15.57) with fixed t;
and s forms the basis of a
representation we consider the effect of an arbitrary group operation. In doing
this we shall need to use the result (15.27) that an arbitrary Lorentz
transformation may be written as a product of a pure Lorentz transformation
and a rotation. We have
T(e,L)lksm s > = T(e)T(L)T (Q[h(k)] )Ikosms> using equation (15.57)
= T(f)T(L')lkosms > where L' = LQ[h(k)]
= T(f)T(Q')T(R')lkosms) using equation (15.27)

= T(6)T(Q ')L:D~!m
m S J
(R ')lko sm~ ) using equation (15.56)
J

= T(~LD(S)m,m.• (R ')Ik'sm~ > using equation (15.57) with k' = Q'ko


m;
= exp(ik' . e) L D(S~mImI (R')Ik' sm~ > using equation (15.49) . (15.59)
m;
Notice that the vector k' is given simply by k' = Lk because
k ' = Q'k o = L'(R '( k o = L'k o = LQ[h(k)]k o = Lk
1

but this result was to be expected from equation (15.55). The parameters of the
rotation R' may be deduced for any given k and L from the defining equations
L' =Q'R ' (15.60)
and
L' = LQ[h(k)] (15.61)
but we shall not go into any details.
354 Space and Time 15.4.2

The equation (15.59) shows that the set (15.57) of basis vectors is invariant
under the general Poincare transformation, and so forms the basis for a
representation which we might denote by plto.S). It may be shown, as for the
Euclidean group in subsection 15.1.3, that representations generated from any
future time-l ike vecto r k with the same ' length' as ko are equivalent to plto..').
The inequivalent representations are therefore denoted simply by pCk.s) where
the index k denotes the 'length' of all the vecto rs k which occur in the basis
vectors Jksms ) ' The proof that the representations p tk . S) are irreducible
follows the argument referred to in subsection 15.1.3 for the group cff J . Because
of the infinite var iety of k with fixed length, these representations are infinite-
dimensional.

Regions (4) and (5), null k (the representations P(o. ml)


Following step (3), for convenience we again choose a simple vector ko' this
time in the future null region , namely ko = (0,0, I, I). (The past null region
may be treated in an identical way.) To deduce the irreducible representations
we follow the same general procedure as above, but the group which leaves ko
unchanged is no longer the group ~ J . Our first task is to find this group and
for this purpose we turn to the infinitesimal operators. If t; is to be unchanged
°
then Ak o = in the notation of equation (15.30). Carrying out the matrix
multiplications with ko = (0, 0, 1, 1) and the matrices (15.31) leads to the
conditions ax = - bp a y = bx ' b, = 0. The three independent infinitesimal
operators which satisfy these conditions may be taken as
Xl ' X, - Y y and X, + Yx (15.62)
Their commutation relations are , from equations (15.32),

and
[X, - Y)" X}' + Y,] = °
The operator X, generates the group ~ 2 of rotations about the z-axis and we
now show that the commutation relations (15.63) are precisely those of the
Euclidean group Iff2 which describes rotations and translations in a plane.
From equation (15.15) and since translations commute with each other, we
deduce the following commutation relations for the infinitesimal operators
P " P, and X, of Iff2'
(15.64)
These are identical with equations (15.63) if we make the following associ-
ations: X, --+ Xl' X, - y y --+ P x' X, + Yx -+ p}" Thus the subgroup that we are
seeking, which leaves ko invariant, is isomorphic with Iff2 so far as infinitesimals
are concerned, and hence has the same irreducible representations. Physically,
the two groups are quite different although they do have one operator X, in
common.
15.4.2 Space and Time 355

In subsection 15.1.4 we deduced that the irreducible representations of C 2


could be labelled by ElIkl) and E(O.m). We shall not pursue the consequences of
choosing E(lk !J for the Iff2 group representation since as we shall see later in
subsection 15.8.3 this has, as yet , no physical application. It would be relevant
in describing a particle with continuous spin but no such particle has been
observed. We therefore move directly to the other alternative E(O.m) . The fact
that ko is invariant under C 2 means that the set of basis vectors transforming
like ko under translation is also invariant under Iff2' They may therefore be
labelled both by t; and the representation E(O .m) of Iff2 ' so we use the notation
>.
Ikom From the definition of E(O,m) in subsection 15.1.4 these basis vectors
must have the properties
T(Rz(a) )Ikom > = exp( - ima)lkom >
(X, - Yy)l~om >= 0
(X y + Yx)lkom> =0 (15.65)
as well as (15.49). The last two equations express the 'translation invariance' of
the vectors of EIO .m) for 'translations' in the Iff2 group.
In constructing the representation we shall need two more properties of
Lorentz transformations. Firstly that there is a unique Lorentz transformation
of the kind R,yQz which carries t; into any given future null Ii. The role of the
boost Qz is to change the magnitudes of the space and time components of ko
without affecting the space direction while the rotation R,y about an axis in the
xy-plane simply c~anges this direction. It is not difficult to see that R.,y and Q,
are unique, given k. The second property says that any Lorentz transformation
may be expressed in the form

L = a., o, L (15.66)
where L belongs to the Iff2 subgroup which leaves t;
= (0,0, 1, 1) invariant. To
show this result, which is analogous to equation (15.27), we define k = Lk o and
define R,yQz by k = R,yQzko. Then we see that
(R,yQz)-1 Lk o = (R,yQz)-1 k = ko
showing that the product (R,yQz)-1 L leaves ko invariant and therefore that L
in equation (15.66) belongs to the C 2 subgroup.
Following step (4), we now construct the representation of [Jj! corresponding
to null k by generating the set of basis vectors
A Ikm> = T(RxyQz) Ikom > (15.67)
with any future null k. To show that this set provides a representation we
consider the effect of a general Poincare operation
T(f, L)lkm>= T(f)T(L)T(RxyQz)lkom > from equation (15.67)
= T(f)T(L')jkom > defining L' = LR,yQz
= T(f)T(R~yQ ~)T(L')lkom > using equation (15.66)
356 Space and Time 15.4.3

= T(f)T(R~yQ ~)exp( - ima)lkom > from equation (15.65), defining a


= T(f)exp( - ima)lk'm > using equation (15.67) with k' = R~yQ~ko
=exp(ik'.f-ima)!k'm > from equation Il Svts), (15.68)

One quickly deduces that k' is given simply by k' = L k but the value of the q( 2
rotation angle a must be deduced in detail from its defining equations. L' is
known from k and L and the factorisation (15.66) of L' then gives L', an
element of ,ff 2 ' But elements of S 2 are products of a 'translation' with an q( 2
rotation. The 'translation' is irrelevant here since Ikom > is 'translation'
invariant, see equation (15.65).
The equation (15.68) shows that the set (15.67) of basis vectors forms a
representation, which we denote by pIO .m). It may be shown, as indicated
before, that this representation is irreducible and that the representation
generated from an arbitrary future null vector k o is equivalent to the pIO ,m)
deduced above for the simple choice k o = (0,0, 1, 1). Remember that the index
m, which arose from the group q( 2 may take the usual values
m = O,±!-,± 1, . . " etc.

Region (1) space-like Ii


The representations of this type may be deduced using the same technique.
However we shall not describe them here since, as we shall see in subsection
15.7.1, they would seem to ha ve ph ysical application only for particles with
imaginary mass !

15.4.3 Casimir operators

In other continuous groups we have seen the usefulness of the Casimir


operator which was introduced in section 7.5. It is constructed from the
infinitesimal operators and commutes with them so that in an irreducible
representation it is a simple multiple of the unit operator. The multiple relates
directly to the labels of the irreducible representation. The two Casimir
operators for the Lorentz group 51! were given in equation (15.35). We now
construct the Casimir operators for the Poincare group. We have at our
disposal the six operators Xq , Y, of 51! and the four infinitesimal translation
operators P q and PI' where q = x, y, z and P, denotes the time translation. The
commutation relations of the X, and Y, are given in equation (15.32), the
translations commute among themselves and the commutation relations
between translations and Lorentz transformations may be deduced from
equation (15.53),see problem 15.7. The results are, in add ition to those given in
equation (15.15),
[P"Xq] =0, [Pq,Yq,] = -Oq,q'P" [P"Yq]= -Pq (15.69)
15.4.3 Space and Time 357

Since P is a four-vector it follows that the operator


2 2 2 2
P .P - -P,-Py-Pz+P,
A A_

is a Lorentz invariant and since it obviously commutes with the translations it


is a Poincare invariant. To deduce its value in a given irreducible repre-
sentation of f!J we note from subsection 15.4.1 that if Ik ) denotes any basis
vector transforming like T (k) under translations then
(15.70)
so that immediately
P.Plk ) =(k.~+k;'+k;-k,2)jk ) = -k2Ik ) (15.71)
t:
where k = (k. k) t is the 'length' of Hence, in the irreducible representation
plk . s) of the group f!J, the Casimir operator P.P will simply be a multiple - k 2
of the unit operator. In the representation pIO .m) the basis vectors all transform
like null k ~nder translation so that k = 0 and P.Phas the value zero , although
of course k "# O.
It is not possible to form a second Casimir operator by taking only quadratic
functions of the infinitesimal operators. This may be seen by noting that, under
Lorentz transformations, P transforms according to the representation U!·!l,
while X ± iY transform according to UI .O) and L (0. I) , respectively, see
subsection 15.2.5. The rule (15.37) for the reduction of product representations
tells us that the only quadratic Lorentz invariants are obtained by taking
products of anyone of these representations with itself, leading to P.P and
X 2 - y 2 ± 2iX. Y. We already have the first of these and no combination of the
others can be made translationally invariant. Another Casimir operator may
however be formed by first constructing a translationally invariant four-vector
Wand then forming the Lorentz invariant scalar product W .W. We have
already seen in subsection 15.1.5 that the operator P .X is invariant under
translations and rotations and therefore serves as the t-component of W. To
find the other components we note that, if Wis to be a four-vector operator
then its components must sat isfy the same commutation relations with X and
Y as those of P. Thus defining W, = P .X we find from equation (15.69), see
problem 15.8,
(15.72)
The structure of Wcould equally well have been found by using the yt 3 vector
coupling coefficients to produce a resultant Lit.!) from the product
L(Hl x U 1 . 0 ). The operator W.P would seem to provide a third Casimir
operator but one sees from the definition (15.72) that it vanishes identically
(15.73)
To find the value of W . Win eaeh of the irreducible representations of the
Poincare' group, it is sufficient to choose any basis vector (since Casimir
operators are a multiple of the unit operator in any irreducible representation).
358 Space and Time 15.4.3

Thus for p(l. s) we may choose the basis vector IkoSms ) corresponding to
ko = (0,0,0, k). Then Pqlkosms ) = 0 for q = x, y and z, while
P,lkosms) = - iklkosms ) ' Hence from equation (15.72)
W,lkosms ) = 0
and Wqlkosms ) = -ikXqlko sms )
so that
W. Wlkosms ) = k 2X 2lkosms ) = - k 2s(s + 1) Ikosms ) (15.74)
2
We have used the fact that X is just the Casimir operator for the .eJt 3 subgroup
whose irreducible representations were labelled by s. The value s(s + 1) for the
91 3 Casimir operator was deduced in section 7.4 and the negative sign comes
from the factor i introduced in section 7.4 when the Hermitian operators J,
were defined as J, = iXq. The result (15.74) is valid for all vectors belonging to
p(k ,s),

For the irreducible representations p(O.m) both Casimir operators P.P and
W.W have the value zero. The latter result follows by choosing the particular
basis vector with k o = (0,0, 1, I) whereupon
W,lkom ) = -i(X,-YY)lkom ) =0,
W y Ikom ) = -i(X y + Y,) Ikom ) = 0,
Wzlkom ) = -iXzlkom ) = -mlkom)
W,lkom ) = -iXzlkom) = -mlkom ) (15.75)
using equation (15.65) and hence W.WI~o m ) = O.
Since both Casimir operators are zero they cannot serve to distinguish
representations plO ,m) with different m. However, in a sense, it is possible to use
the ratio of W to P to distinguish between them. For the particular vector
jkom ) the equations (15.75) imply that, since Plkom ) = - iklkom ) ,
Wlkom ) = - imPlkom )
In other words we may write
(W + imP) Ikom ) = 0 (15.76)
However since all basis vectors in the representation are generated by Lorentz
transformations (15.67) from Ikom) and since both Wand Pare four-vectors,
transforming in the same way under Lorentz transformations, it follows that
the relation (15.76) is true not only for the special vector Ikom), but for all
Ikm) in the representation p(O ,m) (see problem 15.9). Thus the relation (15.76)
may be used , in addition to P.P = 0, to characterise the representation p(O ,m) .
It says that the four-vector operators Wand Pare in the same 'direction' with a
proportionality constant - im which depends on the representation. Notice
that the relation (15.76) came from a consideration of the future null vector
ko = (0,0, 1, 1) and it is therefore only applicable to the representation
15.4.4 Space and Time 359

t.
referring to future null For the representation relating to past null ii we must
choose (0,0, I, - 1) in place of (0,0,1 ,1). This slightly modifies equation
(15.76) and leads to the relati on
(W -imP) lii m > = 0 (15.77)
for all past null Ii.
The proportional ity that we have established between Wand P is not
unexpected since, ifW and P were ord inary four-vectors rather th an operators,
the relation s W.W= 1'.1' = 0 together with W.P = 0 impl y that
(W - .,;1') .( W- .,;1') = 0 for all x, so that by taking z = Wr/P, we conclude
that (W - '1:P) .( W - .,; P ) = 0 a nd hence that W = .,; P a nd ind eed W= '1:1' for
thi s value of '1: .

15.4.4 Definition of scalar product

The scalar product for two vectors belonging to an irreducible repre-


sentation of the Poincare group may be defined by an extension of the method
used in subsection 15.1.6 for the group Iff3' There is the same problem of the
continuum of bas is vectors with two sma ll complications. For the repre-
sentation ptk. s l and follow ing equation (15.19), we write an arbitrary vector as

I"' > =k f"'m,(ii)lii sms >k,-ldk (15.78)

Taken together, the sum and integral run over all basis vectors Iii sms > of p lk. s )
a nd the v; (ii) are the numbers which determine I'" >. To co ver all ii in four
dimension~ with fixed 'length' k , one needs a triple integral a nd there is
naturally some freedom in choosing the three integration va ria bles. It is
t.
convenient to choose the three components of th e space part k of From the
relation k 1 = - Ik 11 + k,1 a nd th e cond ition of future time-lik e ii one sees that
ii is determined by k and th at k mu st range over all three-space. The factor k,- 1
has been included in equation (15.78) for convenience because, as we show
below, the volume element i ; 1 dk is Lorentz invariant. One might have
omitted the factor k,- 1 , essentially absorbing it into the coefficients e m (ii).
The scala r product is now defined, following equat ion (15.20) as '

<4>1"'> =~, .r4>: ,(ii)"'m,lii)k,-ldk (15.79)

with the same volume element as above. The proof that the representation
p tk.s ) given by equation (15.59) is unitary then follow s as in subsection 15.1.6
but we shall not write it out. In that subsection we made use of the fact that if
k ' = Rk then the su rface element df2 k could be replaced by df2 k ,. The
analogous problem here is to tran sform from an integral dk to an integral dk ',
where k' is the space part of ii' = Lii . This requires calculation of the Jacobian
?W, k;. k~ ) / c(k , k)" k: ) but since we may write L = QR, from equation (15.27),
and the Jacobian for a rotation is unity, we need only consider a boost in an y
direct ion . If we choose the z-d irection then x and yare unchanged so that the
360 Space and Time 15 .5

Jacobian is simpl y ak ~ /akz ' From the explicit transformation (15.25) for such a
boost
(15.80)
Thus

but since we ha ve
giving
a k~ /a kz = (I-fJk z/k ,)/(I-fJ 2)1
= k,- Ik; using equation (15.80).
Thus the volume element dk ' may be replaced by k,- Ik;dk or , in other words,
k,- Idk ' may be replaced by k,- Idk. Essentially the same scala r product may be
used for the representations p to. m., except that the sum over m s is no longer
required and m, is replaced by m.
As we have seen in subsection 15.4.1 the vector k label s a sta te with definite
momentum and energy. Thus if th e numbers 1/1 m. (k) ar e normalised so that
<1/1 LI/I > = 1 we may interpret 11/1 m. tk W as the probability densi~y a s a function
A

of k and ms for given k. One so metimes refers to the 1/1 m (k) as the wave
function s in momentum space. We warn that in non-relati vistic treatments it is
usual to define momentum space wave function s normalised with the k,- I
factor missing from (15.79 ). Thi s is a matter of con venti on.
We note that, so far as the ba sis vecto rs Iksms >themselves a re concerned, th e
above defin ition of scala r product impl ies that

(15.81 )

where (j(k - k ') den ote s the three-dimensional delta-function. The function
1/1 m, (k) which describes one of the basis vecto rs itself, say 11/1 > = Ikrn s >is given
simply, see equation (15.78), by
I/Im.(k) = bm,rh.b(k -k)k, (15.82)

15.5 The Poincare group with space inversions f!J s

The addition of the space inver sion I to the Poincare group makes only a slight
change to the repre sentations. We now study this extension of the group,
following quite closel y the methods used in section 15.3. We have already
di scussed in sectio n 15.3 the fact that I co mmutes with rot ations but not with
Lorentz boosts and the mult iplication rule
IQ(b) = Q( -b)I (15.83 )
was deduced . The mult iplication rule with translations follows directly from
15.5 Space and Time 361

the definitions (15.39) and (15.47) of I and Pre),


IP(e) = P(le)1 (15.84)
The irreducible representations of 9 s may be readily constructed from those
of 9 . We show that to each representation plk. s) of 9 there are two
repre sentations plk, s)± of 9 s . Recall that plk. s) was generated from a special
?asis ~ector Iko s~ >.' wh~re k o =(0,0,0,1 ), using ~q uation (15.57). Since the
mversion I leaves k o invanant and a lso commutes with gp 3 we may choose thi s
special basis vector to have definite parity in addition to the labe ls ko , s a nd ms .
A

Thus we write Ikosms ± >,


where
T(I) lkosms ± >= ± Ikosms± > (15.85)
and it is understood that the upper (or lower) signs must be used throughout.
From each of these possibilities we can now generate an irreducible
representation of :?Js exactly as for the group :?J using equation (15.57) and
defining
(15.86)

The actual representation matrix for proper elements T (t, L) is given as before
by equation (15.59), while for the inversion
T(I) lksms ± ) = T(I)T(Q[b(k)]) 1k osms ± )
= T(Q[ -b(k)])T(I)jkosms± > using equation (15.83)
= ± T(Q[ -b(k)])lkosms± > using equation (15.85)
= ± Ilksms ± > (15.87)

In the last step we have used the fact that if b denotes the boost parameter
required to carry t; into k then - b will carry ko into Ik, see equation (15.26).
The important point about equation (15.87) is that T(I), acting on a general
basis vector Iks ms ± ), leads to another basis vector of the set (15.86). Notice
that T(I) is not a diagonal operator. Not surprisingly, T(I) changes the space
direction of k and , as we shall see later, in section 15.7, the ± refers to an
internal (or intrinsic) parity of the system .
The arguments used above are not applicable to the representations p(O.m),
since the special vector in this case, t; = (0,0, 1, 1), is not invariant under
inversions. In fact we shall find that the inversion operator T(I) transforms a
basis vector of the representation p (O.m) into a basis vector of P(o. -mI. It will
then follow that the representation space of p lO.m) cannot remain invariant
under the extended group :?Js but that, because T(I)T(I) = 1, the sum of the
vector spaces ofplO,m)and P(O, -m ) is invariant under e', and therefore provides
a representation of :?Js which must be irreducible. We denote it by p (o.lmll. To
justify these statements we shall need to use the relations
(15.88)
362 Space and Time 15 .6

which follow directly from equation (15.84) and which lead to the results
w.r = IW q , W,I = -IW" [W . W, I] = [P . P, I] = 0 (15.89)
Now if Ikm > belongs to the representation pI O.m ) of the group f!/ so that
equation (15.76) holds then we see that the vector defined by T(I)lkm > has the
following properties
W . WT(I )lkm > = 0
P .PT(l) lkm > =0
(W q- imPq)T(l ) lkm > = T(l )(Wq + imPq)lkm> = 0
(W,- imP, )T(l) lkm > = -T(l)(WI +imP,)lkm > =0
so that (W - imP)T(l)lkm > = ° (15.90)
Furthermore,
P(f)T(I) lkm > = T(l) P(lf) Ikm >
= exp(ik . If)T(I )lkm >
= exp(ilk . t )T(I) lkm >

so that the vecto r T(l) lkm > transforms under translations according to the
representation TOk) of the translation group. Since by definition, k is future
null , so is Ik and thus, comparing equation (15.90) with equation (15.76), we
conclude that T(l)lkm > belongs to a representation pl O. - m) of f!/ . Thus, unless
m = 0, the vectors Ikm > and T(I)lkm > are linearly independent. We may thus
write T(l) Ikom > = IIko- m > so that the set of Ikm> with Ik - m >, generated
from equation (15.67) for all future null k, provide the basis for the
representation p (O.lmp of the extended group f!/s ' It follows generally that
T(l ) lkm > = Il k -m ) (15 .91)

and hence T(I) Ik - m ) = IIkm ). Together with equation (15.68), used within
each of the representations p I O.m ) and p lO. - m) for group elements T(f, L), the

°
representation p (o.lml ) of f!/s is thus determined.
The case m = is an exception since T(l)lkm ) and Ik - m ) both transform
according to the same representation pIO.O) of f!/ . Then, as with the
representations plk. S), one obtains two representations pI O.O) :!: of f!/s depending
on whether T(l)lkoO) = ± llk oO ). We shall not go into any details.

15.6 The Poincare' group with time inversion, &',

Having discussed the space inversion I let us now consider the time inversion
operator I, defined by l,e = e' = (x, y, z, - ct). As in the previous section the
addition of this operation to the Poincare group generates an extended group
which we denote by .'lJ,. We see directly from its definition that the products of
15.7 Space and Time 363

translations with I, satisfy the relation


P(e)I, = I,P(l,e) (15.92)
Thus, if we denote by Ik) a basis vector which transforms according to the
representation T(k) of the translation group, we see that
T(e)T(l,)lk) = T(I,)T(I,e)lk)
= exp(itI,e)T(l,)lk)
= exp(iI,k.e)T(l,)lk) (15.93)

showing that the new vector T(l,)lk ) transforms according to T(I,k) under
translations. In other words, the time component of k is reversed. But we have
seen in section 15.4.1 that the physical significance of this component k, of k is
the energy, E = lick, . Thus the time inversion operation as given by equation
(15.93) interchanges positive and negative energy states. There is however no
evidence that this is a symmetry operation or indeed any evidence for the
existence of negative energy states. For this reason we shall not pursue the
representations based on (15.93).
We shall nevertheless return to the group ;J}I, in subsection 15.7.4 in
discussing time reversal, which is known experimentally to be an almost
universally valid symmetry operation. This operation, which corresponds
physically to a reversal of the direction of motion, is obtained by seeking a
representation of the group element I, which avoids the difficulties which
follow from equation (15.93). In deriving (15.93) we have assumed that T(l,) is
a linear operator in the representation space of the vectors Ik) . If one takes the
operators T(l.) to be 'a ntilinea r' then the appearance 0 f negative energy states
is avoided. Because new concepts are involved we defer the discussion of time
reversal until subsection 15.7.4.
If we wish to include both the space inversion I and the time inversion I,
defined by I,e = (x, y, z, - ct) then to form a group we must also include the
product II, which changes the sign of all four components, II,e = - e. Let us
denote the group formed in this way by .9's,. We note that the product II,
commutes with all Lorentz transformations but not with translations.

15.7 Physical interpretation of the irreducible repre-


sentations of the Poincare group

We have already discussed the physical significance of many of the symmetry


operations in the Poincare group: translations in three-space in subsection
15.1.1, translations and rotations in three-space (the group 0"3) in subsection
15.1.5 and translations in four-space in subsection 15.4.I. We now investigate
the additional consequences which result from the full Poincare group.
It was already clear in subsection 15.1.5 that the group 0" 3 was relevant only
for a free particle or for the external motion of a system of particles and the
364 Space and Time 15 .7.1

representation labels corresponded to various properties of the system. Thus


E(lkl.m) specified the magnitude Ikl of the momentum and the internal angular
momentum (or spin) m about the direction of motion. States corresponding to
different directions of k were equivalent in the sense that they could be
obtained one from another by rotations. The invariant quantities were Ikl and
m. Since these remarks relate to any free system they must also relate even to
elementary particles whose structure we may know nothing about. Thus we are
able to characterise different elementary particles by different values of these
invariants. By this argument however and using only the group ,g3 we should
refer to particles with different magnitudes Ik Iof their momentum as different
particles, and likewise for different projections m of their internal angular
momenta on to the direction of motion. This would be a rather improbable use
of the expression 'd ifferent particle' but we can now see that its unreasonable
aspects are due to our use of ,g3 rather than :!Ji as the symmetry group. The
inclusion of transformations between systems moving with uniform relative
velocity would remove the improbable features mentioned above and these are
precisely the transformations added to ,g3 to form the group f!J'. We therefore
expect to find some more reasonable fundamental invariants from the study of
the Poincare' group :!Ji.

15.7.1 Mass
The irreducible representations of the Poincare' group are characterised
firstly by the Casimir operator P.P whose value, from section 15.4.3, is given
by - k 2 in representations p 1k • s ) and by zero in p<O .m l . In the representations
corresponding to space-like t, which we did not describe in detail, P.P would
be positive. For reasons which become clear as we proceed we associate the
operator P.P with the square of the mass . Precisely, the operator for the
'sq uare of the mass' is defined, including appropriate dimensional factors, to be
- Ji2 P.P/c • Thus if M
2 2
denotes the value of the 'square of the mass' then in
representations p(k .s) we have M 2 = Ji 2k 2/C 2, while in p(O .ml we have M 2 = 0
and for space-like t, M 2 < O. With this interpretation, the mass would be
imaginary for space-like ii, would be zero for pIO .m ) and take the value
M = hk]: in plk.s). (Notice that it is M 2 which has significance, rather than M;
a situation rather similar to angular momentum where J .J is the Casimir
operator with value J(J + 1) but we use the label J.)
Let us now give some justification for using the word 'mass'. Since we have
earlier given physical significance, momentum and energy, to the four
components of P, the very definition P. P = P;- L p~ implies a relation
q
between the eigenvalues of these operators. In plk.s) the relation becomes

_c 2M 2/ Ji 2 = _E 2/Ji 2C2+ LP~ /Ji2


q

i.e. (15.94)
15.7.1 Space and Time 365

making use of the definitions of p and E in subsections 15.1.1 and 15.4.1 and
with M defined above. (This is the familiar relation between energy and
momentum in classical relativity theory, see section 16.1, and the word 'rest-
mass' is sometimes used for M .) If the momentum is small, p ~ Me, then

(15.95)

which gives the energy as the sum of the rest-energy Me 2 and the kinetic energy
p2/2M = tMt: 2.
So far as the label k is concerned, the different vectors Iksm s > in the
representation plk. S) correspond to different states of motion of a particle (or
system) with mass M = ilk/c. Not only may the momentum p = hk be in any
direction, but it may have any magnitude with corresponding energy given by
equation (15.94). Thus, even if we restrict to s = 0, the single representation
plk .O) encompasses all possible states of motion of a classical particle of mass

M. We shall see presently that the label s relates to the internal angular
momentum or spin of the particle.
Before discussing zero mass we briefly consider the representation cor-
responding to k = 0 and referred to in subsection 15.4.2 under region (6). This
would not only have zero mass but also zero energy and momentum in all
states, i.e. in all frames of reference. No such object has been observed and we
discuss the possibility no further, noting however that the point k = 0 must be
excluded from the representation p lO.m l .
Let us now turn our attention to the representation p I O.m) corresponding to
zero mass . This seems a rather unl ikely description for a particle until we notice
that it may nevertheless have finite energy and momentum satisfying, from
equation (15.94), the relation

E = pc (15.96)

Thus, although having zero mass , the object acquires some physical properties
and in fact such particles are well known in physics. In the quantum treatment
of the electromagnetic field, the interaction takes place through the exchange
of photons which need to have zero mass . Also, in fJ-decay one finds neutral
particles emitted which have zero mass and are called neutrinos.
Whereas for a particle with finite mass, there is always a state with p = 0, in
which the particle is at rest, with E = Me 2 , this is not the case for a zero-mass
particle. One cannot set p = 0 without moving to the limit point k = 0 which
defines a different irreducible representation in which all states of the particle
have zero energy and momentum as described above. Classically this implies
that zero-mass particles have velocity t: = e because as we have seen in
subsection 15.2.3, the velocity is then v = e in all frames of reference and no
state of rest is possible. (Another classical argument supporting the statement
that v = e is that with zero mass a particle can acquire finite energy and
momentum only in the limit of v = c, see equation (16.10). )
366 Space and Time 15.7.2

15.7.2 Spin

Having seen how the first Casimir operator P.P is associated with mass and
the components of P are associated with momentum and energy we now turn
our attention to the second Casimir operator W.W defined in subsection
15.4.3. The interpretation of W.W takes rather different forms for finite and
zero mass and we describe them separately.

Finite mass
Consider first the case of finite mass, described by the representations p<k. s).
From equation (15.74) we recall that the number s, which may take values 0, t
1,1, .. . labels the eigenvalues -k 2 s(s+ 1)on".w for given k. In terms of the
mass, introduced in equation (15.94) the eigenvalues of W .Ware
-M 2 c2 s(s + I)/li 2 (15.97)

Since W.W is a Casimir operator, all states belonging to the representation


p1k.S) are eigenfunctions of W.W with the same eigenvalue. Thus s describes
some intrinsic property of the particle, like the mass, which is the same in all
frames of reference.
We shall refer to this intrinsic property as the spin. Strictly, sh is called the
spin, so that s is the spin in units of h. The reason for using this word is clear if
we study the operators Wq for a particle in a state of rest, i.e. when
k = (0,0,0, k). Then from equation (15.72)
W q = -i(Mc /h)X q = -(Mc/h)Jq, W, =0
and
(15.98)
showing that, for a particle at rest, W q is proportional to the angular
momentum Jq • Of course, a classical point-particle at rest has no angular
momentum so that it is appropriate to use the term 'intrinsic spin' to describe
the label s. Furthermore, since Wis translation invariant it follows that, when
operating on a general basis vector Iksm s >ofP(k.S), it cannot change k. Thus W
simply transforms among different ms with fixed t; Notice that the basis
Iksm s> was chosen so that Wz is diagonal on the special vector Ikosms>
through the definition of ms , i.e.
W z Ikosms >= -msklkosms> (15.99)
Since W is a four-vector it follows that, for an arbitrary basis vector Ihm s >'
some transformed component of W will be diagonal. This may be seen by
transforming both sides of equation (15.99) with the boost which carries ko
into k and using equation (15.57)
T(Q[b(~)])Wz T- t (Q[b(~)])I~sm.> = - m.klksms >
which gives, from the explicit form (15.26) for the boost of a four-vector, and
15.7.2 Space and Time 367

using the va lues for b(k) given after equat ion (15.58)
{Wz- (k,+ k)- lkz W,}!ksms >= - msklksms> (15.100 )

In deriving thi s result we ha ve used the property W.P = 0 fro m subsectio n


15.4.3 which enables us to repl ace W . P by W, P,. It is possible to choose a new
ba sis (helicity) in which the component of W in the direct ion k of motion is
a lways diagonal but this would have led to a more complica ted expression than
(15.59) for the general transformation.
From equation (15.59) one sees that the effect of a Lorentz transformation
>
T(O, L) on a n arbitrary vecto r Ik sm s is in two parts. There is a matrix sum
over m~ and a change in the vecto r t. Thus an infinitesimal Lorentz
transformation is described by the sum of a matrix operator acting on the m~
a nd a d ifferential operator which produces an infinitesimal change in k
without affecting ms ' Thi s situ ation is reminiscent of the separation j = s + I
described in section 8.4 for the non-relativistic case . However, the basis Iksms >
does not lead to complete separation because the rotation R' = Q' - I LQ
entering the matrix in equation (15.59) depends on k as well as L. Nevertheless,
by a change of ba sis, it is po ssible to achieve complete separation writing not
only j = s + I for the infinitesimal rotations but also Y = Y. + Y, for the boosts
where sand Y, are 'spin' operators, given by matrices, which commute with the
'orbital' operators I and V" Because thi s sepa ratio n is closel y connected with
the Dirac equation we defer its discu ssion until subsection 15.8.4.

°
Experimentally, one find s that the electron, proton and neutron all have
x = 1 while n-me son s have s = and p-meson s s = I. Co mplex systems like
a toms and nuclei ca n ha ve mu ch greater values for s but al ways an integer or
half-integer.

Zero mass

As we have seen in subsectio n 15.7.1 the representations p IO.m) of the


Poincare group provide a de scription for part icles with zero mass . The
discussion of spin given above for the representations plk.s, is inapplicable for
particles with zero mass because it began by examining the behaviour of the
particle at rest and zero ma ss particles always move with the velocity of light.
However, using equations (15.72) and (15.76) we see that, for any vector Ikm >
of p ' O.ml

-(X . P)Jkm >= - W, Ikm >= imP, lkm >= mk , Ikm >= mlkllkm >
(15.101)

But in phy sical terms, - X . P is th e component of angular momentum in the


d irection of motion multiplied by the magnitude of momentum. Since a
classica l point part icle would ha ve no such component of angular momentum
we again use the word 'spin' to de scribe the number m, but we must realise that
the spin of a zero-mass particle d iffers greatl y from that for finite mass . There is
368 Space and Time 15 .7.3

now no (25 + I)-fold degeneracy of different spin orientations. In fact, even the
state with opposite spin - m would belong to a different irreducible
representation of the group f!J! namely P'? : - ml rather than p<o .m). A particle
with positive m is said to have positive ' helicity' or to be right-handed in the
usual sense ofa screw . We discuss the significance of heIicity in the next section
in connection with parity. However, there is a simple physical argument which
shows how this right- or left-handedness may exist in zero mas s particles but
not in finite mass particles. The helicity of a particle will be rever sed in the
frame of an observer travelling faster than the particle, compared with that of
an observer moving more slowly . Thus, for a finite mass particle which travels
at speed v < c anyone-handedness would violate Lorentz invariance. How-
ever, for zero mass , the particle must travel at speed v = c so that no observers
may travel faster than the particle and the contradiction cannot arise .
Of the known part icles with zero ma ss, the photon is found to have both
m = ± 1 while the neutrino and the antineutrino have m = - i and m = i,
respectively. (See also the discussion of antiparticles in subsection 16.3.5.)

15.7.3 Parity

The representations discussed so far have related to the proper Poincare


group f!J! without inversions in space or time . Let us now suppose that the space
inversion I is also a symmetry element so that the full symmetry group
becomes f!J's whose representations were constructed in section ·15.5. Again we
study the consequences separately for finite and zero mass .

Finite mass
It was shown in section 15.5 that, when the group:!J is extended to fl/ s there
are two irreducible representations p(k,s) ± of f!I's for each representation pCk , s )
of f!J'. The inversion operator was shown in equation (15.87) to have two effects ;
it changed the direction of k and it gave a factor ± 1 in p(k,s) ± , respectively. The
first of these effect s gives the expected reversal in the direction of motion while
the second applies even to a particle at rest. Thus one refers to the ± factor as
the 'intrinsic parity' of the particle, even or odd, which , like the spin, is
independent of the state of motion and is a characteristic of the particle. It may
be imagined to refer to some 'internal' structure of the particle.
In fact it is possible to determine the absolute intrinsic parity of a particle
only in rare cases like the nO. Usually, two or more changes occur in a reaction
so that only the relative intrinsic parity may be deduced . Furthermore, because
of what are believed to be absolute conservation laws. such as charge
conservation, certain classes of transition are forbidden , preventing the
determination of even the relative parities between certain sets of particles. The
absolute intrinsic parity of the n O may be found from a reaction like
p + p -+ p + P + nO, and a detailed analysis of the results of such experiments
shows that the nO has odd parity. Since they have a charge it is not possible to
devise an experiment to measure the absolute intrinsic parity of the n ± but
15.7.4 Space and Time 369

because they are part of the isospin triplet (T = 1) with the nO (see sect ion 10.2),
it is natural to assign an odd parity also to the 7l - . The relative parity of
neutron and proton is then found to be positive from the analysis of reactions
like p + p -+ n + p + 7l + and it is convention to choose them both to have
positive parity. It is crucial to the argument above that the reactions proceed
via the strong nuclear forces which are known to conserve parity.

Zero mass
For a particle with zero mass we recall from sectio n 15.5 that the irreducible
representations of the extended group # , are made up of two irreducible
representation s of !J, plO, lmll = pIO. m) -+- plO. - m) and the space inversion
carries a vector o f one into a vector of the other. This is to be expected since the
inversion commutes with the spin and thus a change in direction of k must
imply a change in helicity. One cannot assign an intrinsic parity here as we did
for finite mass because the irr educible representations PIO. lml)of the group :Y's
do not carry such a label ±. Physically, the intrinsic parity cannot be studied in
the sta te of rest since no such state exists for zero ma ss.
The fact that the photon is found experimentally in both m = ± 1 states
suggests that it is described by the representation plO.11 -+ p lO . - I ) of the group
.:J's or in other words that it has :1/, as its symmetry group. On the other hand
for the neutrino which is found always in the m = -1 state, the appropriate
group is q> excluding the inversion. Thus whereas parity is conserved in
electromagnetic interactions, involving the exchange of photons, one finds
experimentally that parity is not conserved in processes like fJ-decay which
involve neutrinos. It is true that antineutrinos, with m =1, exist also but they
must be regarded as different particles since they do not occur in the same
reactions as neutrinos, see subsection 16.3.7.

15.7.4 Time-reversal

If we run a movie film backwards we shall see all movements reversed; a


falling stone will be replaced by a stone projected upwards. The reversed
motion is also consistent with physical laws in the sense that a stone actually
projected would behave (in a vacuum) just like the backwards run film of a
falling sto ne. This is true because Newton's equation of motion in a uniform
gravitational field is unchanged when the time t is replaced by - t. However, in
the Schrodinger picture of quantum mechanics, the simple time inversion
operator t -+ - t also changes the sign of the energy, as we have seen in section
15.6, because the energy operator is given by iii 2j ct.
Thus in quantum
mechanics, the time inversion operation has a mu ch more drastic effect than
simply reversing the motion . Since the idea of reversing the motion is perfectly
sensible in classical mechanics one would expect to find a corresponding
operation in quantum mechanics, i.e. one which reverses directions of motion
but does not change the sign of the energy. In this section we construct such an
operation, called 'time-reversal' and denoted by T . It will turn out that T is a
370 Space and Time 15 .7.4

very unusual operator, be ing non-linear, so we must digress to explain the


meaning of such operators. (A brief mention of this operator has already been
made in sections 5.10 and 9.8.)
In all the examples of symmetry in quantum mechan ics given so far in this
book, the transformations in the wave functions have been unitary and linear.
Let us look more closely into the reason for th is choice and we shall find that it
is not the only possibility: let It/J > and I¢ > denote two states of a system. One
of the postulates of quantum mechanics is that, if a system is in state It/J > and a
measurement is made to determine whether it is in state I¢ ), then the
probability that it will be found to be in state I¢ > is given by the squared
modulus I<¢ It/J >12 of the scalar p roduct. Ifwe denote by It/J' > = Tlt/J > and
I¢' > = TI¢ > the new states following a symmetry transformation T, then
since the result of the measurement must be unaltered we have
(15.102)
This mu st follow if the primed system is to have the same phy sical properties as
before. It is clear that this equation is sat isfied if T is unitary, i.e.
<T ¢ITt/J > = <¢I t/J > for alii t/J > and I¢ > but, because of the modulus sign s,
there are other possibilities. It may be deduced from equation (15.102) that,
apart from trivial pha se changes. the operator may either be un itary and linear
as we have assumed hitherto in the book, i.e.

<T ¢ IT t/J> = <¢ It/J > (15.103)


and
T(al¢ > +blt/J » = aTI¢ > +bTIt/J >
or that T is anti-unitary and antilinear, i.e.

<T ¢ IT t/J> = <t/J I¢ > = <¢\t/J >*


and
T(a l¢ > +blt/J » = a*TI¢ > +b*TIt/J > (15.104)
For continuous groups we are led naturally to the choice (15.103) since we take
the unit operator to represent the identity and the unit operator satisfies this
equation. Again, when considering time inversion in section 15.6, we chose
the more usual possibility (15.103). By taking now the second possibility
(15.104)we shall find a transformation which represents time-reversal and does
not change the sign of the energy.
Consider the representations of the group &, obtained by adding the time
inversion operator I, to & but where we use an antilinear operator Y to
represent I, . Thus we retain equation (15.92) for the multiplication of group
elements but, using the non-linear Y in place of T(I,) the equation (15.93) is
replaced by
T(€) Ilk >= YT(lr€)lk > = Yexp(ik . Ir€)lk >
= exp( -ik . I,€)Ilk >= exp( -iI,k . €)Ylk >

= exp(iIk. €)Ylk > (15.105)


15.7.4 Space and Time 371

where we have used the second of equations (15.104) and the elementary result
I,k = - Ik, where I is the space inversion. Thus, under translations, the vector
Ylk >transforms according to TOk) and the d ifficulty with negative energies is
avoided. The operator Y also has the desired effect of revers ing the direction of
k . Although time-reversal is sim ilar to space inversion we cannot assign a 'time-
reversal parity' with any phy sical significance. This stems directly from the
I
antilinear nature of Y. For suppose that r/J > was an eigenfunction of I with
;.1
eigenvalue X, so that II r/J > = r/J >. The antilinear nature of Y now implies
that for any phase '1
Texp (iIOI r/J > = exp( - ilJ)II r/J > = exp( - il/)}·I r/J >
= [exp( - 2il/p.] exp(ilJ)Ir/J )

Hence the function exp(iIJ)Ir/J >, which must physically be indistinguishable


I
from r/J ) , has eigenvalue exp ( - 2ilJ)}. for I and so the eigenvalue of I can
have no physical meaning within a phase. In particular, the sign of ;. would be
meaningless. These remarks apply both to finite and zero mass but to study the
irreducible representations of f!J, any further we must treat the two cases
separately. In both cases we shall find that the representation spaces do not
need to be increased in dimension when the extra element I is added to make
f!J,. To make any progress we must use the multiplication law between time
inversion and Lorentz transformations
I, L(a, b) = L(a , - b)I, (15.106)

wh ich follows from the 4 x 4 matrices. We shall assume as usual that the
representation operators satisfy the same law. In words, thi s says that the time-
reversal operator Y commutes with rotations and changes the direction of the
boosts . From the very nature of the physical operation of time-reversal, the
square 1 2 must reproduce the original ph ysical system. Thus
y 21r/J > = exp(itx)Ir/J ) , where tx is any phase. Combining this with the anti-
unitary condition (15.104) leads to the further restriction (see problem 15.10)
that 1 2 \r/J ) = ± Ir/J >. For a strict, single-valued representation we should
take only the plus sign since the group element satisfies 1,2 = E. However, if we
allow double-valued representations (like the half-integer representation of
9!' 3) then the minus sign is permitted .

Finite mass
In studying the behaviour of the basis vectors Iksms >of the representation
plk.s) of f!J when the new element I is added we follow the technique of section
15.5 and look first at the special vectors Ikosms> with t,
= (0,0,0, I). From
equation (15.105) the vector Ilkosms > has the same label k = ko . Because Y
commutes with rotations one might have expected that the vector I lkosms >
would have the same value of ms but this does not take account of the
a ntilinear nature of I . Becau se the rotation operator X= has an imaginary
372 Space and Time 15.7.4

eigenvalue, X,lk!Jsms>= - imslkosm. >, see for example section 7.3.5, we have
X,Ylkosm. > = YX,lkosm. > = 1'( -im.)lkosm. >
= im.Ylkosm. >
which shows that, under rotations about the z-axis, the vector Y lk osm. >
transforms like - m. although it is easily seen that the value of s is unchanged .
In fact, for a general rot ation
T (R(a ) )Y lkosm. > = YT(R(a) )Ik osm. >
= I . r DIs)
msm,
( a )l ko s m~ >
m,

(15.107)

Thus the set of vectors Y /kosm.> with fixed t;


and s transform like the
complex conjugate of the representation 0(' ) of fJt 3 • However, the complex
conjugate of 0 (' ) is equi valent to 0 (" , see section 7.7, and with the usual
convention D(S)* (a ) = (- 1 )m, - m;D(s) . (a). Thus the set of vectors
A m,rns -mJ -m J A

1'!kosm. > transform preci~ely like th e set (-1 )m, lkos - m' >'AThere are now
two possibilities: either 1'lkosm.> is linearly independent of Ikos - m.> or we
may write Ylk osm.> = y( -l)m,!kos - m.> with so me con stant y. Because we
are looking for an irreducible representation of gJ, the first possibility can be
dismissed since it can be shown, following the technique of section 15.3, that
the representation of gJ, generated from Ikosm.> would not then be
irreducible. Returning to the second pos sibilit y it is usual to choose the phase
factor y = ( - 1)' so that, even for half-integers, the factor ( - 1)' + m, is real.
Having seen that the (2s + 1) vecto rs Ikosm. > with fixed ko and s form an
inva ria nt space under rotations and time-reversal, we generate the remaining
bas is vectors of the representation of gJ, by usin g the boost equation (15.57). It
then follow s that for the general basis vector of the representation,
Ylk sm.> = YT(Q[b(k)] )/kosm.>
= T(Q[ -b(k)])( -l)s+m,lkos - m.>
= (- 1 )s+m,IIks - m.> (15.108)
showing how l' changes the sign of both the momentum and the spin. We see
from thi s equation that
1'2lksm. > = (_l)s+m,( _l)s-m' lksm. > = ( - 1)2Slksm. > (15.109)
showing that 1'2 = 1 in a representation with integer spin and 1'2 = - 1 for
half-integer spin. This conclusion is independent of the choice of the phase
factor y. From equation (15.108) we may immediately deduce the effect of T on
an arbitrary vector 11/1 > as defined in equation (15.78); the new vector
15 .7.5 Space and Time 373

11/1 ' > = 111/1> is given by the coefficients 1/1'm, (k) = (_1),,-m'I/I ~ m..(-Ik). The
time-reversal operator may be written as a product of a unitary operator and
an operator K which takes the complex conjugate of the c-nurnbers, i.e. the
expansion coefficients in some basis , but the explicit form of I expressed in this
way must clearly depend on the basis . Ifwe note from equation (20.40) that the
spin rotation R SY ( - z ) transforms m" into - ms with a phase change ( - 1)" + m,
A A

and if we define T, (I )asan operator which changes k into Ik, i.e. which changes
t.
the sign of the space part of then in the basis Iksm,,> the transformation
(15.108) shows the equivalence I = K R ~( - n )Tdl). However, in a coordinate
representation it is given by I = KR ~,( -n)T(I,) where T(I,) replaces t by -1.
=
When s = 1. R~, ( - rr) 2is)..
Zero mass
The multiplication law (15.106) implies that IX = XI and IY = - YI,
while equation (15.92) gives IP = PI and IP, = - P, I . Thus YW = - WY
and YW, = W, I . Now consider a basis vector belonging to the representation
p'O.m) of &>, which isAappropric;,te for zero mass . Then starting from equation
(15.76) we have Y(W + imP) Ikm > = 0 so that, making use of the multipli-
cation rules given above and the antilinearity of I ,
(-W- imP)llkm > =0
a nd
(W,+imP,)llkm > =0
or in other words
(W+imP)Ylkm > =0 (15.110)
Thus we conclude that the vector Ilkm > also belongs to a representation
p,O ,ml of &> and, using equation (15.105), that it transforms like Ilkm > under the
group flJJ. Again we can argue that if Ylkm ) and IIkm > are linearly
independent then the representation of f!J, generated from them will reduce.
Thus for an irreducible representation of f!J, we must have
(15.111)
2
where we have chosen the phase factor to be + 1. Notice that y = + 1 in this
case, irrespective of the choice of phase. The interpretation of equation (15.111)
is again that both momentum. and spin are reversed but, since m is the spin in
the direction of motion, its value is unchanged . From the point of view of the
representations we conclude that the vector space of p'O .ml remains invariant
when Y is added to the group f!J and therefore provides also the basis for an
irreducible representation of f!J, .

15.7.5 Some consequences of time-reversal symmetry

(1) Kramer's theorem


As a first example of the consequences of time-reversal we deduce Kramer's
374 Space and Time 15 .7.5

theorem, first enunciated in 1930, which says that, in a system governed by a


time -reversal invariant Hamiltonian, any states of an odd number of particles
with half-integer spin and finite mass must be degenerate in pairs. For a state

operations on each particle. Now consider the operator ,2.


11/1 >of n part icles the time-reversal operator is defined as the product of such
For each particle it
contributes a factor ( - I )25 from equation (15.109), irrespective of the state of
motion of the particle. Thus if 11/1 > is a state of n particles with half-integer
spm ,
for odd n
Now if 11/1 > is an eigenfunction of a ,-invariant Hamiltonian then so is the
function 1<I> > = ,11/1 >. But I<I> >can be shown to be independent of I1/1 >, in fact
orthogonal to 11/1 ), since from the equation before (15.104)
( 1/1 1<1» = ( 1/1 1' 1/1> = ( , 21/11 ' 1/1> = - (1/11,1/1 > = - (1/11<1»
Thus we conclude that ( 1/1 I<I> > = 0, that I<I> > and 11/1 > are independent and
hence that they give a two-fold degeneracy. The addition of , to a space
symmetry does not alway s increase the degeneracy, for example in a spherical
system this two-fold degeneracy relates to ± m, where m is the z-projection of
angular momentum and is just a part of the familiar (2j + I)-fold degeneracy.

(2) Reactions
In a collision process, the time-reversal operation will interchange the role of
the incoming and outgoing particles. An assumption of time-reversal in-
variance therefore relates the transition probabilities of inverse processes, a
property known as the 'principle of detailed balance'. In an interaction process
which may be treated by perturbation theory it is po ssible to deduce relations
between matrix elements within the same process. To illustrate this point we
must classify operators Q as even or odd under time-reversal, according as
'Q = ± Q', respectively. For example we have seen above that X and Pare
even while Y is odd. Remembering the factor i which enters in passing from
these infinitesimal operators to the Hermitian operators representing angular
momentum and momentum , we see that j, sand p are odd, as one would
expect , while for example rand s .1 are even . (Remember however that one
cannot cla ssify wave functions as even or odd under time-reversal.) For the
matrix elements of even or odd operators, respectively ,

By making measurements which relate matrix elements to those in time


reversed states one can deduce whether Q is even or odd, or indeed some
mixture.
As a particular example, consider the electric dipole moment of a particle at
rest and with half-integer spin and finite mass. This moment is defined through
an energy change proportional to the applied electric field E, i.e. through a
term of the kind D.E in the interaction Hamiltonian. where D is the dipole
15 .8 Space and Time 375

moment operator. In quantum mechanics D is a vector operator and for a


particle with spin s we shall be concerned with matrix elements <m.IDzlm.> if
we take our z-axis in the direction of the field E. Since D is a vector operator,
the Wigner --Eckart theorem, (see equation (7.54), gives
<msl Dzi ms>= ( _I)s+m.<- msl Dzi - ms> (15.113)
But from equation (15.108), I Ims >= (- 1) s + m s 1-ms >so that, depending on
whether D, is even or odd under time-reversal,

<ms \Dzlms >= ± <- ms IDzl- ms>* = ± <- ms \Dzl - ms >


since D, must be Hermitian to give a real energy change. On comparing this
equation with equation (15.113) we see that the dipole moment must vanish
unless D, is an odd operator. In the accepted time -reversal invariant theory of
the electromagnetic interaction D, is even under time-reversal and this is
consistent with the fact that no dipole moment has been observed for any
particle. At the present time very accurate measurements are being made on the
neutron to see if even a very small dipole moment exists . Any such observation
would indicate a breakdown of time -reversal symmetry in the theory. It would
also indicate a breakdown of space inversion symmet ry because accepted space
inversion invariant theory has D, as an odd parity operator. Thus if the particle
has definite parity the diagonal matrix elements of D, would also vanish for
th is reason .
In spite of numerous experiments no firm evidence for any breakdown in
time-reversal invariance has been found . Th is contrasts with the space
inversion symmetry where, because of the left-handedness of the neutrino, the
weak interactions like f3-decay show large violations of space inversion
symmetry.

The group e;
Since the extension from the group & to !?I, did not lead to any change in the
vector space of an irreducible representation, it follows that the extension from
&, to !?Is, which includes both space inversion and time-reversal, will be just
like the extension from !?I to !?Is described in section 15.5. In other words we
may simply include with the results of section 15.5 the transformations
deduced in this section for the operator I . The transformations for the new
product element IT(I) follow directly from those for the two factors.

15.8 Single-particle wave functions and the wave


equations

Our approach throughout this chapter has been to study the irreducible
representations of the Poincare group. This tells us the transformation
properties of the functions which belong to each representation. A belief in
Poincare' invariance then led us to identify the representation labels with
376 Space and Time 15.8.1

invariant properties of the particles, like mass and spin . By comparing with the
properties of observed particles we associated each particle with an irreducible
representation in the same way that in non-relativistic theory we associated
each eigenstate of the Schrodinger equation with a representation of the
symmetry group of the Hamiltonian. However we have never constructed the
wave function explicitly . The philosophy here follows that used in discussing
the group f7l 3 , namely that one can deduce the properties of the basis vectors
Ijrn> of the representations Olj) without specifying their detailed form . One
thereby achieved the greater generality that the results were applicable to any
spherically symmetric system . For example the wave functions Urn> have the
same rotational properties whether they refer to one particle or any number of
particles. In the case ofa single particle, however, one may construct explicitly
the angular dependence of the wave functions by introducing the explicit
differential form for the rotation operators. The Casimir operator is then a
differential operator and its eigenvalue equation provides a wave equation
whose solutions are the basis vectors Ijrn> of the irreducible representations.
In this section we first elaborate the above remarks for fJi 3 and then use the
same idea in the Euclidean and Poincare' groups to deduce the famous
equations of Klein - Gordon, Dirac, Weyl and Maxwell. Although for free
particles these equations carry no more information than we have already in
the irreducible representations they are useful when interactions are in-
troduced, see subsection 15.8.2, and when a theory of quantised fields is
constructed, see section 16.3.

15.8.1 The group 9f 3

Recall from section 7.5 that the fJi 3 Casimir operator is J 2 and it has
eigenvalue jU + 1) in the representation Ow, where j = 0,1, 1,~, . . .. The
explicit differential form of the infinitesimal operators J, = - i(r 1\ V)q was
given in subsection 7.4.5. Thus the basis vectors Ijrn>
must satisfy the 'wave
equation'
(15.114)
Since rotations leave the length of r unchanged we may regard r as a constant in
this study of the basis vectors of Ow. In this case, from equation (7.57), we see
that the equation reduces to
(15.115)
Thus the behaviour of the basis vectors Urn> could be found by solving the
wave equation (15.115). In fact this is the Schrodinger equation for the free
motion of a particle on a sphere of radius r and its solutions,see section 7.5, are
the spherical harmonics for integer j, namely Y ~)((),¢).
To construct wave functions with half-integer j we must extend equation
(15.115) by introducing a framework with several components as described in
section 8.4. If, in particular, we take two components, transforming like Om,
15.8.2 Space and Time 377

then the explicit forms for the rotation operators are J = 1+ s = - ir 1\ V + s


where s denotes the set of three 2 x 2 matrices Sq given in equation (8.15).
Following the same steps as before leads to the new wave equation
(15.116)

where we have inserted the value i for 5 2. The terms which do not involve 5 are
understood to contain the 2 x 2 unit matrix and the wave function Um> has
two components

in a notation similar to that of section 8.4. Thus equation (15.116) provides a


pair of coupled differential equations for the two functions t/J + (0, </» . This
sounds formidable but we have already solved the problem in section 8.4. The
two differential operators in equation (15.116) commute with each other so
that they may be diagonalised simultaneously and we know that the first leads
to spherical harmonics Y~J" Furthermore from the vector coupling in section
8.4 we know that for givenj and with s = 1the only possible values of Iarej ± 1.
By using the properties (7.40) of the raising and lowering operators I ± one can
show (see problem 15.11) that these two solutions are given by

( {(I + t+ m)/(21 + I)}! yt~-l (0, </»)


I)'m > -- {(I + t - m)/(21 + I)}! yto 1 (0, </»
for I =J' +-21
m +,
and by

for I = j-1

The numerical coefficients in front of the spherical harmonics are simply


vector-coupling coefficients and in fact the solutions above are precisely those
given by equation (8.24) for s = 1.

15.8.2 The group 8 3


We have seen in section 15.1 that the symmetry group g 3 has irreducible
representations E(\kl,m) with m = 0, ± 1, ± 1, . " and Ikl > 0, and
E(o. S) == 0(5) . The Casimir operators are P . P and P . X with eigenvalues -Ik 1
2

and -Iklm, in ElIkl. ). If we use the explicit differential forms P = - V and


m

X = - r 1\ V then P. X vanishes so that we can only construct the repre-


sentation E(lkl,O) with m = O. The corresponding wave equation is then
(15.117)
which is simply the Schrodinger equation for a free particle with kinetic energy
Ikl 2
in units of rz2/2M, and without spin .
To construct a wave equation which is appropriate for m # 0 we again
378 Space and Time 15 .8.2

introduce several components. Then, wr iting the rotation operator


X = - iJ = - i(l + s) = - r /\ V - is as before, the Casimir operator P . X is
given by -iP.s = iV.s. Taking two components, with s again given by the
familiar 2 x 2 matrices, we shall expect to produce basis vectors with m = ± 1-
This is soon verified by writ ing down the new wave equation
(iV .s+lklm)It/J) =0
which implies that
(15.118)

but since the spin matrices satisfy s; = t, S,Sy+ sys, = 0, etc., it follows that
(V .S)2 =tv2 so that by comparison with equation (15.117) m 2 =t
and
m = ± 1. Thus, in the two-component framework , we may construct basis
vectors belonging to Elikl.!) and E( lkl. - I). From the explicit forms (8.15) for the
spin matrices one deduces that the appropriate basis vectors (un-normalised)
are
(-kx+ik y ) e''k .r
and Im = - -1 )
2
=
Ikl+k z
(15.119)
with the usual convention for the two components, i.e. s, diagonal.

The spin gyromagnetic ratio


We are now in a position to deduce the value gs = 2 which we quoted in
section 8.4 for the spin g-factor of the electron. It is often said that this value is a
consequence of the Dirac equation but such statements are misleading. It is
true that this value of gs comes naturally from the Dirac equation but it comes
equally naturally from the wave equation (15.118) which is non-relativistic.
The introduction of a magnetic field to the free motion of a particle with charge
- e may be accomplished in classical mechanics by the substitution of
p + eA ]c for the momentum p in the Hamiltonian, where A is the vector
potential ofthe field. The same thing may be done in quantum mechanics in the
approximation where the field itself is not quantised. With this modification
and remembering that p = - iiiV, the equation (15.118) for m = ± 1 becomes
(15.120)

which is sometimes called the Pauli equation. For a uniform magnetic field we
may take A = 1 B /\ r, where B = V /\ A denotes the constant strength and
direction of the field. Then, using the properties of the spin matrices again and
remembering that V does not commute with A this equation (15.120) reduces
to
[V 2 + Ikl2 -(ejlie) {B .(1 + 2s)} Ilv >
= 0 (15.121)
where we have omitted terms of second order in B. Comparing this with the
15.8.3 Space and Time 379

Schrodinger equation (15.117) one sees the appearance of interaction terms


between the field B and the orbital and spin angular momenta of the electron.
Ifwe multiply equation (15.121) through by the constant - n2 /2M, to bring it
into the standard form of a Schrodinger equation with the correct dimensions
of energy, then the interaction term may be written as - Jl. B, where Jl
= - Jlo (I + gss) with the Bohr magneton JlB defined by JlB = ehfl M c and the
spin g-factor given by gs = 2. This result is in agreement with the observed
value for gs' which is very close to 2 and was discu ssed in section 8A.
lt should be emphasised that we have not proved that every particle with
t
s = has gs = 2 but simply that this value occurs naturally. lt is alway s possible
to add an extra term to the Hamiltonian to produce any value of gs. Indeed, for
the proton which also has s = t, gs = 5.58 in unit s of the nuclear magneton.
This value is said to be 'anomalous' and one supposes that the difference from
the natural value of 2 is due to the internal motion of its constituent parts, see
section 12.3. A particle like the electron or the muon for which gs is close to its
natural value is regarded as more 'elementary' than those for which it is not.
Even the electron has gs = 2.0023 but this small departure may be understood
when the electromagnetic field is properly quantised.

15.8.3 The Poincare group with s=o-the Klein-Gordon


equation

We now move on to consider the wave equations governing the motion of


free particles moving at relativistic speeds. Thus we must seek equations whose
solutions belong to irreducible representations of the Poincare group. First we
study the particles with finite mass so that we are interested in the
representations p1k . S). In this subsection we concentrate on the spinless
particles, s = 0, and in the next we investigate the case s = t.
The representations plk . S) are distinguished by the eigenvalues - k 2 and
- k 2 s(s + I) of the two Casimir operators P. P and W.W, see subsection
15.4.3. Following the procedure in earlier sections we introduce the explicit
differential forms P = - V, P, = (I /c 'iJ/iJt, X = - r /\ V and
Y = ctV + (r/c "'j!/ct for the infinitesimal operators. Then from the definition
(15.72) one sees that the operator Wvanishes identically. For non-zero k thi s
implies that s = 0, and the Casimir operator P.P leads to the simple
Klein -Gordon wave equation

(15.122)

If we introduce the notation 0 2 = - V 2 + c - 2 8 2/ at2 the form


(02+k 2)jl/t >= 0
of this equation is very similar to the Schrodinger equation (15.117). But
Ik
whereas in the Schrodinger equation the constant 12 has physical significance
as energy E = n2Ik\ 2/2M, the constant k 2 in the Klein -Gordon equation is
380 Space and Time 15.8.4

related to the mass M = hk]c. The plane-wave solutions of equation (15.122)


are
It/! > = exp [ i(k. r - k,ct)] (15.123)
and we have discussed them already in subsection 15.4.1. Recall also from
subsection 15.4.2 that the representations ptk . s ) have been chosen to have
future time-like ii, i.e. k, > 0, so that the energy E = lick, is always positive.
Thus we are imposing a restriction on the solutions of equation (15.122) which
clearly has solutions with either sign for k,.
As in classical mechanics, the inclusion of an electromagnetic field in the
relativistic Hamiltonian may be accomplished by replacing the energy-
momentum four-vector ft = (p, £ /c) by p+ eA /e,for a particle with charge - e,
where A is the four-vector A = (A , ¢) describing the field with ¢ the usual
scalar potential (see subsection 16.1.3). The Klein - Gordon equation (15.122)
is thus modified with a/at - ie ¢ /Ii replacing a/ (It and V + ie A /eli replacing
V.

15.8.4 The Poincare group with s = ~- the Dirac equation

°
To construct a wave equation which is appropriate for s # we must again
introduce several components. In group-theoretical language this is equivalent
to the construction of product representations. For example in discussing the
representations ElIkl,±1) of g 3 in subsection 15.8.2 we were effectively
constructing the product representation ElIkl.O)x E(0.1) == E(lkl,O)x 0(1) so that
the infinitesimal operators were sums of the differential operators relating to
E(lkl,O) and the matrix (spin) operators relating to 0(1). We found that the
+
product reduced to the sum ElIkl.1) ElIkl, -t). The same technique will now be
followed to describe the representations P(k.!) of the Poincare group and to
deduce the Dirac equation.
A natural choice for the product representation in the s = 1 case will be

°
ptk . O)X U 1,0) where the second factor is the tr anslation-invariant repre-

sentation corresponding to ii = and referred to under the subheading region


(6) in subsection 15.4.2. Being translation-invariant it is a representation of the
Lorentz group 5£, see subsection 15.2.5. (The choice of UO.1) instead of U t.O)
would be an equally natural choice but leads to precisely the same results.) The
basis vectors in this product representation are labelled by Ih >, where ~ refers
to p<k, 0) and n = ± 1distinguishes the two components of L H. 0). The Poincare
transformations in this basis are thus given simply by

T(e, L)I~n > = exp(i~'.&) L L~t~O)(L)I~'n'> (15.124)


where Ii' = i.t:
The matrix elements of the representation U1,O) can be deduced
from the matrix of 0(1) by noting that in Ut,O) the infinitesimal operators take
the form X = A + B = - is and Y = - irA - B) = - s in the notation of
subsection 15.2.4 and where s is the usual spin matrix. However, we shall not
need such details here. The essential feature of equation (15.124) is the simple
15.8.4 Space and Time 381

t.
factorisation in which the matrix is independent of This contrasts with the
basis Ik!m s> used for the representation p(k.1) in subsection 15.4.2, where the
corresponding equation (15.59) may appear to factorise but the rotation R'
depends on k through equations (15.60) and (15.61). In fact we may transform
from the basis Ih > to the basis I~!ms > and this verifies that the product
representation p(k. 01 X Ut. 0) is equivalent to p<k.!l. The transformation is

I~!ms > = I L~!';'~)(Q)lh > (15.125)


"
where the boost Q was defined in equation (15.58). To verify this result we use
equations (15.124) and (15.125) to see that

T(e, L)I~!ms > = exp(i~ ' .e)I [U!.O)(L)LH.O)(Q)]nm,lk 'n >


"
= exp(i~ ' .e) I [LH. OJ -I (Q')LH. O)(L)L(!·O)(Q) ]m;m.I~'!m~ >
m'.,

using also the inverse of equation (15.125). But because U !·O) is a group
representation the matrix here is simply u tO)(Q'-1 LQ) and from equation
(15.60) this isjust U 1'0)(R') = D(!)(R'). Thus the states constructed in equation
(15.125) have the correct transformation property (15.59).
The basis I~n > thus provides a simple two-component description
for a particle with spin s = 1- The wave equation would be simply the
Klein-Gordon equation
(15.126)
where II/J > is now a two-component wave function . The role of this equation is
to provide the plane-wave factor (15.123) appropriate to the factor p(k . 0) in the
product representation. Note that the wave equation (15.126) contains no spin
operators which says that no constraint is being put on the 'direction' ofli/J > in
the two-component space.
As far as the second Casimir operator W.Wis concerned it is easily shown
(see problem 15.12) that in the product representation plk .O)X U t.O) one has the
equivalence W.W = i P.Pwhich is to be expected for the representation plk. Jl
from equation (15.74). However, in the presence of an electromagnetic field the
substitution of ~ + eA/c for' in W.W gives rise to a new term in the
Klein - Gordon equation which couples the electron spin to the field.
The two-component description given above for a particle with spin s = ! is
unsatisfactory for two reasons. Firstly the transformation (15.124) is not
unitary (recall from subsection 15.2.5 that the representation U 1.O) of the
Lorentz group was not unitary). The unitary representation could only be
recovered by returning, through the transformation (15.125) to the original
basis Ik!m s> for the representation p(k.!) and thus losing the factorisation. The
second reason is that the two-component space does not remain invariant
when the space inversion is introduced (recall from section 15.3 that the space
382 Space and Time 15 .8.4

inversion carries vectors of ut,O) into vectors of UO,l) and these two
representations of !fl may be added to form a single four-dimensional
+
irreducible representation ut,O) UO,\) of the Lorentz group with space
inversions !fl s ) and is therefore unsuitable for a description of particles with
definite parity. We now show that a four-component description based on the
+
product representation plk,O)x (Ut,O) UO,J) overcomes both these difficulties
and the restriction in the number of spin states from four to the value two ,
which is appropriate for the representation ptk,l), is achieved by solving what is
known as the Dirac wave equation.
Following a procedure very similar to that above we introduce basis vectors
+
Ika >for the product representation p(k,O)X (Ul,O) UO'\), where a = 1,2,3,4
runs over the four components of the second factor. The Poincare transfor-
mations are then given by
4
T(€, L) Ika >= exp(ik' , €) L M /ia(L)lk' P> (15.127)
/1=1

where M /i~ denotes the matrix elements of the 4 x 4 matrix

_ (U l,O) 0 )
M- o UO,,1 ) (15.128)

in partitioned form with Ul, OJ, etc ., denoting 2 x 2 matrices. Now the four-
component space is clearly too big for describing the representation p(k,l); in
fact p(k,O)X (U l,O) + UO,I)) = 2p(k,lJ. The previous two-component description,
based on the U l.O) piece, was one way of selecting a single p(k,t) representation
but there are other ways which involve the mixing of Ul,OJ and UO,I). To discuss
them we mu st first define some operators which couple these two two-
dimensional subspaces. From the rule (15.37) for reducing product repre-
sentatio ns in the group !f such operators must transform according to U l.l)
and thus form a four-vector which we denote by y. It is soon verified (see
problem 15.13) that such a set is given by the four 4 x 4 matrices

Y, = (0 I)
1 0' (0
Yq = 2 Sq Soq) with q = x, y or z (15.129)

in partitioned form where 1 denotes the 2 x 2 unit matrix and Sq are the 2 x 2
spin matrices. We may now construct the Poincare invariant operator y.Pand
diagonalise it. To find its eigenvalues we note that

(y.p)2 = L giiYiPigjj yjPj = L Y?P?+ L giigjjPjPAYi Yj + YjYd


~j i*j

while, from the definitions (15.129) and the properties of the spin matrices, the
y-matrices satisfy the anticommutation relations
(15.130)
15.8.4 Space and Time 383

Thus we have the equivalence <y.p)2 = P . P so that, within the product


representation p(/<. 0) x (U!,O) -i- L1o'!)),the operator (y.P)2 is just - k 2 times
the unit operator. Hence the possible eigenvalues of y. Pare ± ik and if we
classify the basis vectors of the product representation according to which of
these two eigenvalues they have, the representation space will have broken into
two invariant subspaces each of which forms the basis of a representation
p(k.!l. Thus by restricting our wave functions to one of these subspaces we have
constructed a basis for p(k.!). It is unimportant which of the two we choose and
if we choose - ik the restriction is imposed by demanding that the four-
component wave-function Ir/J ) should satisfy the condition
(15.131 )

which is called the Dirac equation. This acquires a more familiar form if we
write k = eM 11i in terms of the mass M. (see subsection 15.7.1) and introduce
the differential form P = (- V, ( L'c) alat). Then

or, multiplying through by YI and using Y ~ = I ,

[iii :t+ ihcv,»: V - MC2 y] Ir/J ) = 0 (15.132)

For our continuing discussion of the Dirac equation we prefer to use the form
[15.131).
To appreciate the form of the solutions of the Dirac equation and to show
that for given k there are j ust two solutions rather than four we first note that,
for a particl e at rest, PIr/J = 0 and P,Ir/J ) = - ikl r/J ), so that equation (15 .131)
becom es
(15.133)

and from the explicit form (15.129) for YI it is easily seen that the solutions are
any linear combinations of the two solutions

Ik, > and Ik,t-t>~2-(~)lkO)


(15.134)

Ik
where o ) denotes the plane wave (15.123) with k = O. For general k the
matrix equation corresponding to (15.133) is obtained simply from equation
(15.131) and (15.129) by using plr/J ) = - ik'r/J ). The solutions are then
384 Space and Time 15.8.4

expressed as linear combinations of

kx-iky )
Ik 12 _ 12 ) = l f k(k + k ) } - t
2 I t
k+kt-kz
_ k + ik
Ik ) (15.135)
( x y

k+k t +k z
this particular pair having been chosen so that they reduce to (15.134) in the
non-relativistic limit k --+ 0. The normalisation factor will be explained later.
Recalling the structure U t O)-+- uo.
tJ of the four components it is clear that in
this limit the states Ik o ± t t)correspond to the two spin states ms = ± of a t
particle at rest. It may also be shown (see problem 15.14)that the general wave
function (15.135) may be obtained from the solutions (15.134) for a particle at
rest by the operation of a boost Q[ b(k)] , as in equation (15.57) for the vectors
of plk.t) . Finally we stress that, although we are working in a four-component
space, there are only two independent states for a particle with fixed k given by
equation (15.135) and that any state of the particle is expressible as a linear
combination of these two . This is consistent with the conclusions reached from
studying the Poincare representations where there were two states ms = ± t
for each k.
The basis which one uses in the four-component space is clearly a matter of
choice and our choice was determined by equation (15.128), i.e. the first two
rows transform like Ut.O). In this basis we saw that, even in the non-relativistic
limit, there was mixing of components, see equation (15.134). When discussing
phenomena near this limit it would clearly simplify the algebra if one chose a
new basis in which the states (15.134)are given by the column vectors (1,0,0,0)
and (0, 1, 0, 0). In the new basis one refers to the first two components as 'large'
and the last two as 'small' because of their behaviour in this non-relativistic
limit, when Ikl/k is small. We shall not use this new basis .
The solutions (15.135) could also have been obtained directly by construct-
ing a state of definite parity from the state (15.125) and the corresponding state
generated from Lt°·!). If we denote by Ikn )' the two states transforming like
the product plk .01 X Uo.!) then we find, as in (15.125), that the two solutions
(15.135) may be written as

(15.135a)

as may be verified by using the explicit form


Ut.O)(Q) = {(k+k t)+2k .s} / {2k(k + kt )}t
15.8.4 Space and Time 385

together with VO,!l[Q(k)] = V!,Ol[Q(Ik)] which follow from equation (15.58)


and subsection 15.2.5.
We now show how to define a scalar product in the four-component space
which is invariant with respect to Poincare transformations. In other words we
construct a scalar product with respect to which the Poincare transformations
are unitary. Such a scalar product is then suitable for use in calculating values
of physical observables. We write an arbitrary solution as

,"') = ~ f",.(k)lkcx ) k,-l dk (15.136)

which is very like equation (15.78). However, the scalar product is now defined
as

<¢I"' ) = If¢t(k)(Y,)/h ",.(k)k,-I dk (15.137)


•. p

which differs from the corresponding equation (15.79) by the presence of the
matrix 'I,. The reason for including this matrix is to make the scalar product
invariant under the transformations (15.127). As we have remarked before, the
matrix M in that transformation is not unit ary so that, had we omitted the
matrix 'I" the scalar product would not have been invariant. To show the
invariance of the definition (15.137) one must show the matrix equality
Mty,M = 'I, which from equations (15.128) and (15.129) reduces to
(VO,1l) t = (u!.o))- 1. But this relat ion follows directly from the matrices ofVO,!l
and V!,O) for pure rotations and for pure boosts and thus for all Lorentz
transformations. The set of four functions

(15.138)

is sometimes called the adjoint of the set ",.(k) and it is clear that the expression
(15.137) is simpler with this notation. An alternative, but precisely equivalent,
form for the scalar product is obtained from equation (15.137) by using the
Dirac equation (15.131),

<¢I"') = ~ f ¢.*{k)",.(k)kk,- 2dk (15.139)

If we choose for I"' ) a particular plane wave solution Ik 'H) given by


(15.135) then the appropriate coefficients ",.(k) are given by

"'1(k) = (k + k, + k z)k t8(k -k)Hk(k +k,))-!


1jJ2(k) = (kx+ik y)k,8(k-k)t{k(k+kt ) ) - !
1jJ3(k) = (k + k, - k z ) k,8(k -k)t{ k(k + kt ) } - !
1jJ4(k) = (- k; - iky)k,8(k -kH{k(k +k,)}-!
386 Space and Time 15 .8.4

with correspond ing expressions for the second solution 11. 1 - 1). It is soon
verified that these two solutions are orthogonal with our definition of scalar
product and their normalisation is given by
< k'1ms l k1m~) = kI 8 (k' - k ) 8 msms' (15.140)
So far as space inversion and parity is concerned the same results as in
section 15.5 may be arrived at in the four-component description by writing
the inversion operator T(I) as the product of three factors
T(I) = T intr (I)Tk(l)Ts(l)
where the last two factors refer to the two factors in the product representation
p1k. 0) x (U !·O) +- LlO,!)) and the first factor is the intrinsic parity. Now, as we have
seen before, Tk(I) will simply change the direction of the space part of k while
from section 15.3 the matrix form ofTs(I) is simply YI' Thus for a particle with
intrinsic parity 7t = ± 1 we have

T(I)lkct ) = 7tL(Yt)PaII~P ) (15.141)


p
It is soon verified, see problem 15.15, that T(I) commutes with Y.1l so that the
space inversion carries one solution of the Dirac equation into another
solution of the same equation.
The inclusion of an electromagnetic field is achieved as in subsection 15.8.3
by replacing ~ by ~ + eA/c where as usual ~ = inP.
We conclude this section on the Dirac equation by emphasising that,
in the four-component framework, the transformations separate into orbital
and spin factors and consequently the infinitesimal operators are sums
X = X, + X,; Y = Y, + Y, of differential operators and 4 x 4 matrix oper-
ators. From the matrix M given in equation (15.128) and the explicit form for
the infinitesimal operators in the Lorentz group (see subsection 15.2.4) we
have, for the matrix operators,

0) =
S
x
1
2YyYz , etc. (15.142)

and

.
(Y s)x=-I(Ax-B x)=- 0
(s., -sx0) =- 2Yt Yx '
1
etc,

For the differential part we need an operator to represent the infinitesimal


change from Ik ) to ILk ) when the Lorentz transformation L is close to the
identity. For this part of the operator we may ignore the component labels ct.
Thus for a general state 11/1 ) , using equation (15.136), the transformed state is
given by
15.8 .5 Space and Time 387

where we ha ve changed variables for the integration and made use of the
inva ria nce of the vo lume element, see subsect ion 15.4.4. Thus the function !/J(k)
>
describing the state I!/J is carried into !/J(L - 1 k). Ifwe now write L = 1 + A, as
in subs~tion 15.2.4, for infinitesimal A, and develop a Taylor expansion for
!/J(L - 1 k) as we did for rotations in subsection 7.3.5 then we deduce the explicit
differential forms.

( X' )x = -(k y ~(k;


( z
-k= I(k~
(y
)' etc.

( Y,)x = (k,D~l + kl -j)~J' etc.

The first of these is ju st th e familiar form fo r the rotation operator (7.21) when
expressed in the momentum representat ion , while the second gives the
corresponding expression for the infinitesimal boosts.

15.8.5 Particles with zero mass and spin 1m/ =-t - the Weyl
equation

The object here is to construct a factorised form for the representation pIO,m)
of the Poincare group, see subsections 15.4.2 and 15.4.3, when Iml = t. We
naturally con sider the product of the represent ation pto. O) with u t O ) or UO. 1),
where we stress that pto.o, do es not den ote the identity representation but
p lk.O ' with k = 0; in other words k = 0 but k t= O. To be definite, we consider
the product ptO.OIX Li o. ~ ) which we shall find to be appropriate for the case
m = - t corresponding to the neutrino. Since p' O.m l has only one basis vector
for each k while p'o. 0 , x U o. ~ ) has two , it is clear that these two representations
ca nno t be eq uiva lent but we sha ll a im to co nstruct the former fro m vectors of
>.
the latter. If we use th e nota tion I}(m as in subsection 15.4.2 for vectors of
p tO.m), and IklJ. >0 with IJ. = +
- 12 for basis vectors of p,o,o'x Lio.ll then we show
that

Ik - t > = I L~o:,,!I (RxyQz) lklJ. > o (15.143)


x

where, as in subsectio n 15.4.2, the Lorentz transformation RxyQz carries the


t.
vector (0, 0, 1, 1) into This equation is analogous to equation (15.125) and
generalises immediately for any (negative) m. For m = + we must use Ll!·Ol t
instead for reason s which are explained below, The ju stificat ion for equation
(15.143) follow s tha t given for equation (15.125), with
T(t', L) Ik - t>= exp(ik' . t') Ll~}~![ ( R ~y Q~ ) - l LRxyQz] !k' - t>
(15.144)
The absence of any sum on the right ar ises from the fact that the group
operation (R~ y Q: )- 1 LRxyQ= is an clement of the subgroupS 2' see subsection
15.4.2, whose infinitesimal operators were given in equation (15.62) by
X=, X, - Y)' and X). + YX" But in U o.!), X = - iY so that X, - Y)' = X, - iX).
388 Space and Time 15 .8.5

and X, + Y, = i(X., - iX y ) which are both lowering operators for m and give no
contribution for m = - 1.This leaves the single fJt 2 operator X, which leads to
the simple exponential transformation (15.68) for Ik -1).
We now show how it is possible to arrive at this same result (15.143)
by imposing a wave equation constraint on the two-component space. We
recall from section 15.4 that p(O .m) is characterised by equation (15.76),
CW+imP)IIjJ )=0 in addition to W.W=P .P=O. It is clear that the
product representation P(o.O)x Uo.!) satisfies the latter pair of equations but
the constraint (W + imP) 11jJ ) = 0 leads to a 2 x 2 matrix equation which must
be satisfied . To find the matrix form for W we can start from the definition
(15.72)ofW and use the second two rows and columns of equation (15.142) for
the matrix forms of X and Y. (Remember that the differential operator part of
W vanishes identically , see subsection 15.8.3). In this way the cond ition (15.76)
becomes
(15.145)
and
(15.145a)
This is really a set of four equations since (15.76) is a four-vector equation.
However, if the time-component equation (15.145) is satisfied then it follows
that for m = -1 the other three equations (15.145a) are also satisfied, as may
be seen by multiplying equation (15.145) by s. Thus the vital condition to be
sat isfied in the case m = - 1 is equation (15.145) and this is known as the Weyl
equation. Inserting the differential operator forms for P it takes the more
familiar form

{:t - 2eV.s }11jJ) = 0 (15.146)

To find a solution for the case m = +1 which, see subsection 16.3.7,


corresponds to the antineutrino one must use the representation U!·O) but the
method is otherwise unchanged.
We now return to the form (15.145) to study a plane-wave solution inserting
the values - iii for P. With the explicit matrices for s it is simple to solve the
matrix equation (15.145) and obtain a solution which is unique, apart from
normalisation,

Iii - 1 ) = {2(k t + kz }} -! ( -k~~+k:ky) Iii ) (15.147)

This is consistent with the properties of the representation pO. -!) in which, for
given ii, there is only one basis vector. Here we have only one acceptable vector
Iii -1) but it is expressed in a two-component framework . In fact, equation
(15.145) may be written as (k . s) !IjJ ) = mlkllljJ ) which says that 11jJ) is the
eigenstate of the component of spin in the k-direction, with eigenvalue m. The
expression (15.147) is seen to be identical with (15.143) if one inserts the explicit
15 .8.6 Space and Time 389

form for the uo.!) matrix, see subsection 15.2.5.


It is possible to arrive at this same result by starting from the four-
component Dirac framework in the limit k --+ 0 and imposing the additional
constraint ('W + imP) IIjJ ) = 0 with m= -1. Considering again just the time
component of this equation and using the expressions (15.142) for X this
constraint reduces to
(15.148)
where for convenience we have introduced the symbol )' 5 to denote the matrix
= i )',)'x )'y )'z' Since one quickly verifies that the explicit form of this 4 x 4
)' 5
matrix is v, = (b _?), where ± 1 are 2 x 2 unit matrices, it is clear that the
effect of the constraint (15.148) is simply to restrict !1jJ ) to those vectors with
zero in their first and second components. In other words, it takes us back to
the simpler product representation plO . 0) x uo.!) used above.
For the scalar product we may make use of the discussions for the Dirac
equation, since the solutions (15.147) of the Weyl equation are a subset of the
Dirac solutions for k --+ O. The form (15.137) for ( cP IIjJ ) would be unsuitable
since it would always vanish, but a slight modification of the equivalent form
(15.139) by the removal of the invariant constant k leads to a non-vanishing
definition for ( cP IIjJ ) . In fact, one might have removed this constant from both
the earlier forms for ( cP I1jJ ). Notice that the sum in (15.139) now extends only
over ex = 3 and 4 since all first and second components vanish . The
normalisation constant in equation (15.147) was cho sen to give
( k'mlkm) = k/)(k'-k)
Note the similarity between the solution (15.147) and the solution (15.119)
for the Euclidean group, remembering that, for zero mass, k, = Ikl. This
similarity is, of course, to be expected in view of the interpretation given in
subsection 15.7.2 that zero mass particles spin about their direction of
motion. To illustrate this , take k; = ky = 0 and k, > 0 so that k, = kz ' giving
Ik - 1) = (?) Ik) and hence the spin about the direction of motion is negative.
If, on the other hand, one takes k; = k; = 0 with k: < 0 so that k, = - k; then
Ik-1 ) ex (b)ik ), making the spin about the z-axis to be positive, but since the
particle is now moving in the negative direction it still has negative spin about
its direction of motion.

15.8.6 Particles with zero mass and spin Iml = I -the


Maxwell equations

For this case we follow the previous section very closely with m = 1 in place
of m = t, using the product representation P(o. 0) x Lu , 0). This gives us a
three-component space and we must impose the constraint (W + imP>IIjJ) = 0
with m = l. In constructing the 3 x 3 matrices for Wwe must now use the
matrices for X anti Y in the L(\ ' 0) representation. However since in Lll, 0) we
have B = Oin the notation ofsubsection 15.2.4 it follows from equation (15.33)
390 Space and Time

that Y = iX. But so far as X is concerned, L(l· 0) is simply the representation


D' 1) of .CJt 3 . Thus the matrices of X are as given in equation (7.24) and this also
determines the matrices of Y and W. One has the choice of two familiar bases
for 0(1) and we prefer the Cartesian basis, as used in equation (7.24). The
alternative would have been the m-basis as used, for example, in equation
(8.15). As in the previous section there is some redundancy in the four
equations (W + iP) I'" > = 0 and after a little elementary algebra one finds that
they take a concise form in a vector notation, with the three-component wave
function written as a vector I"' >,

(15.149)
and
P .I"' > = 0 (15.150)
If we work in configuration space, introducing the differential forms for P,
these equations become
i a
-;;-1'" >
curl'''' > = c ot

div I'" > = 0


Ifwe use a plane wave basis so that for definite Ii the solution has the form
I"' > =(Olli > then the equations (15.149) and (15.150) provide a matrix
c f'
equation for the coefficients a, band c by writing, as before, P == - ik . If for
A

simplicity we choose k along the positive z-direction, so that k, = k, then one


finds that the three-component structure of It/! > is given by 0)= (t) . Since
this vector is clearly orthogonal to the direction k of motion one refers to the
solution as 'transverse'. This is, of course, a general property following directly
from equation (15.150). As we should expect for m = 1, it represents a state of
positive helicity with spin = 1 a bout the direction of motion.
Only a trivial change of sign is needed to adapt the above to the case
m = - 1. In nature, the photons, which are the field particles responsible for
electromagnetic interactions, have zero- mass and m = ± 1. Their motion is
therefore governed by the Maxwell equations given above (see also subsection
16.3.7).

Bibliography

For more details on special relativity and references to original papers see

Muirhead, H. (1973). The Special Theory of Relativity (Macmillan, London)

A rigorous discussion of the Lorentz and Poincare groups runs into deep
water but for further reading we suggest Boerner (1963), see bibliography to
chapter 4, and the article by T. O. Newton in
Space and Time 391

Kahan, T. (1965). Theory of Groups in Classical and Quantum Physics (Oliver


and Boyd, Edinburgh)

and the references which they contain. The original papers of E. P. Wigner
are conveniently collected as reprints in

Dyson, F. J. (1966). S ymmetry Groups in Nuclear and Particle Physics


(Benjamin, New York)

An account of the use of the Poincare group including a careful discussion


of phases is given by

Halpern, F. R. (1968). Spe cial Relativity and Quantum Mechan ics (Prentice
Hall, Englewood Cliffs, N. J.)

The role of the wave equations is also discussed in

Omnes, R. (1969). Introduction to Particle Physics (Wiley-Interscience,


London)

A convenient summary of the properties of the Lorentz and Poincare


groups and their representations is given in

Lomont, J. S. (1959). Applications of Finite Groups (Academic Press, New


York)

Problems

15.1 WritingP(p)= 1+ ~ P< + I/ P" +( Pz and R(a) = I+La q X q f o r small p and a,


q
use equation (15.5) together with equat ion (7.22) to deduce the commutation
relations (15.15). (Keep terms to first order in a and p.)
15.2 From the matrix definition (15.23) of the lorentz transformations, show that
they form a group.
15.3 From equation (15.26) show that the boost matrix element Q,., is given by
-uysinh b.
15.4 Show that , if Cl > z, it is not possible to find a proper lorentz transformation
which carr ies (0, 0, z, cr) into (0,0, Z, - cr).
15.5 Write down the matrices of Az and B, in the representation l(H) using the basis
e~ .. Hence deduce the matrices of X, and Yz and show that they agree with
equation (15.30) with a suitable change of basis.
15.6 If the translation PI€)of equation (15.48)is expressed in term s of the components
e';
Bi of t with respect t~ a new basis = lei ' show that the corresponding
infinitesimal operators P i are related to the Pi in the original basis by the usual
transformation P- i= L(l -1 )ijP j . In other words , the Pi transform like the
j

components of a four-vector. (Use equation (15.23)and the relation for Bi given


just before it.)
392 Space and Time

15.7 Use equation (15.53) for infinitesima l a, band t , together with equation (15.48)
and the matrix (15.30) of !I. = L - 1, to deduce the co mmuta tion relations (15.69).
15.8 Use the commutation relations (15.32) and (15.69) to ju st ify equation (15.72).
15.9 The components P, of the four- vector P satisfy the relati on T(L)PjT -1(L )
= L (L - I )jj P j with a similar relation for W. Hence sho w that equation (15.76)
j "' '' A "' .......
implies that (W + imP)lkm ) = 0, where Ikm) = T(L) lkom).
15.10 Given that y211/1 ) = exp(iex)I"') , Show that exp(iex) = ± 1. (Consider the
operation of y2 on the state 11/1 ) + Ie/», where Ie/» = Y jl/l ) , and use the
antiunitary nature of Y.)
.
15.11 Show that the 2 x 2 matrix form for the operator 2(s.l) IS. (1 L) 0

1+ -1 0
and hence
verify that the eigenvectors Um ) of equation (15.116) are as given.
15.12 Use equation (15.33) to show that, in the L(!D) representation, X = iY = - isand
hence from equation (15.72) that W; = !p2 , W 2 =!p 2 +~P; so that
(W.W) = ~P .p).
15.13 Show that the matrix Y, defined in equation (15.129) commutes with the X, and
hence that it is the time component of a representation L(! i). Use equation
(15.69) to generate from Y, the other three components Yq.
15.14 Use the 4 x 4 matrices of Y in the basis of U!,Of-+- UD,!) to show that the finite
boost Q(b)=exp(b.Y) takes the form cosh!b+2b -l(b ,Y) sinh}b. Hence
deduce the solutions (15.135) from equation (15.134), taking b(k) as given ju st
after equation (15.58).
15,15 Show that y,yqy,- I = - Yq and hence that the i~version operator T(I) defined just
before equation (15.141) commutes with (y.P).
16
Particles, Fields
and Antiparticles

The concepts of momentum, energy, mass and spin were introduced in the
previous chapter entirely from the transformation properties of the Poincare
group. We worked in the algebraic framework of state vectors, which is
appropriate in quantum mechanics, and we associated different elementary
particles with different irreducible representations of the group. In introducing
these concepts we did not make use of any fundamental equation, like the
Schrodinger equation, which dominated the discussion of symmetry in the first
twelve chapters. In fact in section 15.8 we were able to deduce the forms of the
wave equations necessary to reproduce the transformation properties pre-
viously described. These comments only relate to free particles, however, and
when it comes to interactions it is necessary to have some dynamical
foundation. The foundation which seems to have universal validity is known as
Lagrangian theory or equivalently as Hamiltonian theory and rests on
Hamilton's principle. This principle in its simplest form is applicable to the
motion of particles in non-relativistic classical mechanics. It extends naturally
to relativistic classical mechanics of particles. In the description of elec-
tromagnetic phenomena it extends to incorporate the concept of interaction
with a field and the behaviour of the field itself. Finally it is generalised into a
393
394 Particles, Fields and Antiparticles 16 .1

quantum mechanical framework in which the concepts of particles and fields


are unified and one finds the need to introduce 'antiparticles'. This brings us to
the point in quantum field theory where many problems remain unsolved .
Necessarily therefore this chapter will be somewhat more tentative than those
which have gone before.

16.1 Classical mechanics of particles

16.1.1 lagrange formalism

The motion of particles in classical mechanics is governed by the principle of


least action, or Hamilton's principle. This asserts that to every mechanical
system one may associate a Lagrangian funct ion Liq. , qj, r) which depends on
the coordinate s qj, their time derivatives qj a nd the time . The motion of the
system is then given by those functions qj(t) which lead to the integral
l,

S =
I
I,
Let (16.1)

being a minimum for any two times t 1 and t 2 ' This is a variational principle. To
ensure a minimum (or at least an extremum), the change in S due to a small
change in the functional form of any of the qj(t) must be zero and we write this
condition as ~S = O. The quantity S is called the 'action'. Using the variational
calculus it may be shown that the condition ~S = 0 leads to the Lagrange
equations

:t G~) -:~ = 0
( 16.2)

Given the form of L, these are a set of differential equations for the coordinates
qj(t) as funct ions of time . There are as many equations as coordinates and,
given the initial conditions, a unique solution for the q j(t) may be found. Hence
the motion of the system is determined directly from the Lagrangian. One
system differs from a nother by the form of L as a function of q., qj and t .

16.1.2 Hamiltonian formalism

The Lagrangian formalism refers to coordinates and velocities but it is


sometimes more convenient to work in terms of coordinates and momenta.
This is particularly so in connection with symmetry and the link between
classical and quantum mechanics. There is in fact an alternative, but precisely
equivalent, formalism due to Hamilton which uses momenta rather than
velocities. One first defines the generalised momenta
eL (16.3)
pj = eqj
16 .1.2 Particles. Fields and Antiparticles 395

Using the Lagrange equation (16.2) the differential dL may then be written

oL
= ~ pjdqj + ~ p;dq j + Jt d t

but d (L Pjqj) = L qjdp i + L pjdq j


I I I

so that the difference

does not contain differentials dqj of the velocities. Hence the quantity
t Pjqj - L does not contain the velocities explic itly and we call it the
I

Hamiltonian

(16.5)

From equation (16.4) we have


oH oH
- = -po
° Pi = qj, (16.6)
oqj I

These are the Hamilton equations. Whereas the Lagrange equations will
involve the second time-derivatives of qj (since 0L joqj will generally contain
q;), the Hamilton equ ations involve only first time-derivatives of Pi and qj.
However there are now twice as many equations. Naturally, the two techniques
lead to precisely the same solution.
From the point of view of symmetry one sees that if L (or H) is independent
of a certain coordinate qj so that oL jo qj = 0, then the corresponding
generalised momentum Pi is constant, since from either equations (16.2) or
(16.6) we have Pi = O. Notice also that, from equation (16.4),

(16.7)

Thus if L (or H) is independent of time (explicitly) then H is constant. As one


might expect, this constant which is associated with homogeneity in time, may
be identified with the energy of the system in the same way that the space
homogeneity 0 L loq, = 0 leads to constancy of the momenta Pi'
We have already seen a simple example of the use of the Lagrange equations
in describing molecular vibrations in section 6.2.
396 Particles, Fields and Antiparticles 16.1.3

16.1.3 Examples from relativistic mechanics

Consider first a free particle of mass M . We postulate that S should be a


Poincare invaria nt, which is ju st the statement that the mot ion does not
depend on the position , or ientatio n or uniform velocity o f the system in space.
Since the integral (16.1)appli es to all intervals this invariance mu st apply to the
integrand. In the not at ion o f sectio n 15.2 the coo rdi nates o f a pa rticle are
denoted by
e = (x , y, Z, et )
and de = (dx, d y , d z , ed t) = (o, ; v V
y, z, e)d t
so that ldel = (de. de)1 = (e2 - v2 )1dt
where v is the velocity. From its manner of construction, Idel is an invariant so
that at leas' -~om the invariance point of view the expre ssion (e2 - v 2 )! would
be a suitable form for L. The quantity Idel is the 'interval' between
neighbouring points e and e+ de on the path of the particle in four-
dimensional space. It is like the small distance d s = (dx? + d y 2)1along a curve
in two-dimensional space. In fact we see that the choice
L= -M e(c 2- v 2)1 (16.8)
leads precisely to the usual equat ions of motion for a particle of mass M . In
this formal ism, the mass would be defined as the constant M appearing in the
expression (16.8) for L. From the Lagrange equation (16.2) the ab sence of x, y
o r z in L lead s di rec tly to the co nsta ncy of t' L /iJ.x, e tc. giving

p" = M v., (I- v 2/ e 2) - ! = constant, etc. (16.9)


and for the energy

E=H = L P;l/ j -
i
L = M v 2(l- v 2/e 2)- !+ M e 2(1- v 2/ c 2)!

(16.1 0)

From equations (16.9) and (16.10) we deduce the familiar energy-momentum


formula
E 2_p2e2=M 2e4

as in the quantum case, see equation (15.94). For small ole, equation (16.10)
reduces to the familiar expression
(16.11)
containing the rest-energy M e and the kinetic energy.
2

As a second simple example we introduce an electromagnetic field which we


assume to be described by a four- vector potential A = (A.o A y , A Z' ¢). To
16.1 .3 Particles, Fields and Antiparticles 397

describe the motion of a particle with charge e in this field we must modify the
Lagrangian (16.8) to include A but to remain invariant. The most obvious
invariant to include is the scalar product
A.M = (- A.v+e4»dt
The conventional equations describing the motion of a charged particle in an
electromagnetic field are obtained by including just this term, with a suitable
constant - etc so that
L= -M e(e 2-v2)!+(e l e)(A .v)-e4> (16.12)
The generalised momenta are now given by

aL
P.• = ---at = M v..(1 - v [c )-, + (ele) A,
2 2 1
(16.13)

and the Hamiltonian by


H = M c 2(1- v2Ie 2)-!+e4>
=M e2{1+(p-eA le)2IM2e2}!+e4> (16.14)
Application of the Lagrange equation (16.2) to the Lagrangian (16.12) now
leads, after some algebraic manipulation, to the familiar equations of motion
for a charge in a field. In the non-relativistic limit v ~c one finds
.. a
mv = - (el e) alA - egrad 4> + (ele) v 1\ curl A (16.15)

which is often written in the form


mif=eE+(ele)v I\B (16.16)
by defining

E
a
= - - A le - grad 4> B = curl A (16.17)
at '
The vectors E and B are called the electric and magnetic field intensities. Since
it is equated to (mass) x (acceleration), the right-hand side of equation (16.16) is
called the Lorentz force.
The differential operator 0= (0 lox , 010 y, 0lo z,- oloct ) is a four -vector, see
subsection 15.4, so that the sixteen products O JA j transform like the product
representation U H)x L(t.!l of the Lorentz group. Being first derivatives of the
components of A, it is clear that E and B are contained in these sixteen
products and using the reduction (15.37) of the product representation one can
identify E ± iB with the representations U 1.0) and Uo. 11, respectively, see
problem 16.1.
An electromagnetic field has meaning only by the way in which it influences
the motion ofa charged particle, through equation (16.16). Thus any two fields
A which lead to the same E and B are physically the same field. From equation
398 Particles, Fields and Antiparticles 16 .2

(16.17) one sees that this leaves A undetermined to the extent that iff(r, t) is
any scalar function then the vector potential A' = (A', ¢'), where

A' = A + gradf, ¢' = ¢_~ of


c at
will describe the same physical situation as A. The freedom to choose f is
referred to as the choice of gauge .

16.2 Classical mechanics of fields

The concept of a field is introduced to describe interaction at a distance. Thus if


we have one charge e fixed at the origin and another charge e which is free to
move we may say that the first charge produces a field of force with strength
e2 /r 2 and directed away from the origin and the second charge moves in this
field. In this case the field is static but more generally a field may also depend on
time . Furthermore a field of force, such as the example above, has a direction at
every point. In fact in this example the direction is always given simply by the
vector r but more general fields may be imagined. As is well known the force
field in this example may be described by a potential energy field V(r) = e2/r
with the force given in both magnitude and direction as F = - gradV(r).
Notice that the field V( r) does not have a direction and is an example of a scalar
field while F is an example ofa vector field. The small displacements q which we
used in studying molecular vibrations in chapter 6 are another example of a
vector field, except that it is not a funct ion of a continuous variable r but
associates a vector displacement only with the discrete set of equilibrium
positions of the vibrating atoms. We now define more precisely the concepts of
scalar, vector and more general fields.

16.2.1 The transformation of fields

If under a rotation R a field ¢(r, t) is transformed into


¢ '(r, t) = ¢(R - 1 r, t)

then ¢ is said to be a scalar field, with respect to rotations. If, on the other hand,
the field has 25 + I components ¢m(r , t) with m = 5, 5 - I, ... . - 5 such that
the m-component of the rotated field is given by

¢~(r, t) =~ D~~ (R)¢,;,(R - 1 r, t) (16.18)


m

then the field is said to be a tensor field of degree 5. This is the same
transformation that we deduced in equation (8.20) for the components of a
wave function with spin 5 . The most familiar example is the vector field with
5 = I so that at each point in space the three components define a direction for
the field. It is of course possible to construct fields which are mixtures of
16 .2.2 Particles, Fields and Antiparticles 399

different s. A field with fixed x is irreducible with respect to rotations and


transforms like the irreducible representation D's).
In a four-dimensional framework, fields may be classified by their behaviour
under Poincare transformations. In the not ation of subsection 15.4.2 we would
call ¢(e) a scalar field if the Poincare transformation produces a new field

and tensor fields would be classified by the representation L ci .r , to which they


belong, in a natural extension of equation (16.18),

(16.19)

16.2.2 The Lagrange equation for fields

In the second example of subsection 16.1.3 we included a field as part of the


Lagrangian but it was assumed that the field was given externally to the
problem. We were concerned only with the motion of a particle in the given
field. We now consider the way in which fields may behave on their own, i.e. in
free space. The more general problem, in which the fields interact with particles
whose motion in turn affects the fields, is clearly more complicated and we
discuss it only very briefly . The equations governing the behaviour of a field
may be deduced from the variational principle (16.1) with a suitably chosen
Lagrangian. The Lagrangian will now depend on the field components ¢ mm ' (e)
rather than the coordinates qj and will be written as an integral over all space.
For a scalar field,

(16.20)

where !f' , which is a function of ¢ and its space and time derivatives V ¢ and
<P ( = c ¢ Ii" t) is called the Lagrangian density. Application of the variational
principle 8S = 0 now leads to a generalisation of the Lagrange equation

-[1 ( c c!f' - c!f'


- ,2 ' ) +V · _ ,-=O (16.21)
ct?¢ ?V¢ r ¢

which is called the field equation.


Just as equation (16.2) gave the equations of motion for a particle, so
equation (16.21) is the equation describing the field. If, as is generally the case,
we choose !f' to be a quadratic function of ¢ , V¢ and <P then equation (16.21) is
a linear partial differential equation of, at most, second order for the fields
¢(r, t). Thus the ordinary differential equations of section 16.1 for the particle
coordinates qj(t) are replaced by a partial differential equation for the field
¢(r. t) . The extension of the field equations to tensor fields is straightforward.
400 Particles, Fields and Antiparticles 16.2.3

The symmetry of fields


In constructing the Lagrangians for particles in subsection 16.1.3 we made
use of the Poincare invariance of the integrand Ldt in the action integral. The
same principle applied to field Lagrangians, like that in equation (16.20), will
lead to the invariance of the Lagrangian density !fl because the volume
elementd Vdt is an invariant. (The determinant of the Lorentz transformation
has the numerical value 1.) In other words !fl must satisfy an equation which
may be written formally as!fl [¢(e)] =!fl [¢ '(Le +~)], where for brevity we
have omitted derivatives. A scalar field as defined in subsection 16.2.1 will
satisfy this requirement but for tensor fields and for the derivatives V ¢ and ;P,
the Lagrangian density must contain sums over components to form a scalar.

Conservation laws for fields


Just as in subsection 16.1.2 we defined the Hamiltonian and found it to be a
constant of the motion, so for fields we shall be able to define the Hamiltonian
density

Then, from the invariance of :£ under time translations we are able to show
that H = Itf d V is constant. In very similar fashion, if the Lagrangian density
is translation invariant in space , then the vector P with components
I
r, = Y'qd V is constant where
etc.

The proof of these results is outlined in problem 16.2. It is natural to call & q the
components of the momentum density of the field because they are the
constant quantities arising from translation invariance. The conservation of
angular momentum follows in a similar way from the rotational invariance of
!fl .

16.2.3 The electromagnetic field

As an example, consider an electromagnetic field which, as we have seen in


subsection 16.1.3, is described by the four-vector potential A or, equivalently,
by the field intensities E and B. The Lagrangian density must be invariant and
one obvious choice would be A. A. However, this would lead to a trivial field
equation and furthermore we have seen in subsection 16.1.3 that it is the
derivatives E and B of A which have physical significance. From the fact that
E ± iB transform like LC 1.0) and LCo. 1 ), respectively, it follows that E 2 - B 2
and E . B are both Lorentz invariant. Since the latter is not invariant under
space inversion we might prefer the former as our Lagrangian density and this
16 .3 Particles , Fields and Antiparticles 401

is found to agree with experiment. Conventionally, we wr ite


(E 2 _ B2)
.;e = - - - (16.22)
8n
We cannot immediately apply equation (16.21) since that referred to a scalar
field. However if 4> were replaced by the four-vector A in (16.20) the field-
equation (16.21) would be found to apply separately to each of the four
components of A. Noting the definitions (16.17) of E and B we see that the
expression (16.22) for .;e is a quadratic in the first derivatives of A. After some
vector algebra the four field equations reduce to
I (CE
curl B = - - .
(' 1'1
and
(16.23)
div E =0
These are just the Maxwell equations which describe the behaviour of a field in
the absence of charged particles. The other two Maxwell equations were
contained in the definitions (16.17) and follow directly from them,

div B =0 and
i es
curl E = _ . - - (16.24)
(' at
We discussed in subsection 16.1.3 the behaviour of a particle in a field which
was regarded as fixed, in the sense that it was not influenced by the particle.
Above we have described the behaviour of the field in the absence of charges. In
general we mu st allow for the effect of the charged particles on the field and
this is done formally by taking a Lagrangian density which is a sum of the field
term (16.22) and a term like (16.12) integrated over all charged particles
present. One then deduces field equations which are generalised from equation
(16.23) by the inclusion of terms arising from the presence of A in (16.12) and
which represent the charge and current density due to the charged particles.
Thus the motion of the particles, through equation (16.16), and the motion of
the field are coupled.

16.3 Quantum fields

Classically, a field is introduced to describe the interaction between particles.


For example, a charged particle will produce a field and another charged
particle will interact with the field thus effectively producing an interaction
between the two particles. It is well known that, at small distances, one must
use quantum mechanics for describing the motion of particles and this impl ies
some imprecision, as given by the uncertainty principle. There is therefore
some inconsistency if one describes the motion of a particle in a field by using a
quantum mechanical treatment for the particle and a precise classical
402 Particles. Fields and Antiparticles 16.3.1

description for the field. We therefore seek a quantum treatment of the field.
The technique is to replace the field by a field operator and to consider matrix
elements of this operator as is the usual practice in quantum mechanics. The
form which the operator takes is a combination of what are called creation and
annihilation operators offield particles. Thus in addition to the particles which
a re interacting via the field we introduce the new field particles. For example,
the photon is the field particle of the electromagnetic field and the n-meson the
principal field particle of the strong nuclear forces. The introduction of
operators which create and annihilate particles is not simply a mathematical
device . Such processes are commonplace both with the zero mass photons in
the emission of light and for finite mass particles such as the n-meson in the
high-energy collision of protons p + p -+ p + P + nO . The creation of mass does
not violate the Poincare invariance of the system, see problem 16.4. In a
complete theory one should also describe the interacting particles themselves
by a field so that, in the example above, one would have a mechanism which
allowed the creation of more protons as well as n-mesons and aga in such
processes are observed.

16.3.1 Second quantisation

This subsection introduces the algebraic framework used to describe the


creation and annihilation of particles. We have already noted in subsection
8.6.2 that the wave functions of a collection of identical particles are found to
be either totally symmetric or totally antisymmetric under exchange of
particles. One observes that particles with integer spin (called 'bosons') are
symmetric, while those with half-integer spin (called 'fermions') are anti-
symmetric. This feature is also discussed in section 5.9 and the connection
between the spin value and the symmetry (or statistics) is explained in
subsection 16.3.4. Because of this complete symmetry or antisymmetry of the
wave functions and the symmetry of all physical operators, the numbering of
particles has no physical significance and serves only as a device for ensuring
the relevant symmetry. It is therefore not too surprising that one can construct
a technique in which the particle numbers do not appear at all and the
symmetry is achieved by insisting on certain multiplication rules for the
operators which create and annihilate particles.
To be definite, and for simplicity, let us consider the case ofbosons with zero
spin. The modifications necessary for the case offermions is given at the end of
the section and the extension to non -zero spin is straightforward. For particles
offixed mass M = hk]«: the complete set of single-particle states is denoted by
Ik) where the length k of the four-vector k is fixed. A symmetric state of n
particles may be denoted by IkI k2' . . . , kn ) , where the set of n labels k;
describe the n states occupied by the particles (the labels need not all be
different). The state with no particles present is called the 'vacuum' and
<
denoted by 10) . It is normalised 010) = 1. (There should be no confusion
between the notation 10) for the vacuum and Ik) for the single particle states
16 .3.1 Particles , Fields and Antiparticles 403

since, as remarked in subsection 15.7.1, k = 0 is not acceptable for a particle.)


The vector space (called a 'Fock space') in which we are working is not now
confined to a fixed n but includes all possible n. The number of particles mu st
thus be regarded as an operato r N which will not generally be di agonal and n is
the eigenvalue of N in a state of n part icles. T o exploit thi s larg er vector space
we mu st clearl y introduce opera tors which couple part s of the space with
d ifferent n. The simplest of the se is an operator, denoted by a' (k ) with the
propert y

a t(k) IO> = Ik > (16.25)


>.
and called the creation operator for a particle in state Ik On a more general
sta te of n particles the d efiniti on is
t(k)
a C1kk
lk tk 2 , . • . .k.> = t k 2 • ••• ,k. > (\6.26)
and we shall not concern ourselves with the normalisation factor C. Because of
the assumed symmetry of the wave function it follows from equation (16.26)
that the commutator
[ at(kJ, at(k 'l] = 0 (\6.27)
The Hermitian adjoint a(k ) of at (k) must reduce the number o f particles by
one and is called the annihilation (or destruction) operator. It therefore gives
zero when operating on the vacuum, a(k) 10 > = 0. By tak ing the adjoint of
equation (16.27) we have
[ a(k ), a(k) ] =0 (16.28)
For the commutator between crea tio n and annih ilation operators it is quickly
seen that

so that in the vac uum sta te we may write


[ a(k),at(k ')] = ( fi lk'> (16.29)

With the help of equation (16.26) it can be sh own that this co m muta tio n
relation is valid not only in the vacuum but generally. With the normalisation
(15.81), it takes the form
[ a(k), at(k'lJ = k,8(k -k ') (16.29a)

Physical operators may now be expressed in terms o f these crea tio n and
annihilation operators. For example, an operator which d oes not change the
number of particles must co ntain a n equal number of a and a" factors. In
particular, the number operator is given (see problem 16.5) by

N = J at (k) a(k)k ,- I dk (16.30)


404 Particles, Fields and Antiparticles 16 .3.2

As in subsection 15.4.4 the range of integration extends over all k with fixed
length (mass) and positive k, so that k, = (k 2 +jkI2)!.
From the symmetry point of view the Poincare transformations will not
change the number of particles so that, if we assume the vacuum to be unique,
it must be invariant. Thus one would expect that the creation operators at (I,)
transform like the single-particle states, according to equation (15.59), and this
may soon be justified using equation (16.26). For the zero-spin case being
considered here,
(16.31)
In the case of ferrnions, for which the wave functions must be antisymmetric
rather than symmetric, the results of this subsection are modified simply by the
appearance of the anticommutator [A, B] + = AB + BA in place of the
commutator [A, B] = AB - BA. The need for this modification may be seen
in the step from equation (16.26) to (16.27).

16.3.2 Field operators

We now seek a linear combination of the operators a' (k) which represents
e
the creation ofa particle at a point in space - time rather than in a state Ik).
This new operator is denoted by ¢' (e) and is called a 'field operator'. From
this description the transformed operator must be the operator at the
transformed point Le so that we shall require
(16.32)
Using the property (16.31) one quickly verifies that the necessary linear
combination of the a '(k) is given by

¢t(e) = fex p (ik.e)a'(k)k/- I dk(2rr) - l 2- 1 (16.33)

with the conventional normalisation. The verification is straightforward,

TlE, L)¢' (e)T-l~, L) = feXP[i(k.e + Lk.E)]a' (Lk)k,-ldk(2rr) -j 2-1

= fexP[ i Lk.(Le + E)]a' (Lk)k,-ldk(2rr) -12 -!

= feXP[ik.(Le + E)]a' (k)k,-ldk(2rr) - 12 -1

= ¢'(Le+ E)

where we have changed integration variables Lk ..... k and used the invariance
(see subsection 15.4.4) of the volume element k r-I st:
Bearing in mind the
discussions in subsection 16.2.1 we shall say that the operators ¢ t (e) describe a
scalar field. (The reason why we have Le appearing in equation (16.32) and
16 .3.3 Particles, Fields and Antiparticles 405

L - Ii in subsection 16.2.1 is that ¢ \i) is an operator whereas ¢(i) in


subsection 16.2.1 denoted the value of a classical field at i .)
In the case of non-zero spin one may similarly construct field operators ¢a (i)
with several components (X which transform according to

(16.34)

where M is a representation of the Lorentz group. This equation implies a


factorisation between the transformation of the indices (X and the transfor-
mation of the fields (i.e. the fact that M does not depend on i) which is the same
feature that was discussed in subsection 15.8.4 in connection with the Dirac
equation. Thus again for a finite-mass particle with spin s = 1. four com-
ponents are required in a field of the type (16.34). We shall discuss the non-zero
spin case briefly in sub section 16.3.7.

16.3.3 The physical role of field operators

Before continuing with the discussion of some of the consequences of


symmetry in quantum fields we br iefly illustrate the way in which the field
operators are used in the physical theory.
The first point to clarify is that whereas it is more usual in quantum
mechanics for the operators which represent physical ob servables to be time
independent and the time development of the system to be contained in the
wave funct ion, it is also possible to employ a time- independent wave function
and to have the time development of the system descr ibed by a n explicit time
dependence in the operators. One refers to these two approaches as the
Schrodinger and the Heisenberg pictures, respectively. In relati vistic quantum
mechanics the Heisenberg picture has the ad vantage that the relativistic
invariance may more readily be described when the operators depend on the
four coordinates x, y, z and I. The lack of time dependence in the wave function
is no disad vantage since this acts as the 'initial conditions' on the problem.
Mathematically, in the Schrodinger picture the Hamiltonian H(p , r) is a
function of coordinates r and momenta p with the equation of motion

e
H(p , r) ljJ(r, I) = ih ailjJ(r, I) (16.35)

The Heisenberg picture is obtained, formally, by introducing the transfor-


mation S = exp(iHI/h ) so that, given any initial state ljJ(r, 0), the solution at
subseq uent time 1 is given by
ljJ(r, I) = S - IIjJ(r, 0) (16.36)

< >
The expectation value ljJ(r, 1)1 Q IIjJ(r, I) of an operator Q(P , r) in the
Schrodinger picture may then be reproduced in the Heisenberg picture by
406 Particles, Fields and Antiparticles 16 .3 .3

taking the expectation value of the transformed operator Q ' = SQS - 1 in the
initial state since
< l/J(r , O)I Q ' Il/J(r , O) = < l/J(r, t)I S- l (SQ S- l )SIl/J(r , t)
= < l/J (r, t )IQ Il/J(r, t )
using the fact that S is unitary for Hermitian H. Thus, in the Heisenberg
picture if the time-dependent operators Q ' are known, we may deduce any
required expectation value, given the init ial state l/J(r, 0). Instead of having to
find the wave function l/J(r , t), the problem in the Heisenberg picture is to find
the Q'(P , r , t). But the 'equation of motion' for Q ' follows immediately from
its definition since if we assume Q to be time independent then
dQ '/dt = iH Q '/ h - iQ'H/ n, so that

[ H, Q'] = - in :t Q' (16.37)

which is the Heisenberg equivalent of the Schrodinger equation (16.35). The


transformation S is in fact a finite translation by an amount t in the time
direction, with -iH jnc the infinitesimal operator, see subsection 15.4.1.
Let us now consider the simplest kind of field, a Hermitian (real) field
$ = ¢ + ¢ t, where ¢ t was defined in equation (16.33) in terms of the particle
operators at(k) for particles with mass M = hk ]c and zero spin . (In fact ¢ + ¢ t
is the only combination of these two fields, apart from phase, which satisfies
the causality requirement discussed in the next section.) We also take the
particles to be bosons so that the commutation relation (16.29a) holds. We
must regard the time-dependent field operators $ as operators, like Q ', in the
Heisenberg sense. Picking up the Lagrangian formalism of subsection 16.2.2
for classical fields, we now construct a simple Poincare invariant Lagrangian
density operator
(16.38)

from the field operator $ and its first derivatives 0$. Since $ is a scalar field,
the second term satisfies the Poincare invariance condition on !i' and by
taking the scalar product of derivatives we have ensured that the first term is
invariant also . The reason for having both terms in fi' with the factor k 2 will be
explained below . Following the procedure in subsection 16.2.2 for construct-
ing a Hamiltonian from the Lagrangian we see that
£=!(cD 2c - 2 + V$ .V$+k 2( 2)nc (16.39)
and the momentum density is given by its components
fljJq = cD Vq $nc- 1

From the definition (16.33) of the field operator ¢ t some algebra leads to the
following expressions for Hand P in terms of the creation and annihilation
16 .3.3 Particles. Fields and Antiparticles 407

operators for the field particles of mass M = hk]«,

H = ilic fk, [at (k )a(k) + a(k)at (k) ]k,- 'dk

p = tli fk[ at (k )a(k) + a(k)at tk) ]k,- 'dk


Most field theories suffer from divergence problems and the simple example
above is no exception. In fact the vacuum expectation value <01HIO ) is
infinite. This may be seen by writing a(k)at(k') = a t(k') at(k) + klJ(k - k ') from
equation (16.29a) and using the property a(k)O) = 0, leaving an infinity from
the 8-function. However if one assigns physical significance only to the
difference between the values of H and its value in the vacuum then we obtain
finite results for the operators

J t
H = H - ( 0 1H 10) = lie k,a (k)a(klk,-'dk,
(16.40)
p = p- <OIP IO) = li Jkat(k)a(k)k,-Idk
Comparing the expression (16.40) with the number operator (16.30) one sees
that H may be interpreted as a sum over the possible states k of the number of
particles in that state multiplied by the energy lick, of the particle in that state.
A similar interpretation is valid for the momentum operators P. Thus, starting
from the field description we have arrived at a particle description.
Since $ contains both at and a operators one would have expected H to
contain terms like atat . In fact there is cancellation between such terms arising
from the two pieces of the Lagrangian. This explains our choice of the factor k 2
in !f, since only the absence of terms like a t a t enabled us to interpret H as the
Hamiltonian for a system of non-interacting particles and this was our object.
Naturally a theory of non-interacting particles such as we have given above is
rather empty but by adding other terms to the Lagrangian, as we did in the
classical problems in subsection 16.1.3, one may introduce interactions
between fields and thus between particles.
The conservation of P and H may be shown to follow from the invariance of
!f with respect to space and time translations as was the case for classical
fields, see subsection 16.2.2.
The field described above is appropriate for the neutral nO mesons. Because
some of the arguments may appear to be circular we briefly summarise the
philosophy and content of the discussion of quantum fields so far. In section
15.4 we classified all possible particles according to irreducible representations
of the Poincare group. Using the creation and annihilation operators we next
constructed a scalar field operator ¢J t(i) with the transformation property
(16.32), using the operators for zero-spin particles. From these we constructed
a Lagrangian and hence a Hamiltonian whose eigenstates corresponded to
408 Particles, Fields and Antiparticles 16 .3.4

definite numbers of particles. In doing this we did not make use of the
variational principle descr ibed in subsection 16.1.1 or the field equations
which, as shown in subsection 16.2.2, result from thi s principle. However, if we
were to apply the field equation (16.21 ) to the Lagrangian den sity operator
(16.38) we obtain the di fferenti al equation
(0 2 + k 2 )<I>(e) = 0
This is the familiar Klein - Gordon equation which we met in subsection 15.8.3
in connection with the wave function s for particles. Here it enters in a different
role as a condition on the field operators. From the expression (16.33) for ¢ \e)
one sees that <I>(e) does indeed satisfy this field equation and hence satisfies the
variational principle. (Note the similarity here with our discussion of wave
equations for particles in section 15.8). When interactions are included the field
equations playa more significant role.

16.3.4 Causality and the spin-statistics theorem

In classical relativity theory, no particle or signal may travel with a speed


greater than c. Thus one would expect also that quantum mechanical
measurements at two space - time points eand e' would not interfere with each
other if the d ifference e' - eis space-like, in other words if Ir - r'1> cl t ' - t I.
This property is called 'microscopic causality'. The mathematical statement of
non-interference in quantum mechanics is that the operators shall commute so
that we shall expect
[Q(€), Q(e')] =0 (16.41 )

if e - e is space-like and Q(e) is the operator for some ob servable at e. The


operators Q(e) are usually given , like £ and P/'q in equation (16.39) by
quadratic expressions in the fields and their derivatives. Hence it follows that
the causality condition (16.41) will be satisfied if the field operators either
commute or anticommute for space-like separations. Let us first see whether
the field <I> = ¢ + ¢ t discussed in the previous section satisfies the condition.
From the property (16.27) and the definition (16,33) it is clear that
[ ¢ t(e), ¢ t(e)] = [ ¢(e) , ¢(e)] = 0 so that

[<I>(e), <I>(e ')] = [¢(e) , ¢ \e')] + [¢ \n ¢(e')]


=~ ff k,-I k'(-Idkdk ' { exp [ i(k ' .e -k .el]

- exp[i(k . e- t : e')] } k,8(k - k ')(2n) - 3

= if k r
- I dk "sin] k .(e ' - e) }(2n:) - 3 (16.42)

The integral clearly depends only on the difference e - € and furthermore


16.3.5 Particles, Fields and Antiparticles 409

it depends only on the 'length ' [c 2 ( ( - t)2 -Ir'


_rI 2] ! of e' -e. To see this
we note that any other difference with the same ' length' may be written
as L(e' - e), where L is a Lorentz transformation, but since
t. L(e' - e) = L - I t .(€" - e) and the volume element k I- I dk is invariant (see
subsection 15.4.4) the integral (16.42) is unchanged when e' - e is replaced by
lie' - e). Hence if (16.42) were to vani sh for one space-like difference e' - e it
would vanish for all space-like differences. The simplest case to study would be
t ' = t for which we have simply

[<1>(e), <1>(e')] = <i fkr-Jdksin { k . (r' - r ) }( 2n )- J (16.43)

which clearly vanishes because the sine function is odd and the integral extends
over all three-space of k ,
Thus we conclude that the field <1> built from zero-spin particles with
symmetric many-particle states, and therefore with commutation relations
(16.27) -( 16.29) between c rea tio n and annihilation operators, satisfies the
causality condition. Had we chosen to construct <1> from zero-spin particles
with antisymmetric many-particle states, i.e. with anticommutation relations
in place of (16.27)-( 16.29) then a repetition of the argument above leads to a
cosine in equation (16.43) in place of the sine function . The integral will not
then vanish and causality is not satisfied. Thus, if we accept the causality
condition, we conclude that zero-spin particles cannot obey Fermi statistics,
i.e. have antisymmetric wave functions. The arguments above can be
generalised to the case of any integer spin with precisely the same conclusion.
For half-integer spins a change of sign occurs in the algebra and one concludes
that Bose statistics, i.e. symmetric states, would violate causality. One refers to
these results as the 'spin-statistics theorem'.

16.3.5 Antiparticles

We first remark th at the Hermitian field <1> = 4> + 4> t discussed above is the
only combination (apart from phase) of 4> and 4> t which satisfies causality. A
more general, non-Hermitian field may be constructed only by introducing a
new field operator Xt defined just like 4> t in equat ion (16.33) except that in place
of at(k) we have bt(k) which creates a particle with the same Poincare
behaviour as before, i.e. same mass and spin, but which differs in some other,
internal, property as yet unspecified, e.g. charge or hypercharge. It then follows
as before that the non-Hermitian field <1> defined by
<1> = 4> + X t, <1>t = 4> t + X (16.44)
t
satisfies causality, assuming that the new particle operators b and b satisfy the
sa me commutation relations (16.27H 16.29) as the at and a and that all the new
operators commute with all the old . (For half-integer spin particles the
anticommutation relations apply instead.) The new particles are called the
'a ntipa rticles' of the old ones..
410 Particles , Fields and Antiparticles 16.3.5

A simple Lagrangian like (16.38) may again be con structed, with terms like
<1>t<1> replacing <1>2 to en sure that Y is Hermitian. The expressions for Hand P
then come out to be the sa me as before except that there are additional terms
with the b-operators in place of the a-o pera to rs. Th ere is a lso a new conserved
qu antity

which, when expre ssed in terms of a- and b-operators, becomes

and thus measures the difference between the number of particles and the
number of antiparticles. In a single-particle state Iii > the value of Q is
<IiIQ IIi >= 1, while for a single antiparticle <Q >= - I. The operator Q
clearly distinguishes particle a nd antiparticle and in the case that the field
particles and antiparticles are the IT + and IT - meson s, respectively , thi s
operator will serve to represent the electric charge of the field. On the other
hand if the field is used for the neutral K 0 and K 0 meson s, which differ in
hav ing hypercharges of ± I respecti vely, then Q will represent the hyper-
charge. For the pure K ± field, which has both charge and hypercharge, Q will
be able to represent both cha rges simultaneously. (When severa l different
particles are pre sent, the operator expre ssion s for Q becom e more elaborate
than that given above, with additional terms ari sing from the other particles.)
In this situation the expressions for the electric charge and for the hypercharge
will naturally become different from each other. There are indeed charges
other than these two like the baryon number and the lepton number (see
sectio n 11.3).
Notice that the field <1> has simple properties with respect to Q, with <1>t
increasing the value of Q by one unit and <1> decreasing it by one unit. As a
result we may retain charge conservation even when interaction terms are
included in the Lagrangian. This can be achieved by adding to Y the product
of <1> with a corresponding term of another field which has the opposite effect
on Q . The conservation of charge relates to an intrinsic property of the particle
but we may nevertheless associate charge conservation with a transformation
of the field (as momentum conservation is associated with translations) by
introducing the transformation <1> -+ U<1>U -I, where U = exp(if:Q). Then
since <1> decreases the value of Q by unity, we have [ Q , <1>] = - <1> and hence
the transformation is given by <1>-+ exp( - je)<1>, a con stant phase change. This
is called a gauge transformation (see also subsection 16.1.3). Thus Q is the
infinitesimal operator for U, as P is for translations. In fact it is a general
property that the invariance of the Lagrangian with respect to transformations
of thi s kind leads to a conserved charge. Furthermore there is an associated
16.3.6 Particles. Fields and Antiparticles 4 11

'conserved current' which, in the example above, is defined as the four-vector


j = i {(V<I>t)<I>_<I>t(V<I» }, j, = _i {q, t<l>_<I>tq, }

It is trivial to verify, using the Klein - Gordon eq uation, that


V.j +ih /at = 0 which is the property which gives rise to the name
conserved current. Integrating this equation over all three-space, using Gauss's
theorem JV .j dV = Jj . dS and assumingj to drop off more rap idly than - : 2 at
large r so that the surface integral vanishes we have (a/at)Jj,dV = O. This is the
conservation of charge since, comparing with the expre ssions above,
Q = - Jj,dV.
Perhaps the best known antiparticle is the positron, a positively charged
particle with the mass of the electron and therefore regarded as the anti-
electro n. In high-energy collisions the antiproton, negatively charged, has been
observed. The fields for these particles are similar to the one described above
except that they have a spin of tn. We discus s non-zero spin fields briefly in
subsection 16.3.7.

16.3.6 Charge conjugation and the PCT theorem

We continue this very br ief discussion of quantum fields by introducing the


PCT theorem. This concern s the product of the three operations space
inversion P, which in th is book (see sections 15.3 and 15.5) we have denoted by
I, the time-re versal operation T, which (see subsection 15.7.4) we have denoted
by Y , and a new operation C known as 'charge conjugation' which transforms
a particle into an antiparticle and vice versa. The reason for considering the
product PCT is that, under very general conditions, it may be shown that PCT
is a symmetry operation. For electromagnetic and the strong nuclear
interactions this is not of great interest because experiment shows that the
three operations P, C and T are each symmetry operations. However for the
weak nuclear interactions neither P nor C are symmetry operations separately
but the evidence suggests that the product PC and also T are symmetry
operations, which is con sistent with the PCT theorem. (There may be some
experimental evidence for a tiny violation of PC symmetry in weak in-
teractions but this could still be consistent with PCT symmetry.)
Let us then consider the effect of these transformations on the field <1>
introduced in subsection 16.3.5. We shall restrict our discussion to this zero-
spin case although the arguments may be extended to any spin. For
con sistency with previous chapters the operators representing P and Twill
be denoted by T(I) and Y , respectively. From the work of section 15.5
the beha viour of the particle operators under space inversion is
T(I)at(k)T -1(I) = ±at(lk ), where the ± denotes the intrinsic parity of the
particle. Also, from equation (15.108) for zero spin , time-reversal gives
Yat (k jy - 1 = at (Ik). We now define the unitary charge -conjugation operator
t At A t
C by the relations C 2 = 1 a nd Ca (k )C - 1 = b (k ), where b refers to the
412 Particles, Fields and Antiparticles 16 .3.6

antiparticle. It then follows directly that


T(I)CY<1>t (e) (T(I)Cn- 1 = ± <1>( - e) (16.45)

using the anti-unitary property of Y . Thus the effect of PCT (or T(I)CY in our
notation) on the field is to take the Hermitian adjoint and to replace i by - i.
The phase ± in equation (16.45) is not significant because of the arbitrariness
in choosing the phase for Y, see subsection 15.7.4. The transformation i-+ - i
which reflects both space and time is sometimes called 'strong reflection' and
because of the Hermitian property of the simple Lagrangian density !E ,
constructed for the field <1>, as in equation (16.38), it follows that the PCT
transformation induces the change !E(i)-+!E( -i). When the Hamiltonian
density .Yf (i) is constructed from !E(i), as in equation (16.39), the same change
occurs, yt (i) -+ .Yt ( -i) and finally the Hamiltonian operator H = J.Yt' (i)dV
formed by integrating over all three-space satisfies Hu) -+ H( -l). But since H
must be time independent this means that H is unchanged by the PCT
operation, which completes the argument that PCT is a symmetry operation.
The power of this result lies in the fact that the arguments outlined above are
equally applicable when the Lagrangian contains complicated interaction
terms and is dependent only on the Lagrangian being Hermitian and invariant
under the proper Poincare transformations. (There is also the implicit
assumption of point particles, without structure, and our use of differential
equations which implies a continuity in space, without any granularity.)
A more formal proof of the PCT theorem may be constructed by
arguing that invariance of the field theory with respect to proper Poincare
transformations implies invariance with respect to proper complex Poincare
transformations and that this complex group contains the strong reflection
e-: -e.
To illustrate the PCT symmetry we give two examples where there is no
symmetry with respect to P or C separately but where the product PC is
conserved . In subsection 15.7.3 it was argued that the neutrino did not possess
definite parity. In fact the parity operation produced a right-handed spinning
particle which is not found in nature. However the operation CP will take the
left-handed neutrino into the right-handed antineutrino which is found to
exist. One finds here that, although P is not conserved, the product CP is
conserved as also is T.
The KO meson with finite mass, zero spin and definite strangeness -1 has
as its antiparticle the K ° meson with strangeness + 1. These particles are
produced via the strong nuclear interaction which conserves strangeness.
However they decay via the weak interaction which does not conserve
strangeness, which does not conserve C or P separately but which to a high
degree conserves the product CPoThe KO is observed to decay in two ways
either to two n-mesons with a lifetime of about (10)- 10 sec or to three n-
mesons with a lifetime of about (10)-7 sec. These results may be understood
as follows . Since the KO particle has definite parity we may write
16 .3.7 Particles, Fields and Antiparticles 413

CPIK ° >= 1K ° >so that it is necessary to take linear combinations


K ~ >= 2 -! ( K ° >+ K ° >)
1 1 I

I K ~ >= 2 - ! ( I K ° >- 1K ° >)


to produce states with definite CP, i.e.

CP I K~ > = K~ >
1

CPI K~ >= -I K~ >


One °
can then write I K > as the sum of two components
I K ° >= 2 - l ( I K~ >+ I K~ >). Now the lowest state of the two-a system, with
no relative angular momentum, has definite CP = + 1 so that in a CP-
conserving process only the component I K~ > may decay in this way. The
other component IK~ >must decay by a more complex mode involving three
n-mesons combined to give CP = - 1. Not unnaturally, the two-a mode of
decay has the shorter decay time and the suffices Sand L on K ° refer to short
and long decay times . Experimentally thi s implie s the following, rather
amusing situ ation. Consider the collisions of n - mesons on protons which can
produce a beam of K 0 particles, tt - + P = K 0 + A. After travelling a short
distance the K ~ component has decayed leaving K ~ which is a mixture of K 0
and its antiparticle K '' . In other word s a beam which originally contained only
K O particles gradually acquires a K 0 component through the decay ofK~ . The
presence of K 0 in the beam has been demonstrated by observing the reaction
K O+p=n ++A (16.46)
which could not be produced by K ° particles because of strangeness
conservation. In other words this reaction (16.46) is not produced by the
original beam but is produced once the K~ component in the beam has
decayed.
Finally, we should remark that recent experiments show evidence for a very
small degree of CP violation in the K 0 decay and it has been suggested that
there may be some 'superweak' interactions which violate both CP and T but
st ill conserve CPT. Their strength would need to be no greater than (10)- ~ that
of the weak interactions. One seem s therefore to have very few absolute
symmetries but rather a hierarchy of interactions with the strength dropping
sharply as each symmetry is violated.

16 .3.7 Field for particles with non-zero spin

In discussing fields in the present chapter we have so far restricted our


attention to the case of zero spin . The technique extends naturally to the more
general case and we carry out this extension for three interesting cases s = !
with finite mass (the Dirac field) and for zero mass with Iml = ! (the Weyl,
neutrino, field) and Iml = I (electromagnetic field).
414 Particles, Fields and Antiparticles 16.3.7
The Dirac field
Recall from subsection 15.8.4 that, although there are only two independent
particle states ms = ± ~ for each k, it was necessary to introduce a four-
component space in order to achieve a factorisation between the transfor-
mation of k and the transformation of the components. The four-component
space was effectively reduced to two by the imposition of the Dirac equation
(15.132). We shall follow the sa me approach here by constructing a four-
component field. In place of the simple transformation law (16.32) for the
scalar field we shall now require the four components <1>.(e) of the field
operators to satisfy equation (16.34) which may be written
(16.47)
where M is the 4x 4 matrix introduced in equation (15.128)and(J) now denotes
a column vector. This requirement is consistent with the transformation
properties (16.19) of the classical fields since the expectation values
<'" I'" )
1<1>. (e) are associated with the classical field <P. (e). Remember that
T(~, L) is a unitary operator acting on the state vectors I'" ) , although M is not
a unitary matrix but satisfies M t)lt M = )It , as was shown in subsection 15.8.4.
Thus, taking the Hermitian adjoint of equation (16.47) we have
T(E , L)<1>t(e)T - I (E, L) = <1>t(Le + E)(M - l) t = <Ilt(Le + Eh't MYt
It is convenient to introduce the notation (J)(e) by writing (J)(e) = (J)\e) I', so
that

(16.48)

showing that the row-vector (J) transforms by post-multiplication with the


matrix M(L).
We now construct the four-component field operator di in a fashion similar
to equation (16.44) with equation (16.33) but making use of creation operators
at (km s ) for particles with spin s = 1 and the corresponding antiparticle
t A

operators b (km,) . We write

<1>. (e) = k i(2n) - i fk,- I dk L { U am, (k)a(kms)exp( - ik.;)


m,

+v ':Xm~
(k)b\km,)exp(ik .el) (16.49)
where the normalisation factor is conventional and the coefficients U and vare
determined to ensure the transformation property (16.47) and causality.
Regarding the u and vas 4 x 2 matrices, it is argued below that they are given by
(U !.Ol(Q)) (-U I.Ul(Q) ) ( 0 I)
= U O. !I(Q) , v(k ) =
A A

u(k) LIU. II(Q) x - I 0 (16.50)

where the boost Q in the 2 x 2 matrices L carries the vector (0 0 0 1) into ~, see
equation (15.125). (In fact , the two columns of u(k) are the column vectors in
the two single-pa rt icle sta tes Ik ± ) given in equat ion (15.135a ).)
16 .3.7 Particles, Fields and Antiparticles 415

To show that cD has the required transformation properties we must start


from the known transformation (15.59) of the particle (and antiparticle) states

T(E, L)at (km,)T - I(E, L) = exp(ik' . E) LDH)


, m,m,
(R')at (k'm:)
m,

where k' = Lk and R' = Q ' - ILQ from equation (15.60) with Q' the boost for
t: Taking the Hermitian adjoint of this equation gives
T(E, L)a(km,)T -I(E, L) = exp( -ik' .E)~D~~;(R,-1 )a(k'm:)
m,

using the unitarity of the D-matrix. Since, for rotations, D(1) = U 1. 0) = UO,1),
we may write , as a matrix product,
D(\)(R , -l) = U J,O)(Q -I)U1,O)(L - I)U J, O)(Q ') (16.51)
= Lt0, 1J(Q - 1 IUD,1)(L - I)UO,1l(Q ') (16.51al
From equation (16.49) we now have

T(E, L)C1l.(e)T- I (E, L) = k!(2n)-t fk'- Idk ~ {exp( - ik' .E - ik .e)


m,

x[u(k)D!(R,-I)] am,,a (k' m: ) +exp(ik' .E + ik.e)


l
x [ v(k )D (1 )* ( R ' - 1)] am,, b (k ' m:n (16.52)
where we have again used the unitarity of D to write Dl!) (R') =
D(!)*, (R'- I). The matrix products uD and vD* simplify consid~ribly on
m,m,
using (16.50) and the pair of equations (16.51),
A _(U J,OI(L - I) Ld.OJ(Q ')) A

u(k)D(\J(R' 1)= U O,1l(L - I) UO,1)(Q') =M -1u(k')

-1)D(1)(R' -I)( 0 1)
o -1 0

_ (-Ld, O)(L - 1) Lt1. 0)(Q')) ( 0 1) _ M -IV k'


- Lt0,1l(L - 1) Lt0, IJ(Q') -1 0 - ()
(16.53)
where the form (15.128) has been used for M. The role of the matrix
( _ ~ ~) = ia; is to transform D* into D since these representations are
equivalent, see section 7.7. The desired result (16.47) now follows immediately
on putting the expressions (16.53) into (16.52) and making a change of
integration variable from k to k '. Note that this derivation would still be valid
had we not included the negative sign in the first two rows of v in equation
(16.50). However, this choice is cruci al in ensuring causal ity, as may be seen by
416 Particles, Fields and Antiparticles 16.3.7

calculating the anticommutator of <b T(e) and <b(e') following subsection 16.3.4.
An interesting consequence of the structure of the Dirac field is that the
intrinsic parity of particle and ant iparticle must be opposite. This differs from
the zero-spin case where the two parities were the same. Let us return briefly to
that ca se before discussing the Dirac field. Let T(I) denote the unitary
transformation induced by the space inversion 1. Then if n denotes the intrinsic
parity of the particle we have , from equation (15.87),
T(l )a\k)T - 1 (I) = nat (lk)
with

Let us denote the intrinsic parity of the antiparticle by ti ' so that


T(l)bt(k)T -1(l) = n 'bt(lk)
Now if we require the field <1> to have definite parity so that
T(l)<1>(e)T- 1(l)
= ±<1>(Ie) (16.54)

then from the definition (16.44) of <1> it follows quickly, using the properties of a
and b above, that equation (16.54) will be satisfied only if n = tt ' and then
t

T(I) <1>( en - 1 (I ) = n<1>( Ie) (16.55)


showing that the parity of the particle determines the parity of the field . For the
x-mesons the parity is n = - 1 so that the field is odd and since it is also scalar
it is called pseudo-scalar. (We apologise for the double use of the symbol n in
this sentence!)
Turning to the Dirac field we must start from the inversion transformation
in the particle states,
( 16.56)
and let us write
(16.57)

Now, recalling from subsection 15.8.4 that in the four-component space the
inversion operator is represented by the matrix y" we shall require a field with
definite parity to satisfy
(16.58)

where again <b(e) denotes a column vector. But , using the results (16.56) and
(16.57) in (16.49), we have

T(I)<1>.(e)T- 1(I) = k1(2n) - 1fk - 1 dk


I
L { U m (k)na (lkm. )exp( -
m,
~ ,
ik .e)

+ V!Xm, (k)n' b' (lkm.)exp(ik . e)} (16.59)


16.3.7 Particles, Fields and Antiparticles 417

To relate this expression to cD(li) we need to change the notation for the
integration variable, writing Ik -+ t.
We must also use the property
u(lk) = }', u(k) which may be seen directly from the explicit form (15.135a) for
u(k), and the consequent property v(lk) = - }', v(k). Because of this sign
change it follows that the expression (16.59) reduces to the form (16.58) only
when n ' = - n: and then
(16.60)
Experimentally, the st udy of the electron - positron system confirms this result
that for s = !. the antiparticle has a parity opposite to that of the particle.
The Dirac equation (15.131) referred to the one-particle states and contained
operators Pwhich described infinitesimal translations in the state vectors. We
now see that the field operators cD(i) satisfy a Dirac equation
(Y .D+ik)cD(i) = 0 (16.61)
where 0 = (- V, % ct ) is a differential operator acting on the posinon
variables i . This result follows directly from the definition (16.49) by operating
with 0 on the exponentials and using equation (16.60). The Dirac equation
(16.61) is again the field equat ion which arises from the Lagrangian. In this case
the appropriate Lagrangian density which gives rise to the particle picture, as
in subsection 16.3.3, and also leads to this field equation is

2) = cD(i)( Y. 0 + ik)cD(i)

The neutrino field


In subsection 15.8.5 we showed that the single-particle state for the neutrino,
zero mass and m = - i , could be expressed in two-component form using the
representation LID.I) together with the Wey1 equation. Thus we shall seek a
two-component field operator cD(i) similar to that given in equation (16.49).
We shall expect it to transform according to
T(g, L)cD(i)T -1(g, L) = (VD.!l) - lcD(Le + g) (16.62)
The field will be constructed from the neutrino operators a'(k -) and the
antineutrino operators b'(k +). The negative sign in the operator at serves to
remind us that the neutrino has m = -i and so transforms according to
P(O, -t) as given in equation (15.68) or equivalently in equation (15.144). We
have assigned the antineutrino a positive helicity but this will be seen to be
necessary if the field is to satisfy the invariance and causality conditions. We

f
shall write

<1>,(i) = (2n:)- i k,-l dkw,(k){ a(k - )exp( - ik .i )


+ bt(k + )exp(ik.e)} (16.63)
and show that the coefficients w, which in this case form a 2 x 1 matrix, are
given by the second column of the matrix Lt°· !l( R,,. Q:), where the Lorentz
418 Particles , Fields and Antiparticles 16.3.7

transformation R, y.Q : was defined in subsection 15.4.2 as that which carries


the vector (0,0, 1, 1) into Ii. To justify this expression we use the particle
transformat ion (15.144) in equation (16.63) giving

T(E, l)<1>,(e )T- 1 (E, L ) = (2n) - ! f k,- ldklr.( kH exp( -ik'. E- ik.e)
x L (~'ll~ l [ ( R~y. Q~) - 1 LRxyQJ*a (k' - ) + exp (ik' . E + ik .e)
x q\i O )[ (R ~)'Q ~ ) - ILRxyQJbt(k' +):
(16.64)
Writing w(ii) = U O.1)(R, y.Q :) ( y) as the seco nd column of U O.I) and using the
unitarity of the particle transformation (15.144) we see that
w(k)L(~i~ I[(R :, )Q~) -1 lR"QJ*

= { uo. I)(R".Q=) Uo. ll[ (Rq Q=)-1 t. - I R~yQJ } (~)


= U O.I'(l - I)w(ii')
where we ha ve also used the fact that qO~\;[(R~yQ' :) - llRxy.Q:] = 0, see
the discu ssion after equation (15.144). A simila r result
w(k )q\ O)[ ( R ~ y Q ~) - 1 lRxyQJ= UO.I)(l - 1 )w(k')

hold s for the second term in equ at ion (16.64) using the equ ivalence of U l. ol and
L1o. H*, see subsection 15.2.5. Putt ing the se two result s into equation (16.64)
and making the change Ii' -+ Ii in the integration variable leads to the des ired
result (16.62).
As in previous examples one sees that the field (J)( €) satisfies the appropriate
field equation which, in thi s case , is the Weyl equation (15.146)

1
( s.V - -2c - at
0) (J)(e) =
A
°
The Lagrangian density for the free neutrino field is

!f' = cl>t(e) ( s. V - L:t )(J)(€)

Note that (J)t transforms as.cordin~ to U 1,O) and the Weyl operator is a
component of a four-vector W + imP. It is rather surprising that one does not
need to include the other components to ensure the invariance of !f' but this is
due to the special relat ions between these components as pointed out after
eq uation (15.145).
Although the neutrino has no electric charge we can define an operator Q as
in subsection 16.3.5 which measures the difference between the numbers of
neutrinos and anti neutrinos. It is called the 'lepton charge' and if a similar
16.3.7 Particles, Fields and Antiparticles 419

charge is assigned to electrons and muons then one finds experimentally that
this lepton charge is conserved, like the electric charge. This is illustrated in the
fJ-decay of a neutron, n --+ p + e + v into a proton, electron and antineutrino V,
since one assigns a zero lepton charge to the heavier particles nand p.
Confirmation of these ideas is found in many experiments; for example the
process v + n --+ p + e does not take place whereas v + p --+ n + e, with e
denoting the anti-electron (or positron), is observed. We might briefly remark
that the heavier particles like nand p belong to a class of particles called
'baryons' with a baryon charge of + I and that this baryon charge provides a
further conserved charge.
We have seen in subsection 15.7.3 that space inversion changes the helicity,
i.e. changes the sign of m. Thus the neutrino field cannot have definite parity
like the Dirac field. The space inversion would carry the neutrino (m = -t)
into a neutrino with m = ! but such particles do not occur in nature and are
therefore not contained in the field. However if we introduce the charge
conjugation operator C as in subsection 16.3.6 then the existence of
antineutrinos with m = ! suggests that the field may transform simply under
the combined operation of charge conjugation and space inversion which we
denote by CP. Defining this unitary operator by the relations
t
CPat(k - )(CP) ~ I = '7(k)b (Ik +), (CP)2 = I
where '7(k) is a phase factor, we are able to show , by using the properties of
w(k), that CP transforms the components of(J)(€) into those of (J)t(I€). In fact

CP<1>,(€)(CP) -1 = Ip (iO) ,p<1>; (Ie)

where CT y = 2sy is the 2 x 2 spin matrix. One can therefore construct a


Lagrangian containing the neutrino field and which is invariant under CP but
not separately under C or P. (It is perhaps worth remarking that charge
conservation and the invariance under charge conjugation are two quite
distinct properties.)

The electromagnetic field


The electromagnetic field is a familiar classical concept with consequences
over macroscopic distances. This contrasts with the nuclear fields whose
interactions drop off exponentially with distance with dimensions of order
10- 13 em. Nevertheless, for a consistent theory and indeed to agree with
accurate experiments, the electromagnetic field must be constructed in
quantised form . We shall see below that if we construct field operators
appropriate to particles with zero mass and m = ± I, the Poincare repre-
sentations P(o. ± I) , then such a field has the familiar properties of the
electromagnetic field. The particles are called photons. As in the passage from
quantum to classical mechanics for particles one can recover the classical field
interpretation. The method is very similar to that of the previous discussion of
420 Particles , Fields and Antiparticles 16 .3.7

the neutrino field with the trivial change from Iml


=! to Iml
= 1 and two
significant changes. In the first place the photon exists in both helicity states
m = ± 1 so that the field may be constructed with definite parity and the
electromagnetic interactions may conserve parity. The second change is that
the photon has no charge of any kind and thus has no distinct antiparticle.
Taking these two changes together and noting that, in a sense, both helicity
states were present in the neutrino field, one for particle and one for
antiparticle, we have as a natural extension of equation (16.63) to the case
Iml = 1,

<l>,(f) = (27<) - j f
k,-t dkw,(k){ a(k - )exp( - ik . f )

+ at(k + jexp(ik.e)} (16.65)


where at(k ±) creates photons with helicity ± 1 and the fields
and the
coefficients w(k) are now three-vectors. As before we can ensure that the fields
satisfy the transformation properties
re, L)«Il(f)T -1(t, L) = (Llo. 11) - I(J)( L f + t )
T(e, L)cI>t(e)T - 1(e, L) = cI>t(Le + e)LlI. 01 (16.66)
by choosing w(k) to be the three-vector appearing in the third column of
Llo. tl(R,yQz) corresponding to m = - 1. The justification for this choice
follows that given for the neutrino so closely that we shall not give it. We do,
however, need to understand the structure of the vector w(k) which is given by

wlf) ~ L'" "IR',o,{~) ~ L'" "IR".)L" "IO'{~)


~ D"'IR,,{~)
k, (16.67)
where we have used , for rotations, L'? : 1)(R,y) = D(l'(R,y), while for the
boost in L' ": II(Q:),

Q z = exp(bY z ) = exp(ibX z ) = exp(bs z )


making Llo. I/(Qz) diagonal with L(~·t' ~l (Qz) = exp( - b) = k,. The relation
between band k, comes about because Qz(O 01 1) = (00 k, k,) so that from
equation (15.24), cosh b - sinh b = k. . Now recall from subsection 7.4.5 that

the vector ( ~) in the m-basis is given by (e, - iey)/2i in the Cartesian basis so

that equation (16.67) tells us that w(k) is a vector oflength k., of the kind (~)
16 .3.7 Particles , Fields and Antiparticles 421

but rotated by R, y- Since R, ycarries the z-axis into the d irection k, the vector
w(k) may therefore be written
w(k ) = k,[e dk)- ie2(k)] /2 1 (16.68)
where e I and e 2are a pair of o rthogonal unit vectors in the plane perpendicular
to k . The pos ition of e 1 in th is plane is unimportant since an y rotation in this
plane changes onl y the phase of w(k ). However, by choosing e2 to be the axi s
of the rotation Rx }" the se unit vectors acquire the simple properties
e2 (1 k ) = e2(k)and e 1 (lk ) = -e dk),since Ik is arrived at by a n extra rotation
o f n: in R,}" Hence from (16.68), w(Ik) = - w*(k).
From equation (15.91) the behaviour of the particle operators under space
inversion is given by

(16.69)
where the choice of negative sign is only a convention, since the relative phase
o f the two helicity states is free. (The po sitive sign was used in equation
(15.91).) However, o nce thi s cho ice is made, the po siti ve sign between a(k -)
and at(k + ) in (16.65) acq uires significa nce. Using (16.69) we see that the field
satis fies the equat ion
T(I)<I>(e)T - 1(1) = cIlt(le )
We may therefore form linear combinations of«Dand «D t which have even and
odd parity, respe cti vely:
t(e)] t(
B(e)= [«D(e) + «D / 2l and E(e ) = - i[ cIl(e) - cIl e )] / 2!
The numerical factors a re conventional and the factor i makes E Hermitian as
well as B. By using the symbols E and B we are anticipating their association
with th e famil iar magnet ic and electric fields of classical theory, see subsectio n
16.2.3. To construct the q uantised field corresponding to the classical vector
potential A it is convenient to introduce the particle operators
t A t l' t l' 1
a1(k) = [a (K+) -a (K -)] /2 1
tA t l' t l' I
a 2 (k ) = -i[a (K+)+a (K-)] /2 I ( 16.70)
which are linear combinations of the two helicity states ± and describe states
of plane polarisation. In terms of these new operators and using the form
(16.68) for w, the expressions (16.65) lead to

E(e) = (2n:) - l2 -I i fdk {[ edk)adk) + e 2(k) a 2 (k)]exp( - ik . e)

- [e l (i~)a: (k) +e2(k)a~(k )]exp(ik.e) }

f
B(e) = (2n:) -l2 -li dk {[ e2(k) adk) - edk)a 2(k)]exp( - ik .e)

t
- [e2(k )a 1 (k ) - e 1 (k)a ; (k )]exp (ik. e) } (16.71 )
422 Particles, Fields and Antiparticles 16.3.7

From the form of the expressions (16.71) we see that both E and B may be
expressed in terms of a single vector field

A(e) = (27t)-~r1 f k,- 1 dk i. J; 2 e(k, i.J{ a(kJ)exp( - ik. €)

+at(k, A)exp Ut e )} (16.72)

by writing

E(€) =
I c
- - ---;-A(€), B(i) = curIA(€) (16.73)
e ot
and using the trivial properties e. ». e 1 = e z, e. r. e z = -e 1 for a right-
handed set of orthonormal vectors.
It is quickly verified that, as in the other examples, the field satisfies the
appropriate wave equat ions deduced in subsection 15.8.6. In fact it is «I» \2),
transforming like L(t,O) , which satisfies the equations given there while «1»(2)
needs a change of sign in (15.149). Thus
i a
curl «I»(e) = - - --;-«I»(l),
t
curl «I» (e)
ic
= - --;-«1»
t
(e)
c at e M

div«l»(e) = 0,
and these equations lead to the more famili~r form of Maxwell's equations
when they are expressed in terms of E and B,
l e i 0
cur l 8 = - - E curl E = - - - 8
c ct ' c at
div B = div E = ° (16.74)
One therefore identifies this field, constructed from the zero-mass particles
with Iml = I (the photons), with the famil iar electromagnetic field of E and B
which it approaches in the classical limit. As one would expect, the way in
which E and B are obtained from the single field A is the same as in the classical
theory, see subsection 16.1.3. Recall, however, that in the classical problem we
used a four-vector potential A = (A , (/J) to describe the Lorentz invariance of
the theory, whereas in equation (16.72) we have only a three-vector. However
in subsection 16.1.3 we showed that any two potentials A which led to the same
E and B were physically equivalent. There was therefore freedom in the
choice of A and one possibility, called the ' rad ia tio n (or Coulomb) gauge' is to
set ¢ = 0. The simple choice of A(e) in equation (16.72) corresponds to this
gauge.
We have constructed the field A(~) by a somewhat circuitous path, starting
from the product representations P(O, 0) X L(1 , oi and p10, 0) X Lt O, I) . This had
the advantage of similarity with the neutrino field. However it would have been
possible to arrive more directly at E(e) by using the product representation
plO, 0) x L H,!I which also contains p(O, ± I).
Particles, Fields and Antiparticles 423

Finally we note that, since the fields E and B were constructed to have even
and odd parity, respectively, it follows from equation (16.72) th at A must have
odd parity,
T(I)A(e)T-1(l) = -A(Ie)
For this reason the photon is said to have odd intrinsic parity. By taking the
opposite relative sign of alii - ) and at(ii + ) in <1.l,\e) in equation (16.65) we
could have constructed a field with positive parity but this would not have been
consistent with experiment for the electromagnetic field.

Bibliography

For further reading in classical mechanics and fields we suggest

Landau, L. D. and Lifshitz, E. M. (1972). A Shorter Cour se of Theoretical


Phy sics, vol. 1, Mechanics and electrodynamics (Pergamon Press, London)

A more complete introduction to relativistic quantum fields may be found


in one of the following:

Bjorken, J. D. and Drell, S. D . (1965). Relativistic Quantum Fields (McGraw-


Hill, New York)
Schweber, S. (1961). An Introduction to Relativistic Quantum Field Theory
(Harper and Row, New York)

For a more mathematical treatment of field theory we suggest

Streater, R. F. and Wightman, A. S. (1964). PCT, Spin and Statistics and all that
(Benjamin, New York)
For details of the experimental data on which the new concepts like
charge conjugation are based , one should refer to the following particle
physics references: Perkins (1972) of chapter 11 and Dyson (1966) of
chapter 15.

Problems

16.1 Show that the fields E ± i B defined in equation (16.17) transform according to
the representations U 10) and U °\), respectively, of the Lorentz group. (Use the
matrices (15.30) of X and Yand for L11 0l to sho w that (X - iY)(E + iB) = O.

16.2 If the Lagrangian density is invariant under time translation (dd~ = 0 ) show
J
that H = Jf'dV is constant. (Use the field equation (16.21) and Gauss's
theorem, assuming the fields vanish at infinity.)
16.3 In the same way, show that the field momentum P defined in subsection 16.2.2 is
constant when 5f is invariant under space translations.
424 Particles. Fields and Antiparticles

16.4 Two particles with rest mass M and equal and opposite momenta ±pcollideand
come to rest with the production, also at rest, of a third particle. Show that its rest

mass is given by 2 { ( M 2 + :: Y- M } .
16.5 · Show that the number operato r defined by equation (16.30) has the properties
NIO) = 0. Nlk ) = Ik) a nd generally Nlk.k2' " k. ) = nlk 1k 2 • • • k.).
17
The Symmetric Group ~

The symmetric group Y n ' which is composed of all permutations of n objects,


plays a curious role in the theory of symmetry. It is possible to obtain a very
good understanding of symmetry without any knowledge of the properties of
Y n but , on the other hand , this group is to be found in the background of many
physical problems. It is connected with the po int groups ofchapter 9 through a
result , called Cayley's theorem, that every point group is isomorphic with a
group of permutat ions, though not necessarily with the group Y n of all
permutations. For example, we have seen the isomo rphism between D 3 and Y 3
in chapter 2. The symmetric group is also connected with the rotation and
unitary groups because the con struction of the irreducible representations of
those groups may be accomplished by building products of elementary factors
and studying the behaviour of the products under permutations of the factors.
We elaborate this point in chapter 18. In applications to physical problems the
Bose or Fermi statistics ensure that the overall wave function is either totally
symmetric or totally antisymmetric, see section 5.9, but when this wave
function is built up from separate wave functions describing the position
coordinates, the spin coordinates and other degrees of freedom the behaviour
of the separate factors under permutation can be much more elaborate. In fact,
we have already made use of the symmetric group Y 3 in describing the
structure of atoms in subsection 8.6.4 and the structure of nuclei and
425
426 The Symmetric Group /:/) 17 .1
n

elementary particles in chapter 12. We were able to do this without much


discussion because of the simplicity for small n and in particular because of the
isomorphism with the group DJ which we studied in some detail early in the
book. For higher values of n, the way in which the group is applied to physical
problems is exactly the same as for small n so we give no more applications in
this chapter. Our aim here is to describe generally the properties of the group
Y n and its representations.
One simplifying feature of the group Y n is that, in contrast to the point
groups, its properties may be described for general n instead of having to
consider each group separately. We shall often refer to the n objects being
permuted as particles, since this is the most usual situation in applications. The
first three sections describe some properties of the permutations themselves.
Sections 17.4-17.9 are concerned with the irreducible representations and
characters of Y n and a convenient method of labelling both irreducible
representations and basis vectors by Young diagrams and tableaux is
introduced. A rule for writing down the general representation matrix is
deferred until section 17.13. Following a discussion of the familiar direct
product in section 17.10 we introduce in section 17.11 the concept of an outer
product which willalso have relevance in chapter 18. The emphasis throughout
this chapter is on explanation and understanding rather than rigour. For this
reason we do not prove all results generally for all n. Neither do we deduce
general formula for such things as the characters, although we do describe how
they may be deduced for any particular n.

17.1 Cycles

A cycle is a rather special type of permutation, defined as follows . Suppose that


the integers PI' P2' . . ., P, are placed in order, at equal intervals of Zn ]! around
the circumference ofa circle. In the notation ofsubsection 2.2(10),a rotation of
2rr/1 will induce the permutation

PI P2PJ . . . P,-I PI)


( P2PJP4' . . PI PI

Such a permutation is called a 'cycle' of length I and IS denoted by


(PIP2PJ . . . P,),

An arbitary permutation P =( 1 2 ... n) may be broken up into a


PIP2 .. · Pn
product of cycles. Thus, for example, we may write the particular permutation

12 3 4 5 678910)=(1358)(27)( 6910)
( 3754810216 9 3581 72 106 9
= (1358) (6109)(2 7)(4),
17.2 The Symmetric Group Y' 427
n

showing that it breaks into a product of four cycles of lengths 4, 3, 2 and 1. In


the arbitrary permutation P, the number I is replaced by PI' while the number
PI itself occurs somewhere in the first row of P and is replaced by the number
below it, say q I ' Continuing in the same way q I will be replaced by some
number r l . We therefore have a sequence 1,PI ' q I JI , . .. , which will continue
until the number I appears again, prod ucing a cycle denoted by (1 PI q 1 r I . • . ).
It is possible that this cycle contains all the numbers 1 to n but generally it does
not. In the latter case, choose a number not in the cycle and it will generate, in
the same way, a new cycle containing no common element with the first. Hence
the permutation P is expressible, uniquely, as a product of cycles. Since there
are no numbers common to any two cycles, the permutations represented by
each cycle clearly commute. Notice that if a number is unmoved by the
permutation P, then it will contribute a cycle of length I = I containing just
that single number. The sum of the lengths Ii of the cycles in a permutation P
must be equ al to n. Thus the breakdown into cycles associates a partition
n= L Ii of n into integers Ii ' To
i
be definite, we always order the cycles by
decreasing length so that l, ~ Ii + I and we use the notation [11 12 .•• In]' In the
example given above for n = 10, the associated partition is [432 I].
A cycle of length I = 2 is called a 'transposition' since it simply transposes
two numbers. It is denoted by P i j . In particular, P j j + I is called an 'adjacent
transposition'.
Any cycle, and therefore any permutation, may be written as a product of
transpositions since, as is quickly verified,

12 3 k-Ik)
(123 . . . k) == ( 234 k I = PI2P23P34 '" Pk- I k (17.1)

and any transposition may be written as a product of adjacent transpositions


using, successively, relations like PI k = P 12 P 2k P 12'

17.2 The parity of a permutation

It is possible to assign a 'parity' of + I or - 1 to a permutation according to the


following definition: (1) a cycle of odd length has even parity + I; (2) a cycle of
even length has odd parity - 1;(3) the parity of a permutation is equal to the
product of the parities of each of its cycles. This association of even parity with
odd length is not deliberately perverse but is in fact natural, implying that a
cycle of length one, which causes no permutation, has even parity.
There are two other precisely equivalent definitions of the parity of a
permutation. In the first we write each cycle as a product of transpositions and
note that a cycle of odd length involves an even number of transpositions and
vice versa. Thus a permutation with even parity involves an even number of
428 The Symmetric Group Y' 17 .3
n

transpositions and vice versa. In the second equ ivalent definition we make use
of the function

(17.2)

which is the determinant formed from n different functions ¢q of n different


variables XI ' X2' . • • , X n • Any permutation of variables will produce a
determinant with the rows of <I> A interchanged in some way. From the theory
of determinants this simply introduces a factor ± 1 depending on whether the
number of interchanges is even or odd. But this is precisely the parity, giving
P<I> A = n(P)<1> A (17.3)
where we have introduced the notation n(P) for the parity ofP. Ofcourse n(P)
may only take values ± 1. Thus n(P) is defined as the result of P operating on
<1> A '

17.3 Classes

For the rotation groups it was shown in section 2.7 that if two rotations Rand
R' are in the same class then the group must contain a rotation which carries
the axis of R into the axis of R'. The analogous result for permutations is that if
two permutations P and P' are in the same class of Y' n then there must be a
permutation Q which, when carried out on both rows of P, produces P' . To
prove this statement we first define the permutations

p_(1 2 oo .n) (17.4)


PIP 2· oopn

and

(17.5)

where the last equality defines the r j given the pj and q.. Hence

(17.6)

but the permutation P' is nothing more than the permutation P of(17.4) with
both rows permuted by Q of (17.5), i.e. the numbers 1 2 . . . n are replaced by
ql q2°o ·qn and the numbers PIP2 °o .Pn are replaced by r l r 2 °o.r n •
17.3 The Symmetric Group Y' 429
n

In terms of cycles this result implies that all elements of g" assoc iated with
the same partition [111 2 • • • ] are in the same class and conversely that all
elements in the same class have the same structure in terms of cycle length and
are thus associated with the same partition. For example, in the group g 5 the
two permutations

P= (~~~:~) == (132)(45) (17.7)

and

P' =(~~~~~)=(12)(345)=(345)(12) (17.8)

are in the same class. To find a Q such that P' = Q PQ - lone needs simply to
permute the cycle labels of (17.7) into those of (17.8), giving

Q=(12345)
35412
It is quickly verified that

Q PQ _ 1 = ( 12 34 5) (1 2 3 4 5) (3 54 1 2) = (3 5 4 12 )
35412 31254 12345 43521

== ( I 2 3 4 5) = P'
21453

Of course, Q is not unique, but that is always so if it is defined only by the


relation P' = Q PQ - 1, because Q may be post-rnutiplied by any permutation
which commutes with P, without disturbing the relation.
As an illustration we see that the six elements of g 3 fall into three classes

C(j 1 = E (the identity)


Cfj 2 = (1 2 3), (1 3 2)

Cfj 3 = (12), (13), (23)

associated with the three partitions [ 1 1 1 J.[ 3 ] and [2 1], respectively. It is


usual to omit the cycles oflength one , as in class Cfj 3 above. No ambiguity arises
here, the convention being simply that the numbers not written down are
unmoved. In this example the ordering of classes Cfj l ' Cfj 2 and Cfj 3 was made to
agree with that of examples 2.7.2 and 2.7.3. More systematically, one would
order the classes by the length of the ir longest cycle, starting with the shortest.
Classes whose longest cycles have the same length would be ordered by their
next longest cycles and so on .
430 The Symmetric Group Y 17.4
n

17.4 The identity and alternating


representations-symmetric and
antisymmetric functions

The remainder of th is chapter is devoted mainl y to a description of th e


irred ucible representation s of Y n' culminating in the construction of the actual
matrices T(P ) in section 17.13. In this sectio n we restrict our att ention to two
simple on e-dimensional repre sent ations which occur for any n. Firstly there is,
as for any group, the on e-dim ension al identity representation which associates
+ 1 with every group element. A fun ction l1>s which transform s according to
this representation is therefore unchanged by an y permutat ion a nd is usually
ca lled a totally symmetric funct ion . Thus for all P,
Pl1>s = l1>s (17.9)
For any n there is also another one-dimensional representation called the
'a lternating representation' which associates the parity n(P) = ± 1 with each
element P. That this is in fact a representation follow s from the result that the
parity of the product of permutations is the product of the parities. To prove
this result we consider the effect of a product R = Q P on the function l1> A of
equation (17.2)
Rl1> A = QPl1> A = Qn(P )l1> A = n(Q ) n(P)l1> A
which gives the desired result , n(R ) = n(Q ) n(P). In fact the function l1> A is a
co nvenient basis function for the a lternating representation
(17.10)
A function like l1> A wh ich transform s according to the alternating repre-
sentation is called totally antisymmetric. In particular it has the property
(17.11)
for all transpositions P ji ,
In section 4.19 we constructed a projection operator, which had the property
of producing a function of particular symmetry type when operating on an
arbitrary function . For the two one-dimensional representations described
above, the projection operators are

S=~LP
n! p

and
A = ~ L n(P)P (17.12)
n! p
although for most purposes the numerical factor l in ! is irrelevant. These two
operators are called the symmetriser and the antisymmetriser, respectively.
According to the general theory of section 4.19 the functions f s = Sf and
fA = Af will be totally symmetric and totally antisymmetric respectively, no
17 .5 The Symmetric Group Y' 431
n

matter what the functionf. It is always possible however, forcertainf, that the
functions f s and J~ may vanish identically.
Consider for example II = 2 with f = x I ' Then
Sf= t(E+PI2)X I =t(x. +X2) =fs
Af= !(E- P 12) X1 = !(x 1 -x 2 ) =fA

Notice here that E = S + A so that f = f~ + J~ showing that any function of


two variables may be expressed as a sum of symmetric and antisymmetric parts.
As a second example take II = 3 and again project from f = x I ' Then

SJ=
' I[
(; E+PI2+PI 3+P2 3+ (123)
231 + (123)J
312 XI

= i (x 1 +X2 +x 3 +x I +x 2 +x 3)
= 1(x 1 +X2 +x 3) =J~

Af= k[ E-P I2 -P 13 -P 23 + C
(~~:) + ~~) JX I
= i(x i -X 2 -X3 - XI +x 2 +X3) =0

This means that the function f = x I is too simple to produce a totally


antisymmetric function; it does not contain a totally antisymmetric part to be
projected out. One also sees that for II > 2 we cannot write E = S + A,
implying that for II > 2 it is necessary to introduce other symmetry operators in
addition to S and A in order to allow expansion of an arbitrary function . In
other words for II > 2 the identity and alternat ing representations do not
exhaust the possible irreducible representations.

17.5 The character table for irreducible


representations
For any II, the classes are labelled by the partitions of II, as shown in section
17.3. Hence the general result of section 4.13, that the number of irreducible
representations is equal to the number of classes, tells us that the number of
irreducible representations of the group Y'o is given by the number of
partitions [II I 112" ' ] of II . Thus for II = 2 there are two irreducible repre-
sentations corresponding to the partitions [2] and [ II] while for II = 3 there
are three corresponding to [3], [21] and [111] and for /I = 4 there are five.
The character table for Y' 3 has already been deduced in section 4.10.
Actually we deduced the character table for the group D3 but we noted in
section 2.3 an isomorphism between D3 and Y' 3 ' The methods described in
section 4.15 may also be used to deduce character tables for other small II, see
problem 17.2. There is, however, a general procedure which we now describe
432 The Symmetric Group Y
n
17 .5

which leads to the character tables and also gives some insight into the
properties of the basis vectors of the irreducible representations. We shall see
that not only is the number of irreducible representations equal to the number
of partitions, but also that each irreducible representation may be labelled by a
partition in a natural way.
Consider a number n of particles which may be put into a number n of
single-particle states cPi so that cP;(k) signifies that particle k is in state i. To each
partition [n l n2"'] of n we associate the product function
$([n 1n2 "']) = cPd1)cPd2) ... cPl(n.)cP2(n l + l)cP2(n l +2) .. .
. . . cP2(n l +n2)cP3(nl +n2 + 1) . . . (17.13)

in which the first n 1 particles are in the state cP l' the next n 2 particles are in state
cP2 and so on . For each partition we now generate the set of n!/(n l ! n 2! . .. )
independent product functions P$( [n 1 n2 • • . ] ), where P is any permutation
of the particles. (Although there are n! permutations P, this number must be
divided by n l ! n 2 ! .. • to allow for those permutations which leave (17.13)
unchanged.) This set of states defines, for each partition, a vector space
L([n 1 n2 •• • J) which is invariant under permutations and therefore provides a
representation of the group Y n • In general these representations are not
irreducible but their characters are relatively easy to calculate. The characters
of the irreducible representations may then be deduced by a process of
subtraction, starting from the simple partition [n]. In this case the function
(17.13) is invariant under all permutations so that the corresponding
representation is the identity representation.
To calculate the character for a permutation Q in the representation
generated by the vector space L([n l n 2 • •• J) we must sum the diagonal matrix
elements of Q in this space. However, the result of operating with Q on a
general basis vector P$( [n 1 n2 ••• ] ) is simply another basis vector because
QP is just another permutation. Thus every column of the matrix of Q will
contain one entry of + 1 and zero in every other position. The only
contribution to the character will occur when this entry + 1 occurs on the
diagonal of the matrix. In other words there is a contribution of + 1 for every
basis vector which is unchanged by Q and hence the character for the
permutation Q is given by the number of basis vectors of the kind
P$( [n 1 n2 •• • ] ) which are unchanged by Q. It is of course only necessary to
choose one permutation Q from each class. For the identity element E the
character is simply the dimension n!/n l! n 2! • . . of the space L( [n l n 2 ... ]).
If Q is a single transposition P 12 the character is given by the number of basis
vectors for which the particles 1 and 2 are in the same state, i.e. either they are
both in state cPI or both in state cP2, etc. This number is deduced by counting
the number of ways of arranging the remaining (n - 2) particles, namely
(n-2)! (n-2)!
+, (17.14)
(n l - 2)'. n 2 ·I n 3 I• • • • n l · (n2 - 2)'.n 3 I• • • •
+ .. .
17 .5 The Symmetric Group [I' 433
n

If Q is a single cycle of length I then a similar argument leads to the expression


(n-/)! (n-/)!
(n l I)'.n Z·n3
"
· ·· ·
+ nl ·(n
I II I + ... (17.15)
- z- ).n 3· · · ·
for the character of Q. When Q is a product of cycles one must allow for all
possible distributions of the cycles between the states allowing any number of
complete cycles to lie in any state. It is possible to deduce a cumbersome
general formula but we shall not give it here. However, we can see from the
above argument that the character will vanish unless the numbers ni in the
partition are large enough to contain the cycles Ii of Q in some order. This
explains the triangular blocks of zeros in the tables below.
Application of these results for the characters of Q in the spaces
L([n1nz . .. J) for n~4leads to table 17.1. We use the notation I/I[nln, ] for
the character, while the classes of the group [l'n are denoted by (/1 I z ),
where Ii are the lengths of the cycles of any permutation in that class. Having

Table 17.1

[l'z (11) (2) [/'4 (1111) (211) (22) (31) (4)


I/I[Z] 1 I 1/1[4] I I I I I
1/1[11] 2 0 1/1[31] 4 2 0 I 0
1/1[22] 6 2 2 0 0
x[2] I 1/1[211] 12 2 0 0 0
X[II] -I 1/1[1111] 24 0 0 0 0

l4] I I I I 1
X[3I] 3 1 -1 0 -1
[/'3 (Ill) (21) (3) X[22] 2 0 2 -I 0
X[211] 3 -I -1 0 1
1/1[3] 1 I 1 X[III1] 1 -I 1 1 -1
1/1[21] 3 I 0
1/1[111] 6 0 0

x[3] I 1 I
X[21] 2 0 -1
X[III] 1 -I 1

calculated the characters I/I[n, n, . . . ] in this way, the characters X of the


irreducible representations may be deduced as follows . In the first place I/I[n] is
clearly irreducible, being one-dimensional, so we call it X[n] . Now proceed to
the second row, considering I/I[n - 1. 1). Using the techniques described in
section 4.11 we can show that X[n] appears once in the reduction of l/I[n-I.1]
and the remaining character is irreducible by the criterion (4.29). This
remainder is denoted by X[n -I, I]. In this way we may proceed down the table
and at each stage it is found that I/I[n, n, . . . ] reduces into a sum of irreducible
434 The Symmetric Group Y 17 .6
n

characters deduced earlier and that the remainder is also irreducible. The
number of times that each of the earlier characters appears and the verification
that the remainder is irreducible may be obtained with the techniques of
section 4.11, see also problem 17.3. The remainder is denoted by X[n, n, . . . ] so
that we obtain an irreducible character corresponding to each partition. The
results of this analysis are shown in table 17.1 for n ~ 4. It is interesting to note
that ljI[ III ... ] is the character of the regular representation, see section 4.13,
which contains every irreducible representation a number of times equal to its
dimension. In terms of the basis functions (17.13), the progression down the
table corresponds to the introduction of a greater variety of functions ¢q and
hence there is greater scope for antisymmetry. Finally in the last row there are
as many functions ¢q as particles and every symmetry type may be constructed,
including the totally antisymmetric (alternating) representation X[III . . . ]
already described in section 17.4.

17.6 Young diagrams

Although we did not give a general proof, it was shown in the preceding section
that the irreducible representations of Y n were labelled by a partition
[n 1 n2 • • • ] of the number n. For many purposes it is convenient to use a more
visual notation for a partition as follows. To each partition we construct a
diagram of n small squares of which n, are in the first row, n2 are in the second
row, and so on, with the left-hand ends of all rows being in the same vertical
line. Such a diagram is named after the mathematician A. Young. As an
illustration, the five partitions of n = 4 are given by the Young diagrams
below:

[4] [31] [22] [211 ] [1111 ]


I i
EfTIEB
From the result s of the previous section we may now label each irreducible
representation of Y n by a Young diagram containing n squares. Although at
first sight this not ation may seem clumsy it will be seen in section 17.8 to lead to
a way of describing not only the irreducible representations of Y n but also the
properties of the basis vectors of the representations.

17.7 The restriction from s: to 9;;-1


Consider the symmetric group Y n _, which permutes the objects 1, 2, . . .,
n -1, leaving object n unmoved. It is clearly a subgroup of Y n • Hence an
irreducible representation of Y n will also be a representation of Y n _ , though
17 .7 The Symmetric Group .Cf' 435
n

generally not irreducible with respect to [fl" _ l ' We may therefore ask which
irreducible representations of [fl" _ 1 will occur in the reduction of an
irreducible representation of [fl" and the answer comes immediately from the
character table.
Let T(2.1 denote the irreducible representation of [fl" corresponding to a
partition oc" of n. The reduction with respe ct to [fl n- l is then wr itten

(17.16)
)';,, -1

where T(" - I) is an irreducible representation of .51'" _ 1 and the sum runs over all
partitions (1." _ 1 of n - I .
The character ofT( ") for those elements which lie in [fl" _ 1 is obtained from
the character table of Y " by selecting only those classes with at least one cycle
of length unity. Then with the help of the full character table for [fl" _ 1 the
formula (4.28) enables us to deduce the coefficients m(oc" _ d.
For an example, consider n = 4. The only classes of [fl 4 which contain
elements of.Cf' J will be (1111), (211) and (31) and they contain, respectively, the
elements of the classes (Ill), (21) and (3) of Y J . We give, in table 17.2 the
characters for TI,.I and T(' ,)taken from table 17.1. The results of the reduction
are also given on the right , where it is seen that the m(ocJ) are either zero or one.

Table 17.2

(III) /21) (3)


--_ . _ -_ . _. _ -- ~~~ -- - - - - - -- - - _ . _--

I I 1
2 o - I
I - 1 1
~- - -_ ._ ~ -- --- ~ -- -- - _ . _ - - - _. _-- - -- ------

I I 1
3 1 o
2 o -1
3 -1 o
I -1 1
-~--- - - - - - -

It may be shown generally that the coefficients m(ocn _ 1) in equation (17.16)


are always either zero or one and a very simple rule may be given to deduce
which m( (1.n _ 1) are equal to one . In the language of Young diagrams it says
simply that the Young diagrams (1." _ 1 for which m( (1." _ 1 ) = I are obtained
from the Young diagram (1." by the removal of one square. For example the
EJ=cJ
§.
removal of one square from the Young diagram can lead to the

diagrams lID and Efl bnt not to


436 The Symmetric Group Y 17.8
n

17.8 The basis vectors of the irreducible


representations
The problem of labelling the basis vectors of the irreducible representation
T(~·lof Y. may be solved by studying their behaviour under the chain of
subgroups Y. --+ Y . _ 1 --+ Y . - 2 • • . --+ Y 2' At each stage we use the reduction
(17.16) described in the preceding section. The chain of subgroups is obtained
by first removing object n, then object (n - I), and so on, until finally only
objects 2 and 1 remain in Y 2 . Each basis vector of T(~') is assigned a sequence
of labels (X., (X. _ l ' . . •, (X2 describing its behaviour under this succession of
subgroups. The argument is conveniently divided into four steps .

(1) It is trivial that all basis vectors of the irreducible representation 1'2.) of
Y. may be assigned, as a first label, the same symbol (X••
(2) In the reduction of T(~'), on restricting to the subgroup Y . _ l ' the basis
vectors may be chosen to belong to one or another of the irreducible
representations T(~' -1) of Y . _ 1 which occur in (17.16). As a second label the
basis functions may therefore carry the appropriate symbol :x._ 1. Since no
representation T (~" - I) occurs more than once, the number of basis vectors
carrying the symbol (X. _ 1 will be equal to the dimension of the representation
T(a.- al.
(3) Proceeding from Y . _ 1 to Y . _ 2' each irreducible representation T(~' I )
will reduce to a sum of irreducible representations T (a. -, )of Y . _ 2 . The set of
basis vectors with the same pair oflabels (X. (X. _ 1 may now be chosen to belong
to one or another of the T(a. - , J. Thus the basis vectors acquire a further label
(X. _ 2 to include with (X. and (X. _ 1 . Again the number of basis vectors carrying
the sequence of labels (X. (X. _ 1 (X. _ 2 will be equal to the dimension of the
irreducible representation rv- -zl
(4) Clearly this process may be continued until eventually the basis vectors
carry the sequence of labels (X. (X. _ 1 (I. _ 2 •• • (X2 . Again, the number of basis
vectors with a given sequence will be equal to the dimension of the last
irreducible representation in the chain. But this is now T(2,)and the group Y 2
has only one-dimensional irreducible representations. Hence the labelling by
the sequence (X. (X. _ 1 a. _ 2 • •. a2 is complete in the sense that no two basis
vectors of T(~") will have the same sequence and that, to every seq uence, there
corresponds a basis vector ofT(~·). It must ofcourse be remembered that in any
such sequence the Young diagram a. _ 1 must be obtainable from a. by the
removal of one square, that a. _ 2 mu st be obtainable from a._ 1 by the removal
of one square, and so on .

The sequence (X. a. _ 1 (X. _ 2 ••• (X2 may seem to be a very complicated label
since each a, represents a Young diagram of r squares. It is possible however to
denote the set by a single Young diagram a. in which the numbers 1 to n have
been entered into the squares. The resulting picture is called a Young tableau.
17.8 The Symmetric Group Y
n
437

The procedure for entering the numbers is to put the number n into that square
which, when removed from (X., gives the Young diagram (X. _I ' The number
n - 1is put into that square of (X. _ I which , when removed, produces the Young
diagram (X. _ 2, and so on . By the manner of construction it is clear that the
numbers in the squares of a Young tableau must decrease to the left and
upwards.
Because of the completeness of this labelling system it follows immediately
that the dimension of the irreducible representation T(~·) is equal to the
number of different tableaux obtainable from the Young diagram (Xe - This
provides a simple way of deducing, see Littlewood (1950) in the bibliography,
the dimension of any irreducible representation ofY., although the dimension
also appears in the character table as the character of the identity element.
Table 17.3 gives the Young tableaux for each irreducible representation of
Y 2' Y 3' Y 4 . For example, the representation [31] of Y 4 gives rise to three

Table 17.3 Young tableaux for n = 2, 3 and 4

Partition Young tableau Partition Young tableau

[II] [31]

r
[22]

[21] [211] 4
[ill]
~ ITEJ
~ 2
~ ~ 3
[Ill] [1111 ]

W
tableaux corresponding to three basis vectors. This agrees with the dimension
given in the character table 17.1. The equivalence between the clumsy notation
(X4 (X3 (X2 used earlier and the Young tableau is illustrated by the example

CD == [31) [21) [2)

An even more convenient notation, for pnntmg, was introduced by


Yamanouchi. He defined a Yamanouchi symbol (rnr._ I . . . r.) in place of the
Young tableau with rk denoting the row in which the number k appears. The
438 The Symmetric Group /1'
n
17 .9

Yamanouchi symbol in the example given above would be (1 21 1). That this
notation is sufficient to specify the tableau is soon clear if one starts at r 1 and
moves left, remembering that the numbers in the tableau must increase to the
right and downwards. Incidentally, it must always happen that r 1 = I, r 2 = I
or 2, etc. One advantage of the Yamanouchi notation is that by removing rn
one obtains the Yamanouchi symbol specifying the behaviour of the basis
vector of T(~·) under the group Y n _ 1 and the successive removal of rn - l '
rn _ 2' . . ., r3 specifies the behaviour of the basis vector under the chain of
subgroups Y n - 2 , Y n - 3 , · · · , Y 2 •

17.9 Examples of basis vectors and representation


matrices
The labelling system introduced in the previous section for the basis vectors of
an irreducible representation is beautifully systematic but rather formal. We
now construct some concrete examples of basis vectors for small n, starting
from the funct ion (17.13). This will also lead to the actual matrices of the
irreducible representations. An elegant general prescription for writing down
these matrices is given later, in section 17.13.
For any n, the one-dimensional representations [n] and [111 . . . ] clearly
correspond to the totally symmetric and totally antisymmetric representations
described in section 17.4. Starting from small n the first representation which is
not of this kind corresponds to the partition [21] of n = 3 and the appropriate
function (17.13) is
(17.17)
and we abbreviate the notation by writing $( [21]) = 1112 > where, in general,
Iij k > is an abbreviation for 4>i(1) 4>P) 4>k(3). In the notation of section 17.5
the space L([21]) generated by permutations on $([21]) will be three-
dimensional with basis vectors 1112 >, 1121> and 1211 >. The character table
17.1 in section 17.5 shows that it reduces into a sum of the two irreducible
representations T(3 J and T[21 J. The totally symmetric vector is clearly

0([3]) = {1112 > +1121 > +1 211 >} / J 3 (17.18)


with the J 3 included for normalisation and the orthogonal complement, the
two-dimensional space of vectors orthogonal to (17.18), must provide the
representation T[21 J. There is naturally some freedom in choosing a pair of
basis vectors in this space and if we take the pair
O( [21] a) = PI112 > -1121 > -1211 > }/ J6 (17.19)
8([ 21] b) = {1121 > -1211 >}/ J2
it is seen that they are not only orthogonal to (17.18) and to each other but also
belong to irreducible representations of the subgroup Y 2 on particles 1 and 2.
The first, 8( [21 ]a), is symmetric under P 12 while 8( [21 ]b) is antisymmetric.
17 .10 The Symmetric Group Y' 439
n

Thus, in terms of tableaux they correspond to ffi2l and f±PJ' respectively.


The matrix of P 12 in this basis of T[2I] is immediately given by

and from (17.19) the matrix of P 2 3 is soon found by carrying out the
permutation, for example
P 2 3 O( [21] a) = PII 21> -1112> -1 2 11> }/ J 6
= -to([21]a)+AO([21]b)
leading to the matrix

( -t -.Ii )
P23 = -.Ii t
It was shown in section 17.1 that any permutation could be expressed as a
product of adjacent transpositions so that , having found matrices for P 12 and
P 2 3 , all other matrices in the representation T[21] will follow from the known
rules for multiplying group elements. Thus, for example, P 13 = P 12 P 23 P 12 so
that the matrix of P 13 is given by the product of matrices

-I
0) (-t-.Ii -.Ii )
t
(I° 0) = (- -.Ii-t
-1
A similar exercise for n = 4 is suggested in problem 17.5. In section 17.13 we
give a simple prescription for writing down the matrices of adjacent
transpositions in any irreducible representation but the derivation of the
prescription is not simple.
We note that, as a result of the isomorphism between Y' 3 and the point
group D 3 , see section 2.3, the matrices given above are the same as those given
in table 4.1 with a re-ordering of basis vectors.

17.10 The direct product of two representations

Given any two representations T(<X') and T(<X') we can form the (direct) product
representation T(<X,)x T(<X') whose dimension is the product of the dimensions
ofT(<Xd and T(<X,J. This concept was described generally in section 4.17 and may
of course be used in the group Y' e- The reduction of the product representation
into a sum of irreducible representations was studied by using the character
table, see the formula (4.45). Thus, for example, table 17.1 of section 17.5 tells
+
us that T[2I]X T(21] = T(3] T(21] T[111]. +
There are two important particular examples of such products which are
worth describing here . Consider first the product T(<XJ x T[111 .. . ] of any
irreducible representation with the totally antisymmetric representation. Since
440 The Symmetric Group Y 17 .10
n

T[ III ... J is one-dimen sional the product has the same dimension as T I"l.
Furthermore if Xl" )denotes the charact er of T'" for a permutation in a class C(; P
then the character of the product representation is simply 7rpX~)' where 7rp
denotes the parity of elements in the class C(/ p ' Here we have used the result
(17.10) that "» is the character of T[ III . . . J. Now if T'" is irreducible it follows
immediately from the criterio n (4.29) that the product representation is also
irreducible using 7r; = I. Hence we may write
T l"lx T[ III ... ] = T(~l with X (~P l = 7r X (" )
p p
(17.20)
where ti labels another irreducible representation of Y n' For example if
(X =
[n] then a = [III . . . ] so that, in the language of Young diagrams, the
two representations Tt")and T(1) are related by an interchange of the role of
rows and columns. Inspection of table 17.1 shows the same result for the
example, (X = [31 ] with & = [211]. In fact it may be sho wn generally that the
Young diagram for (X and & are obtained from each other by interchanging the
roles of rows and columns or , in other words, by a reflection in the downward
diagonal at 45° from the top left-hand corner of the diagram. We say that the
representation T (1)is the 'adjoint' ofT'a)and, naturall y, some representations
like T [21] and T [22] are 'self-adjoint'. In a self-adjoint representation the
character must va nish for a class with odd parity, as is verified fro m table 17.1.
Consider next the reduction
T(alx T IP) = I myTIYl (17.21)
r
of the product of two irreducible representations and let us determine two
rather special coefficients m y, namely m[n ] and m[1 J I . .. ]- From the general
formula (4.45) we have
m[n] = (n!) -I I cp x ~") x~) (17.22)
p

m[11 1 ... ] = (n!)- I "L ,Cp7rp XI")


p
XIP)
p
= (n!)-I " c Xl1lXlPl
L,p p p
p p

having inserted the known characters 1 and 7rp for T[n] and T[III ...] and used
equation (17.20). Finally, using the orthogonality formula (4.25b) for charac-
ters and the known property that the characters of Y n are real, we have from
equations (17.22)
m[n]=8",lb m[III . . . ]=8;,p (17.23)
The meaning of these results is that the totally symmetric representation can be
produced onl y if the two irreducible factors are the same, or at least equivalent
in the sense defined in section 4.6. In the same way, total ant isymmetry may be
obtained only in the product of a representation with its adjoint. We ha ve
made use of the latter result in constructing antisymmetric wave functions in
atomic and nuclear problems, see subsection 8.6.4 and chapter 12. For
example, in the atomic structure problems of chapter 8, an orbital state with
symmetry (X must be combined with a spin state of symmetry &. Furthermore,
17 ,11 The Symmetric Group Y' 441
n

since an electron with spin 5 = t has only two spin states, the Young diagrams
(i may have , at most, two rows , Thus the Young diagrams (X for the orbital part
must have, at most, two columns. The precise relat ion between the total spin S
and the Young diagram a is discussed in section 18.10. The result is that S
= tOil - il 2 ), where ill and il 2 are the lengths of the two row s of a,
To actually produce such totally symmetric or a ntisymmetric functions
from the products it is of course necessary to sum over products of different
basis vectors of the two factors with appropriate Clebsch - Gordan coefficients
following the general arguments of section 4.17. These coefficients are easily
deduced, see problem 17.6,

17.11 The outer product of two irreducible


representations

In this section and the next we study a new kind of product of two
representations called the outer prod uct. Starting from a representation T(2) of
Y'n and a representation T(2') of Y'n' we generate a representation T of Y' n+ n' .
From the physical point of view this process is useful in building up wave
functions for (n + n') particles, given those of the smaller systems of nand n'
particles, but we shall not give examples. From the mathematical point of view
there is an interesting connection between this outer product for the symmetric
groups and the usual inner product for the unitary groups. This connection is
described in chapter 18.
Consider products of the type
};(I, 2, . , ., n) gj(n+ 1,n+2" . .,n+n') (17.24)
where the}; transform according to some irreducible representation T(2) of the
group Y'n on particles 1 to n and the gj transform according to some
representation T(2') of Y' n' on particles n + 1 to n + n'. The set of products
(17.24), where j runs through the s" basis vectors of the representation T(2) andj
runs through the 52' basis vectors of T(2'), will not form an invariant space
under Y'n+ n' but if one enlarges the space to allow all permutations between
the particles inland those in 9 then this larger space will be invariant. Allowing
for the number (n + n')! jn!n'! of ways of selecting n particles from (n + n'), the
dimension of the larger space is given by
(17.25)
Hence this large space will provide a representation T of Y' n + n' but not, in
general, an irreducible one . The representation T of Y' n + n ' is called the 'outer
product' of the representations T"I of Y'nand T" ') of Y' n' and we use the notation
T = T(2) ® T(2'). We are therefore faced with an interesting problem: given the
irreducible representation labels (X and (x' which are partitions of nand n' ,
respectively, can we deduce the possible irreducible representations of Y' n + n'
which appear in the reduction ofT? Let us first consider a few simple examples.
442 The Symmetric Group yo
n
17 .11

If n = n' = I, then the dimension of T is two, with the two basis vectors
j( I )g(2) andj(2)g( I). It is clear that by taking the symmetric and antisymmetric
combinations, thi s two-dimensional representation reduces with respect to the
group yo2 ' In terms of Young diagrams we would write

For the product rn ® D with n = 2, n' = I there are three basis vectors
and, if we write the symmetric two-particle functionf(l , 2) of equat ion (17.24)
as ¢(1)¢(2), then they may be taken as ¢(I)¢(2)g(3), ¢(I)¢(3)g(2) and
¢(2)¢(3)g(l). This set is identical with the set generated from the function
<1>( [21] ) of equation (I7.13)and we have seen in table 17.1 of section 17.5 and
in section 17.9 how this three-dimensional space breaks up into a single
symmetric vector and a pair of mixed symmetry [21]. Thus
rn @ D = OIJ + BJ (17.26)

This result may also be deduced in a way which leads to a simple


generalisation. Expand the product ¢( I) ¢(2) g(3) into components cor-
responding to the tableaux of the Young diagrams of n = 3. Since it is
symmetric in I and 2 it can only involve the two tableaux IJ.:TI.TI] and HPJ
in which I and 2 are in the same row . If now we generate all possible functions

ffiIl
P¢(I)¢(2)g(3) as P runs through Y03 the only new component which can be
produced will be the remaining basis function of the representation

[21]. Thus we have equation (17.26) which says that the outer product contains
all irreducible representations of yo3 which may be formed by adding one
square to the original diagram m
This argument may be extended to the more general reduction
0 =t
T(2) ®
.. T(7') (17.27)

where the sum runs over those tableaux a' which are obtainable by adding one
square to the tableau a. It is even possible to deduce a rule for obtaining the
coefficients m(f3; a, a') for the most general case
T = T 12 ) ® T 12 ' ) = tm(f3 ; a, a')TI/l) (17.28)
(i

where f3 is a partition of n + n'. The rule is stated below without proof:


(1) Place a letter a in each square of the first row of the diagram a', a letter b
in each square of the second row and so on.
(2) Add the squares of a', complete with their letters, to the Young diagram
a subject to the conditions (i) that the a's are added before the b's which in turn
17 .12 The Symmetric Group Y
n
443

are added before the c's and so on, (ii) that a Young diagram is obtained at
every stage, (iii) that no two identical letters appear in the same column, (iv)that
if the added letters are finally read off from the completed diagram beginning
with the first row and reading from right to left, then at no stage must the
number of b's exceed the number of a's or the number of c's exceed the number
of b's, etc.
(3) The final Young diagrams 13 which result from this process when all
letters have been added will appear in the sum (17.28) with non-zero
coefficients. If, by the given procedure, a particular diagram 13 may be built in a
number p of different ways then for that 13, the coefficient m(f3;a, rx') = p in
equation (17.28).
Consider two examples.
(i) IIIII@~ L-.L..--'--'------'-I_o,-I0--,-10-JI + ~ [ I I0 I0 I

so that

(ii)

so that
T(2I] @ T[21] = T[42] + T[411] + T[33] + 2T[321]
+ T[3I.II] + T[222]+ T[22J I] (17.29)
In example (ii) we see that the partition [321] arises twice in the process,
meaning that the representation T[321] appears with m = 2 in the reduction
(17.28). It is worth making a dimension check on such reductions using
equation (17.25). The dimensions in the two cases are I x I x 7!/4!3! = 35 in
example (i) and 2 x 2 x 6!/3!3! = 80 in example (ii). These numbers are soon
verified to be equal to the sums of the dimensions of the representations TtP)
obtained in the reductions, including the factors m(f3; rx, «'),

17.12 Restriction to a subgroup and the outer


product
In section 17.7 we studied the reduction of irreducible representations of Y n
following the restriction to the subgroup Y n _ ,. More generally, the product
444 The Symmetric Group .'I' 17 .12
n

group [/. x [/. ' is a subgroup of .<f'. +.' (e.g. in section 17.11, [/. could refer to
particles numbered I, 2, . .., nand [/. ' refer to particles n + I, n + 2, .. .,
n + n'.) Consequently an irr ed ucible representation T ' fl )of[/ n + n' will generally
reduce as a representation of the subgroup .<f'. x Y n' ,

(17.30)

where the two Young diagram labels a. and ::x.' are necessary to specify an
irreducible representation of the product group [/. x [/ n' . The coefficients
m(p;::1., a.') in this reduction may be deduced in the usual way with the help of
character tables and using the result (4.66) for the characters of a product
group.
The reader may have noticed that the same notation m(p ; a., a.') has been used
for the coefficients in two apparently unrelated equations, (17.28) and (17.30).
This is deliberate because we now proceed to show that the coefficients are the
same, as the notation implies . From equation (17.30) we have

XIP) = L m(p ; a, a.')x(' d )


:L.~ '

and using the orthogonality of the irreducible characters of [/. x [/n the
coefficients are given by
I
-,
n!n!
LG.G '
XIJlJ(GG ')X(, x,'l(GG') = m({3;a.,a.') (17.31)

where G is an element of [/. and G' is an element of [/ n ' The corresponding


formula for the coefficient s in the outer product (17.28) is
I
(-n + n'-)1. L xfl(H)x(H) = m(p;a.,a.') (17.32)
H

where H is an element of Y n + n and X is the character of the outer product


representation T = T I') e Tl" ). Remembering the type of function (17.24)
from which T was generated it is seen that any class of [/ n + n' whose elements
do not belong to a subgroup Y . x [/ n must have zero character in T, i.e.
X(H) = 0 for elements H in such classes . For the remaining elements it can be
shown that the character is given by
(n + n')! - ,
NHX(H) = , N GG'x"X ' )(G G') (17.33)
n!n!
where GG' is an element of [/. x [/ n belonging to the same class as H, where
N H denotes the number of elements in this class of Y n + .. and NGG ' denotes the
number of elements in the corresponding class of [/. x [/n ' Finally, inserting
this result (17.33) into equation (17.32) and remembering that all elements in
the same class give the same contribution in the sums (17.31) and (17.32) we see
the equivalence of these two equations and hence the justification for using the
same notation m(p ; a., a.') in equations (17.28) and (17.30).
17.13 The Symmetric Group Y
n
445

For the particular case n' = 1, the reduction (17.30) is none other than the
reduction Y n + 1 --+ Y n described in section 17.7. It was seen there that the
coefficients m were all unity which provides another way of deducing equation
(17.27).
As an example of the reduction (17.30) consider the group Y 4 and its
subgroup Y 2 X Y 2 for which the character table is reproduced in table 17.4:

Table 17.4

(l IXII) (1IX 2) (2X 11) (2X2)

[/2 X [/ 2 [2] x [2]


[2] x [ l l] - I - I
[1I] x[ 2] - I - I
[11] x[lI] - I - I

[/. [4] =[ 2] x [2]


[ 31] - 1 = [2] x [ 2] -i- [2] x [ II] +[11] x [ 2]
[22] o o = [ 2] x [2] + [ 11] x [ I I]
[211] - I -I - I = [ 2] x [ 11 ]+ [11] x [2] + [ 11] x [ 11 ]
[ 111 1] - I - I =[ II] x[l I]

The classes for the product group are just the possible products of the classes
(11) and (2) for each Y 2 group. Likewise the irreducible representations, all
one-dimensional, are the possible products of the representations [2] and
[11] of each Y 2 group. For the irreducible representations of Y 4 the
appropriate characters are read ofT from the character table for Y 4 ' given in
table 17.1 of section 17.5, by identifying the classes (11)(2) and (2)(11) as
belonging to the class (211) of Y 4 ' while the product (2)(2) belongs to the class
(22).The reductions given to the right of the table are deducible by inspection,
although of course one may use the orthogonality formula (4.28) if in doubt.
From this example we may deduce the corresponding reduction (17.28) for
the outer product
[2] <9 [2] = [4]+[31]+[22] (17.34)
by quoting the appropriate coefficients m(p; [2][2]) given in table 17.4.
One soon verifies the dimension check on equation (17.34), namely
1 x I x 4!/2!2! = 1 + 3 + 2.

17.13 The standard matrices of the irreducible


representations
In section 17.8 we described a labelling system for the basis vectors of
irreducible representations of Y n based on a chain of subgroups and using the
Young tableaux. Some explicit examples of such vectors were given in section
446 The Symmetric Group /1' 17 .13
n

17.9 and the corresponding representation matrices were deduced. In this


section we first give a completely general prescription, see Rutherford (1948) in
the bibliography, for writing down such matrices and after some examples the
prescription is later justified. Because of equivalence, these matrices are not
unique but they are the ones most commonly used so that we call them
standard matrices. They have the advantage of being orthogonal. Once the
matrices are known, the projection operators of section 4.19 (called Young
operators in the case of Y n) may be used to construct a set of basis vectors from
any given function . In a way, this is the reverse of the procedure followed in
section 17.9, where in simple cases the basis vectors were constructed first and
the matrices deduced from them.
Consider an irreducible representation T(iX,) of Y n • The partition :1n of n
defines the Young diagram and from section 17.8 the basis vectors are labelled
by the possible tableaux IXn IXn _ 1 • . • IX2 ' (As an example, the nine tableaux
corresponding to the partition [42] are drawn in table 17.5.) It was shown in
section 17.1 that any permutation may be written in terms of adjacent
transpositions P, _ I. i : If we call P i _ I. i the last transposition in the subgroup
Y j it follows that a general description of the matrix of the last transposition
P, - l .n in Y n will enable us to deduce matrices for all the adjacent
transpositions and hence for all permutations. Here we rely on the fact that the
tableau label ::tn IXn _ 1 ••• 1X 2 in Y n is at the same time a tableau label for the
chain of subgroups Y n - 1 , Y n - 2 , •• •, Y 2 by removing successively IXn ,
IXn - 1 , · · ·, 1X 3 •
Since P n _ l.n is the ident ity so far as objects 1, 2, . . ., n - 2 are concerned the
only non-zero matrix elements of'P, _ l. n will occur between those basi s vectors
for which the tableaux of numbers 1, 2 . . ., n - 2 are identical. We must
therefore enquire which different tableaux, for a given partition [n. n2 . .• ] ,
contain the same tableau of the numbers 1, 2, .. ., n - 2. We introduce a
notation y to denote a tableau for the partition [ n 1 n 2 • • • ] with 5'denoting the
tableau obtained by removing the numbers nand n - 1 from y. It is trivial that
the possible tableaux y may be divided into four types which we denote by: (a)
in which the numbers nand n - 1 lie in the same row ; (b) in which the numbers
nand n - 1 lie in the same column; (c 1) in which the numbers nand n - 1 are
neither in the same row nor the same column and in which n lies in a row lower
than n - 1;(c 2) which is like (c 1) except that n lies in a row higher than n - 1.
Thus in table 17.5 the first, eighth and ninth rows are of type (a), the second,
third and fourth are of type (c 1) and the fifth, sixth and seventh are of type (c 2)'
One sees immediately that the tableau 5' appearing in any tableau y of the
type (a) or (b) cannot appear in any other tableau of the partition [ n 1 n2 •• • ].
The tableau 5' appearing in a tableau of type (c 1 ) will however also appear in
the corresponding tableau of type (c 2 ) in which nand n - 1 have been
interchanged. Hence the only non-zero matrix elements of P n _ l.n are on the
diagonal in positions labelled by tableaux of type (a) or (b) and in the form of
2 x 2 matrices coupling pairs of tableaux of types (c.) and (C2)'
As an example, the non-zero elements of the matrix P 56 for the part ition
17.13 The Symmetric Group g
n
447

~- N

~- M

~
'-
0
;:::;'
....
f::.
-o

iJ
- N
,,"',-
?

~
c:
0
<;;
c
.'"
II)
- M

.
II)

0..

~
II)

Il)

:0
'u
.
::l
"0 - "'1"
Il)

~
II)

-5
c: -I"'
I
-o
0..'" - N
'-
0
se

~
' 1:
<;; 001<>
E ?
II)
s: - M
f-
.,...
-
-,;

~
~
-C>
~
- "'1"

lJ
- V)
448 The Symmetric Group Y 17.13
n

[42] are shown in table 17.5. The derivation of the numerical values is
described later. Notice how the first row is labelled by the tableau CiTiT3T4l
~
which is of type (a) . The tableau y for this row is 111 213141 which does not
occur again , implying that the only non-zero element in the first row must be
on the diagonal. Notice how the second row is of type (cd and has for its
partner of type (c 2 ) the fifth row, these two and none others having the tableau
ffilill in common. There are therefore non-zero matrix elements coupling

the second row and fifth column and the fifth row and second column but,
apart from the diagonal positions, all other matrix elements in the second and
fifth rows are zero . The rules for deducing the few non-zero matrix elements
are as follows:

(1) The diagonal matrix element for a tableau of type (a) is + l.


(2) The diagonal matrix element for a tableau of type (b) is -l.
(3) The diagonal matrix elements for tableaux of types (c.) and (c 2 ) are
given by :+ 1- 1, respectively , where 1is called the 'axial distance' from (n - 1) to
n in the tableau of type c 1 and is defined as the smallest number of steps
required to reach n from (n - 1), moving in horizontal and vertical steps from
one square to the next. A movement of one square to the left or downwards
counts as a positive step. In fact, the results (1) and (2) above are special cases
of (3).
(4) The off-diagonal matrix elements coupling a pair of tableaux of types
(c.) and ( c 2 ) are given by +(1-1 - 2 )!, thus ensuring that the matrix is
orthogonal.

These rules may be used, as explained above, to deduce not only the matrix
of P 5 6 ' as given in table 17.5, but also the matrices for all other adjacent
transpositions. Thus, for example, the matrix of P 34 in the representation T[ 42]
is as follows:

-1 J&1
J~ 3

-1
17 .13 The Symmetric Group Y 449
n

As a further example, the matrices of the transpositions P 12 and P 23 in the


representation [21] of Y 3 which were given in section 17.9 may be verified by
the rules above.
We close this section by justifying the rules given above . Denoting the
matrix of P n- I • n by T(Pn-l. n) we have T 2 = 1 because P~-l.n = E the
t
identity. We are also free to demand that T be a unitary matrix TT = 1
because of the result in section 4.6 that any representation of a finite group is
equivalent to a unitary representation. These two conditions on T show that
the diagonal elements for tableaux of types (a) and (b) must be ± I, while the
2 x 2 submatrix corresponding to a pair of tableaux of types (c 1) and (c 2) must
have the form
-a ~( I - a2) )
( ~ ( I - a2) (17.35)
a
with some real a. The choice of the positive square root simply defines the
relative phase of the basis vectors. It remains to show that a = 1- I , where I is
the axial distance from (n - I) to n in the tableau of type (c I) and we do this by
the method of induction. In other words we show that if the rules are valid for
(n - 2) and (n - 1) then they are valid for n so that knowing the validity in the
special cases n = 2 and 3 enables us to deduce the validity for all n. The proof
makes use of the identity

Pn- l.n Pn- 2,n- I Pn- l. n = Pn- 2.n- I Pn- l.n Pn- 2.n - I (17.36)

which is soon verified. Suppose that the pair of tableaux of types (c I) and (c 2 )
occur in rows rand s of the representation T and let us construct the matrix
element in row r and column s, of the two equivalent expressions in equation
(17.36). We note that P, _ 2,n _ 1 must have a zero matrix element in this position
since the Young diagrams obtained by removing the number n from these two
tableaux have different shape and Pn- 2,n- I belongs to the subgroup Y n- I '
The diagonal matrix elements of Pn_ 2, n_ I are now assumed to follow the rules
set out above so that

1
Tss(Pn- 2.n-l)= -/ 2
where II is the axial distance from n - 2 to n - 1 in the tableau of row rand 12 is
the axial distance from n - 2 to n - 1 in the tableau of row s. Using (17.35) for
T(Pn-l.n) the identity (17.36) gives

a/~1 ~ ( I - a 2 ) - ~ ( I - a 2 ) / 2 I a = I~I ~ (1 - a 2) / 2 1

by considering row r and column s and this simplifies to a = (/2-/1)-1 .


However, since the tableaux of rows rand s differ only through the exchange of
nand n - 1 it follows that /2 is also the axial distance from n - 2 to n in the
tableau of row r. Hence the difference (/2 -II) must be the axial distance from n
- 1 to n in the tableau of row r, in other words a = 1- I , Thus having assumed
450 The Symmetric Group f:l'n 17.14

the rules to be valid for P n _ 2.n _ I we have shown that they are then valid for
P, - \,n and hence the inductive proof is complete. The rules (1) and (2) may be
regarded as special cases with axial distances of =+= 1, respectively.

17.14 The class operator L T(P i)


i<i
It can be shown (see appendix 3.3) that in any irreducible representation T(a)
the sum 1: T1a)(G ) of the representation operators over all group elementsG.
y y ,

in a chosen class C(j is a multiple of the unit matrix and that this multiple ). is
related to the character in that class by the formula ).. . = ~x.~) [s; where ~ is
the number of elements in the class and Sa is the dimension of the
representation. In the symmetric group f:l'n the transpositions P i j form a class
(211 .. . ), in the notation of section 17.3, so that as a particular case the sum

,L .r a)(P i j· ) of operators in an irreducible representation Tla) will be some


I <}

multiple l, of the unit operator and z, may be deduced from the character table.
This operator provides a convenient measure of the symmetry of the
representation. For example in the totally symmetric representation it has the
value 1n(n - 1) and is the negative of this in the totally antisymmetric
representation. In a self-adjoint representation it vanishes for the reason given
in section 17.10. We may regard this class operator as measuring the difference
between the number of symmetric and antisymmetric pairs remembering of
course that individual operators T(Pij) are not generally diagonal. There is a
simple formula for the numbers J, which enables them to be deduced quickly
from the Young diagram for T(a). Ifnpdenotes the length of the pth row and mq
denotes the length of the qth column then it may be shown, see problem 17.7,
that

).. , = 1L np(np -1) -1 L mq(mq -1) (17.37)


p q

Bibliography

A very complete account of the symmetric group, its representations and


projection operators is given by

Rutherford, D. E. (1948). Substitutional Analysis (Edinburgh University Press)

The Clebsch-Gordan coefficients for this group are discussed in some


detail by

Hamermesh, M. (1962). Group Theory and its Application to Physical Problems


(Addison Wesley, Reading, Mass.)
The Symmetric Group !I' n 451

The relation between the group !I' n and the continuous groups U Nand!Yi N
is described and useful characters and formulae are given by

Littlewood, D. E. (1950). The Theory of Group Characters (Oxford University


Press)

Problems

17.1 Write the permutation (~~ii~~ las a product of cycles and hence show that it
belongs to the class [321] and has odd parity.
17.2 Construct the character table for the group // 4 using the methods of section 4.15
and check your result with table 17.1 of section 17.5.
17.3 Construct the table 17.1 of values oftjJ[n,n, ... ] and X[n,n, . . . ] for the case
n = 4.
17.4 Find the dimension of the representation [32] of Y 5 by listing the possible
Young tableaux.
17.5 Use the methods of section 17.9 to construct representation matrices for the
permutations P I Z ' P 23 and P 3 4 in the representation T[3I ] Or.'/'4 and compare
your results with those given by the rules of section 17.13.
17.6 Show that if ; ; and gi denote sets of functions transforming according to an
orthogonal matrix representation T(~)of y
Y n then the sum F s = r;;.
gi is totally
symmetric. If hi denotes a set transforming according to the adjoint repre-
sentation T[ II ... J X T''" then show that Fa = r;; hi is totally antisymmetric.
17.7 Deduce the value(17.37)fortheclass operator 0 ~ r T(Pij)inarepresentation
i <t
T[ n, n , . . J by using the following steps: (a) Let I be the basis vector whose
Young tableau y has numbers 1 to n l in the first row , n 1 + 1 to n 1 +n z in the
second row , etc, and construct the function 9 = AI, where A is the product of the
antisymmetrisers for the columns of y. (b) Write 0 = Or + 0, + Om , where Or
contains pairs ij in the same row of y, Or contains pairs in the same column and
Om the remaining pairs. (c) Show that O,y = - 1L !!!q(m q - I )y. (d) Show that
(Or + Om)AI = A(Or + Om)}. (e) Show that OJ"=- ! r mp(m p - I)}. (f) Show
that AOmI = O. P
18
The Unitary Group UN

The unit ary groups SU 2 and SU 3 have been studied in so me detail in chapters
10 and 11 and their irreducible representations were found by constructing
diagrams in one and two dimensions, respectively, as in figure s 11.1 and 11.5 of
section 11.6. This method clearl y becomes difficult for larger N , as we found in
chapter 12 in the brief encounter with SU 4 and SU 6 . In the present chapter we
introduce a simple method for describing the irreducible representations of
U N for general N which makes use of some properties of the symmetric group
Y o obtained in the preceding chapter. In fact these two very different groups
are related in an intimate way which we shall demonstrate in this chapter. The
relation is seen most clearly in the physical situation where it is of most interest,
namely in the classification of the product wave functions for n particles, where
each single particle may be in one of N different states. Thus one may consider
both the permutations of the particles and the unitary transformations in the
N-dimensional vector space of states of each single particle. Although we shall
be deducing properties of the group U N quite generally we shall use the
language of particles and sta tes.
From the physical point of view we ha ve seen in chapters lO, 11 and 12 how
the groups SU 2 , SU 3 , SU 4 and SU 6 have significance as approximate
symmetry groups of the nuclear and elementary particle Hamiltonians. The
discussion of U N in this chapter for general N helps to unify those earlier
452
18 .1 The Unitary Group U 453
I N

treatments. We shall also see how the groups V 2/ + 1 and V 2 j + 1 may be used to
classify the complicated many-electron wave functions which may be built
from n electrons in valence orbits with angular momentum lor j . In this latter
problem the unitary groups are being used more as a mathematical device than
as a symmetry group of the Hamiltonian.
We first describe the irreducible representations of V N and then make use of
a chain V N --> V N _ I --> •• . V 1 of subgroups to label the basis vectors of these
representations. In section 18.7 the general results are related to the simple
cases of S V 2' S V 3 and S V 4 discussed in earlier chapters. The infinitesimal
operators for V N are described in section 18.8. Some applications of V N in the
study of many-particle wave functions are discussed in section 18.10.
Sections 18.11 and 18.12 are concerned with the characters of the irreducible
representations of V N and the last section discusses the connection between
SV 2 and Yf3 ·

18.1 The irreducible representations ofUN

In chapter 7 we used the commutation relations of the infinitesimal operators


to construct the irred ucible representations of the group Yf3 and the same
technique was used in chapter 11 for SV 3 ' In this chapter for the group V N of
unitary transformations in N dimensions we adopt a different approach.
Starting from the N basis vectors in the original N-dimensional space in which
V N is defined , we construct the product spaces with n factors. In this way new
representations are built up from the original N x N unitary matrices and by
suitably symmetrising the product spaces a complete set of irreducible
representations of V N will be constructed.
Let cP l' cP 2' . . . , cP N denote a set of basis vectors in the original N-
dimensional space so that the group V N is defined by the set of unitary
transformations
N
cPj = UcPj = I V;jcP; (18.1)
i = 1

where Vij is an arbitrary N x N unitary matrix. We now construct the product


space defined by the set of N" products of the kind
(18.2)
Each factor carries two labels , a suffix i, j, etc., with in the range from I to N
which describes its properties under the unitary transformation (18.1) and a
second labell , 2, . .. , n which distinguishes the factors and a llows the freedom
to form antisymmetric products . (Without this second label a product of the
kind cP;(I) cPP) - cPj(l)cP;(2) would vanish.) The order of the products is
irrelevant; for example cP;(I) cPj(2) means the same as cPj(2) cP;(I), and we usually
write them in increasing order I, 2, .. . , n of the second label. Notice that
whereas the second label must be different in each factor the suffices on
454 The Unitary Group UN 18.1

different factors may be the same. We shall refer to the second labell , 2, .. . , n
as the 'particle' label and the suffix as the 'state' label. (This language is taken
from the physical problem of constructing a product wave function for n
electrons, each of which may be in one of N degenerate single-particle states.
For example N = 2 would relate to the pos sible spin states of n electrons each
with spin s = t, while N = 21 + 1 would relate to the possible orbital states ofn
electrons in a valence orbit with angular momentum 1and projections m, = I,
1-1, . . . ,-1.)
The transformation induced in the product space (18.2) by the unitary
transformation (18.1) on all particles is denoted by T(U) and given by
T(U)<I> = ¢;(I)¢j(2)¢~(3) . .. ¢~(n)

L U ;,;U/Pk'k" , U P'p¢j' (1) ¢/ (2) ¢ k, (3) . . . ¢p,(n) (18.3)


i'j'k' ...p'
Hence the N°·dimensional space L of the products <I> is invariant under T(U)
and provides a representation T of U N given by the direct product
Ux Ux Ux . . . x U of the N x N matrices U with n factors. However, the
representation T is not irreducible and we now show how to reduce L into a
number of invariant subspaces by making use of permutations in the particle
labels. Clearly the space L is also invariant under the group g o of all
permutations of the particle labels and these permutations commute with the
unitary transformations T(U) since, for example,
T(U) P\ 2<I> = T(U)¢j(I)¢i(2)¢k(3) .. . ¢p(n)

L UfjU j'jUk'k ' " U p'p¢j'(I) ¢j'(2) ¢d3) . .. ¢p{n)


I'j'k' ... p'

while

and the right-hand sides of these equations are identical. Hence T(U) will
commute with the projection operators pl:z) for g o, where (IX) == [n\ n 2 •• •] is
an irreducible representation label for go ' Thus if we reduce L into subspaces
Ll:z) corresponding to each projection

L=L Ll:z)
12)

then it follows that each subspace Ll:z) is invariant under T(U) because
T(U) pl:z)<I> = pl:z) T(U)<I> = pl:z)<I>'

showing that T(U) operating on any member of LI:zI produces a vector also
lying within i». Furthermore, each space Ltd is clearly invariant under the
group go so that, since the permutations commute with the unitary
18.1 The Unitary Group UN 455

transformations, the space L'(I.) is invariant under the product group 9'. xU N.
Thus we may introduce a notation <1>~.j for the basis vectors of L'(I.) where, for a
unitary transformation,

T(U)<1>~~ = Ix'
U~!x<1>~~, (18.4)

and for a permutation P,

P<1>~~ = I,. ry; (P)<1>~), (18,5)

Here the index t is the usual row label for the irreducible representation T'(I.) of
9'" described in chapter 17, while x is a new index introduced to provide a basis
for all vectors of L'(I.) with the same t. Thus U'(I.) is a representation of UN and it
is this which interests us in the present chapter. Now the representation T'(I.)(P)
of 9'. is necessarily irreducible by its method of construction through the
projection operator P'(I.) . It is a remarkable fact that the representation U'(I.) of
UN generated in this way is also irreducible but we shall not give the proof. The
set of representations U'(I.) for all partitions of all integers n is not actually a
complete set but with a small extension described in section 18,9 they become
complete, The set of basis vectors <1>~~ of L'(I.) for fixed Ct. is conveniently pictured
in the rectangular array shown below
<1> '(I.) <1>'(1.) <1>'(1.)
11 12 13 <1>\~
<1>'(1.) <1>\7)
21 22 <1>~~,
<1>(~)
3\ (18 ,6)

<1>\71 <1>(~) <1>(~)


S",1 s~ 2 .\..d 3

The vectors in each column transform according to the (same) irreducible


representation T'(I.) of 9'., while each row gives the (same) representation U'(I.) of
UN . We use the same notation So as in chapter 17 for the dimension of T'" and
d; is introduced for the dimension of U'(I.) . To summarise the results so far ; the
separability of the labels is a consequence of the invariance of L'(I.) under the
product group 9'. x UN, the irreducibility of T'(I.) was imposed by using the
appropriate projection operator and the claim that the representation U'(I.) of
UN is irreducible has yet to be proved. For the representation T of 9'. xU N
generated in L we may write

T =I T'(I.)x U'(I.) (18.7)


o

and by equating the dimension on both sides of this equation we get the
relation

N· = I sod, (18.8)
456 The Unitary Group UN 18.2

between the dimensions Sa. and da of'T'" and u(a), respectively. It is important to
realise that the partition IX serves a double purpose of labelling irreducible
representations of both the finite group Y n and the continuous group UN . For
fixed N we may choose any positive integer n, corresponding to a different
number of factors in the products (18.2). Thus, as expected for a continuous
group, we have an infinite number of irreducible representations. There is,
however, one limitation on the possible partitions [n 1n2 .. .]. Remember
from chapter 17 that a partition into p parts implies total antisymmetry
between sets of p particles out of n. However if p > N this is impossible because
there are only N different single-particle states. Hence the irreducible
representations U'" of U N are labelled by the partitions [n 1 n2 • • •] of any
integer n into not more than N parts. In other words they may be labelled by
the Young diagrams with not more than N rows.

18.2 Some examples

To illustrate the rather abstract statements in the preceding section let us


consider some small values of Nand n.

(1) N = 2, n = 2 There are N" = 4 product functions


1" ) , 112 ),1 21 ) and J22 )
using the notation tP i(l)tPP)tPk(3) . . . tPp(n) == lijk . . . p ) . The irred ucible
representations of Y' 2 are labeIled by the two possible partit ions [2] and [II]
and the corresponding subspaces are defined by the following basis vectors,
L[l] : Il l ), (112 )+121 ») / J2, 122 )
L[ll]: (1 12 ) -1 2 1 » )/ J 2
Thus, for example, the partition [2] is seen to label a three-dimensional
representation of U 2 and, from sectio n 17.5,a one-dimensional representation
of Y 2 •
(2) N = 3, n = 2 Using the same notation as above, the 3 2 = 9-
dimensional space L reduces into
L(2]: 111 ) ,122 ),133 ) , (112)+121 ») / )2, (\ 13) + 13 I ) )/ ) 2,
(1 23 ) + 132 ) )/v!2
L[II] : (112 ) - 121 »)/ )2, (113 ) -131 ) )/ )2,
(123) -1 32) )/ }2
showing that the partitions [2] and [II] now label six-dimensional and three-
dimensional representations of U 3 '
(3) N = 2, n = 3 The 23 = 8 product functions are given by

II11 ) , IIl2 ) , 1121 ),1211 ) ,1122 ), 1212),1221 ) and 1222 )


18 .3 The Unitary Group UN 457

The subspace L(3] is simply the set of totally symmetric combinations of these
and is easily constructed,
L(3]: 1111), (1 112 ) + 1121) + 1211) )/ J 3,
1222), (i221) + 1212) + 1122»)/)3

The representation of Y 3 labelled by the partition [21J is two-dimensional, so


the corresponding array (18.6) has two rows. From the example considered in
section 17.9 or from the projection operator we find the array of basis vectors
may be taken as
L[21]: (21 112) -1 121) -1211»)/ -/6, (21221 ) -1212) -1122) )/ -/6
(i 121) -1 211 »)/ -/2, (1212) -1 122») /-/2
In each column we have chosen the vectors to transform according to the
standard representation T(21] described in section 17.9. We do not yet have a
standard for the representation U[21] but the two vectors in the first row are
distinguished by the number of 'particles' in each 'state'. They are clearly
orthogonal but any orthogonal pair formed from them by linear combination
would serve equally well. The setting up of a standard basis for the
representations UIz) is discussed in section 18.4. The fact that the same
representation of UN is generated by each row is demonstrated by showing
that the matrices of the general unitary transformation T(U) are the same in
these two spaces, making use of the result
<i'j'k'IT(U)lijk) = Uj"jUrPn
from equation (18.3). This completes the reduction of the eight-dimensional
space L and notice that there is no space L[lll] because N = 2. Any attempt to
project such a space would give zero.
(4) N = 3,n = 3 We leave this example as an exercise. The space Lnowhas
3 3 = 27 dimensions and is seen to reduce into LL3J with a dimension often, a
sixteen-dimensional space L[21] described by an array with two rows and eight
columns and a one-dimensional space L[I11].

18.3 The chain of subgroups


U N-+UN-1-+ U N-2-+' •• -+ U 2-+U 1

It is always convenient to identify a chain of subgroups since this helps in the


labelling of the basis vectors of representations. For example, the use of the
subgroup Yt 2 of 9t 3 enabled us in chapter 7 to label the basis vectors of the
representation OUI of 9t 3 by the number m. Again, in chapter 11 we labelled the
basis vectors of the SU 3 representations (AJl) by the numbers T, M r and Y
relating to the subgroups SU 2' 9t 2 and U I of SU 3' This device was also used for
the symmetric group Y ft in section 17.8 using the chain
Yft -+ Y ft- I -+ ... -+Y 2 •
458 The Unitary Group UN 18.3

For the unitary groups we use the chain of subgroups


UN - UN - I - . . . - U 2 - U 1 obtained by successively restricting the trans-
formations to matrices with one less dimension. We need only consider the
first step in th is reduction and it sim plifies the d iscussion if we consider the
restriction U N- U N_ I X U 1 which retains the sim ple U 1 transformations in
the one excluded dimension. The U 1 group consists of 1 x 1 matrices
exp( -iO) and, as for the group (}ft2 in section 7.3, has one-dimensional
irreducible representations exp( -imO) labelled by the integer m. To be
definite, let us take ¢ N as the basis vector of this excluded dimension in the
notation of sectio n 18.1 so that U N_ I refers to the basi s vectors ¢ l ' ¢2' .. .,
¢ N - I ' Thu s a product of the kind <1> in equation (18.2) will transform
acco rd ing to the representation exp ( - irO) of U 1 if it contains the vector ¢ N
exactly r times, i.e. if r of the n particles are in the sta te ¢ .'I/' The problem is to
study the reduction of the representation U(~·) of UN following restriction to
the subgroup U N_ I X U i - It is convenient to include the suffix n on IX. to
indicate that IXn is a partition of n or, in other words, a Young diagram with n
sq ua res. Let us first sta te the result and then ju stify it.
n

U(<I·) = t t' y ( ~. -,) x exp( - ir O) (18.9)


r= 0 2" _ r

where v e. .» denotes an irreducible representation of U N-I and the prime on


the summation means that only those Young diagrams C(n _ r occur wtucn may
be obtained from ern by the removal of r squares, no two of which come from
the same column. Before giving the derivation we illustrate the result with the
examples from sectio n 18.2 for the restriction U 3 - U 2 X Uland n = 2,

U[2] = y[2]-texp( -iO)Y[I] -texp(-2iO)Y[ O]


U[II] = y[ll] -texp( -iO)y[l]

Noting that the value of r correspond s to the number of particles occupying


the state ¢ 3 it is ea sy to identify which of the basis vectors given under example
(2) of section 18.2 corresponds to each Y(~ 2 -' ),

r = 0 y(2] : 111 ) , 122 ), (112 )+121 »)/ j2


U(2) r = 1 y[I) : (i13)+j31 »)/ j2, (i23)+132 »)/ j2
{ r = 2 y[O]: 133 )

r=0 Y[IIJ: (112 )-121») / j2


U[11]
{ r=1 y[l} (i13) -131 ) )/ j2, (\23)-\32 »)/ j2
18 .4 The Unitary Group UN 459

The vectors given here for y[2] and vl II] are precisely those given in example ( I)
of section 18.2. The two pairs of vectors given for the y[I] representation are
different but transform in the same way under the group U 2 on states <PI and
<P2' In fact they transform like the defining 2 x 2 matrices of U 2 corresponding
> >.
to n = I and the vectors II and 12 The remaining vector 133 is clearly >
invariant under U 2 '
To prove the reduction formula (18.9), or 'branching law' as it is sometimes
called, we make use of the result given in section 17.11 for the outer product of
irreducible representations of g n' First we reduce the space L of all products
(18.2) into subspaces L = i.
,= 0
L, corresponding to behaviour under the

subgroup U t - In other words, L, is defined by those products which contain r


particles in the state <P N and n-r in the remaining states <PI to <PN-I . To
analyse the structure of L, we may regard it as built up from products

(18.10)

by (a) taking all possible choices of i, j , .. " p from I, 2, . . ., N - I and


(b) carrying out all permutations which mix particle numbers I to n - rand
(n - r + I) to n. This product has a similar structure to the form (17.24) used in
discussing outer products. The step (a) is identical with the original reduction
of L described in section 18.1 except that nand N are replaced by n - rand
N - I. Thus. so far as particles I to 11- r are concerned, we can generate all
Young diagrams an r with n - r squares and no more than N - I rows . These
-r

diagrams an _ , will describe the behaviour both under UN _ I and under g n_, '
To find the behaviour of the products (18.10) under Y'n' we need to form the
outer product of 'J.n r of g n_ r with the totally symmetric representation [r] of
- r

.'/, which relates to the last r particles in (18.10). But this can be done as a
specia l case of the rule (17.28) with the result that for given 'J.n _ r the possible
Young diagrams an relating to ''/n are obtained by adding r squares to the
diagram an_ r in such a way that no two added squares lie in the same column.
Reversing this analysis we may now deduce the result (18.9) that , given anand r,
the possible representations of UN - I correspond to those Young diagrams
an-r which may be obtained by the removal ofr squares, no two of which lie in
the same column.

18.4 A labelling system for the basis vectors

In section 17.8 we used the chain of subgroups g n- 1 -+ g n- 2 . . . -+ g 2 of g n


to provide a unique labelling system for the basis vectors of the irreducible
representations T(2.) of g n' Exactly similar arguments may be used to obtain a
unique labelling system for the basis vectors of the representations U(2.)of U N
by using the chain of subgroups U N _ 1-+ U N _ 2 -+ . . . -+ U 1 described in the
460 The Unitary Group UN 18.4

preceding section. The result is that the basis vectors ofUl".) are labelled by the
sequence

an , CCn-rN '(Xn-r N -r N _I' . . 0' CXn -r N -r N _I- . . . -'2


(18.11)

of Young diagrams where an- ' N defines the representation of the subgroup
U N- I to which that vector belongs, an- 'N - ' N _ I refers likewise to the subgroup
UN _ 2 and so on . The justification that this labelling system is unique, i.e. that
no two basis vectors of U "·) have the same set of labels , follows so closely the
arguments in section 17.8 that we do not repeat them. As an illustration, the six
basis vectors of the representation U[2] of U 3 which were discussed in the
previous section would be assigned the sequences of partitions shown in the
second column of table 18.1.

Table 18.1
III ) [2] , [2], [2] [ill]
(j12 )+121 »)/ J2 [2] , [2], [I] [JTIJ
122 ) [2], [2], [0] 0IJ
(113 )+131 »)/ J2 [2], [I], [I] OJIJ
(123 )+132 »)/ J2 [2] , [I], [0] 0IJ
133 ) [2], [0] , [0] [ill]

As in section 17.8, the sequence of'Youqg diagrams, or partitions, is a clumsy


notation and a more concise labelling is achieved by drawing, for each basis
vector, the single Young diagram IX n with the 'state labels' 1 to N entered in the
squares in the following way. The number N is placed in those squares of IX n
which are absent in IXn -'IV' the number N - 1 is placed in those squares of IXn -'IV
which are absent in an-'IV-'IV _I and so on . This labelling is illustrated in the
third column of table 18.1. As a further example, the eight basis vectors of the
[21] representation of U 3 would be labelled by

HPBfJffiTIHPJBFJBfJBPJHPJ (18.12)

in which, for example, the first corresponds to the chain of partitions


[21] -+ [21] -+ [2], the sixth to the chain [21] -+ [11] -+ [1] and the eighth to
the chain [21] -+ [1] -+ [0].
18 .5 The Unitary Group UN 461

It must be stressed that this device of writing state labels into the Young
diagrams for U N is quite distinct from the device of writing in particle labels
which was used in section 17.8 for Y n'

18.5 The direct product of representations of UN

Given any two representations U(~·) and ut~'·) of U N we may form the direct
product as defined generally in section 4.17. This provides a larger repre-
sentation of dimension equal to the product of the dimensions of the two
original representations. We are interested in the reduction of this product
representation into its irreducible constituents and once again the result is
simple to state.

U(~n)x U(~~)= I m(f3; an,a~') UUI) (18.13)


Ii
where the sum runs over the Young diagrams f3 of n + n' squares and the
coefficients m(f3; an, a~') are given by the same rules which were stated in section
17.11 for the reduction of the outer product of representations of Y n and Y n "
Diagrams f3 with more than N rows are disregarded since no such repre-
sentations of U Ncan exist , see section 18.1. In fact, this is the only way in which
the number N enters equation (18.13). Before justifying this result we again
give some examples.
For the particular case n = 2, n' = 1 and an = [2] the coefficients
m(f3; an' a ~, ) were given in equation (17.26) and the reduction (18.13) is then
U(2] x U[ I] = U[3] +U(2I] (18.14)
For the reasons given above, the second term must be omitted when N = 1.
Of course, for N = 1 there is no reduction because the irreducible repre-
sentations are all one-dimensional exp( - inO)and labelled by the integer n. For
N = 2 the left-hand side of equation (18.14) has d imension 3 x 2 = 6 and on
the right the two terms have dimensions 4 and 2, respectively, see examples (1)
and (3) of section 18.2. For N = 3, the examples (2) and (4) of section 18.2 show
that the left-hand side now has dimension 6 x 3 = 18, while the right has term s
with dimension 10 and 8, respectively.
As a more complicated example consider the product U[2I]X U[2I] for which
the outer product reduction was given in equation (17.29). Without repeating
that equation here , we note that for N = 2 the product representation has a
dimension of 4 and reduces into a sum of U[42 ] and U [33] with dimensions 3
and 1, respectively, while for N = 3 the product representation has dimension
64 and reduces into the sum of U[42] , U[4I1] , U[3 3] (twice) and U[222] with
dimensions of 27, 10, 10, 2 x 8 and 1.
To justify the result (18.13) we make use of the property, deduced in section
17.12, that the coefficients m(f3; an, a~') relate also to the reduction (17.30) of a
representation T(P) of Y n x n' following restriction to the subgroup Y n X Y n' .
462 The Unitary Group UN 18.6

Consider the space L of products of the kind (18.2) for n + n' particles. As
explained in section 18.1 this space reduces

T = ! T(Plx UIi )
P
into irreducible representation s of the product group Y n + n ' X UN labell ed by
partitions f3 of n + n', Restriction to the subgroup Y n x Y n' X U N leads to the
further reduction

T = ! m(f3; CX n , CX~') TI~n)x T(7~')X VIP) (18.15)


P' ~n '~'
using equation (17.30). However, by considering the first n and the last n'
particles separately, applying equation (18.7) to each part, we also have the
reduction

T = I T( ~n)x U(<x·)xT( a~'lx V ( ~~' I (18.16)


C( n ' ~ n '

Finally, comparing this with the preceding equation gives the desired result

Ulan )x U(7~') = t m(f3; CX n , cx~.) UI Ii~


Ii
for the reduction of product representations of UN.

18.6 The restriction from UN to its subgroup SUN

The irreducible representations u ta) of U N were constructed in section 18.1


from products like (18.2), where o: denotes a partition of n, the number of
factors in the product. In this section we study the subgroup consisting of the
N x N unitary matrices with determinant equal to unity. It is denoted by SU N
and called the 'special unitary group' or 'unimodular unitary group'. The
relation between U Nand SU N may best be seen by writing U N= U I X SU Nas a
direct product where U 1 is the group of N x N matrices of the form exp( - it/J)
multiplied by the unit matrix and t/J is real. (This group U 1 is not the same as
the groups U 1 used in section 18.3 to transform just one of the N dimensions,
although the groups are isomorphic.)
If A and B are square matrices then it is an elementary property that
det AB = (det A) (det B). Hence the unitary condition VU t = 1 on the matrix
V implies that [det U 12 = 1. For any unitary matrix V we may therefore write
- -
V = exp( - it/J) V where det V = + 1. It is soon verified that the set of all
- -
unitary N x N matrices V with det U = + 1 satisfies the postulates ofa group
so that the group U N has been factorised into a U 1 group of the factor
exp( - it/J) and the group of the U which we call SU N .
From the general theory of representations of product groups in section
18 .6 The Unitary Group UN 463

4.21 we know that the irreducible representations uta) of V N must take the
form of a direct product of irreducible representations of V I and SV N.
The representations of V I are, of course, one-dimensional with the form
exp( - iml/J), where m is any integer. Since the representation UIZ) was
constructed from n factors in section 18.1 it follows that the direct product
form for UIZ) is
U IZ) = exp( - inl/J) x U(:i) (18.17)

where n is the number of squares in the Young tableau a. Hence, the irreducible
representation uta) of S V N has the same dimension as UI» . In other words, U IZ)
does not reduce on restriction from V N to the subgroup SV N. We may
therefore use the same partition notation (a) = (a) = [n l n 2 • • . ] to label the
irreducible representations ofSV N. However, whereas for V Nit is obvious that
two representations with different n are inequivalent that is not the case for
SV N as we now demonstrate.
Consider the representation [11 . . . 1] of V N, where n = N . It must be one-
dimensional since there is only one totally antisymmetric combination which
can be formed from the products (1 8.2), namely

'¥ = Ln(P)PcPdl)cP2(2)
p
... cPN(N) (18.18)

where the sum runs over all permutations of the particle labels and n( P)
denotes the parity of the permutation P. Using equation (18.3) the effect of a
unitary transformation on '¥ is given by

T(U),¥ = LV ii V j 2 • • • VpN Ln(P)PcPi(1)cPj(2) .. . cPp(N) (18.19)


ij . . . p P

Because of the pre sence of the antisymmetrising operator, the only com-
binations of suffices ij ... p to give a non-zero result are those in which no
suffix is repeated so that ij . . . p is a rearrangement of 12 . .. N. Using the
notation ( :~ . . .t ) for the permutation which carries the particle labels ij . .. p
into 12 . . . N the equat ion (18.19) may be written

T( U )'¥ = '\
L. Vii V j 2 · • •
'\
VpNL.n(P)P (i j . . .P ) . . · · · cPp(p)
cPi(l)cPh)
ij .. . P P 12.. . N

.. )
We stress that the permutation C~ · ~ operates only on the particle labels, as
does P. However, since the order of factors is not significant we may rearrange
them so that cP d I) comes first, cP 2 (2) second and so on . Thus we have

T(U) '¥
'\ PVii V
= ij '-:. j2 • • •
'\
V p N~ n(P)P 12
(i j P )
N cPdl)cP2(2) cP N(N)

= I.
ij ... P
Vii V j 2 .. . V PNn(~~' . .N) L n(P ')P'cPdl)cP2(2)
IJ . . . p P'
cPN(N)
464 The Unitary Group UN 18 .7

where we have defined P' as the product of permutations P' = P( ~~ : : :~) and
used the result of section 17.4 that the parity of a product is the product of the
parities. We have also used the fact that as P runs over the group 9' N so does P',
see section 2.9. Finally we note that the sum over P' is just 'P, while the sum
over ij . . . p gives the determinant of the matrix U so that
T(U) 'P = (det U) 'P (18.20)
The significance of the result (18.20) is that 'P is invariant for transfor-
mations of SU N and so we have an equivalence between the representation
labelled by the partition [11 . . . 1] of n = N and the identity representation
given trivially by n = O. In terms of Young diagrams this implies equivalence
between the single complete column of N rows and the trivial diagram with no
squares. Further equivalences now follow using the results of section 18.5 for
multiplying representations. If u(a) is any irreducible representation then from
the rules quoted for equation (18.13)
u(a) X U[II .. . I] = U(/ll

where the Young diagram for f3 is obtained by adding a complete column to


the left of the Young diagram for IX. But since U[II ... I] is equivalent to the
identity, this means that U'"' == U(P). Hence the addition, or subtraction, of one
(or any number of) complete columns gives an equivalent representation. It is
usual to denote the inequivalent irreducible representations by the Young
diagram with least squares, obtained by subtracting all complete columns of
length N . Thus for SU N we need only the Young diagrams with N - 1 rows. It
may be shown that no further equivalences exist.
The reduction of product representations in SU N follows from the rule
(18.13) for UN, making use of the equivalences given above. For example, in
the group SU 2 , the representations have only one row in their Young diagrams
and applying (18.13) for the product U[3]x U[2] of SU 2 gives

[IT]x[TI =1111I1+BTIJ+EtP

=111111+ [IT] + 0
In the more familiar notation D<i) which was used for the representation of
SU 2 in chapters 7 and 10, this is an example of equation (7.44)
Dd)x 0(1) = D(~)+ D(~)+ D(!)
using the identification of urn] with D(!n) which we discuss in the next section.

18.7 The special cases of SU 2 , SU 3 and SU4

We now make contact between the preceding discussion of SUN and the
particular descriptions of SU 2, SU 3 and SU 4 given in chapters 10-12. In the
18 .7 The Unitary Group UN 465

case of SU 2 the irreducible representations are labelled by Young diagrams


with one row of length n, in other words by a single integer n. However in
chapter 10 we used the notation Dei) withj = 0, t, 1,1, . . ., for the group SU 2
because it is homomorphic with :JI3 (see also section 18.13). We can soon see
that the relation j = tn unites the two treatments. This relation may be
deduced in various ways, one of which is to note that in Dei) the numberj refers
to the greatest value of the infinitesimal operator J, which has the matrix (b -i)
in the original two-dimensional space . Thus, on a product <I> of the type (18.2)
this infinitesimal operator is diagonal and its value is made up of contributions
+ t from every particle in state cP 1and - t from every particle in state rP 2' Thus
the greatest value is obtained by putting as many as possible of the n particles
into state cP. and for the totally symmetric functions corresponding to the
single-row Young diagram [n] this number is clearly n, so that j = tn.
Thus the integer nature of n leads to the familiar integer or half-integer nature
of j .
A similar argument enables us to show the identity between the labels (i.jl)
used in chapter 11 for the irreducible representations of SU 3 and the Young
diagram [n 1 nJ with two rows where n. = i. + u, n2 = jl. Remember from
chapter 11 that tri. + 2jl) is the greatest value of the infinitesimal operator Y
l'
which has the matrix 3\0b?0 - 8)2
in the original three-dimensional space and, for
this value ofY, the number F is the greatest value of the infinitesimal operator
T z. which has the matrix t{ b -?8). In the product <1>, the greatest Y is therefore
obtained by putting as fe~ p~~ticles into state rP3 as possible. However, to
produce a function with only a two-rowed Young diagram [n 1n2] it is
necessary only to occupy two different states. Thus all particles may be put into
states rPl and cP2 giving i.+2jl = n. +n 2 • For such functions the maximum
value of T: is obtained by putting as many particles as possible into state cP 1
and for the diagram [n. n 2] this is n•• with n2 particles in state rP2, so that
i. = n 1 - n2 • Combining this with the value for i. + 2jl gives n 1 = i. + u, n 2 = jl
which demonstrates the equivalence

Thus the octet (11) corresponds to the Young diagram B=J and the decuplet

(30) to [ I ] ] . The rules given in section 11.6 for multiplying S U 3


representations are now special cases of equation (18.13). In particular,
equation (ll.ll) follows from equation (l7.29), taking account of the
equivalences in SU 3 '
For the group SU 4 which was used in discussing nuclear structure in chapter
12 we may similarly relate the labels (P P' PO) to the partition. In this case a
three-rowed partition [n 1 n2 n3] is sufficient. Recall that P, P' and P" relate to
466 The Unitary Group UN 18 .8

maximum values of S" T, and Y z z which have explicit matrices

sz ~~e
2
0
I
0 0
0
0
-I
) C
o
o ,Tz =
1
2
0
0
0 0
-I 0
0 I j)
0)
0 0 0 -I 0 0 0

c
0
1 0 -I o 0
v., = 2 ~ 0 -I 0
0 o 1

in the original four-dimensional space of p f, nt, pi, n j in the notation of section


12.1. From these matrices we see that the combinations S, + T,; S, - Yzz and
T, - Yzz are the counting operators for the differences between the number of
particles in states ¢1 ' ¢2 and ¢3 ' respectively, and the number in state ¢4'
Since, for a three-rowed partition, we do not need to put any particles into state
¢4 ,it follows that P + P' = n l , P - P" = n2 and P' - P" = n 3 • Referring to the
examples in equation (12.6) we have the relations

DI ~ H ) =U(3], DIH!1 = U[21], DIH - 1) = U[III]

which agree with the assignment s in equation (12.5).

18.8 The infinitesimal operators of UN

Returning now to the group UN we note that in the original N-dimensional


space there are N 2 infinitesimal operators corresponding to the N 2 inde-
pendent Hermitian N x N matrices. It is convenient to take N of these to be
diagonal, with zero in all positions except for one diagonal matrix element
which has the value + 1. We call this set of operators Hi with i = I to N. The
remaining N(N - 1) operators may be combined, as for the operators T ±, U ±,
V ± of(lI.2), to form !N(N - I) pairs of raising and lowering operators E jj and
Ej j which are zero except for + I in the off-diagonal positions ij and ji,
respectively. They have the effect of shifting a particle from state ¢i to ¢j' so
that Eji¢k = 6ik ¢ j'
We now investigate the matrices of these operators in an arbitrary
irreducible representation U(~). In the first place , the operators H, commute
with each other and are the infinitesimal operators of the string
U 1 x U 1 X •• • x U 1 of N subgroups U 1 referring to each dimension. Thus,
using the results of section 18.3, we see that the matrices H~~) of H, in the
representation U'" are diagonal in the basis chosen in sect ion 18.4 with values
18 .9 The Unitary Group UN 467

rj in the notation used there. The set of numbers (r. t z . .. rN ) is calIed the
' weight' of the basis vector. For example the eight basis vectors given in
equation (18.12) have the folIowing weight s, in the sa me order,
(210 )(1 20 )(201 )(\ 11)(021 )(1 11)(102) (0 12 ) (18.21 )

Notice that the weight s do not pr ovide a unique labelIing system whereas the
sequence of Young diagram s described in sect ion 18.4 does. It is con ven ient to
order the po ssible weight s and we say that (r', r ' 2 . . . r ' N ) is greate r than (r i -
r 2 • • • r N) if r', > r i - For the case r, = r \ the ordering is done by the values of
r ' 2 and r 2 and so on . Thus the greatest weight in the set ( 18.2 1) is (210) . It can be
seen from the definiti on of r I that, in an arbitrar y irreducible representation
U '~ ' where (J. == [n, n2 . . . ], the greatest weight is the sa me as the partition
r. = n\ , r 2 = n2' . . . , etc . Thu s the greatest weight serves a s a label for the
representation U'·) , a result which is a generalisation of the use of the greatest
m( = j) to label the irreducible representations D,j, of .~ J .
The matrices of the ra ising and lowering operators may be obtained by a
generalisation of the method of subsection 7.4.2 making use a lso of the chain
of subgroups. We shall not give any details.

18.9 The complex conjugate representations of UN


and SUN
To every representat ion Ul ~ )of U N there is a corr espo ndi ng co mplex conjugate
representation U (7)* obta ined by taking the complex conjugates of alI the
matrix elements of U I. '. Thu s in particular for the infinitesima l operators th is
implies that the set of weights in U (7)* are the negati ves of those in U I. '.
Immediately we see that U (7)* cannot coincide with an y of the representations
U '· ) defined in sectio n 18.1 because they necessar ily ha ve positive weights.
One consequence of thi s is that the representation s U I. , described by alI
part itions a of an y n are not complete. However it may be sho wn that the
complete set of irreducible represent ations of U .\' is given by (det U) -IU'· ),
where 1 is any positive integer. From the arguments given in sectio n 18.6 it is
clear that the inclusion of the simple factor (det U) - I produces another
representation and if the weights of UI" are given by (r l ' r 2 , .. ., r N ) then the
weights of (det U) -IU I.' are given by (r\ -I, r 2 -I, ..., r » -I). To see this we
need to write U = 1 + ia jH j for an infin itesimal transformation by H j,
where a j is the corresponding small parameter. Then (det U)-I =
(l + ia i ) -I :::: 1- ila, but , by definition of the infinitesimal operator H~') in the
representation U I.I, we ha ve U I.) = 1 + iajH~% 1 so that for sma lI ai '

(det U) -IU(·)= 1 +iaj(H~"-I) (18.22)

which gives the desired result .


468 The Unitary Group UN 18 .9

Making use of this larger set of irreducible representations we can now show
the equivalence
(18.23)

for the complex conjugate representation where, as usual, o: denotes the


partition « = [n tn2 ' " nN]andp = [n l -nN,n l -nN -I> " ., n l -n 2,0].In
terms of Young diagrams, p may be obtained quickly from rx by removing the
diagram rx from the rectangular diagram with n l columns and N rows, as
indicated in figure 18.1. (Remember that some of the nj may be zero and the
first row of p is the last row of the figure, etc .)

nl

n2 I
n3
I
. ,'

N rows
f3

n N_ 1
I
nN I
Figure 18.1

Before justifying this equivalence we take a few examples.

N = 2 U[I]*=: (det ur ' U[l], U(2]*= (det U) - 2 U(2], U[II]* = (det U)-I U[O
U(2I]*=:(det Uj -ZU[I]
N=3 U[I]*=(det U)-IU[lI] , U(2]*=(det U) -2U(22], U[II]*=(det U)-IU[I
U(2I]*= (det U) -2 U[2I]

The expression (18.23) is not unique because from the arguments in


section 18.6, the addition of a full column to p and the inclusion of an extra
factor (det U)- I causes no change. The equivalence (18.23) may be justified by
showing that each side of the equation has the same set of weights. For
example, the maximum weight on the left is (-nN , -nN-" . . . , -n 2, -n 1)
corresponding to the minimum weight in Ual, namely (nN, nN-I, , n 2, nil.
Thus the maximum weight ofU(/I) will be (n l - nN, n l - nN _ I' , n l - n2, 0)
as stated below (18.23). Finally, we use a result that the set of weights defines
the representation.
For the group SUN the property det U = 1 leads to the much simpler
equivalence u(a)* = U(/I), where p is related to rx as before. Remembering also
the equivalences described in section 18.6 we see that for SU 2, u(nj* = u(n) so
that all representations of S U 2 and therefore 9i 3 are self-conjugate. For S U 3 '
18.10 The Unitary Group UN 469

using the ().Jl) notation, we have the general result

(I.Jl)* == ~ * == /IA.'+
IL:J" f == (JlI.)
so that in particular (10)* = (0 1), and (11) is self-conjugate.

18.10 The use of the group UNin classifying


many-particle wave functions

Because of the way in which we have introduced the group UN, through
products of the kind <I> in equ ation (18.2), it should come as no surprise that the
group U N has a role to play when constructing atomic wave functions for
severa l valence electrons. Consider n valence electrons in single-pa rticle states
</J'm with fixed 1and the usual range of possible values of m = I, (1- 1), .. . , -I.
From the arguments of sectio n 18.1 it is clear that the product wave functions
with permutat ion symmetry described by a partition (J. = [n J n2 .• • ] of n will
also transform according to the irreducible representat ion UbI of the group
U 21 + I of unitary tran sformations among the (21 + 1) single-particle states </J'm '
Although the Pauli prin ciple demands that the wave function be totally
antisymmetric with respect to the permutat ion of particles, the pa rtition (J.
refers only to the orbital coord inates. Thus (J. itself need not be the partition
[11 . . . 1] but it mu st be capa ble of yield ing thi s part ition when combined
with the spin part of the wave function . Th is problem was d iscussed in
subsect ion 8.6.4 and in section 17.10, where it was shown that the spin part
must transform according to the adjoint representation Ii of the group Y n of
part icle permutation s in the s pin coord inates. (Remember that Ii is related to (J.
by the interchange of row s and columns in the Young diagram.) However, the
spin part of the wave function is itself a product of the kind (18.2) with N = 2
so that the spin part of the wave function transforms according to the
representation U (:i ) of U 2' Thus the Young diagram Ii may ha ve two rows at
most and hence the orbital part ition (J. has two columns at most and, of course,
21 + 1 row s at mo st.
We have seen in sect ion s 18.6 and 18.7 the relation between representations
of U 2 ' SU 2 and fJ1i 3 so that Ii immediately determines the spin S of the system
of n particles. Ifwe write Ii = [ill ' il 2] then S = -! (ill - il 2) . Thus the label (J. of
the representation U'" of U 21 + I is already specified when nand S are given and
in a sense this new label (J. is superfluo us. A further problem is to deduce the
possible values of total orbital angular momentum L which may be associated
with given nand S. However, since L is the representation label for the group
fJ1i 3 we are asking how the representation Ulzl of U 21 + I reduces on restriction
to its subgroup fJ1i 3 '
U( ' )= ImLD(L ) (18.24)
L
(The coefficient mL should not be confused with the state label m.)
470 The Unitary Group UN 18 .10

But this is a familiar group-theoretical problem which may be answered in


various ways. In fact, a method for deducing the coefficients mt. was developed
in subsection 8.6.4 using the group Y n , without any reference to UN . An
alternative method is now available using the set of weights of Vlz l. If we label
our N single-particle states by m = 1,1- 1, . .. , - I, in that order, then the total
I
angular momentum projection operator L, is given by L, = I mH m in
m= - I
term s of the diagonal operators Hi introduced in sectio n 18.8. Thus the values
1
of M a re given simply in terms of the weights, M = I mr m ' For example, in
m = - I
the eight-dimensional representation [21] of U 3 , the set of weight s (18.21) lead
to the following set of M-values: 2,1, 1,0, -1,0, -1 , -2. From the known
multiplet structure of O (L ), i.e. M = L, L - 1, . .. , - L, we therefore conclude
that for the group U 3 ' the reduction (18.24) becomes U[2I] = 0(2) -i- O( 1). This
method is essentially using the group characters of U N which we discuss in
detail later in section 18.11; see problem 18.5 for another example. A third
alternative method for deducing the coefficients mL , which is less tedious for
large I, is explained in problem 18.6.
As an example of this method of classifying sta tes, let us return to the simple
problem of three p-electrons (/ = 1) discussed in subsectio n 8.6.4. Table 18.2
lists first the possible partitions IX of n = 3, followed in column two by the
adjoint partitions a. Notice that the partition IX = [3] must be excluded since
the corresponding a = [Ill] has more than two rows. The value of S follows
directly from S = t(n 1 - n2 ) . The value s of L for IX = [21] were deduced above
and the same technique shows that the one-dimensional representation [111]
has the single weight (Ill), implying that M = 0 only and hence that L = O.

Table 18.2

Orbita l symmet ry Sp in sy mmetry and


S L
and U 3 label rl U 2 label IX

[111J [3J J
2 0
[21J [21J t 2, I

The role of the group U N in atomic multiplets discussed above may be


extended immediately to the nuclear supermultiplets which were introduced in
chapter 12 using also the interpretation of the SU4 labels (P P' PO) which was
described in section 18.8. We give no details. As remarked in section 12.1 the
spin dependence of the nuclea r force is sufficiently strong that the SU 4
symmetry is strongly broken and the jj-coupling approximation makes a better
first approximation (see subsection 8.6.1 for a mention ofjj-coupling in atoms).
In this case each particle carries a definite total angular momentum i. which
includes both spin and orbital contributions, and the usual isospin of 1. When
18.10.1 The Unitary Group UN 471

n valence nucleons are considered it is clear that the group U 2j+ I takes the
place of U 21+ I ' The reduction from U 2j + I to 91 3 gives the values of J as in
equation (18.24). The association between J and isospin Tis analogous to the
association between Land S in the atomic case; for a given representation U II )
of U 2j+ I the isospin part of the wave function must have adjoint symmetry a
t
giving T = (ill - 11 2 ),

18.10.1 The use of subgroups of UN

The introduction of the groups U 2/+ I and U 2j+ I into these problems of
atomic and nuclear structure is in a sense disappointing because they carry no
new physical information. The representation labels IX are , as we have seen,
equivalent to the spin, supermultiplet or isospin labels, depending on the
system . However, they may be used as a first step in defining a new group
which, depending on the nature of the interaction between particles, may
carry useful physical information. To illustrate this idea, consider three
electrons in the atomicf-shell, i.e. I = 3. In the reduction (18.24) corresponding
to IX = [21] the non-zero values of m L are given by m I = m 6 = m 7 = m s = I,
m 2 = m3 = m4 = m s = 2. Unlike the case shown in table 18.2 there are now L-
values for which mi. > 1,so that the quantum numbers S, L , M sand M L are no
longer sufficient to define the possible states. Ofcourse physically this problem
is resolved in perturbation theory by diagonalising the Coulomb interaction
numerically. A 2 x 2 matrix occurs for L = 2, 3, 4 and 5. From the group-
theoretical point of view there is a possibility of finding a labelling system for
these pairs of states if one can find-see Racah (1965) in the bibliography-a
new group '§ which is a subgroup of U 2/ + I and which contains 91 3 as a
subgroup. One may then carry out the restriction from U 7 to 91 3 in two steps
U 7 -> '§ -> 91 3 , In this way the basis vectors (wave functions) would carry also
a representation label for the group '§. In some , but not necessarily all, cases
the pairs of states referred to above will carry different labels and therefore be
distinguished by their properties under this intermediate group '§. The
introduction of such a group '§ may be useful on two levels. It may simply be
used to provide a convenient, system atic labelling scheme within which to set
up the matrix of the interaction. It may also be significant physically , if the
interaction has approximate invariance with respect to '§. In this case the use
of a basis labelled by irreducible representations of '§ will approximately
diagonalise the energy matrix and the labels may be associated with physical
states.
To find the intermediate group '§ we must Impose restrictions on the unitary
transformations without excluding any of the ~ 3 transformations. A system-
atic way of doing this is first to write down the commutation relations of the
infinitesimal operators of U 2/ + I ' One then searches for a subset of the
operators of U 2/+ I which contains the operators of 91 3 and which is closed
under commutation. By this we mean that the commutator of any pair in the
subset gives a linear combination of operators in the subset. From the general
472 The Unitary Group UN 18 .10 .1

definitions of section 7.2 and from equation (7.7) in particular, such a subset
defines a group which is necessarily a subgroup of U Z/+ 1 and which contains
Yl 3 as a subgroup. By comparing the structure constants obtained for <§ in this
way with those of all known groups one may identify <§. This technique shows
that, for arbitrary I, the group Yl z1+ 1 may be chosen for <§ but that, for the
small 1of interest in physical problems, very few other possibilities exist. The
more direct way to introducefii ZI+ 1 is to restrict the unitary transformations U
by the condition that the two-particle function

(18.25)

in the notation of section 18.1, should be invariant. This implies that

<1> = T(U)<1> = L c; u; ¢j( 1) ¢k(2)


i.j.k
so that
L Uj i u; = s,
i

In other words, Uk; = (U- 1 b but since U is unitary th is implies that U is a real
orthogonal matrix, which defines the group Yl z1+ r- To show that the physical
Yl 3 is contained in this group Yl z/ + 1 we must find a basis ¢; for the (21 + 1)
states with angular momentum 1such that the function <1> defined in equation
(18.25) is invariant under rotations. One sees immediately that the usual m-
basis I/Im would not suffice. However if we define a new basis

¢:, = [I/Im +(- l)m 1/1 -m] /2!


¢,;; = i[ I/Im -( _l)m 1/1 - m]/ 2! with ¢o = 1/10
then we have

<1> = L [¢:'(1) ¢:'(2) + 4>';;(1) ¢ ';;(2) + ¢0(1) ¢0(2)]


m >O

= L (_l)m [I/Im(l) 1/1 - m(2)+ 1/1 - mOl I/Im(2)] + 1/10(1) 1/10(2)


m >0

(18.26)
m

This expression is recognised as the S-state (L = 0) of the two particles with


the usual sign conventions for the I/Im (see subsection AA.2), and is therefore
invariant under rotations.
We shall not go into any detail here concerning the irreducible repre-
sentations of Yl z/ + 1 except to remark that they are denoted by a set of 1integers
18.10.1 The Unitary Group U 473
N

(r 1 r 2 • • • r,) and we give in table 18.3 a few examples of the reductions for the
case 1 = 3. It is seen that the new ~ 7 label has succeeded in distinguishing the
two L = 3 states for n = 3 and S = ! but has failed to do so for the pairs of
states with L = 2, 4 and 5.

Table 18.3

n U7 Jt 7 Jt 3 (L)

[I] (I (0) 3

2 [2] (000) 0
(2(0) 2,4,6
[II] (110) I, 3, 5

3 [21] (I (0) 3
(210) 1, 2, 2, 3, 4, 4, 5, 5, 6, 7, 8
[111 ] (111) 0,2,3,4,6

In the jj-coupli ng problem where the group U 2j ~ 1 is encountered, one can


carry out a similar search (Racah, 1965) for a group '§ which contains ~ 3 and is
contained in U 2j ~ l : The result of such a sea rch leads to what is known as the
'symplectic group' SP2j~ r - One might have expected to find ~ 2j ~ 1 by the same
arguments as before but this is not the case. The last step in arriving at equation
(18.26) assumed that (- l )" = (- 1)- rn which is only true for integer m and
hence for integer I. The group SP2j + 1 may be identified by its infinitesimal
operators, as outlined earlier, but the most direct way to visualise the group is
to note that there is still only one invariant two-particle state in jj-coupling,
namely the state with J = 0, which is given as in equation (18.26) by
j
<1>= L (-ly -rnt/Jrn(I)t/J_m(2)
m = - j

= L (-ly -m[t/Jm(1)t/J -m(2)-t/J -m(1)t/Jm(2)] (18.27)


m >0

Notice that <1> is now a ntisymmetric under particle permutation so that it


cannot be written in the form (18.25) by any choice of basis . The set of matrices
which leaves this state <1> invariant may then be taken as the definition of the
symplectic group in (2j + I) dimensions. Its irreducible representations are
described by a set of U+!) integers, very like the group ~ 2/ ~ i - Again we give
no details except to quote some examples for the case j = 1 in table 18.4.
The word 'seniority' is sometimes used to describe the representations of
~ 2/~ 1 or SP2j+ i : One says that a state has seniority v if the smallest number of
particles for which that representation occurs is n = v. Thus for example the
two L = 3 states with S =! (i.e. [21]) for three particles shown in table 18.3
will have seniorities v = 1 and 3, respectively ; for (100) and (210). In table 18.4
474 The Unitary Group UN 18.10 .1

Table 18.4

n Vb SPb 9l 3 (J )

[I] (100) ~
2 [2] (200) 1,3,5
[11] (000) 0
(110) 2,4

3 [21] (100) t
(210) !, i, }, (W, t 1[-, ¥
[Ill] (100) 5
I
(111) J. 2
2, 2

one also sees that, for odd n and each U 6 symmetry, there is a state with v = 1
which necessarily has the same spin J = t as a single particle. For even n there
is always a state with seniority v = 0 and hence J = O.
From the physical point of view, the group 9l l /+ 1 has significance in
describing the levels of the configuration In if the perturbing interaction
between pairs of valence electrons exhibits degeneracies appropriate to the
group. In other words , for the case of 1= 3 illustrated in table 18.3 we would
need to find the L = 2,4 and 6 states for n = 2 at about the same energy and the
states with L = 1,3 and 5 lying together at some other energy. In practice with
the Coulomb forces this does not happen since the states with high L are
depressed, see subsection 8.6.5. Thus, in atoms, the group 9l z/ + 1 is useful only
as a means of setting up a systematic framework in which to diagonalise the
perturbation.
In nuclei, however, the strong spin - orbit force leads to jj-coupling and the
attractive force between valence nucleons then depresses the state with low J .
This is the reverse of the atomic situation since the force between electrons is
repulsive. In fact, the short-range attractive nuclear force not only depresses
the J = 0 state but also leaves the J = 2 and 4 states, shown in table 18.4 with
labels [11] (10),close together in energy . Hence in nuclei, the group SPlj + 1 has
real physical significance. As a result, the ground states of nuclei would be
expected to be described by irreducible representations of SPlj + i - Because in
the two-particle system the invariant (000) lies lowest in energy one might
expect the ground state for general n to contain as many such pairs as possible,
consistent with the Pauli principle. This may be verified in a more detailed
model which shows that for odd n the ground state has seniority v = 1, the
representation (100), while, for even numbers of neutrons and protons
separately, the ground state has v = 0 with representation (000). These
conclusions agree with, and hence provide justification for, the simple single-
particle nuclear shell-model which assumes that all pairs of neutrons and
protons couple to J = 0 leaving the spin of the nucleus equal to that of the last
odd nucleon. (See also Section 19.2.)
18.11 The Unitary Group U 475
N

The groups U 2/ ~ 1 and U 2 j ~ 1 are not symmetry groups of the Hamiltonian


in the way that the group !Jf3 was, in chapter 7. The operators of U 21+ 1 are
defined only within the space of the configurations I· . Even within this space
the operators of U 21 + 1 do not commute with the residual interaction between
electrons but in a very crude approximation they might be assumed to do so. If
one were to assume that, in the states of the two-particle configuration 12 , all
symmetric states (i.e. even Land S = 0) had the same energy and all
antisymmetric states (i.e. odd Land S = 1) had another common value for
their energy then it would follow that the Hamiltonian must be invariant under
U 2/ + i - This provides some physical justification for the group although in
practice the splitt ing of the energies for different L is large. The same remarks
apply to the nuclear structure problems mentioned above and in section 12.1.

18.11 Characters

In the next two sections of this chapter we return to the mathematical


properties of the group U N, calculating the characters of the irreducible
representations and describing the integration over the group volume. Little
practical use is made of the results but they are relevant to some of the proofs
we have omitted and they provide another link with our treatment of the group
!Jf3 in chapter 7.
The first step in deducing the character is to identify the classes of UN, since
the character of a representation is a class property, see section 4.9. G iven any
unitary matrix U it is possible to bring it to diagonal form W by a un itary
transformation V: VUV - 1 = W. Hence U is in the same class as the diagonal
matrix Wand we need only consider the characters for the diagonal matrices.
Since U is also unitary the diagonal matrix elements have the form exp( - i(}) so
that only n real parameters () l' ()2 ' . . . , 0. lying between 0 and 2n are needed to
define the diagonal unitary matrix. (For the group SU N there is the additional
constraint 0 1 + O2 + . . . + (}N = 0.) Hence , although U N is a group with n 2
parameters the character depends on only n of them . Th is feature was present
in the group £!Il 3 discussed in chapter 7, where the character was a function of
the rotation angle only and was independent of the other two parameters
defining the axis of rotation.
Consider now the representation U For the diagonal group element U the
2
).

matrix of U') in the basis described in section 18.4will also be diagonal and the
matrix element corresponding to a basis vector with weight (r 1 r 2 • • • r.) will be

(-i ,f
simply
exp rjo j )
J = 1

Thus the character of UU) is given by

x' 1 =
2
L
'1'2 , 'N
exp (- if. r OJ)
j = l
j
(18.28)
476 The Un itary Group UN 18 .12

where the sum runs over all the weights belonging to the representation IX, as
described in section 18.4. With the help of some rather elaborate algebra this
sum may be given in closed form, as the quotient of two determinants,

f."! + N - I
'I
£,I1 ,+N - 2
'I " e~~ I f,'~ - I e'i - 2
X(·) =
e1! + !\i - 1 e1,+ N-2 e1"'- l e'~ - 1 eN2 - -)
(18.29)

e "! +N - I
e l~ "'- fN - I
#N
' f..-

where e j = exp ( - ifl j ).


As a simple illustration of this general result let us take N = 2 and the
partition IX = [n] with only one row , The possible weights are simply (n, 0) ,
(n -I, I), , , " (0, n) so that from equation (18.28)

in)=exp(-in0tl+exp[-i(n-I)Ot-i02]+ . . . +exp(-infl 2 )
=exp[ -tin(O) +02)]sin{t(n+l)(0)-02)} /sint(O)-02)
a result which also follows from equation (18.29). On restricting to the group
SU 2' putting 0 1 + O2 = 0, this reduces to the familiar expression (7.42) with
n = 2j.

18.12 Group integration and orthogonality

From the general theory of chapter 4 the characters of irreducible repre-


sentations of U Nshould have orthogonality relations when integrated over the
group parameters. Since the character is a class fun ction it is necessary to
integrate only over the parameters OJ> which distinguish the classes. Thus from
equation (4.25) and the remarks in section 7.1 we expect a relation of the kind

(18.30)

where 0 is an abbreviation for the set 0 10 2 " , fiN and similarly°


dO:; dOl d0 2 • • • dON' the range of integration being from 0 to 2rr for each
angle OJ . The volume Vis defined by V = Jp(O)dOand it remains to determine
the appropriate weight factor p(O) using the results of appendix 4.3. Before
giving the derivation we quote the result , p (0) = 1 ~ 12 , where ~ is the
denominator in equation (18.29).
To obtain th is result we start from equation (A.4.15), the object being to
calculate the Jacobian (oel oa)\u = 0 ' where a and c refer to the N 2 parameters of
the unitary transformations and U(c) = U(a) U(b). For an arbitrary unitary
transformation U we take N of the N 2 parameters to be the angles OJ obtained
as in section 18.11 when U is diagonalised and the remaining N 2 - N
parameters to define the matrix which brings U to diagonal form. Without loss
18.12 The Unitary Group UN 477

of generality we may take U(b) to be diagonal (this defines the basis) with
matrix elements Uj j(b) = ej in the notation of the previous section. Since we are
only interested in the limit a -+ 0 we need consider only small a for which the
arbitrary U(a) differs from the identit y by a skew -Hermitian matrix

I-iall a l 2 - ia 21 . . . )
U(a) = -~12 + ia 21 1- ia 22
(

where the N 2 parameters aj j a re real and independent. By ordinary matrix


multiplication, Ute) is then given by

U«) ~ c~~::'~);d" )" ~7'~~::;;~" )


Since U(c) is nearly diagonal, it is easily written in the form U = v : : WY,
where W is d iagonal, and we find

Y (c) =

(1 - ia l l )el O. . . )
W(c) = 0 (1 - ia 22 )e2
(

ignoring second-order terms in the aij' The matrix Y is not uniquely


determined by the equation U = y -I Wy since it may be multiplied by any
diagonal matrix and we have made use of this freedom by taking the diagonal
matrix elements to be + 1. If we now take as the N 2 parameters for U(c) the
N angles 0; given by W(c), i.e. 0; = OJ + a., and the real and imaginary
parts of the off-diagonal matrix elements of y(c), namely

C12 = a 12 9l {(e l / e2 - I) - I } - a 21 J { (el / e2 - 1)- I }

C21 = a\2 + a 21 9l { (e\ / e2 - 1) - \ }


..I { (e l / e2 - I) - I }

then the Jacobian (oc/oa) is easily calculated. Since oO;/oaj k = bijbj k the first
478 The Unitary Group U
N
18 .13

N rows and columns of the determinant of ( Gcl Ga) corresponding to the 0;


and ajj is just the unit matrix. The remaining part of the determinant is a
product of !N(N - 1) factors of the type

GC12 I cia 12

I GC21I Ga 12
Thus, finally, the weight factor is given by

p(lJ) =( aG: )-1


a=O

N
= In
;<j
( f,j- CjW = Itll
2

since the product is the same as the determinant In the denominator of


equation (18.29).
The orthogonality relation (18.30) may now be verified for the characters x"
given in equation (18.29) and this also leads to proofs of irreducibility and
completeness using a technique similar to that in section 4.12. Notice that the
weight factor p(lJ) exactly cancels with the denominators of x* and X in the
integrand of (18.30).

18.13 The groupsSU 2 and ~3

We now consider in more detail the special case of SU 2 which was first
introduced in section 10.1 to describe isospin. In that section we argued briefly
that SU 2 had three infinitesimal operators with the same commutation
relations as those of :!It 3 and hence we were able to take over all the
representation theory developed in chapter 7 for :!It 3 ' Here we describe more
carefully the relation between the parameters of the SU 2 matrices and the
rotation parameters. The fact that this relation is two-to-one rather than one-
to-one, a homomorphism rather than an isomorphism, throws light on the
double-valued representations of :!It3 ' The use of SU 2 also suggests a new
(Cayley - Klein) set of parameters to describe rotations and these are most
convenient in deducing the parameters of a product of two rotations.

18.13.1 The parameters of SU 2


Let us denote an orthonormal pair of basis vectors in a two-dimensional
space by e I and e2, although we stress that they do not refer to the x- and y-axes
in physical space. A typical group element of U 2 is thus given by a 2 x 2 unitary
matrix in this basis and it is soon verified that the group postulates are satisfied.
For example, the product of two unitary matrices is itself unitary. An arbitrary
2 x 2 matrix with complex elements contains 8 parameters but the unitary
condition imposes four real equations to be satisfied leaving 4 parameters to
18.13 .1 The Unitary Group UN 479

describe the general group element of U 2 ' It is convenient to write the general
element as

(18.31)

with the parameters satisfying IX/)- py = 1. From the unitary condition


I I
U U t = I it is clear that det U = I and the form (18.31) exhibits U as the
product of a matrix with determinant = + I and a phase factor. From the
unitary condition we have the relations IXIX* + PP* = I, yy* + /)/)* = 1,
exy* + P/)* = 0, YIX * + /)P* = 0, the last two of which may be regarded either as a
single complex equation or two real equat ions . Taking these equations
together with the condition exl) - py = 1 leads to the conclusion I) = ex*,
y = - P* so that

°
with the one remaining condition IXIX* + PP* = 1.
The set of matrices U with ¢ = and hence with determinant = + 1 from a
subgroup of U 2 which is called the ' unimod ular' or 'specia l unitary' group and
is denoted by SU 2 ' In fact we may write U 2 = U 1 X SU 2 as the product of SU 2
with the Abelian, one-parameter group U 1 which multiplies by the phase
factor exp(i¢) and is seen to be isomorphic with :fl 2 . It will therefore be
sufficient to study the group SU 2 which has only three parameters. Henceforth
we shall consider only SU 2 with elements

(18.32)

and the condition IXIX* + PP* = 1. We now introduce three convenient real
parameters in place of the two complex numbers IX, Pwith one real condition.
The notation that we use may seem rather strange but it is chosen to simplify
the discussion in subsection 18.13.2 of the connection between SU 2 and :fl 3 •
Without loss of generality we may write

ex = cos ta - ik, sin ta, (18.33)

°
with k = (k" k y, k z ) any unit vector and ~ a ~ 2n, see problem 18.8. The
three real parameters are a and the two angles required to fix the unit vector k .
As a further simplification we define ax = ak; ay = ak; and az = ak, and then
the matrix U is determined by the three real parameters aq which lie in the
range L: a: ~ 4n 2
, i.e. the vector a = (ax, ay, az ) lies inside a sphere of
q
radius 2n.
480 The Unitary Group UN 18.13.2

18.13.2 Infinitesimal operators and irreducible


representations of SU 2

To find the infinitesimal operators corresponding to the three parameters aq


we must study U close to the un it operator, in other words for small a. Then we
have IX = I - tiak z = I- tia z and {3 = - t ay- t iax giving

so that the three infinite simal operators are simply - iSq in the notation of
equation (8.15) where Sq are the familiar spin matrices for s = t.
We therefore conclude th at the commutation relations of the infinitesimal
operators of SU 2 are the same as those for the group 9f 3 since the spin matrices
are infinitesimal operators of 9f 3 in the representation D(!). Because the
irreducible representations of 9f 3 were deduced in section 7.4 from these
commutation relations, it follows that the irreducible representations of SU 2
will be the same as those of 9f 3' However we shall see in subsectio n 18.13.3 that
whereas the half-integer representations of 9f 3 were double-valued they will be
single-valued representations of S U 2 ' Indeed, the representation D (!) is
identica l with the group elements themselves, the 2 x 2 matrices (18.32).

18.13.3 Connection between the groups 9f 3 and SU 2

The fact that two groups like 9f 3 and SU 2 have the same commutation
relat ions for their infinitesimal operators, the same set ofstruc ture constants in
the language of section 7.2, doe s not mean that the two groups are necessarily
isomorphic (one-to-one co rrespo ndence between group elements). To study
this point it is necessar y to go beyond the infinitesimal operations to the group
elements for finite va lues of the par ameters. Consider first the group 9f 3 • For
rotations about th e z-axis we have seen in section 7.3 and equation (7.6) that the
finite rot ation is related to th e infinite simal operator by a n exponential
Rz(a) = exp(aX z )' In thi s relation there is nothing specia l about the z-axis so
that for a rotation RI< (a) about an arbitrary axis k we may also write

R,.(a) = ex p( ~aqXq) = exp(ak .X) (18.34)

where in the notation of equation (7.22), aq = ak, with kq the component of k


and we have introduced the vecto r operator X = (X" X y , X.) . In making this
appa rent jump from the group 9f 2 to 9f 3 we are saying only that an arbitrary
rotation is a member of the one-parameter subgroup 9f 2 of rotations about the
axis k and is hence generated from the appropriate small change Iq aqXq in the
same way that the rotation about the z-axis was generated from aX, in
section 7.2.
18.13.3 The Un itary Group UN 481

The operator expression (18.34) may be applied in any representation and


for the two-dimensional representation O!t) it gives a very simple explicit way
of deducing the matrices Oll ) (a." ay, az ) for a finite rotation from the matrices
(8.15) already deduced for the infinitesimal operators. Remembering the
difference X, = - iJ q between the infinitesimal operators X, and the angular
momentum operators Jq , equation (8.15) gives

X.. =
1.(01 01)'
-I' X, =
1(01
I

from which we see that X; = - t and X, X, + X, X, = 0 for p "# q. Thus the


scalar product in the exponent of equation (18.34) has the property

(k .X)2 = L kqXqkpX p = -tLk; = -t


q.p q
using the normalisation of the vector k. Thus the representation matrix for
j= t is
O (! ) (k, a) = exp(ak.X)

= { 1- t (a2j4 )+ -h(a4 j I6)+ . .. } + (k .X){a -!(a 3 j4)+ .. .}


= cos ta + 2(k.X) sin t a (18.36)
and using the matrices (18.35) for the X, this becomes
Il) k ) = 1 -,'k z Sin
' Ia 'kx )Sin
' Ia I )
l
cos Ia (-k y -,
D ( ,a . 1 .l. .l. (18.37)
( (k y- ik.,) Sin Ia cos 2a + ikzsin 2a

This 2 x 2 matrix is identical with (18.32}-if we identify the rotation


parameters k and a with the SU 2 parameters k and a introduced in equation
(18.33). The correspondence is not one-to-one however. To produce all unitary
matrices the unit vector k may point in any direction and the angle ta must
cover the interval 0 ~ ta ~ n (Notice that it is not necessary to allow
ta to lie between nand 2n, i.e. for a to lie between 2n and 4n, since this merely
changes the sign of sin ta which is the same as a change in sign of the vector k .)
However to produce all rotations the angle a need only cover the interval
o ~ a ~ n since Rk(n + a) = R - k(n - a). Thus two unitary transformations
with their a-values differing by n will correspond to the same rotation. The
groups ~ 3 and SU 2 are said to be homomorphic. The correspondence may be
made one-to-one only if the group 9t 3 is extended, abstractly, to include
rotations through angles from 0 to 4n. The name 'universal covering group' is
used to describe such an extension.
We note that the double group which was introduced in section 7.6 with
482 The Un itary Grop U N 18.13.4

elemen ts R,,(a ) a nd E R,,(a) whe re 0 ~ a ~ n is isom orphic with the gro u p SU 2


through th e corres po nde nce
Rd a) <-> D (! )(k , a)
ER _,,(a ) <-> DI11(k, 2n - a)

The matrix ( 18.37) illustrates clearly that th e hal f-int eger represent ations a re
not co nti nuo us represent at ion s of fA 3 . As one rot at es thro ugh an angle a = 2n
the physical system returns to its sta rting po int but th e mat rix ( 18.37) is
cha nged by a facto r - I. In th e cove ring gro up as a increases to 4n the matrix
( 18.37) does return to its o rig inal value and co nti nuity is resto red. Thus th e
half-integer repr esent at ion s a re co ntin uo us as representa tio ns of SU 2.

18.13.4 Explicit formula for the parameters of


a product of rotations

cos tc = cos ta cos tb - kl sin ta sin t b (18. 38)

(sin tc) m = (sin t a cos t b)k + (sin tb cos t a)1+ (sin t a sin t b)k /\ I
T he fo ur real num bers cos ta, k, sin t a, k; sin t a and k, sin t a a re sometimes
ca lled the Cay ley - Klein par am eters of th e rot a tion R,,(a). T he formulae
( 18.38) ar e the eq uivalent for the group :Jl3 of th e relation c = a + b fo r
co mbining ro tations in 3f 2 .

Th e for m (18.36) may be used to ded uce the a ngle and axis of th e pr oduct of
two rotat ion s. Su ppo se Rm(c ) = R 1 ( a )R, ( b ), then from equa tion (18.36) we
ha ve
D l!)(m, c) = (cos ta +2k . Xsinta)(costb+2 I.Xsintb)
= (cos t a co s tb - k .l sin ta sin tb)
+ 2 {(sin ta cos t b)k + (sin t b cos ta) 1+ ( sin ta sin t b)k /\ I }. X
Eq ua ting thi s to (cos tc + 2m .X sin tc) we deduce the angle c and ax is m from
the two equations

18.13.5 Examples of SU2 basis vectors

Let us introd uce th e coord ina tes ~, '1 of an a rbitra ry vecto r e in the two-
d imen sional space of th e group $U 2 so that fro m eq ua tio n (3.38) th e
tra nsforma tion of functio ns is given by th e eq ua tion

T (U )ljJ( ~ , '1 ) = ljJ( ~, ij)


-
with ~ = rx* ~ - /3'1

Ii = /3 * ~ + rx'l (18.39)
The Unitary Group UN 483

from equations (3.39) and (18.32). Consider the transformation of the (2j + 1)
functions defined by
(18.40)
Since (18.39) is a linear transformation it is clear that this set of functions wiIl
form an invariant space and thus provide a representation of SU 2 with
dimension (2j + 1). For the particular unitary transformation corresponding to
a rotation of angle a about the z-axis, the parameters a and f3 are given by
a = exp( - ! ia)and f3 = O. In this case the tran sformation (18.39) is very simple
and the character of the representation provided by the set offunct ions (18.40)
is given by

(a*)2j + (a*)2j - la + .. . + a 2i = exp(iaj) + exp[ iaU - 1)] + . . . + exp( - iaj)

Comparison with section 7.4.3 shows thi s to be just the character of the
irreducible representation O(jl of 9i 3 ' Hence we conclude that the set (18.40)
provides a basis for the representation O (j) of S U 2 ' In particular, the three
functions e,
~ 17 and r]2 provide a basis for the vector representation 0(11 .

Bibliography

More details of the group U N may be found in Boerner (1963), see


bibliography to chapter 4, or Littlewood (1950), see chapter 17 bibliog-
raphy. The spectroscopic applications introduced in section 18.10 are
discussed further by Judd (1963), see chapter 8 bibliography, in atomic
structure, while the nuclear structure appl ications are reviewed by

Hecht, K. T. (1973). Symm etry in Nucl ei: Annual reviews of nuclear science
(Annual Reviews Inc., Palo Alto , California)

Much of the early work on the use of the unitary groups and their
subgroups in spectroscopy was done by Racah and some of this is described
In

Racah, G . (1965). Group theory and spectroscopy, Ergebn. exakt. N alUrw.,


37,28

Problems

18.1 Construct product functions belonging to the representations [3], [21] and
[III] of U 3 , as suggested in example (4) of section 18.2.
18.2 Use the method of section 18.4 to label the basis vectors of the following
representations: (a) [3] of U J • (b) [21] of U 4 .
18.3 Reduce the direct product [21] x [2] of U J . How does this result simplify if,
484 The Unitary Group UN

instead of U 3, one is using (a) the group SU 3 (use the notation (i. ,~)), '(b) the
group SU 2?
18.4 Deduce the weights of the basis vectors in part (a) of problem 2 and hence show
that the reduction (18.24) for this case is U[3] = 0(3) -+ 0(1).
18.5 Find the L-values for the totally antisymmetric states of two and three d-
particles. (There is one antisymmetric state for each combination of different m-
values of the particles. Hence list the number of independent states for each value
of M = L m and so deduce the possible values of L.)
18.6 Reduce the products [1] x [1], [1] x [11] and [1] x [2] for the group U 5 '
Hence, knowing the reduction of product representations (7.44) of ,Jf3 and the
results of the previous problem, deduce by subtraction the L -values of the [2],
[3] and [21] representations of two and three d-particles.
18,7 Follow the method of problems 5 and 6 to list the possible combinations of J and
T for three nucleons in a j = ~ shel l.
18.8 Show that there is a one-to-one correspondence between the parameters k and a
of equation (18.33), with 0 < a < 2rr and k a unit vector, and the two complex
parameters !x, # with the condition 'l'l* + P#* = I.
19
Two Familiar 'Accidental'
Degeneracies-the Oscillator
and Coulomb Potentials

The appearance of degeneracies in the spectrum of energy levels as a


co nseq uence of the existence of a gr oup of symmetry operat ions on the
Ham iltonian of the system, has been centra l to our discussion s of symmetry in
qu antum mechanics. However, as we explained in sect ion 5.3, the full
degeneracies will be understood only if the full symmetry group has been
identified . A degeneracy may alwa ys happen by cha nce or by de sign . For
exa mple a numerical parameter in the Hamiltonian may be adj usted until two
energy levels cross and thi s is usually called an 'accidenta l' degeneracy. If
however one finds what appear to be systematic accide nta l degeneracies, not
ju st in a single level but in man y levels, then it is usually a sign that the
degeneracies are not accidental, but ar e due to the existence of so me hitherto
unsuspected symmetry operati on s which enlarge the symmetry gro up. In such
a situa tion one sho uld strictly not use the wo rd 'accidenta l'. In th is cha pter we
describe two famou s exa mples (see Jau ch and Hill, 1940, in th e bibliography)
of such accidental degen eracies which a re not strictly accidenta l but are due to
a larger symmetry group. Both exampl es relate to the mot ion of a single
485
486 'Accidental' Degeneracies 19 .1

particle in a spherically symmetric potential so that one expects the usual


(21 + 1)-fold degeneracies due to rotational symmetry as described in sections
7.5 and 8.1. For general forms of the radial dependence of the potential V(r)
this is all the degeneracy one finds , but for two particular cases V(r)ocr 2 and
V(r)ocr- 1 , the detailed solution of the Schrodinger equation is found to lead
to additional degeneracies of a systematic kind . These two cases are well
known as the harmonic oscillator and the Coulomb potential and they both
have important roles to play in physics. The oscillator is of general importance
in describing small departures from an equilibrium position and is also of
importance as a first approximation to the motion of individual nucleons in a
nucleus, the origin being at the centre of mass of the nucleus. The Coulomb
potential describes the motion of the electron in the hydrogen atom. We shall
show that the symmetry groups in the two cases are U 3 and rJi 4 , respectively,
both of which contain the obvious rotation group rJi 3 as a subgroup. As a
straightforward generalisation, we show that the harmonic oscillator in n-
dimensions has symmetry group Un'

19.1 The three-dimensional harmonic oscillator for


one particle

It is conventional to write the oscillator potential as V(r) = !M w 2 r 2 , where


w/27C is the classical frequency of oscillation for a particle of mass M. In
quantum mechanics we must solve the Schrodinger equation
(_h 2 V 2 /2M +-tMw 2r 2 ) ljJ(r,O,cP) = EIjJ(r,O,cP)

For convenience, let us measure energies in un its of hco and distances in units
of (h/ M w)i which we call the oscillator 'length parameter' and denote by b. In
these units, the Schrodinger equation becomes-a - V 2 + r 2)1jJ = EIjJ and the
Hamiltonian is
(19.1)

This Hamiltonian is clearly spherically symmetric, a fact which is expressed


through the three commutation relations [H, Lq ] = 0, where q = x, y or z and
L, = -i(xe/oy- yO/ex), etc ., are the infinitesimal rotation operators, see
sections 5.6 and 8.2. We shall presently construct a total of nine operators
which all commute with H and which are then shown to be the infinitesimal
operators of a group U 3' It is convenient to begin by defining three operators
ax = (x+%x) / j2, a}. = (y+ %y)/ j2, a, = (z+ 0/oz) / j2
and their Hermitian adjoints
a; = (x -0/ox)/j2, a; = (y -0/oy)/ j2, a; = (z -0/oz) / j2 (19.2)
(Recall that Px= - iii%x is Hermitian so that the Hermitian adjoint of%x is
19 .1 .Accidental' Degeneracies 487

- 0/ ox .) One immediately verifies the commutation relations


[aq ,aq,] = [a:,a:,] = 0, [aq ,a:,] = Oq.q' (19.3)
t • t
[aq ,H] = aq , [aq,H] = -a q (19.4)
using the explicit form (19.1) for H. The Hamiltonian may also be written

H = L (a:aq +1) (19.5)


q

This algebraic technique is often used in quantum mechanics textbooks to


solve the Schrodinger equation for the oscillator because from (19.4) we may
conclude that if 1/1 is an eigenfunction with energy E then the functions aql/l are
also eigenfunctions of H with energy (E -1), In other words, aq reduces the
energy by one unit and similarly a: increases the energy by one unit. The wave
function 1/1 0 of the lowest energy level must satisfy the conditions aq1/1 0 = 0 so
that from (19.5) the lowest eigenvalue has the value 1. Using the explicit
differential form for the aq these conditions pose simple different ial equations
which lead to the explicit solution 1/10 = exp( -1r 2), apart from normalisation,
The excited states then appear at integer intervals and the wave functions
corresponding to an excitation N are given by the set of functions
t N t N t
(ax) .(a)') ,(a:) N'l/Io (19.6)
where N " N)' and N z are any integers satisfying N,+N )'+N: = N . In
particular, for N = 1 there is a three-fold degeneracy, for N = 2 a six-fold
degeneracy and in general a degeneracy of order 1( N + 1) (N + 2), The possible
values of N x ' N; and N z are shown in table 19.1 for smaIl N . One sees
immediately that these degeneracies are not of the kind (21 + 1) which were
Table 19.1

N 2 3
N, 100 20 0 1 10 3 002 2 1 100 1
N). 010 020 1 0 1 030 1 0 202 1 1
N, 001 0020 1 1 003 0 1 021 2 1
2,0 3,1

expected. They must therefore be the sum of several (21 + 1)-fold degeneracies
and the appropriate values of 1are shown in the last row of table 19.1. To see
how these I-values are deduced we notice that the three operators aJ transform
like the vector representation 0(1) of the group 91 3, Thus, since 1/10 is
sphericaIly symmetric, the excited states (19.6) with fixed N will transform like
the symmetrised product 0(1) x O(l) X • • • X O(l) with N factors, as defined in
appendix 3.1. Only the symmetrised product may occur because the a:
commute with each other. There is no scope for any antisymmetry with only a
single particle. The formulae of appendix 3,1 for the character of symmetrised
488 'Accidental' Degeneracies 19 .1

products together with the standard formula (7.42) for the gt3 characters give
the reduction of this product and hence the resulting values of I, as shown in
the last row of table 19.1. In fact one may deduce generally that I = N,
N - 2, . . ., 1 or 0 depending on whether N is odd or even, see problem 19.2.
Having established in this explicit way that the three-dimensional harmonic
oscillator has a greater degeneracy than expected on the basis of spherical
symmetry we now return to the problem of finding the larger symmetry group.
From the commutation relations (19.4) one may verify directly that , for any q
and q', [a; a q., H] = O. This result may al so be seen from the fact that, whereas
a q reduces the eigenvalue by one unit, a; increases it by the same amount so
that the product aqa q. cannot change the eigenvalue but merel y transforms
within the vector space of each set of degenerate eigenfunctions. Thus we have
found a set of nine operators a;a q. which all commute with the Hamiltonian H
and from these may be constructed a set of unitary symmetry operators

U = exp ( I . cqq.a;aq, )
a,«
provided that the coefficients C qq' satisfy the skew-Hermitian condition
c:q' = - cq' q' This condition leaves the freedom of nine real numbers in the
coefficients Cqq' and so we ha ve a nine-parameter set of unitary transformations
which commute with the Hamiltonian H. To confirm that this is in fact the
group U 3 we consider its effect on the set of three operators a;, finding

(as')' = U as' U - 1 = I(expC)s' sas~ (19.7)


s'

where C is the matrix Cqq" To obtain this result we use the operator identity
e A Be - A = B + [ A, B] + i! [A, [ A, B] ] + tr [ A, [ A, [A, B] ] ] +
together with the relation
t t
[aqaq.,a s] = oq'sa q
which follows from equation (19.3). The significance of the result (19.7) is that,
for skew-Hermitian C, it is the most general unitary transformation in the
three-dimensional space of the operators a; . Thus the operators U are in one-
to-one correspondence with the matrices exp C of the group U 3' In this sense,
the matrix of the operator a;a q. is given by (a;a q. ),s = Oq·sOq/.
The infinitesimal operators are the nine independent skew-Hermitian
combinations of the operators a;a q ' , corresponding to the nine skew-
Hermitian 3 x 3 matrices described in section 11.4. For some purposes it is
convenient to take linear combinations which are irreducible tensor operators
with respect to rotations, see subsection 7.4.6. Since both a; and a q transform
like a vector, i.e. the representation 0 (1 ), the set of nine products a;a q'
transforms like the product representation
O(1)x 0(1) = 0(0) -i- 0(1) -i- 0(2)
using the familiar rule (7.44) for its reduction. The invariant 0(0) is essentially
19.1 'Accidental' Degeneracies 489

the Hamiltonian, since from equation (19.5), ~ a qta q


Q
=H- ¥
The vector operators D( II are the angular momentum operators L, since
L z = -i(x iJ j(!y-y Dj Dx) = -i(at,a).-a:a.,),etc.

The remaining five independent combinations may now be written as a


tensor operator of rank 2 (quadrupole), like a spherical harmonic y~21. In the
conventional notation of subsection 7.4.6 we write

(19.8)

Since H has this larger symmetry group U 3 one expects each eigenfunction
to be labelled by an irr ed ucible representation of U 3 ' We first note that
U 3 = U 1 X SU 3 following the argument given in section 10.4. In the present
problem the infinitesimal operator for U 1 is simpl y the invariant
~ a:a q =H- ~ and the eight remaining traceless operators L, and Q~2)
q
describe SU 3 ' Thus the U 1 representation is labelled by N and it remains to
identify the SU 3 representation corresponding to each energy level. Recalling
the study of SU 3 operators in section 11.6 it is seen from the matrices (11.2)
that we may associate the operators !(a:a x+a:.a y -2a:a z ) with Yand
-!(a;a, - a}:a y) with T, so that we may relate the numbers N " Nya nd N z used
above to the numbers Yand M T used in chapter II by the equations

(19.9)

The diagrams in figure 11.6 then show immediately that the three eigenfun-
ctions with N = I belong to the representation (10), those with N = 2 belong
to (20) and those with N = 3 belong to (30). As an illustration, in figure 19.1 we
reproduce from figure 11.6 the diagram for D( 3 0) showing the values of
(N,N ).N z) corresponding to each lattice point. Because the products (19.6) are
symmetric in the N factors a: and thus belong to the U 3 representation [N],
see section 18.7, it follow s generally that the set of eigenfunctions with energy
( N + ~) belongs to the irreducible representat ion ( NO).
The question of which values of angular momentum I occur for given N is,
group theoretically, the question of restricting from the group SU 3 to its
subgroup rJt 3' We stress that this is a different process from the restriction
from SU 3 to SU 2 described in section 11.5. The infinitesimal operators of the
490 'Accidental' Degeneracies 19.1

(030) (120) (210) (300)


x x x x

x\
\'020' (I~I)
120,,1
/X- M
r = '12 (N, -Ny)

\/
(012) (102)

X
(003)

Figure 19.1

physical rotations flt 3 are the operators L q , see equation (.19.7), while the
operators for the SU 2 subgroup in the xy-plane would be

(19.10)

It is clear that, for example, T + is a 'raising operator' which reduces N ; by one


unit and increases N. by one unit, thus moving one step to the right in figures
like figure 19.1. One soon verifies, using equation (19.3), that the three
operators (19.10) satis fy the standard commutation relations (7.30) of SU 2 '
The operators defined in equation (19.10) do not, of course, have anything to
do with isospin but we use the same notation to emphasise the mathematical
connection with chapter 11. The a ngular momentum L, may now be written
simply as
L z = -i(a;a).-a>.)= -i(T + -T_)
= -i(T.+ iT y-T. + iT)")
= 2T)" (19.11)
Hence the set of eigenvalues of L, within a representation (}.J1) of SU 3 is the
same as the set of eigenvalues of2T)", which is necessarily the same as the set of
eigenvalues of 2T z ' which is given by the set of values of 2M T = N.• -N y •
These may be deduced immediately from the figures like figure 19.1. Knowing
the eigenvalues of L, enables us finally to deduce the possible values of L For
example, figure 19.1 tells us that the representation (30) contains the following
eigenvalues for L z ,
- 3, -1, 1, 3, - 2, 0, 2, - 1, 1, 0
which implies that I = 3 and I = 1. In general one deduces that I = N, N - 2,
19.2 'Accidental' Degeneracies 491

N -4, . ..,1 orO for the representation (NO), which is the familiar result for the
I-values in an oscillator energy level.

19.2 The three-dimensional harmonic oscillator for


many particles
In the preceding section we considered the motion of a single particle in a
spherically symmetric oscillator potential and we found additional de-
generacies beyond those expected for an arbitrary spherically symmetric
potential. Consider now a system of particles moving in an oscillator potential
and subject to a perturbing interaction potential 1: V(r i j ) between particles.
i <]
This situation is similar to the system of electrons in an atom which we
described in section 8.6. There is was seen to lead to an LS coupling result,
although in our present discussion the spin need not be considered. From the
group theory point of view the total angular momentum L reflects the fact that
the wave function transforms like the representation O(L) of the group fJi3 of
simultaneous rotations of all particles and is additional to the labels I which
describe the behaviour of the single-particle wave functions under separate
rotations. The Hamiltonian, including the perturbation 1: V(r i j ) is invariant
i <]
under the group ::Jl 3 of simultaneous rotations. Let us now study this process
of extension from one to many particles for the group U 3' following the steps
taken for rYt 3 . Just as we define the total angular momentum operators
L, = 1: Lq(i) as sums over particles i, so we can define the nine total U 3
I

operators as 1: a;(i)aq.(i). It is clear that the commutators of these operators


I

are the same as those for the single-particle operators since operators for
different particles commute. Hence they describe the group U 3 of simul-
taneous U 3 transformations on all particles, which contains the group of
simultaneous rotations, described by the total angular momentum L, as a
subgroup. As with the single-particle operators, the nine U 3 operators may be
taken as H = 1: H(i), where H(i) is the oscillator Hamiltonian for particle i,
i
L, = 1: Lq(i) the total angular momentum and the five components of the
I

quadrupole operator Q~2) = 1: Q~2)(i) . From general theory it follows that if


I

the perturbing potential 1: V(r j j ) commutes with these nine operators then
i <i
the eigenfunctions of the system are labelled by irreducible representations of
U 3, in other words by N which determines the unperturbed oscillator energy
and a label (I.Jl) of SU 3 ' We shall not go into detail regarding the possible (I.Jl)
which appear with each N, except to remark that the procedure is similar to the
problem of deducing the possible values of L in an atomic configuration which
we described in subsection 8.6.4. In such a many-particle system, one no longer
has only the simple representations with Jl = 0 which arose for a single particle.
492 'Accidental' Degenerac ies 19 .3

A perturbation which is invariant under SV 3 could be constructed from the


Casimir operators which are, by defin ition, invariant. For the group of
simultaneous SV 3 transformations, the Casimir operator C 2 defined in section
11.10 will involve both one- and two-body parts. In terms of the operators L,
and Q~21 it may be written as

r. j i, i
(19.12)
Since the eigenvalues of C 2 are known, see equation (11.21), and the
eigenvalues of (L. L) are simply L(L + 1), it follows that the eigenvalues of the
quadrupole term (Q(2) .Q(2)) may be deduced. The two-body part of this is
sometimes called the 'quadrupole force' and has been used as a model for
understanding the development of deformations and rotational motion in
nuclei-see Harvey (1968) in the bibliography.
The interesting general aspect of this use of SV 3 is the way in which the
symmetry is extended from the single-particle to the many-body system .

19.3 The harmonic oscillator in n dimensions ,

In section 19.1 we chose to discuss the three-dimensional harmonic oscillator


because of its obvious physical significance in the real three-dimensional space
of the physical world . However from the algebraic point of view the fact that
the number of dimensions was three was not important. We could equally well
have started with operators v 2 =
q
r.= 1
021 or: and r
2
=
q
i:= r: defining an n-
1
dimensional oscillator. The extension from three to n dimensions is
straightforward leading to the conclusion that the symmetry group is V •.
In particular the two -dimensional oscillator has degeneracies corresponding
to irreducible representations of SV 2 which, as we have seen in section 18.13,
are the same as those of 9t 3 ' Hence the two-dimensional oscillator, with
energies (N + 1), has degeneracies corresponding to an angular momentum-
like label tN, the order of the degeneracy being 2(tN) + 1 = N + 1.The pair of
operators a.: and a; in this case transform like the representation oW of the
group SV 2 •

19.4 The symmetry group of the Coulomb potential

The Hamiltonian for a particle of charge - e in a Coulomb potential elr, like


the electron in the hydrogen atom, is given by (see section 8.5)
H = p2 /2M -e 2 lr = -1i 2 V 2 / 2M -e 2 lr (19.13)
19.4 'Accidental' Degeneracies 493

Ifwe measure length in units of 1i 2 / Me 2 (the Bohr radius) and energy in units
of e4 / 21i 2 (= Rh, where R is the Rydberg constant), then the Hamiltonian
takes the form
H= -(V 2+2 /r) (19.14)
and the Schrodinger equation is given as usual by H ljJ = EljJ , with E the energy
and ljJ (r) the wave function . It is well known that the energy levels are labelled
by an integer n = I, 2, 3, .. ., with energy - n - 2 and th at for given n the
angular momentum takes the values 1=0, I, 2, . . ., (n - I). Thus there is a
greater degeneracy than one would expect from the obvious spherical
symmetry.
Because it is spherically symmetric, the Hamiltonian will again commute
with the three infinitesimal operators L, of ;Ji 3 ' Although it is by no means
obvious, some detailed but straightforward algebra shows that H also
commutes with the three components of the vector operator
A = t:(V f\ (r f\ V))-((r f\ V) f\ V): + r /r (19.15)

In terms of physical operators this may be written


A = - [ {(p f\ L) -(L r; p) }/2M -e 2 r / r ] / e2
where the quantity in the square bracket has been called the Runge-Lentz
vector and is known in the classical Coulomb problem to be a constant of the
motion. The existence of these three further Hermitian symmetry operators A q
suggests that, together with the L q , they may be the infinitesimal operators of
some continuous group which contains fYt 3 . We therefore study the com-
mutation relations between the six operators A q and L, to see firstly if they
satisfy relations of the general type (7.7) and secondly if the structure constants
are those of a well-known Lie group.
Since the three operators Aq form a vector their commutation relations with
the rotat ion operators L, are simply
(19.16)
see equation (7.26) and subsection 7.4.6. From the definition (19.15) the
commutator of A q with A q . may be calculated, giving ultimately
[A"A y] = (V 2+2 /r)(x iJ /iJy - yiJ /iJ x ) = -iHLz,etc. (19.17)
Thus the commutator is not given precisely in the form (7.7) as a sum of the
symmetry operators L, and A q . but involves also the Hamiltonian H.
However, since H commutes with A and L we may incorpo rate the factor - H
into the operators by defining

so that
(19.17a)
494 .Accidental' Degeneracies 19 .4.1

The use of - H rather than H in the defin ition of A' will be convenient in
considering bound states, for which the eigenvalues of H are negative. In thi s
case A' is Hermitian but it become s skew-Hermitian for positi ve energies. Thus
for negati ve energie s the operators L and A' define unitar y transformations
ex p { -i(a . L ) -i (c. A') } where a and c are real vector parameters. The six
infinitesimal operators are - iL a nd - iA' . The minus sign here ensures that
the rotation parameters a are the same as those defined in sub section 7.4.1. We
sha ll now identify the commutation relations (19.16) and (19.17a) as those for
the group 9i 4 •

19.4.1 The groups 9i 4 and 5e

The rotation group 9i 4 in four dimensions is defined as the set of (real)


orthogonal 4 x 4 matrices with determinant of + 1. If R denotes such a matrix
we may write R = 1 + M where the matrix elements of M are small if R is close
to the identity. The orthogonality condition Rt R = 1 implie s that, for small M,
M + M = 0 so that the infinitesimal operator M is skew-symmetric and may
t

be written

(19.18)

Th is defines the six inde pendent infinitesimal operators Xq , Zq of the group


where the notation X, follows that used in subsectio n 7.4.1 for the group fJt3 ,
exhibiting the subgroup 9i 3 of 9i 4 obtained by putting cq = O. The com-
mutation relat ions follow fro m the definitions in equ ation (19.18),
[X ." X y] = X, ; etc .
[X"Zx] = 0, [X ." Z y] = Zz, [X"Zz] = - Z)', etc. (19.19)
[Z" z .] = X" etc.
where the others follow from cyclic rearrangement of indices. Comparison
with equations (19.16) and (l9.17a) shows that X and Z have the same
commutation relations as the infinitesima l symmetry operators - iL and
-iA'.
Thus, so far as infinitesimals are concerned, we can identify the symmetry
group of the bound states with 9i 4 . A more careful stud y of the range of the
parameters would be necessary to establish that the finite symmetry operations
could be put into one -to-one correspondence (isomorphism) with fJi4 . (In fact,
the 9i 4 invariance of H may be shown explicitly by mapping the momentum p
on to a point which lies on a unit sphere in a four-dimensional space and ha s
coordinates
2), IJ = 2Popy/(pfJ + p 2)
~ = 2Pop., /(PfJ + p
( = 2Popz/(p fJ + p2), X = (p fJ - p2)/(p fJ + p2)
19.4.2 .Accidental' Degeneracies 495

where Po = ( - E) - !. If the Schrodinger equation is written as an integral


equation in momentum space it exhibits an invariance with respect to rotation
in this four-dimensional space.)
For positive energy states it is more convenient to define A" = (H) - !A
instead of A ', so that A " is Hermitian with the commutation relation (19.17a),
replaced by
(19.17b)
The commutation relations of the infinitesimal symmetry operators - iL and
-iA" now differ from those given in equation (19.19) for the group 9t 4 .
However, they now agree with the commutation relations (15.32) for the
Lorentz group 5f'. This result is to be expected since the Lorentz group is a
four-dimensional rotation group with a sign change in one component of the
metric and this corresponds to the change in the sign of H.

19.4.2 The classification of states of the Coulomb potential

Having found the symmetry group it now follows from the general theory
that the energy levels may be classified by irreducible representations. We
therefore face the task of finding the po ssible irreducible representations of the
group 9t 4 but, fortunately, thi s may be done very quickly by making use of a
(local) isomorphism between the groups ~ 4 and ~ 3 X Jf 3' To understand th is
we define new infinitesimal operators
F = t(X +Z ) and G = t (X -Z)
which is equ ivalent to the use of (a + c ) and (a - c ) as parameters in equation
(19.18). Then, from equation (19.19),

(19.20)
with all components of F commuting with all components of G . Thus the
operators F and G separately sa tisfy the sa me commutat ion relat ions as X
and therefore de scribe ~ 3 groups. Since the y commute with each other it
follows that together they describe the product group iJt3 X 9f 3 and hence we
have established an isomorphism for the infinitesimals of 9f 4 and 9f 3 X 9i 3 . We
may therefore use the known irreducible representations O UI of 9i 3 withj = 0,
t .. .,
t , I, to conclude th at the irreducible representations of 9f'4 are labelled
by a pair j Ij 2 of such indice s (see sectio n 4.21 for representations of a product
group). The irreducible represent ations of 9i 4 are therefore denoted by O U,j , ),
where j I and i, refer to the two &3 groups. The o perato rs (F. F) and (G. G)
provide two Casimir operator s with va lues - jI Ut +l) and -j2U2+1 ),
respectively (see section 7.5), and the dimension of O U,j , ) is (2j I + I) (2j2 + 1).
In studying the osc illat or in section 19.1 we saw that only some of the
representations of SU 3 co uld be realised with only a single particle and a
similar situation occurs with the Coulomb problem. The restriction in this case
496 'A ccidental' Degeneracies

comes about because (A' . L) = (L . A') = 0, which follows directly from the
definitions of L and A', using some detailed algebra. Thus if we construct the
symmetry operators -1 i( L ± A' ) corresponding to F and G it is clear that
both Casimir operators reduce to the sa me expression - !{(L . L) + (A' . A') }.
Hence the only possible representations are those with i, = l z namely' O W)
t
withj = 0, I, ~ ; . ... Som e more algebra shows that the Ca sim ir operators
simplify further, giving
(F. F ) = (G . G ) = - H(L. L ) + (A ' . A' )} = !(l + H - 1)
In a representation O Uil thi s Casimir operator has the value - jU + I) so that
the corresponding value of the energy is given by
E= <H>= -(2j+ 1) - 2 (19.21)
with a degeneracy of (2j + 1)2. In the standard theory of the hydrogen atom it
is of course conventional to use an integer n = (2j + 1) to label the energies so
thatj = 0, t 1, t ..., corresponds to n = 1,2,3,4, . .., and the energy formula
(19.21) reduces to the more famil iar expre ssion E = - n - 2 in th e energy units
introd uced at the beginning of sectio n 19.4.
To deduce which angular momenta I occur at each energy involves the
reduction
OW) = L c/O(/) (19.22)
I

of the irreducible represent at ion O W ) of fJi4 follow ing the restriction to the
subgroup yt 3 of physical rot ati on s. Notice that it is the opera to r X = F + G
which describes the se rotat ion s. The problem of deducing the non-zero
coefficients in the series (19.22) is now exactl y the same as the vector coupling
of two angular momenta I = j + j which was described in subsection 7.4.4.
Thus from equation (7.44) we conclude that I = 0, 1, 2, . .., 2j( = n - I), which
is the familiar result for the hydrogen at om .
For the positive energy states we simply remark that the Lorentz group has
no finite-dimensional uriitary irreducible representations, see subsectio n
15.2.5, and this is con sistent with the well-known continuum of states which
occurs for positive energy.
The technique described in section 19.2 for extending the use of the group
U 3 to a system of particles may also be applied to a system of particles moving
in a Coulomb potential with an 91 4 symmetric interaction. However, this does
not seem to have much relevan ce to any recognised phy sical problem.

Bibliography

The symmetries of the oscillator and Coulomb potentials for a single


particle were discussed by

Jauch, 1. M. and Hill, E. L. ( 1940). Phys. Re v., 57,641


·Accidental' Degeneracies 497

The exploitation of the oscillator symmetry in the nuclear many-body


problem is reviewed by

Harvey, M . (1968). Advan ces in Nuclear Physics 1. (Plenum, New York)

More recent work on the Coulomb symmetries is described by

Englefield, M . J. (1972). Group Theory and the Coulomb Problem (Wiley-


Interscience, New York)

Problems

19.1 Show that the co nditio ns aqt/lo= 0, with a q defined by equation (19.1), lead to the
solution t/lo = A exp( - 1r2 ).
19.2 Justify the result quoted in section 19.1. that on restriction to the subgroup 9i 3 '
the totally symmetric representation [N] of SU 3 reduces into a sum of
representations Dill , where I= N , N - 2, N - 4, . . .. 1 or O. (Follow the
technique of problem 18.5 by taking symmetric product s of N factors, each of
which has rn = L 0 or - 1. Then show th at there is only one product with
Lorn = N or N - 1, that there are two products with Lorn = N - 2, a nd N - 3.
three with N - 4 and N - 5, etc .)
20
A Miscellany

Having described systematically the theory ofsymmetry and its applications to


a wide variety of physical systems we find that there are a number of related
topics which, usually because they would have drawn attention away from the
main argument, have so far been omitted. These topics, most of which are not
connected with each other, are collected together here. Each section of this
chapter is self-contained and the order of sections is arbitrary. The first two
sections have immediate physical relevance, while the next two are concerned
with general mathematical questions and the last with a point of mathematical
detail.

20.1 Non-invariance groups

Hitherto we have been concerned exclusively with symmetry groups f§, which
by definition leave the Hamiltonian invariant. We now investigate some
groups which do not have this property but which nevertheless are of some
interest. They are often called non-invariance or dynamical groups. Their
significance is that they bring together into a single representation eigenfun-
ctions corresponding to different energies, so that the spectrum and transition
probabilities may be expressed in terms of the group operators. Hence these
498
20 .1 A Miscellany 499

d ynamical quantities may be deduced from the group theory. The application
of these ideas ha s so far been limited to simple systems.
As an illustration, we consider the one-d imensional harmonic oscillator. In
the notation of section 19.1, th e Hamilton ian may be wr itten

(20.1)

a nd the energies are well known to be given by (n + i) in the usual units of hco,
where n = 0, 1,2, . . . . We now construct t he three Hermitian operators

I 2 t2
KI = 4(a +a ), (20.2)

and from the commutation relations (19.3) we find that

It may now be seen th at these are just the commutation relations of a Lorentz
group z" J in three dimensions. Such a group ma y be constructed by removing
o ne of the space d imensions (r) from the usual Lorentz group discussed in
section 15.2. Its infinitesimal operators are then X, ; Y, and Y)" and their
co m mutatio n relations a re given in equations (15.32),

These agree with equat ion s (20.3 ) if we ass oc iate K l ' K 2 , K J with iY" iYy and
iX: . Again, the factors o f i a re necessary here because the infinitesima l
operators for the unitary transformat ions generated by the Hermitian
o pera to rs (20.2) mu st includ e a fact or i, see sectio n 7.2. Hence the non-
invariance group defined by th e operators (20.2) has the same commutation
relations as .!£ J and is therefore (locally) isomorphic with it. We shall not
pursue the question of the global transformations, i.e. the question of the range
of the group parameters. The infin itesimal properties are sufficient to
determine the representations if one is not concerned with questions of many-
valued ness.
We now study the unitary irreducible representations of Y J following the
procedure used in subsection 7.4.2 for 9t J . First choose a basis in which the
operator K J , and hence the energy, is diagonal. We write K J cPn = !-(n + !-)cPn,
where for convenience the eigen values are expressed in the familiar form of the
oscillator energies. There is no loss of generality here since, in this general
discussion of the representations of 2' J , the number n is, as yet, quite arbitrary
and is not restricted to be an integer. The remaining group operators K 1 and
K 2 then couple eigenfunctions with different energy. In fact, if we define
K ± = K. ±iK 2 then
(20.5)
500 A Miscellany 20 .1

showing that K + are raising and lowering operators in the energy. In detail, if
HcPn = (n +!)cP: then
H(K ±cPn) = 2K 3K±cPn = 2K±(K 3 ±1)cPn = (n+t±2)(K±cPn) (20.6)

showing that the operators K + raise and lower the energy by two units of hs».
The Casimir operator for z' 3~ corresponding to J2 for the group Yf3 , is no-w
given by

with a significant change of sign in the third component.


If we now follow the steps used in subsection 7.4.2 to construct the finite
dimensional unitary irreducible representations of :lI3 we are led to a
contradiction because the expression for the normalisation constants An
defined by the equation
(20.8)

takes the form IA n l 2 < O. Thus we conclude that there are no finite dimensional
unitary irreducible representations of !f:' 3 except for the trivial identity
representation. (The same conclusion was reached in subsection 15.2.4. for the
group !e.) We therefore explore the possibility of infinite dimensional
representations. This amounts to releasing the constraints that the possible
values of n in the representation are bounded above and below.
Let us first release only the upper bound so that we assume the existence of a
least n = no which implies that
(20.9)

The value of the Casimir operator in this representation is found by writing


equation(20.7)asC = K . K _ +K 3 -K5and then operating on cP togivethe
n"
value
c = /6(1 + 2no)(3 - 2no) (20.10)

This value is then valid for all basis vectors cPn in the representation. Starting
with cPn o we now use the raising operator K + to generate the set of cPn, where
n = no + 2t with t an integer. (Remember from equation (20.5) that K +
increases the value of K 3 by I and hence the value of n by 2.) To ensure that the
representation is unitary, the infinitesimal operators K l' K 2 and K 3 must have
Hermitian matrices. This implies that n, and therefore no, must be real. It also
implies that K : = K _ and so from equation (20.8)
IA l- 2 = (cPn,K_K + cPn) =
n (cPn ,(C +K5 +K 3)cPn)
= 116{O +2no)(3-2no)+(1 +2n)(5+2n)}
= t(n + no + l)(n - no + 2) (20.11)
20 .1 A Miscellany 501

To avoid contradiction, this expression must be ~ 0 for n ~ no which leads to


the limitation no ~ -1, with the value no = - t producing the identity
representation. For no > -t and n > no the expression (20.10) will not vanish,
so that for each no we have an infinite dimensional unitary irreducible
representation T,no ). However, not all of these representations can be realised
in the oscillator problem because, if we use the explicit form (20.2) for the
group operators and make use of the commutation relations (19.3), the Casimir
operator (20.7) red uces to a constant, C = 3/16 . Putting this value into the
expression (20.10) leads to the two possibilities no = 0 or no = 1. The
eigenfunctions must therefore belong to T IO) with n = 0, 2,4, .. ., or to TO)
with n = 1, 3, 5, . ... This accounts for the entire spectrum of the one-
dimensional oscillator within two irreducible representations of the non-
invariance group z" 3 ' The process by which the value of the Casimir operator
limits the po ssible representations has been met before, in subsection
19.4.2.
The representations Tlno ) derived above are not the only unitary irreducible
representations of!f 3 ' In the first place one could retain an upper bound to the
possible values of K 3 in the representation. This would lead to a kind of mirror
image ofT(no) with values of K 3 extending down to - cc, It is also possible to
generate representations without either upper or lower bounds for K 3 .
However, none of these representations is relevant to the oscillator problem
since they would contain states with unbounded negative energy.
The method used here to generate the non-invariance group!£ 3 of the one-
dimensional oscillator may be applied to the more complica ted three-
dimensional examples of the oscillator and the Coulomb potential. In the
latter case, all eigenfunctions fall into a single irreducible representation of the
group .z' 5, which is like a five-dimensional rotation group with a change in sign
of one component in the metric. In this case the symmetry group fJt4 is
contained as a subgroup within the non-invariance (or dynamical) group Z' 5 '
The choice of dynamical group for a given system is not, however, unique. It
should contain the symmetry group and hence account for the degeneracies
and also contain operators which couple eigenfunctions. The spectrum of
energies should emerge as the allowed values of one of the group operators.
From this point of view the concept of the dynamical group has been of
interest in the theory of elementary particles. Here the masses of different
particles provide a spectrum with apparently unrelated values. One would like
to find a dynamical group which brings them together within a single
representation and hence relates the masses to some group operators.
On a more practical level the non-invariance groups may be used to deduce
simple formulae for transition matrix elements between eigenfunctions. Thus
in the simple one-dimensional oscillator we have, from the definitions of Hand
K I ' the identity x 2 := H + 2K I so that the matrix elements of x 2 follow from
those of the group operators Hand K I which are known for the irreducible
representations using equations (20.8) and (20.11). But, x 2 is the kind of
operator involved in electromagnetic transitions of quadrupole type .
502 A Miscellany 20 .2

20.2 The Jahn - Teller effect and spontaneously


broken symmetries

The Jahn- Teller effect is concerned with symmetries which occur following an
approximate separation of the Hamiltonian. Because the separation is only
approximate there are terms which break the symmetry. The word 'spon-
taneous' is sometimes used to distinguish this effect from the symmetry
breaking caused by the addition of an external perturbation which we
discussed in section 5.8. Although the idea is a general one, it is in the structure
of molecules that the Jahn- Teller effect was first described and in which it is
most clearly seen. More recently the same idea has been used in the theory of
elementary particles. To help visualisation we shall present the theory in the
language of molecules but its generality will be apparent. Brief mention of the
application to elementary particles will be made at the end of the section.

20.2.1 The adiabatic approximation

A molecule is composed of nuclei and electrons interacting via the usual


electrical forces and we denote the coordinates of the nuclei and electrons
symbolically by Rand r, respectively . Because the nuclei are more than a
thousand times heavier than the electrons and interact with the same forces as
the electrons one expects the nuclei to move much more slowly than the
electrons. The crucial approximation is first to calculate the orbital motion of
the electrons on the assumption that the nuclei are fixed, and then to calculate
the equilibrium positions and the rotational and vibrational motion of the
nuclei taking account of con sequent changes in the electron energies. In
general, one calls this an 'adiabatic' approximation, although in molecules it is
more often referred to as the Born -Oppenheimer method.
In more det ail let us write the full Hamiltonian as
H(R ,r) = T N(R ) + T e(r) + V(R,r) (20.12)
where T N and T, denote the kinetic energies of nuclei and electrons,
respectively , and V denotes the potential energy of interaction between the
nuclei, between electrons and between electrons and nuclei. In the adiabatic
approximation one first solves the Schrodinger equation
{Te(r) + V(R ,r) } "'n(R ,r) = En(R)"'n(R,r) (20.13)
for the electrons, keeping R fixed. We have used the symbol n to label the
eigenvalues although in practice n will need to be a composite label
because of the many dimensions of the electron coordinates r . Next, one seeks
an approximate eigenfunction 'P of the Hamiltonian (20.12) by writing
111= "'n(R ,r)<1>n(R). Substitution into the Schrodinger equation
H(R,r)'P = E'P then leads to the equation

{T N(R ) + En(R)} <1>n(R) = E<1>n(R) (20.14)


20.2.2 A Miscellany 503

if we neglect the derivatives of l/Jn(R,r) with respect to R. This neglect really


defines the adiabatic approximation in its simplest form . Knowing the
electronic energies En(R) for any fixed R from equation (20.13) enables us in
principle to solve equation (20.14) and so complete the solution.
In practice the solution of equation (20.14) is obtained by continuing with
the idea of slow motion of the nuclei. To find the lowest energy one first
neglects the kinetic energy T N(R) and simply minimises the function Eo(R)
with respect to R, i.e. for each component R, of R :
(20.15)
This provides the equilibrium positions R = R o of the nuclei. Of course, it
only fixes their relative positions and the inclusion of T N(R) contributes a
rotational energy. Allowing an expansion of the potential function Eo(R)
about R o we can also generate vibrations about the equilibrium position, as
described in chapter 6. In this chapter, however, we are interested only in the
static equilibrium position.

20.2.2 The role of symmetry

The full Hamiltonian H(R,r) is necessarily invariant under all rotations of


nuclei and electrons together, since the interactions depend only on relative
distances . One therefore has the familiar result that the overall wave function
'I' has definite angular momentum J which in the present problem will relate to
the rotational motion. However, there is nothing new about this general result
which has been discussed earlier.
The present interest in symmetry concerns the potential V(R,r) which for
fixed R governs the electronic motion. Regarded as a function of the electron
coordinates r, it will not be invariant under all rotations of r because the
nuclear coordinates R are fixed at their equilibrium values R o • However, if the
equilibrium arrangement of nuclei has symmetry with respect to a point group
<§ then V(Ro,r) must have the same symmetry. When the molecule contains
several identical atoms, as in the examples quoted in chapter 6, one can
reasonably expect to find such symmetry. The consequences of the symmetry
willappear both in the vibrational spectrum, as described in chapter 6, and also
in the electronic energy levels En which will have characteristic degeneracies
corresponding to irreducible representations T(a) of <§, see section 14.4.
Let us now look a little more closely at the consistency of equations (20.13)
and (20.15) for a symmetrical equilibrium configuration. We shall find that in
many cases the distribution of electrons is sufficiently unsymmetrical that the
equilibrium position moves away from the symmetrical one -the Jahn- Teller
effect. Differentiating the Schrodinger equation (20.13)with respect to R, leads
to the result
(20.16)
so that from equation (20.15) these matrix elements should vanish in the
504 A Miscellany 20 .2.2

equilibrium position R = R o . Let us now consider the significance of the


vanishing of these matrix elements using the selection rule technique described
in section 5.4. As we have seen in detail in chapter 6, the 3N coordinates R, of
the N nuclei of a molecule may be defined such that they belong to definite
irreducible representations T CY! of r§ . If R, is a symmetrical coordinate,
belonging to the identity representation T( 1', then since V is invariant , so also is
the operator av/ aRj • Hence, no matter to which representation T "'! the wave
function l/Jo belongs, there is no symmetry reason why the matrix element
should vanish. However the value of any symmetrical coordinate may be
varied indefinitely without destroying the symmetry, so that the vanishing of
the matrix element may be achieved in thi s way and simply determines the size
and the other free parameters of the symmetrica l system. For the remaining
non-symmetrical coordinates no such freedom exists since a non-zero value
for any of them implies a departure from symmetry. Hence if the matrix
elements (20.16)are to vanish they must do so for reasons of symmetry. Since V
is invariant, the set of representations T(Y! to which the derivatives ()V/ aR,
belong is the same as the set to which the coordinates belong. Thus we require
the vanishing of (20.16) for all TIl'), other than TO), which occur for the
coordinates R . (One should exclude here the rotational and translational
coordinates, see section 6.5, since V is necessarily independent of these.) From
the selection rule arguments of section 5.4 this means that T('I must not occur
in the reduction of the products T( Y ) x T(') and in terms of characters this
condition becomes
L z"I * (Gul/I (Gul z" I(Gul = 0
G.
(20.17l
L I x"" (GuJ!lx';" (G u) = 0
G.

For anyone-dimensional (unitary) representation T"", we have Ixt"(GaW= 1


for each element Go so that (20.17) is always satisfied because of the
orthogonality of X(YI with the identity representation. However, for all
representations with dimensions greater than one it is found that except for
linear molecules there is at least one T(Y) for which (20.17) is not satisfied. Thus
if there is a degeneracy in the electronic ground state then the minimum energy
condition (20.15) is not satisfied in the symmetrical position and so the lowest
energy is obtained in a non-symmetrical position of equilibrium in which there
will generally be no degeneracies. It is interesting to note that in the case of a
one-dimensional representation T"" the electron density Il/Jo12 is invariant
since 1x""(G a ) 12 = I (This is true even when T"') is not the identity and l/Jo is
itself not invariant.) Intuitively one would expect the symmetry to be unbroken
in this situation, while for representations with greater dimension no such
argument is generally possible.
As an example, consider the ammonia molecule described in section 6.5,
where it was found that the six internal coordinates belonged to repre-
sentations A 1 and E, each appearing twice. For the sake of argument let us
20 .2.3 A Miscellany 505

suppose that the electron wave function belongs to the two-dimensional


representation E giving a degeneracy. Then the matrix elements (20.16)
corresponding to the two symmetrical A I-type coordinates wil1 not vanish for
symmetry reasons but can be made to do so by choice of the bond angle ¢ and
the NH bond length. For the E-type coordinates the use of the known
characters shows that eq uation (20.17) is not satisfied when T(a) = T(Y) = E and
hence the symmetrical position is not a position of equilibrium.
In practice it is not trivial to find an observable example of the Jahn- Tel1er
effect. This is because the electronic ground states of molecules are usual1y
symmetrical and therefore non-degenerate. Excited states should, however,
demonstrate the effect and probably the best example is seen in the photo-
electron spectrum of methane which indicates (see Sturge, 1967, in the
bibliography) that, in the final ionised state of methane, the molecule is
distorted from the regular tetrahedron.
We have ignored spin in the discussion above. It may be shown that,
although in principle any spin degeneracy (beyond the two-fold time-reversal
degeneracy for an odd number of electrons) leads to a Jahn- Tel1er symmetry
breaking, the effect is very small.

20.2.3 Spontaneous symmetry breaking

There is a very interesting symmetry argument in the Born -Oppenheimer


approximation for molecules which occurs in many other physical systems and
is sometimes referred to as spontaneous symmetry breaking. The complete
Hamiltonian has the ful1:Jf3 symmetry because it depends only on the relative
interactions of the constituent electrons and nuclei . However, in finding an
approximate solution we made use of a subgroup of (fl3, the point group '§
appropriate to the equilibrium arrangement of the nuclei . In this sense, the
symmetry was broken from (fl 3 to '§ and would be restored only when the
rotational motion of the molecule is included, i.e. in addition to the motion of
the electrons and the vibration of the nuclei one must also allow the entire
system to rotate. In a macroscopic crystal, however, the symmetry group '§
acquires greater significance because, with a macroscopic moment of inertia,
the energy quanta for rotational motion become very small.
For a molecule, the restriction from ~3 to a point group '§ was perhaps
natural because of the distinction between electrons and nuclei . However, the
same principle occurs in other physical systems . For example, in the structure
of nuclei the constituent nucleons are all identical but there is strong evidence
that they acquire a non-spherical equilibrium shape which then rotates. The
number of nucleons is sufficently large, of the order of 100, that the subgroup is
not a point group but the continuous group f.lt 2 together with two-fold
rotations about all axes perpendicular to the axis of f.lt 2-a group denoted by
D x: (the limit of Dn ) . Unlike a crystal, there is no macroscopic assembly of
nuclei in which to see the D x; symmetry but one may infer the symmetry from
the properties of the rotational levels of nuclei.
506 A Miscellany 20 .2.3

Another more sophisticated example occurs in the theory of supercon-


ductivity. Here, one is concerned not with rotational symmetry but with the
conservation of the number ofelectrons. This is such an obvious symmetry ofa
non-relativistic system that one does not usually mention it. The group
concerned is the unitary group in one dimension with a single parameter a and
group operators exp( iaN), where N is the number operator. The essence of the
theory is to construct, in a particularly simple form, a trial wave function which
breaks the number symmetry, i.e. which is a superposition of states with
different particle number. For a large number of electrons interacting mainly
in time-reversed pairs of single-particle states, this proves to be a good
approximation. For a small number of particles one would have to recover a
state with definite particle number (the quantity corresponding to angular
momentum in the previous example) and this might be achieved by the action
of a number projection operator on the symmetry-breaking trial wave
function . Again, in a macroscopic system , the symmetry breaking solution
acquires significance and leads to an explanation of the phenomenon of
superconductivity whereby electr ic current passes without resistance . (Exci-
tations from the symmetry breaking state are not possible until the energy
passes a certain energy gap , whereas for free electrons there are excitations at
infinitesimally low energy.)
As a final example, we mention briefly the use of symmetry breaking in the
relativistic field theory ofelementary particles. It arises in attempts to unify the
theories of electromagnetic and weak (p-decay) interactions. Because of the
short range of the weak interactions the relevant field particles (vector mesons)
must have finite mass, in contrast to the photon, and th is introduces
divergencies in the field theory which cannot (apparently) be removed. For
zero-mass field particles such difficulties do not occur and one is able to
construct a unified Lagrangian which is invariant under a set of gauge
transformations (see subsection 16.3.5) and which offers hope of relating
various coupling constants that occur in the weak and electromagnetic
interactions. The vital step of introducing a finite mass for the vector mesons is
now achieved by the device of symmetry breaking. It is supposed that the
vacuum is not an invariant (nor does it belong to any irreducible repre-
sentation) of the full gauge symmetry group but is invariant only under some
subgroup. This introduces, into the Lagrangian, terms which are equivalent to
assigning a mass to some of the vector mesons. By suitable choice of the gauge
group and its subgroup under which the vacuum is invariant, one may arrange
that the photon preserves its zero mass while the vector mesons acquire mass.
The introduction of mass by this device avoids the divergencies referred to
earlier. The mechanism by which the vacuum achieves broken symmetry is a
little obscure but seems to involve the introduction of an additional field of
scalar particles, some of which conspire to provide the third component
needed by the vector mesons in acquiring mass! (Note from subsection 16.3.7
that the photon has only two states m = ± 1, whereas a finite mass particle
with s = 1 needs m = ± 1 and m = 0.)
20.3 A Miscellany 507

20.3 Normal subgroups, semi-direct products and


little groups

In the early parts of this book we limited our study of the theory of groups to
those properties which were vital in understanding the physical applications.
There are many fascinating properties, even offinite groups, which we have not
mentioned. Here we introduce the idea of the normal subgroup and describe
its use in finding representations. In fact this provides a general background
for the introduction oflittle groups which we have used in an ad hoc fashion in
obtaining represent ations of the space groups in chapter 14 and the Euclidean
and Poincare groups in chapter 15.
For the direct product of two groups defined in section 2.5 the characters
and representations are obtained trivially from those of the two separate
groups, as described in section 4.21. Remember that in the direct product
'§ = Yf x X the elements of Yf must commute with those of X . We now
define what is meant by saying that a group '§ is a 'semi-direct product', written
'§ = Yf /\ X . The elements of '§ must be the set of products G j a = HjK a,
where Hi and K, denote, respectively, elements of two subgroups Yf and X
and , although the H, need not commute with the K a , the relati on

(20.18)

must be satisfied where H, is another element of Yf . (It is possible to give a


more general definition of a normal subgroup in which Yf is simply a subgroup
of'§ for which a relation of the kind GaH jG a- 1 = H, exists for all G, in '§.
The factorisation G = HK of the elements of '§ does not then necessarily
occur and although the factor group may still be defined, it is not necessarily a
subgroup of '§.) This definition implies that the elements conjugate to any H,
with respect to any element of '§ will all lie in Yf . Note that the definition
(20.18) is not symmetrical between the two subgroups Yf and X . The group Yf
is called a 'normal' (or invariant or self-conjugate) subgroup of '§. The group
,ff is called the 'factor' (or quotient) group of '§ with respect to .Jf', a relation
often expressed by the notation .X- = '§ j,Jf'. If g, hand k denote the orders
(number of elements) of '§,Jf' and ,ff then clearly g = hk and k is called the
'index' ofJf' in '§.
As examples of semi-direct product groups among the finite groups there is
the tetrahedral group, see section 9.3, which may be written as T = D /\ C 3 and
the simpler example D4 = C4 /\ C 2 ' A careful study of the elements of T
(or D4) and their multiplication table is necessary to convince oneself of these
relations. The space groups of section 14.9 which were formed by including a
group ff of discrete translations with a point group '§ are semi-direct products
ff /\ '§ with the translations forming the normal subgroup because of
equation (14.72). In a similar way the Euclidean group of section 15.1 is given
as iff3 = ff 3/\ 9t 3 and the Poincare group of section 15.4 as f!JJ = ff 4 /\ !f,
508 A Miscellany 20 .3

where :?7 3 and :?7 4 denote the translation groups in three-space and space-
time, respectively.
Let us now see how a knowledge of thi s semi-direct product structure
rs = Yf /\ X of a group can help us to find its irreducible representations. For
simplicity we limit our discussion to the case when the normal subgroup .Yf is
Abelian (the generalisat ion is straightforward) so that its irreducible rep re-
sentations are one-dimensional denoted by T(;·). Since .Yf is a subgroup of rs
the basis vectors of an irreducible representation T of rs may be chosen to
belong to definite irreducible representations of .Yf so that, denoting these
II.
vectors by >we have, for any H in Yf ,
I >= T(/.)(H) IA. >
T(H) I. (20.19)
We now choose a particular II. >and use the elements K, of X to generate
from it the set of vectors denoted by
(20.20)
Each of these new vectors must itself belong to one of the representations of J't'
because from equations (20.18), (20.19) and (20.20)

I
T(H) I.a >= T(H)T(K a) I)· >= T(Ka)T(H ') II. >= T(/.)(H') II.a >
(20.21)
where H' = K,' I HK a. However, notice that the representat ion of ,Y{' defined
by this equation is not necessarily T(/.) since it associates with the group
element H the matrix element T(/.)(H') for a different element H'. Let us
therefore denote this representation of.Yf by T(;,.). Consider first the simplest
possibility, that the set of representations TV.), as G, runs through the elements
of X , are all inequivalent. In this case the vectors II.a > must be linearly
independent and provide a basis for a k-dimensional irreducible repre-
sentation of rs. The explicit form for the representation is quickly found by
operating on I).a >with the general group element G = HK,

I I
T(HK) I.a >= T(H)T(K)T(K a) I. > = T(H)T(K b) I. > I
= T(Kb)T(H ') I)· > = T(/.)(H') I).b > (20.22)

where we have defined K, = KK a and H ' = K; 1 HK b. Thus the matrix


elements of T(HK) are zero except for the numbers T(;')(H') coupling rows a
with the appropriate columns b and these numbers are known from the
irreducible representations T(/.) of the subgroup ,Y{.
Some modification is necessary when any of the representations T( A.) are
equivalent to T(/.). Given I, let us denote by Ka the set of elements of X for
which this equivalence holds. We now show that this set forms a subgroup of
X which we denote by %. It is generally referred to as the 'little group', relative
to rs, Yf and A.. We shall not verify that the set Ka satisfy all the group
postulates but only the multiplication law . Given that Ka and Kb belong to the
20.3 A Miscellany 509

set so that
T().)(K; I HK a ) = ST().)(H)S-I and T(A)(K; I HK b) = RT().)(H)R- 1

for some fixed Sand R, independent of H, we have for KbK a


T(A)(K;IK;IHKbK a ) = T().I(lt./H'K a )= ST(A)(H')S -I = SRT(A)(H)R-1S- 1

where again H' = K; I HK b. Thus the multiplication law is satisfied. The set of
vectors II.a > which transform equivalently to a given A. must therefore form an
invariant space under the little group .% and so we may choose them to
transform according to definite irreducible representations "Pa ) of X-. Thus the
basis vectors will carry both labels ;. and rx and also a row label i if Tla) has
dimension greater than one. To generate the irreducible representations of ~
we start from one such set of vectors II.rxi > with fixed A. and,rx. It is a general
property of any group X of order k that given a subgroup X of order k any
element 0[,)( may be written in the form K = MeK, where Me is one ofa set of
klk elements of.)( (this is Lagrange's theorem, see problem 2.4). Because of
this result the representation T of ~ may be generated by operating on the
vectors IA.rxi> with the set of k/k elements M" T(Me)IArxi > = IArxi, c). The
representation is labelled T().· 0) and has dimension saklk with its matrix
elements being found as in the derivation of equation (20.22), i.e.

T(HK)j).rxi, c > = T(H)T(K)T(M e) II.rxi > = T(H)T(K') IArxi >


= T(K')T(H') I/.rxi > = T(K')T(i·)(H') I).rxi >

= T(i·)(H')T(Md)'nK') I),rxi > = T(i·)(H') LT~a/(K')T(Md) IA.rxj )


j

= T().)(H ') L T)f)(K')IAc<j, d> (20.23)


j

where the new symbols K', H', M d, K' and d are defined by the equations
K' = KM e, H' = K,-I HK', K' = MdK '. Thus the representation is given in
t~rms of the matrix elements of the representations of the subgroups .Jft' and
X.
The question arises whether each choice of ;. leads to a different irreducible
representation of ~ and the answer is generally 'no', because if some of the TV..
are not equivalent to T(i·)then the same representation of ~ would have been
obtained by starting from any of these TV.·) instead of TW. The set of
representations T(A.) are said to be in the same 'orbit'. In constructing all
irreducible representations of ~ one need only choose one I. from each orbit.
(The concept of orbit is closely related to that of the 'star of k' in subsection
14.9.1, since k corresponds to A.. Representations T(k,) of the translation group,
where k , is in the star of k, will be in the same orbit.)
As a simple example of this procedure let us find the representations of the
510 A Miscellany 20.4

semi-direct product group D 4 = C 4 /\ C 2 . We first note geometrically that the


C 2 axis must be perpendicular to the C 4 axis and that the product elements fall
into the five classes of D4 as follows : E; CI; C 4 , C~ ; C 2, C~C2 ; C 4C 2' C~C2 '
There are four irreducible representations T(A) of C 4 from which to begin , see
table A.1 of volume 1. Starting with the identity representation A of C4 it is
clear that the little group here must be the factor group C 2 itself, since for the
identity representation the matrix elements T()·)(H) = 1 are the same for all H.
Thus to each irreducible representation T(') ofC 2, and there are two of these (A
and B in the notation of table A.I ), we get a one-dimensional representation of
D4 . From equation (20.23) and using the character table for C 2 one sees that
this provides the representations of D 4 which are labelled A 1 and A 2 in table
A.I. Turning now to the next representation B of C; the little group is again
seen to be C 2 so that two more representations Bland B 2 of D 4 are produced.
Finally if we take the first of the pair of representations ofC 4labelled by E, the
little group can no longer contain the element C 2 because Ci I C 4 C 2 = C~ and
the matrix elements of C, and C~ are not the same for the representations E of
C 4 • Thus the little group reduces trivially to the identity and we have a two-
dimensional representation (k/k = 2) of D4 given by equation (20.22) and
labelIed by E in table A.I. The use of the second of the pair of E representations
of C4 leads to the same repre sentation of D4 because the pair belong to the
same orbit.

20.4 • he classification of Lie groups

A systematic listing of the finite point groups was given in chapter 9 but for the
continuous Lie groups introduced in chapter 7, we have so far considered only
those examples which have occurred in particular physical problems. The
object of this section is to find a systematic listing of Lie groups so that in any
new symmetry problem one is in a position to recognise the group responsible.
We begin by defining a few new terms. Ifa group has no invariant subgroup
it is calIed 'simple'. If it has no Abelian invariant subgroup it is calIed 'semi-
simple', so that the simple groups form a subset of the semi-simple groups. The
word 'non-semi-simple' is sometimes used for groups which have an Abelian
invariant subgroup. As we have seen in the preceding section the non-semi-
simple groups may be described as semi-direct products so we restrict our
attention to semi-simple groups. Furthermore it may be shown that if a Lie
group is semi-simple but not simple then it may be written as a direct product
of simple groups. Thus we shalI, later on, restrict our attention further to the
classification of the simple Lie groups only .
The discussion starts from the set of r infinitesimal operators X, together
with their commutations relations, see section 7.2,

(20.24)
20 .4 A Miscellany 511

(A set of operators with this law of 'multiplica tio n' is often called a Lie
'a lgebra") If the matrix

gq s = L e~p eft (20.25)


P .t

is formed from the structure constants e~p then the (Cartan) criterion for the
group to be semi-simple is that det 9 #- O. (It is not difficult to deduce this
criterion if we notice that the existence of an Abelian invariant subgroup
implies that a subset Xi' of the Xi satisfy the relations

[Xq .,X p'] = 0, [Xq"X t ] = Ls' c",qr XI


where we use a prime to denote the members of the subset). For any given
group there is much freedom in defining the X, because any real linear
combination of the X, is itself an infinitesimal operator, and in trying to
classify the different Lie groups we shall naturally choose a convenient set. We
first find the largest set of mutually commuting infinitesimal operators,
denoting them by Hi' The number of operators Hi is called the 'rank' of the
group and denoted by f. Necessarily 1 ~ f ~ r since any operator commutes
with itself. One may then choose the remaining operators E.. to be
'eigeno perators' of the Hi ' in the sense that
(20.26)
for some set of 'eigenvalues' a i . That the E.. may be chosen in this way to be
eigenvectors of all the Hi at the sa me time follow s from the fact that the Hi
commute-a familiar result in ordinary eigenvector theory. Given the
structure constants for any set of X, one can search to find the Hi and hence
construct the H i and E...The set of f eigenvalues ai is conveniently denoted by a
vector a = (ex 1 ex 2 . . • ex,) in I dimensions. It is called a 'root vector' and it can be
shown that for semi-simple groups there is no degeneracy, i.e. no two operators
E have the same non-zero root vector. (One may of course regard the Hi
themselves as eigenoperators with a = 0, in which special case there is a
degeneracy of order f). Thus the root vector a is a sufficient label for the E...
To complete the commutation relations among the new set of operators Hi,
E.. we need to study [E.., Ell] but by rewriting the double commutator (Jacobi
identity) we have

[Hi' [E.., Ell] ] = [[ Hi' E..] , Ell] + [ E.., [Hi' Ell] ] = (ai + Pi)[ E.., Ell]
so that the commutator [E.., Ep] is itself an eigenoperator of the Hi with root
vector ex + p. Thus if a + fJ i= 0 we may write
[E.., Ell] = N..6 E" +6 (20.27)
with some normalisation constant which depends on the definition of the E...
When a + fJ = 0 the commutator [E.., Ell] commutes with the Hi and must
512 A Miscellany 20 .4

therefore be a linear combination of them so that we may write

(20.28)

defining the constants !Xi which again depend on the choice of Hi and the
normalisation of E~ . In fact we can conveniently make ai = ai • (To j ustify this
step one first chooses the Hi such that gij = Dij by using a Schmidt process
because g ik = 1:
, ai~ has the form of a scalar product. Next, one defines
Cpqs = ~ gPtC~s and shows it to be antisymmetric in its three indices. Hence we
have

Ck, -a= Igk'C~ -, = ~ -, = ~


t

and also

Ck a -, = C - aka = Ig-21 c L = g -aac:a = a,. ga-a


t

Thus rl is proportional to ~ and a choice of magnitude of the E a will then lead


to the equality.)
In classifying the Lie groups one may first use the rank Iand for each value of
Ifind the possible sets of root vectors ex in I-dimensional space. We shall see that
these vectors must satisfy some rather stringent conditions so that there are
only a few solutions. The first property of the roots is that if ex is a root vector
then so is -ex . (This can be seen by constructing the matrix elements gap from
the definition (20.25) using the vanishing of certain structure constants implied
by equations (20.26H20.28). It is seen that for given ex, all matrix elements
vanish unless f1 = - ex so that, unless a root - (X exists, the ent ire row (X of the
matrix 9 vanishes, implying that det 9 = 0 which violates Cartan's criterion of
semi-simplicity.) This enables us to divide the roots into positive and negative,
according to whether the first component a l > 0 or a l < O. If Cl I = 0 then the
sign of the second component Cll is used to decide whether the root is positive
or negative and so on . The ordering of the H i and hence of the a, is arbitrary.
The number t(r -I) of positive root vectors is generally greater than I so that

vectors (1,0,0 . .. ), (0,1 , °...),


they are not linearly independent. For a basis one may continue to use the I
etc., but it is more convenient to use a basis of
vectors each of which is itself a positive root vector and in particular a basis of
what are called 'simple positive roots'. Such a root is defined to have the
property that it cannot be written as the sum of two other positive roots. It may
be shown that there are just I simple positive roots and that they are linearly
independent. A further property is that if (X and pare two simple positive roots
then it follows that p- (X cannot be a root of any kind. (For suppose p- (X = y,
with ya positive root, then p = (X + y which contradicts the definition of p as a
'simple' positive root. A similar argument disposes of the possibility that y is
negative) .' However, there may be roots of the form fJ +a and indeed there will

t The symbol s y and~ , like ex and p, represent vectors and should have been printed in bold in
thi s secti on .
20.4 A Miscellany 513

generally be strings of roots P + a, P + 2a, .. . , P + na, which terminate at


some integer n, i.e. P+ (n + I)a is not a root. The value of n may be related to
the vectors by the equation
n = - 2(a.p) j(a .a.) (20.29)

where the scalar product is defined as a..P = Lrx;/3 i . This general result is
deduced from a recurrence relation as follows . 'tr m is an integer ~ n then,
from the Jacobi identity,

lv- [E eJ , E _ eJ]] = [[ E p + meJ, EeJ]' E_ eJ] + [ EeJ ' [E p + meJ' E _ eJ]]

r.e.

and so

- N p+meJ. -eJN p+meJ-eJ.eJ )E p+meJ

and if, for brevity, we define Am = N p+meJ .eJN p +m<J+eJ . -eJ then we have
a .p+ma .a = -Am +A m- I (20.30)
Now since, as explained above, P- a. is not a root we have N p_ eJ = 0 so that
A _ I = 0 and from recurrence relation (20.30) we may deduce
- Am = (m + I)a .p +-!m(m + l)« .« (20.31)
If n denotes the termination of the string, as defined above, then N p +neJ,eJ = 0
so that An = 0 and hence from equation (20.31) we deduce the relation (20.29).
From the definition of the scalar product it follows (Schwartz inequality)
that (a..P)2 ~ (a. .a.)(p.P)and hence from equation (20.29), and the correspond-
ing equation p = -2(a .p) j(p .P) for the p-string a+p,a+2p, .. .,a.+pP,
where p is again an integer,
-(a.p) j(a..a.)t(p .P)t = +tl-np) t ~ I (20.32)
This may be said to define the angle between a. and p. From the same two
equations one has the ratio of the lengths
<p'P)/(a..a.) = n/p (20.33)
The values of n (and p) are strictly limited by equation (20.32) which implies
that np = 0, I , 2 or 3. (The value 4 is not allowed since this would mean that
a = - Pcontradicting the definition of a and Pas both positive roots.) Hence
we are able to enumerate the possible sets of simple positive roots and, from
these, all roots are constructed by forming all possible strings. In some cases it
514 A Miscellany 20.4

is possible to build p-strings on roots p + rnex although multiples like rn(p + ex)
with rn > 1 are not allowed.
The enumeration is done very concisely by forming what are known as
' Dynk ln diagrams'. Each simple positive root is represented by a point. For
each pair of such roots the quantity np (which determines the angle between
them) is signified by joining the corresponding points with that number np of
straight lines, dra wn close together. The relat ive lengths of two roots joined in
this way is sometimes denoted by placing the relative values of ex.ex over each
point. In fact these are very limited because if np = 1 then n = p = 1 and from
equation (20 .33) the lengths are equal. If I1p = 2 one is v ' 2 times the other ,
while if np = 3 one is v '3 times the other. The task of drawing all possible
Dynkin diagrams is greatly simplified by noting th e following rule s which
follow from the equations we have derived above:

(1) Only connected diagrams need be considered because if a diagram splits


into two unconnected parts then, because there are no connecting lines, the
roots in one part are orthogonal (a .p = 0) to those in the other and this implies
that the group is a direct product and not 'simple'.
(2) There can be no loops. For example, suppose a, p and y were to form a
triangle of single lines then, from equation (20.32) with np = 1,
ii.iJ = iJ. Y = Y·ii = - t, wherefor brevity we write ii = a/(a.cx)t, etc. Now since
ii, iJ and y have to be linearly independent it must be possible to write
ii = biJ + ey +~, where ~ is orthogonal to p and y and is non-zero. This is
simply the projection of ii on to the subspace defined by iJ and y with ~ lying in
the orthogonal complement space. The coefficients band c are then given by
the equations p.ii = b + ep .y and y. ii = by . P+ cand inserting the value - t
for each scalar product this gives ~.~ = 0, a contradiction. Larger loops are
excluded by a similar argument and the use of double lines only sharpens the
contradiction by giving ~ . ~ < O.
(3) There can be at most three lines radiating from any point. Suppose there
were roots cx l ' a2' . .. cxp all connected to a root p then no two of the roots a j
could be connected to each other without violating rule (2) and so the roots a i
are mutually orthogonal aj .cxj = 0 for i =I j . Thus if as in (2) we were to write

p- = 1=1p
. I: ajti + ~
j with ~ .cX j =0 for all i, then the coefficients are given by

aj = Ii. eXj and p

~ .~ = 1- L af < I - tp
j= 1

But once again ~ . ~ > 0 so that p ~ 3.


(4) Rule (3) excludes diagrams containing any of the following pieces:

~orX
20.4 A Miscellany 515

The arguments in rule (3) may be extended to show that if any number of single

>-<
lines are inserted into these pieces, for example

or

then the same contradiction arises . One therefore concludes that no Dynkin

-<
diagram may contain more than one double-line connection

~ or junction

nor may it contain both a double line and a junction.


We are now in a position to list the only four types of diagrams which may
exist:
(a) A chain of single lines
(Xl (X2 (X,
~-----~ ~.--~

(b) A chain with one double-line connection and p + q + 2 =I


II (X2 (Xp Y Pq + I Pq
. ......
P2
_.---PI
(The unsymmetrical notation is for convenience in the algebra below.)
(c) A chain with one junction and p, q and r roots in each of the three legs
with p + q + r + I = I

»:
}II

(Xl (X~~/
~q
_.-~--

The angles between legs in the diagram have no significance.


(d) A single triple connection

There is little more to be said about types (a) and (d). In type (a) all the simple
roots have the same length and the group may be identified with the special
unitary group SU, + i- To see this equivalence one must use the diagram to
construct the commutation relations of the infinitesimal operators and then
compare with the relations appropriate to the group S U/+ I ' see section 18.8.
The rank I may take any integer value. In type (d) one may not add any more
roots since this would imply more than three lines at one of the points. Thus in
type (d) there is only one group, with rank I = 2, which is conventionally called
G2 ' The simple roots have lengths in the ratio J
3: 1.
In type (b) we must ask where the double connection may lie in the chain, i.e.
what are the values for p and q, having chosen p ~ q without loss ofgenerality. As
before, we write
516 A Miscellany 20 .4

p q + 1
Y = I ajlX j + I bjPj + ~
i i

with ~ .IX; = ~ ';3; = 0, so that


IXj.y = IadlX;.lX j) = aj -t(a j + 1 +Qj-l)

°
j

with a similar result for P/t. However IXj' Y= except for IXp' Y = - so that t
we can deduce Q p = -P/(P+ 1) and similarly bq = - J (2)q/ (q + 1). Using
these numbers in the expression

~ .~ = 1- IQjlXj.y - IbjPj .y = 1- QplXp.'Y -b q + IPq + 1'1'


i i

= 1 +tap+bq + 1/ J 2
reduces the condition ~ . ~ > 0 to the result pq < 2. Thus either q = 0 with any
P or P = q = 1. Diagrams of type (b) may therefore be of the general form
a I a2 a/ _ 2 Y
. . .. .. ,
P~~====o

for any I or of the particular form , for I = 4,

In the first case there are two different Dynkin diagrams because although the
a i and )I all have the same length (since they are joined by single lines) the ratio
of (length)? of Pand y must be 2. Thus either the root Pis a factor J 2 smaller
than all others or a factor J 2 greater. These two possibilities may be identified
with the odd rotation groups 9i 2/ + 1 and the symplectic groups SP2/'
respectively (see section 18.10 for some applications of these groups). The
particular diagram with p = q = 1 corresponds to a group known as F 4 '
The limitations on the values of p, q and r in diagrams of type (c)a re deduced
by the same procedure, expressing 0 as a sum over the a;, Pi and Yj with eits
projection orthogonal to the space of the aj, P; and y;. In this case again
Q p = - p /(p + I) with corresponding expressions forthe other coefficients and

e .~ = 1 + t(Q p + bq + c,). The condition ~ .~ > 0 then becomes

p/(p+I)+q /(q+I)+r /(r+l) < 2 (20.34)

Because of the symmetry of the diagram we may take p ~ q ~r without loss of


generality and r ~ I or else the diagram reduces to type (a). Starting with the
smallest values q = r = 1 the condition (20.34) gives p/(p + I) = 1 which is
satisfied for all p and gives the general diagram
20.4 A Miscellany 517

which corresponds to the even rotation groups 9t 2/ . Moving on to the next


value of q, namely q = 2, r = I, the condition (20.34) becomes p/(p + 1) < i
which is satisfied only if p = 2, 3 or 4. Greater values of q and r lead to no
solutio n. Thus there are ju st three more particular groups, referred to as E6 , E 7
and E 8 with the Dynkin diagrams

It is an amusing coincidence that the inequality (20.34) (sometimes called a


Diophantine equation) is essent ially the same, when r = 1, as equation (9.3)
which arose in classifying the point groups. Hence the enumeration of Lie
groups of type (c) is in one-to-one correspondence with the listing of the point
groups.
Two groups which are isomorphic or homomorphic and so have the same
structure constants, may be identified by having the same Dynkin diagram.
Thus, for example, the groups 9t 6 and S U 4 both have the diagram

The groups Jf) and S U 2 are represented, almost trivially , by a single point. The
group 91 4 does not appear in the classification above because it has the same
structure constants as the direct product Jf) x (]f) and is not therefore a
'simple' group. This structure is in fact suggested by trying to put I = 2 in the
Jf 2/ diagram which would lead to two unconnected dots -the diagram
expected for (]f ) x (]f).
To find all roots and hence the commutation relations, given the Dynkin
diagram one must construct the strings described earlier in this section. We
illustrate the procedure with the group S U) described in detail in sections 11.6
and 18.7. The Dynkin diagram is

which implies that I = 2 with two equal roots satisfying Ii.P= - t. Hence from
equation (20.29) the string contains only (X + P and of course there are
also the corresponding three negative roots, making six in all. Since I = 2 in
this case, we can simply draw all the roots in two dimensions and the scalar
product Ii.p = -t implies an angle of 120' giving the picture shown in figure
20.1. Notice that, because of the angle of 120°, the root (X + P has the same
length as (X and p. This picture of course agrees with figure 11.3 for the six
infinitesimal operators T +, U + , V + in the notation of chapter 11. (The relative
scale of the two axes is different in the two figures by the factor! ) 3 discussed
in section 11.6.) Finally, to deduce the commutation relations (20.26)-(20.28)
we must choose a basis and hence define HI and H 2 • We shall take axis 2 along
(X with axis 1 perpendicular to it, since this is consistent with (X and P being
518 A Miscellany 20.4

a+{J

-a a

-{J

Figure 20.\

simple positive roots. With the vectors taken of unit length this gives
[H 2 , E,,] = E", [H 2 , EfI] = -1EfI , [H 2,E<l+fI] = 1E<l+fI
[HI>E,,] = 0, [H1,EfI] = 1(3)!E fI , [H1 ,E<l+fI] = 1(3)!E<l+fI
together with
[E", E_,,] = H 2 ,

[E<l+Il,E_,, _p] = 1H 2 +1(3)!H 1


The remaining commutation relations are given by equation (20.27) but there
is freedom in fixing the normalisation constants because the relative magni-
tudes of the operators E are not yet defined. In general one may do this so that
the operators H are Hermitian while E; = - E -ra and the constants N are real.
This enables one to deduce generally that N _" _ p = - N "p and , by considering
the double commutator [ [Ea, Ell]' u,.; ] that Nail = Nil. -a -II' Using also
the trivial relation Nail = - N Ila we can then determine the constants Nail from
equation (20.31) apart from phase. In the SU 3 example this gives Nail = so J1
that

[E,, +p,E_,,] =- v!(1)E p, etc.


It is soon verified that all these commutation relations agree with those given in
equations (11.3) with the correspondences
20 .5 A Miscellany 519

In most of this section we have discussed the infinitesimal operators X, or


the more convenient linear combinations Hi and E... The finite group elements
are generated as in section 7.2 as exponentials of real linear combinations of
the X p , although it is possible to define what is known as the complex extension
of a Lie group by allowing complex coefficients of the X p ' In a unitary
representation the X, must be skew-Hermitian, see section 7.2, so that from
each Hermitian operator Hi one may construct a corresponding infinitesimal
operator by including a factor .J -1. To each pair E± .., one has a pair ofskew-
Hermitian infinitesimal operators i(E" + E _,,) and (E" - E _,,). Note that E" is
not strictly an infinitesimal operator. The simplest example of this is the group
9t 3 where, see chapter 7, the three infinitesimal operators are

X~ = -iJ x = -!-i(J+ +L), Xy = - iJ }. = -!-(1+ -L) and X% = -iJ:


with J: playing the role of the only H operator and the raising and lowering
operators J ± corresponding to E ±..as is appropriate for a group of rank I = 1.
The simple positive roots and Dynkin diagrams may also be used to deduce
the irreducible representations of the Lie groups, generalising the method that
we have used in subsection 7.4.2 and section 18.8.

20.5 The rotation matrices

The irreducible representation matrices DU~


mm
(R) for an element R of the group
9t 3 were essentially defined in chapter 7 through the expressions (7.39) and
(7.40) for the matrix elements of the infinitesimal operators J and the
expression
T(R) = expa .X = exp( -ia.J) (20.35)
which relates the finite rotation operator to the infinitesimal operators. Here
the rotation is parameterised by the vector a as in section 7.4. One often refers
to the DU~
mm
(R) as the rotation matrices. Although we have used the properties
of these matrices from time to time we have not tried to calculate them as
functions of the parameters a except for the special case of j = !- in subsection
18.13.3. As this example might suggest , the use of the parameters a is not very
convenient and it is usual to parameterise the rotation by what are known as
the three Euler angles (x , fJ and y. (The Euler angles (x, fJ and y should not be
confused with the matrix elements (X and fJ of the SU 2 matrix (18.32)for which
the same notation was unfortunately used.) To explain this new para-
meterisation we note that an arbitrary rotation is defined by the position of a
'new' set ofaxes x', y' and z', obtained by rotating the original set x, y and z. To
do this we introduce the usual polar angles (like () and 4J for a point r) for the z'-
axis relative to the original set and call them, conventionally, fJ and (X with the
usual range 0 ~ fJ ~ 7t and 0 ~ (X ~ 27t. In other words fJ is the angle between the
z- and z'-axes and (X the angle between the x-axis and the projection of the z'-
axis on to the xy-plane, see figure 20.2. It remains to define the position of the
520 A Miscellany 20 .5
z

Figure 20.2

x' -axis (and hence automatically the y' -axis) and this is done by giving the angle
y between the planes O zz' and O z'x', as shown in the figure. This arbitrary
rotation is con veniently expressed as the product of three simple rotations
about the original axes. We first note that if y = 0 the rotation is achieved by
the product R.( ex )R}"(pj. Here the fJ-rotation carries the x-axis to a latitude of
fJ-south while the ex-ro ta tion gives it a longitude of ex, in ter rest rial notation.
(The order of the se rotat ions is important since otherwise the fJ-rotat ion
would have to be about the axis obtained from O y by the rotation Rz(ex). For
the general case y #- O we mu st first carry out a v-rotatio n about the z-axis
giving the general rotation as
(20.36)
(We could equally well have used afinal rotation of y about the z'-axis but we
sha ll soon see the advantage of having all rotations in (20.36) relating to the
same, original set of axes.)
From the factorisation (20.36) the representation operator (20.35) becomes
T( R) = exp( - icxJ z ) exp( - if3J}.) exp( - iyJ z )
and we recall the definition of the rotation matrix, using the notation Jjm for >
the basis vectors
D~!m(ex, 13, y) = ( j m' IT(R) Urn >
Hence , using the fact that J z is diagonal in this basis , we have the immediate
simplification
D~!m (ex. 13, y) = exp[ - i(ex m' + ym)] G m' Iexp( - if3J},) Ij m > (20.37)
20.5 A Miscellany 521

The remammg matrix element may be found using the relation


J, = (J + ....:. J _ )/2i and the formula (7.40) for the matrix elements of J ±. As a
first step in deriving the general matrix element we first deduce the matrix for
j = t. In this particular case the matrix of J, was given in equation (8.15)

Jy = 1(0
2 i '-Oi)
so that J; = 1and, as in subsection 18.13.3, the series for exp ( - i{3J y) simplifies
2
exp( - ifJJy) = (1 -lfJ + . ..) - iJ y(fJ - z'4fJ3 + .. .) = cos tfJ - 2iJ y sin tfJ
The complete matrix is therefore given as

OdJ( fJ ) = (ex p[ -ti(tx+y)]costfJ - exp[ t i(y - tx) ]sin tfJ)


a, ,y exp[ti(tx- y)]sintfJ exp[ ti(tx + y)] cos t{3
(20.38)
By equating this to the matrix (18.37) we can relate the Euler angles to the
previous parameters a. Hence we can show that the weight function in the
group integration, see appendix 4.3, is given by pt«, fJ, y) = sin fJ.
The general matrix element may be deduced from equation (20.38) taking
Iii > = IH >1 IH >2 . . ·1H >2j as the product of 2j factors, like the wave
function of2jparticles with spinj = t and m = t. Then Ijm hasj + mspins up >
and j - m spins down and will be a sum over the different ways of assigning the
spins to the particles. The number of ways is given by the binomial coefficient
(j ~im)' Including the normalisation constants and using equation (20.37) gives

.
D~!m(tx ,fJ, y)=exp[-i(txm '+ym)] ( . 2j ,
J-m
)-!( )-! . 2j
J-m

where m j denotes the spin projection ± t of particle i and the sum is restricted
by the constraints L mj = m, L mi = m. Because of the symmetry of the
i i
operator and the wave functions we may include only one term on the right
and multiply by the number of terms. Thus

D~!m(tx,fJ, y)=exp[ -i(txm'+ ym)]( . 2j


J-m
,)-1(.t r
2j ) 1
m
j+m j- m
x L (m'lm 2 " . lexp ( - ifJJy) 1 ~ ~>
m;
(20.39)
where the firstj + m particles on the right have m = + t . Since the operator is a
product of corresponding operators for each particle, each matrix element in
522 A Miscellany

(20.39) is a product of 2j matrix elements of exp( - if3J y ) taken from the DOl
matrix (20.38). For each particle whose spin projection is the same on left and
right there is a factor cos tfl, while for those which differ there is a factor
±sin 1/3 corresponding to the two cases <-I 1+)
and re- <+I 1-),
spectively. The only relevant parameter in the sum over m; is the number x of
particles in the first j + m which have spin + t. Thus, including the binomial
coefficients e:m)and CJ,;;,"'-x) to account for the number of ways of arranging
the spins in this way gives finally

)-lG, 2j)1 L(j+m)(j-


D~)m(C( ,p, y)=exp[-i(C(m'+ym)] ( , 2j
J-m -m ., x m-x
I

J
. + m; )

x ( - l)j ~ m ' - .' (cos 1P) 2 ., - m - m' (sin tf3f j - 2 ., ~ m + m ' (20.40)

a result which simplifies a little when the explicit forms are inserted for the
binomial coefficients.

Bibliography

Some more discussion of dynamical groups may be found in Englefield


(1972), in the bibliography to chapter 19, and in

Wybourne, B. G . (1974). Classical Groupsfor Physicists (Wiley-Interscience,


New York)

The Jahn - Teller effect is described in some detail in

Sturge, M. D. (1967). Jahn- Teller effects in solids, in Solid State Physics, vol. 20
(Academic Press, New York) and for an experimental verification of the
effect , see

Dixon, R. N. (1971). Malec. Phys., 20, 113

Spontaneous symmetry breaking is reviewed by

Guralnik, G . S., Hagen, C. R. and Kibble, T . W. B. (1968). Broken symmetries


and the Goldstone theorem, in Advances in Particle Physics, vol. 2
(Interscience, New York)

and· although they concentrate on elementary particle theory, there is a


section on non-relativistic systems. See also the review by

Beg, M. A. B. and Sirlin, A. (1974). Gauge theories of weak interactions


A. Rev. nucl. Sci., 24
A Miscellany 523

A more rigorous account of the use of normal subgroups in constructing


irreducible representations is given in Lomont (1959),in the bibliography to
chapter 15.
The structure of Lie groups and their classification is described with
mathematical rigour in

Gilmore, R. (1974). Lie Groups, Lie Algebras and some of their Applications
(Wiley-Interscience, New York)

For further reading on the representations of Lie groups see Boerner


(1963),in the bibliography to chapter 4; Littlewood (1950) in the chapter 17
bibliography; and Wybourne (1974) above.
The bibliography to chapter 7, especially Brink and Satchler (1968), may
be consulted for further reading on the subject of rotation matrices.

Problems

20.1 Show that D 3 is the semi-direct product group C 3 /\ C 2 . Hence deduce the
character table for D3 by following the method of section 20.3.
20.2 Draw the root diagram for the group G2 introduced in section 20.4.
Appendix 3*: Topics
in Representation Theory

Some useful but more advanced properties of group representations are


collected here rather than in chapter 4 of volume 1.

A.3.1 Symmetrised products of representations

The direct product TI~) x TIP) of two representations of a group t§ was defined
in section 4.17 and a simple formula (4.43) was deduced for its character. This
would extend naturally to a product T = T(a) X TIP) X •. • of n terms and in the
present section we are concerned with the special case when all the factors are
equivalent o: = f3 = . . .. To discuss this representation it is helpful to
construct a set of basis vectors from products of functions, as in section 4.17.
Let <Pi with i = 1, 2, . . ., s" denote a set of functions transforming like the
irreducible representation T(aJ so that the set of (s,,)n products <pj(l) <pj(2) . . .
<pdn) provide a basis for T

TcjJj(l)<pj(2) .. . rh (n) = L Tli~rr<~.l. ·· TlN<p;-(l)<Pr(2) . . . <pIt·(n)


i'.j', .. ..k' ))

• Appendices 1 and 2 appear at the end of volume I.

524
A.3 .1 Appendix 525

The index p in the factors 4>i(P) serves to label the factors and is relevant in the
definition of scalar product, namely
(4)d1)4> i ' (2) . , . 4>k' (n), 4>i (1)4> j(2) , , , 4>dn) )
= (4);-(1), 4>i (1))( 4>/ (2), 4> j(2)) , . , (4)dn), 4>k (n))
= ()j"i()j'j '" (l,.'k
showing that the set of products is orthonormal given that the 4>i are
orthonormal. (In physical applications, p will label the different particles of a
system while i labels the possible states of any of the particles .) It is convenient
to introduce a more concise notation, writing
lij , , , k) == 4>i(l)4>j(2) , . . 4>k(n)
where the index p has been dropped but is inferred by the ordering of the labels
in Iij . . . k ). The scalar product is then written as <i'] ... k'iij ... k) . (This
notation is familiar in quantum mechanics .)
Let us now consider permutations P of the n indices p, defined for example
by
(A.3.1)
The set of all such permutations form the symmetric group Y n described in
chapter 17 and it is clear from equation (A.3.1) that the vector space L of the
products Iij . .. k) will be invariant under the group Y n, In fact, the
permutations commute with the operations of '§ so that L provides a basis for
a representation W of the product group '§ x Y n. Ifwe denote the irreducible
representations of the groups '§, Y nand '§ x Y n by T(Y), VIA) and T(Y) x VIA),
respectively, then the reduction of the representation W may be written as

(A.3.2)

Thus it is possible to choose a basis for L in which the basis vectors are labelled
by y and )., describing their transformation properties with respect to both '§
and Y n. If G a, P and GaP denote arbitrary elements of the groups '§, Y n and
'§ x Y n, then the character equation corresponding to the reduction (A.3.2) is

x(GaP) = Lmy).X(Y)(Ga)X( A)(P) (A.3.3)


Y. i.

using the result (4.66).


In practice, we are often interested in one of the subspaces in the reduction
L = 'LL . where L , contains all the basis vectors of L which transform
A ", A

according to a particular irreducible representation V(i.) of Y n, This space L .


provides a representation T A of '§ which from equation (A.3.2) reduces into"
irreducible representations of '§ according to

T I.· = '\'
L... mYI.· T (y)
Y
526 Appendix A.3.1

and its character is therefore given by

xJG a ) = L my;.x(Y)(G a)
Y

By using the orthogonality (4.25a) of characters of 5I'n this is given from


equation (A.3.3) by

(A.3.4)

where the sum runs over the n! permutations P of 5I'n . The procedure for
finding X;. will be first to calculate X(GaP) and then to use equation (A.3.4)with
the known X()·) from the character table for the symmetric group. Once x·I . is
found, the integers my). follow from the usual orthogonality of characters of CfJ

(A.3 .5)

where 9 is the number of elements in CfJ.


We now calculate the X(GaP) for a few simple permutations and leave the
deduction of a general formula until the end of the section. Consider first the
identity P = E which, from the definition of a character, gives

X(GaE) = L (ij .. . kIT(Ga)!U .. . k)


t.j. .. ., k

L T~)(Ga)Tf}(Ga)'" 71,(G a)
i.], .. ., k

= {L T:~)(GaW = V~)(GaW
(A.3.6)
i

thus expressing X( GaE) in terms of the known irreducible characters of CfJ. In a


similar way, for the permutation P 12 we have, using equation (A.3.1),

X(GaP I 2 ) = L (ij ... kIT(Ga)Ui . . . k)


i.] , . . .,k

L
i,j• . .., k
nY(Ga)T~~/(Ga) '" Tl~)(Ga)

= L T~)(G;) . . . Tl':!(G a )
i. .. ., k

(A.3.7)

which involves just the irreducible characters again, since G; is simply another
element of CfJ. For the more complicated permutation (g~) of three objects, in
A.3.2 Appendix 527

the notation of section 2.2(10), a similar argument shows that


x(G a(1 ~ ~)) = x<a)(G~ ){ia) (Ga )}n - 3 (A.3.8)

With these results we are in a position to deduce all the X.(GA a ) for the
examples n = 2 and n = 3. Firstly for n = 2 there are only two group elements
E and P 12 and there are two irreducible representations corresponding to
symmetric or ant isymmetric basis vectors. From the characters given in table
4.5 and equations (A. 3.4) to (A.3 .7) we have

Xsym(Ga) = H { XI:t) (G a) }2 + l '}(G ; )]


and (A.3 .9)

Xantisym(Ga) = H {XI:t )(G a) }2 - XlJ)(G;)]

For n = 3 we may take the characters from table 4.2 because of the
isomorphism between !/3 and D3 (see section 2.3). There are three classes,
given by E; P 12 ' P 23 , P 3 1 ; n~n, Og)and the three irreducible representations
are usually described as (I) totally symmetric, (2) totally antisymmetric, and (3)
mixed symmetry (two-dimensional). In this case, we find as before

Xsym(Ga) = H {X I:t) (G a l }3 + 3XlJ)(G;) XlJl(Ga ) + 2XlJ )(G ~)]


Xantisym(Ga ) = H { XlJ )(G a ) }3 - 3XlJl(G;) XlJ)(G a ) + 2 XlJ )(G~)](A.3.10)
Xmixed(G a ) = H2 { XlJl(Ga ) }3 - 2X I:t) ( G ~ )]

Notice that there has been no need to specify the group "§ although, finally, to
obtain numbers from these expressions, one must look up the values of the
irreducible characters X(a) of «.
It is possible to obtain a general formula for the characters X(GaP) by using
the result , discussed in section 17.1, that any permutation may be written as a
product of 'cycles' of various lengths. If P contains n t cycles of length 1, n2
cycles oflength 2, and so on , then by a continuation of the method used for the
simple permutations above, one finds that

X(GaP) = Il {X(a)(G~)}n, (A.3.11)


i

A.3.2 The use of a subgroup in reducing product


representations
The way in which a product of two representations of a group "§ reduces into
irreducible pieces was described in section 4.17 and in particular the integer
coefficients m.,. in the reduction (4.44)
528 Appendix A.3 .2

were deduced from equation (4.45) using the tables of characters. There is an
alternative way of deducing these coefficients if the corresponding coefficients
are known for a subgroup Jf of ~ . Let us denote the irreducible repre-
sentations of Jf by T(liJ, T(P), etc., and suppose that we know the coefficients
m;P in the reduction

T(Ii)X TIP) = tm;PT(ji) (A.3.l2)


y
and also that we know the coefficients n~a for the reduction (4.49) of T!J) on
restriction from ~ to its subgroup Jf,

T!J) = 'n
L... ~~· T (Ii)
Ii
Combining these three equations we have

T(') X T(P) = I n, anp'pT( a) x T IP)


a:p. y

(A.3.l3)

and also

T(~) x TIP) = L myT()')


v

(A.3.14)

so that, equating the right-hand sides of equations (A.3.13) and (A.3.14), we


have for each y

(A.3.15)

The only unknowns in this set of equations (A.3.l5) are the required
coefficients my and, in many cases, the equations are sufficient to enable us to
deduce them.
To illustrate the method, we use the same example as in section 4.17, the
product T(3) x T(3) for the group D3 , using now the subgroup C3 as described
in section 4.l8. Thus equation (A.3.l3) becomes
T(3)X T (3) = (,12) + ,(3))X (,(2) + ,(3))
= 2,(1) +,(2) +,( 3)
Comparing with the known reductions T (3) = , (2) +,(3), T(1) = T(2) = ,(1)
from section 4.18 we are able to deduce that
A.3 .3 Appendix 529

where m, + m2 = 2. This example also illustrates the limitations of the method,


i.e. it does not always provide a complete solution. It cannot in this example tell
us that m, = m2 = I but only that m, + m2 = 2. Nevertheless, the method is
increasingly useful in the continuous groups where the characters of the larger
group ~ are more difficult to handle than those of the subgroup .tf.

A.3.3 Class multiplication

The methods used in section 4.15 for finding the irreducible characters were
satisfactory for groups where the number of classes was small. In this section
we derive additional relations between the characters which are sufficient to
determine them for any finite group. We first introduce the sum of all group
elements in a class rej which we denote by the same symbol rei and which is
called a 'class operator':

re =j L
o, in -e,
Ga (A.3.l6)

We next show that the product of two such sums must itself contain all the
elements of a given class the same number of times, so that we can write

reirej = L GaG b = LCjjkrek (A.3.17)


G. in ",. k
G. in '6 j

To prove this result, write


rejrej = nbGb + ncGc + . . . (A.3.l8)
where G, and G, are two elements in the same class rej. For any element G a in
the group we can write
(A.3.19)
since for any Gin C(jj the element GaGG a- 1 is also in rej and is different for each
choice of G . It therefore follows that

rejrej = GareiGa-'GarejG; t = GareirejGa-t


= nbGaGbGa-' + ncGaGcGa- 1 + (A.3.20)

By choosing G, such that GaGbG; 1 = G, and comparing (A.3.20) with


(A.3.18) we see that nb = nc which justifies the expression (A.3.l7).
We now choose some irreducible representation T (') of ~ and construct the
matrix sums

T.I
(.j
L T(')(G a ) (A.3.21)
G. in 't,

Equation (A.3.19) is now written in the form G, C(jj = rejG, which, in the matrix
530 Appendix A.3 .3

representation T(lll, becomes T(")(Gg)T:"' = TI")T(")(Gg). Since this equation


is satisfied for all G, in l§ it follows from Schur's first lemma that TI") is a
constant times the unit matrix, T(ll) = A. ( II) 1.
Equating traces in equation (A~3.21)/now gives .~J(ll)
I
= CiX(lll, so that
I

(A.3.22)

where Cj is the number of elements in the class ~i ' Writing (A.3.l7) in matrix
form, we have

T (ll)T (ll) = "c'.. T(lll


i j 1... I )k k
k

and using (A.3.22) this gives a relation between characters,

(A.3.23)

We note that the coefficients C i j k appearing in equation (A.3.23) are known


from the group multiplication table via equation (A.3.l7). Thus we have
derived a set of equations (A.3.23) for all i and j which can be used in deducing
the characters.
Appendix 4: Some Results
Pertaining to the Group f!l3

The first two sections in this appendix contain some useful properties of the
spherical harmonics, which follow directly from our study of the group 1Ji 3 in
chapter 7. The last section discusses group integration generally and then
deduces the necessary weight function for the group integrals in 1Ji 3 •

A.4.1 An integral over three spherical harmonics

The Wigner - Eckart theorem may be used to evaluate a useful integral. From
equation (7.53), regarding yOtl(O,
m,
¢) as an operator,

f:~o 1=0 y~);(O, ¢)Y~',)(O, ¢) Y~~)(O, ¢ )sinOdOd¢


= C(11121,mlm2m) <IIIYU.)ljl2> (A.4.1)

using the fact that IJi3 is simply reducible so that t is not required and also that
the Clebsch-Gordan coefficients C are real. This result may be used to expand
the product 1"'.)
m
1"1,),
m
regarded as a single
.
function, in the form of equation
l l

531
532 Appendix A.4 .2

(7.49). We have

y~: ,) (0, cP) y~:2 ) (0, cP ) = LC(l1/2 1,m lm 2m) (/IIY(/,)11/2 >y~) (O,cP) (A.4.2)
1m

making use of equation (7.50) and (A.4.1). To find the value of the reduced
matrix element we first invert equation (A.4.2) using the orthogonality of the
Clebsch-Gordan coefficients, see section 4.17,

(A.4.3)

Then, since the reduced matrix element (IIIY(/,)11/ 2 >is independent of 0 and
cP, we may insert any convenient values for 0 and cP in equation (A.4.3). By
taking 0 = 0, the spherical harmonics simplify enormously, becoming inde-
pendent of cP,

yl~(O,cP) =
21+
om.O( ~ 1)1Pk(cosO) = (21~
+ 1)1
om.O (A.4.4)

giving

( /IIy(I,)1112 >= ((2/ 1 + 1)(2/2 + 1))1C(1 I 1000)


4n(21 + 1) I 2 ,
(A.4.5)

The integral (A.4.1) is finally given by

1~0 1=0y~). (0, cP) y~:)(O, cP) y~~)(O, cP )sinOdOdcP

= ((2/ 1 + 1)(2/2 +
4n(21 + 1)
1))tC(I I I OOO)C(I I I m m m)
I 2 , I 2 , I 2 (A.4.6)

It is interesting to note that th is integral vanishes unless II, 12 and I satisfy


triangular conditions, see subsection 7.4.4. It also vanishes unless II + 12 + I is
even, because of parity.

A.4.2 The spherical harmonic addition theorem

Consider the products of spherical harmonics of the coordinates of two


particles Y(/ ,)(Ol cPdy(12)(02cP 2)' In the way described in section 4.17 and
subsection 7~4.4, usingmequation (4.46) it is possible to form an invariant by
summing such products when II = 12 , for then it is possible to have J = 0 in
equation (7.44). The invariant may be written as

f o
lO) = "C(I
L. 10m I -m I O)y(l,)(O
1 I , m. I
cP I )y(/
- m,,) (0 2 cP 2 ) (A.4.7)
ml
A.4.3 Appendix 533

making use of the fact that, for invariance under the subgroup ~ 2 , necessarily
m, + m2 = 0. Notice that althoughf~O) is an invariant under rotations of the
system it will depend on the relative positions of the two particles. In fact we
shall now show that f6°) is a Legendre polynomial of the angle ()'2 between the
radius vectors to the two particles. To show this we note that since fJO) is

°
invariant it is unaffected by a rotation of the system which brings particle 1 on
to the z-axis. In this position, (), = and ()2 = ()'2 so that using equation
(A.4.4) the equation (A.4.7) reduces to

f (0) (2/, +
o =C(l,/10,OOO) ~
1) PI,(cOS(}12)

where we have also used the fact that when m = 0, the spherical harmonic
reduces to a Legendre polynomial. The Clebsch-Gordan coefficients in
equation (A.4.7) have a simple form (see problem 7.14)
C(l,/, O,m, -m, 0) = (-I) /, -m' /(2/, + I)!
giving finally for equation (A.4.7)

~(_1)my~)(o, cj>d~(~(02cj>2) = C~: 1 )p/(COSO ul (A.4.8)

This result is known as the spherical harmonic addition theorem and is used
for example in analysing the interaction potential between two electrons into
factors which depend on the coordinates of the separate electrons, see
appendix 5.1.

A.4.3 Group integration

In section 7.1 we remarked that the group sum f


a= I
,which occurred in the
orthogonality relations for irreducible representations of a finite group, would
be replaced by an integral over the parameters ai' a z, .. ., ar with some weight
function p(a" a2' . . ., arl o In the particular case of the group f?(2 we saw in
equation (7.14) that the orthogonality was given with a weight function
pta) = 1. To deduce the weight function for the group f?(3 it is necessary to
study more carefully the properties which it must possess . In this section we
first consider the general problem and then deduce the weight function for f?(3 '
This enables us to write down the orthogonality relations for the irreducible
representations and characters of 91 3 ,
The essential property of the finite sum

(A.4.9)

wherefis some function depending on the group element G a , is that the sum is
534 Appendix A.4 .3

unchanged if, throughout the sum, Go is replaced by the element G, = GoG b


for some fixed G b • This property was used repeatedly in chapter 4 and its proof
in section 2.9 is simple. From the properties ofa group, as Go runs through the
elements once then so does G, for fixed G, so that the same terms appear in the
sum (AA.9) but in a different order. It is this property which needs to be
explored for continuous groups, namely that p(a) must be chosen to ensure
that, for any function j{a) depending on the group elements G(a) ,

f f(a)p(a)da = f f(c)p(a)da
(AA.lO)
a in <;I a in '§

where c denotes the parameters of the group element


G(c) = G(a)G(b) (AA.11)

for any fixed b. Here we use the notation a for the set of parameters
ai' a 2 , • • •, a" while da is an abbreviation for the volume element da
= da l da2 .. . dar . Now trivially

a in
f '§
f(a)p(a)da = f
c in '§
f(c)p(c)dc
(AA.12)

but if we use equation (AA.II) for fixed b then c is a function of the parameters
a and b, so that introducing the Jacobian
eC I GCI
- -
GC O'!I Ga2 (AA.13)
Ga
OCr OCr
oa I oar
we may change the integration variable from c to a. Notice that if b is fixed and
c runs over the group then a also runs over the group so that equation (AA.12)
becomes

f
a in '§
f(a)p(a)da = f
Gin ~
f(c)p(c) ~: da
Comparing this with the required equation (AA.10) shows that the weight
function p must satisfy the condition
GC
p(c) oa = p(a) (AA.14)

for all a. Assuming that such a p can be found, we may choose any convenient
value of a for evaluating p(c) and by taking a = 0 we have

p(c) = p(O) IG: )a =o (AA.15)


A.4 .3 Appendix 535

The constant p(O) is unimportant and is generally taken as p(O) = 1. To show


that the choice (A.4.15) is sufficient to ensure that equation (A.4.l4) is satisfied
is left as an exercise.
In the finite groups the group sum invariably entered in the form
1
- L f(G a )
9
(A.4.16)
ga =1
The division by g, the number of group elements, ensures that iff is the con-
sta nt functionj'(Gj ) = I for all a then the sum also take s the value unity . For a
continuous group the number 9 must be replaced by the group volume defined
as

v= f p(a)da
a in <§
It is then clear that the integral

~ f
a in '§
f(a)p(a)da (A.4.17)

will take the value un ity iff is the constant functionf(a) = 1. Both definitions
(A.4.16) and (A.4.17) may therefore be regarded as averaging procedures over
the group. It may now be seen that the proofs of chapter.4 are equally valid if
the finite average (A.4.l6) is replaced by the continuous average (A.4.17). Of
course, this assumes that the integral (A.4.17), or group volume, is finite. Such
groups, which include the rotation and unitary groups, are called 'compact'.
Although mo st of the groups of physical interest fall into this category, there
are some which do not, such as the Lorentz and translation groups discussed in
chapter 15.
For the group fit 2 the functional relation between the parameters a. band c
of equation (A.4.11) is simply c = a + b so that trivially ocloa = 1.The volume
is then gnda = 2n, and this justifies the orthogonality relation (7.14) as a
specia l case of equation (4.25).
In the group fit 3 the relation between c, a and b is not so simple and we make
use of the vector representation of these rotations. Let us choose the z-axis in
the direction of the axis of rotation of R(b) which gives to R(b) the simple
matrix

R(b) =
COS b

~nb
-sinb
cosb 0
0)
(
o 1

For R(a) we must then take an arbitrary direction but since the Jacobean is
evaluated as a -> 0 we may treat the rotation angle as small and for its matrix
536 Appendix AA.3

-n
use the unit operator plus the infinitesimal operator from equation (7.24)

R(a) ~ ( _:: -::

Hence the matrix of the product is


COS b - a, sin b - sin b - Gz cos b Y)
R(c) = R(a)R(b)::::: Gzcosb+sin~ - Gz sin b + cos b -G
Gx
(
-Gycosb+a,smb Gy.sinb+Gxcosb I

(A.4.18)
Having found the matrix of R(c) it remains to deduce the parameters c
from the matrix. The angle c is found by equating the trace
of the matrix R( c) to the known trace (I + 2 cos c) for a rotation through angle
c, giving cos c = cos b - G z sin b which, remembering that G z is small, implies
c = b + Gz • To find the direction of c we use the fact that the vector c will be
unchanged by the rotation R(c),
R(c)c = c (A.4.19)
We must take care in using equation (A.4.19) since the form (A.4.18) for the
matrix of R(c) is good only to first order in a. However from equation (A.4.19)
and the unitarity of the rotations we have also R t(c)c = C so that
{R(c) - R \c)}c = 0, which gives explicitly the three equations

-2(sinb+azcosb)cy+[a y.(1 +cosb)-a,sinb] cz = 0


2(sinb+azcosb)cx-[Gx(l +cosb)+aysinb] cz = 0
[-G/l +cosb)+a,sinb]cx+[aysinb+ajl +cosb)]c y = 0
Hence the components c" c y, c, of c are in the ratio
{aysinb +Gx(l +cosb)} :{ -a, sinb + Gy(l +cosb)} : 2{sin b +azcosb },
wh ich to leading order is the same as
1{Gy+Gjl +cosb) /sinb} :1{ -ax+ay(l +cosb) /sinb }: I
Taking into account the magnitude c = b + Gz deduced earlier we have finally
and again to first order in a,
c, = -!-b{ Gy. + GA 1 +cosb) /sin b}, cy = -!-b{ - a, + Gy(l +cosb) /sinb},
Cz = b +Gz
The Jacobian is now evaluated directly

DC
-!-b(l +cosb) /sinb -!-b o
Da
--!-b -!-b(l +cosb) /sinb o =-!-b 2(1+cosb) /sin 2b
o o 1
A.4 ,3 Appendix 537

In the limit a -+ 0 we may write b = c so that the required weight function is


2 sin 2 c 2(1 - cos c) A 0
( )
pc = 2 ( .4.2 )
c (l +cosc) c2

The group volume defined earlier is then

V = f 2(1 - cos a) d
a
2 a, d a yd a:
a in '§
Transforming to polar coordinates for the vector a, and integrating over angle,
gives

v= r" 2(1-~Osa) a2daffdO= 8n2


Ja= a 0

The group average (A.4.17) is then

4~2 if(a)(I-COSa)dadO (A.4.21)

For a functionf(a) which is independent of the angle of a, and this would be


true for any class function like the character, the average becomes

1r"
~l~~)(I-COSa)da (A.4.22)

Using this average to replace the finite sum in equation (4.25a) then gives the
orthogonality relation for characters of 9€ 3,

.~Jar"
t: = 0
XUtlXU' )(1 - cosa)da = {) .
a a }il : (A.4.23)

a result which may be verified from the explicit formula (7.42) for XU). It
should however be noticed that the orthogonality is not valid for the aha lf_
integer representations which are double-valued. Orthogonality for these
double-valued representations may be recovered by extending the range of the
rotation angle to 2n. Over this range the representation is single-valued.
Further discussion of this enlargement of the group 9€ 3 is given in section 7.6
and in section 18.13.3.
If the Euler angles iX , f3 and y are used as parameters for the rotation instead
of aI' a 2 and a3 then it may be shown, see section 20,S,that the weight function
is p( iX, f3, y) = sin f3 so that the average becomes

8n
12 1 r" 1
2
"
.= o Jp= o
2

=/(iX,f3,y)sin f3diXdf3d y
" (A.4.24)

The orthogonality relation for matrix elements of the irreducible repre-


538 Appendix A.4 .3

sentations then takes the form

1 1"1
2

-8 12 2" " u )* u) .
Dm!m (ex. ,fJ, y)Dm~m (ex. ,f3, y)smfJdex.dfJdy
n 0 0 0 " , ,

= OJ,j,Om',m', om,m,(2j\ + 1) - 1 (A.4.25)


using equation (4.23).Again, for half-integer j the range of the parameters must
be doubled and this may be achieved in a variety of ways, for example by
allowing ex. to extend up to 4n .
Appendix 5: Techniques
in Atomic Structure
Calculations

Some detailed calculations of the ordering of atomic energy levels for the
configura tions p2 a nd pJ are given in the first section, to reinforce the more
qu alitative discussion in chapter 8. The second section returns to the theory of
a ngular momentum and the gro up Yi J and introduces the more intricate
problems encountered when more than two angular momenta are coupled
together. Some of these ideas are used in section A.5.3 to calculate atomic
transition strengths. Some more details of the effects of symm etr y in a crystal
field, which was introduced in section 9.9 of the first volume , are given in the
last two sections.

A.5.1 Term energies for p? and p3 configurations

Two valence p-electrons


We first deduce the energies of the three terms J P, IS and I D of two electrons
in a shell with I = I (see subsection 8.6.4) by using the simplest method, that of
539
540 Appendix A.5.1

the 'diagonal sum'. A more sophisticated method which requires the know-
ledge of tabulated coefficients is described in section A.S.2. We make use of the
fact that since the Coulomb repul sion e Z [ r i z is invariant under rotation of all
particles the energies are independent of M L and diagonal in L . Thus we may
select any convenient value of M L' The matrix elements of the Coulomb
repulsion are most easily calculated in the product basis m l mz rather than the
coupled basis L M L' Furthermore we know that the trace of a matrix is
independent of the basis . Let us denote the required term energies by ELand
the diagonal matrix elements in the product basis by Eim, mz). Then
considering separately the matrix for each value of M we are able to deduce a
set of equations relating the E L to the E(mlmz).
Taking first the maximum M = 2 for which there is only one basis state we
have directly ED = E(11). For M = 1 there are two basis states and we have
E p + ED = E(1 0) + E(O 1). For M = °
there are three basis states giving
E s + E p + ED = E(1 - 1) + E(O 0) + E( - 1 1). It remains only to calculate the
diagonal matrix elements
E(m l mz)

= ffdr l dr z u,;,(r I) u,;,(r z )Y(/)*


m
(0 I <P I) y(I)*iOz
m
<Pz) y(/)(
m
0 1 <P I )y(/m) (Oz <P z )(ez[r IZ)
l 1 l 1

We do this for arbitrary [since no more work is required. The interaction (l /r IZ)

°
is separated into parts depending on each electron by first expanding it in
Legendre polynomials of cos 1 2 , where el z is the angle between r , and rz , and
then using the spherical harmonic addition theorem (A4.8). Finally, the
angular integrals are carried out using equation (AA.6). Thus expanding
ex.
e 2/r l Z = L Vd rl ,rZ)Pk(cos 0 12 )
k=O
and writing for the radial integral

F(k) = f fdddrldr2U';/(rtlu,;,(rl)Vk~rl,r1)
we have
1/
E(mlm 1) = L F(k)C2(kll,OOO)C(kll,Omlml)C(kll,Om1m1) (AS.l)
k (even) = 0

The upper limit on k has come about because of the vector coupling of k, [and l.
The coefficients C vanish otherwise. The restriction to even k results from
consideration of parity in the single-particle integral (A4.6). The vector
coupling coefficients C all take the value unity when k = so that in the case of
[ = 1 we have the simple formula
°
E(m l m1) = F(Ol + F(2 lC 2(211 ,OOO)C(211 , Om l m l )C(211, Om1m 1)
A.5 .1 Appendix 541

Using the known values of the coupling coefficients (see problem 7.8) we can
now deduce E(m\ m2)and hence ELin terms ofthe two radial integrals FlO) and
F(2),

Ep = PO) - !P2), ED = FlO) + l5 F (2l, Es = Fl O) + ~F( 2) (A.S.2)


F rom its definition one sees that the integral F(k) is positive since Vk (r \, r2)has
the explicit form vd r\, r2) = r~ /r~+ 1, where r < and r~ denote the lesser and
greater of r\ and r 2 • Without calculating the value of F(2), which would
demand knowledge of the radial wave functions un,(r), we can see that the term
energies satisfy Hund's rule . In fact we may deduce the ratio of the excitation
energies of the \Sand \ D terms above the ground state 3 P term
(Es-Ep)/(ED-E p) = 2.5, which is to be compared with the observed ratio of
2.1 in the carbon (2 = 6) atom which has two valence electrons in the 2p-shell.
The small discrepancy is due to the admixture of higher states such as that
obtained by placing one valence electron in the 3p-shell. We have, after all,
only been using first-order perturbation theory.

Three valence p -electrons


The energies of the three possible terms 4 S, 2P and 2D, see subsection 8.6.4,
may be calculated by a natural extension of the method given for two particles
using a product basis, m\m 2m 3 , but this method soon becomes laborious as the
number of electrons increases beyond three and a more sophisticated method
known as the 'fractio nal parentage' method is generally used . The basic idea is
simple and we shall use it here .
Any antisymmetric function lJI S L ( I 23) of three particles with the same n/
must of necessity be antisymmetric in particles numbered I and 2. It must
therefore be possible to write
lJI S L( 123) = I Q<l>{ <1>(12)1/1(3) } SLM sM L (A.S.3)
<l>
where <1> runs over all the terms for two valence electrons and the curly bracket
denotes the vector coupling of <1> to the single particle wave function of electron
number 3 to give total S and total L .
We wish to calculate the term energies

F SL = (lJIsdI23)II e2 /rjjllJlsd123))
, < )

but since 'P is antisymmetric, the contribution from each pair of electrons is the
same and the sum is equal to three times the contribution of any particular
pair, for example electrons numbered 1 and 2. Thus using also equation (A.5.3)

FS L = 3 I I 12(<1>(1 2) Ie 2 [r \21 <1>( I 2))


Q<l>
<l>

(A.SA)
542 Appendix A.5 .1

where E<J> denotes the term energies for the two-particle system . In deriving
equation (A.5.4) we have integrated over the coordinates of particle number 3
which is trivial and gives a factor unity from normalisation since the operator
is now independent of particle 3. We have also introduced the coupling
coefficients implied in the expression (A.5.3) and carried out the sum over the
Ms a nd M L values of <1>. These sums again give a factor unity from the
normalisation of the coupling coefficients and the fact that E<I> is independent
of these M-values.
Knowing the two-electron term energies from the previous section one
needs only to deduce the 'fractional parentage coefficients' a<l> to obtain the
three-electron term energies F SL' Some elaborate methods have been de-
veloped for this purpose and extensive tables have been published. Here we
shall deduce the coefficients by an elementary method. From the normali-
sation of 'II the coefficients must be normalised I: Ia<l> IZ = 1 and for the
purpose of using (A.5.4) the phase of a<l> is not rele~ant. The coefficients are
conveniently presented in table A.4-with 'II labelling the rows and <I> the

Table A.4
y- ~ 3p IS
4S i
I
o
--,------------;c--- ----:::- -

2D I ex o
2p I ~ _

columns. The zero entries in the table follow directly from the triangular
conditions on vector coupling, i.e. a total S = t cannot be produced by
coupling the single particle spin of! of electron number 3 to the 'parent' spin
S = 0 and a tot al L = 2 cannot be produced by coupling I = 1 to the 1 S parent.
Use of the normalisation condition completes the first row and gives the
energy of the ground state term as
F(4S) = 3E p = 3F(O) - (3/5)F(Z) (A.5.5a)
using the results (A.5.2).To find the second row we use the result (8.44) noting
that equation (A.5.4)is valid for any invariant two-body interaction and hence
for,Ar in (8.44). Since P~j = - I in the 3 P term and + 1 in the 1 D this gives
3( -!Xz + fJz) = 3 -~ - !( ~) = 0
which combined with the normalisation !XZ + fJz = 1 gives !XZ = fJz = 1- Using
these values for (XZ and fJz we deduce the energy of the zD term,
FeD) = ~Ep +~ED = 3F (O) - z6sF(Z) (A.5.5b)
The same argument applied to the third row gives
3( _ yZ +OZ + £z) = 3 _~_!(~)
A.5 .2 Appendix 543

which together with the normalisation gives y2 = 1, (52 + £2 = 1. To find (52 and
£2 we use the two-body operator L [l(i) + IU)]2 in precise analogy with the
i<j
corresponding spin operator used in deriving equation (8.43). The correspond-
ing expression is

L [I(i) +IU)]2 = L2 - 1(2 - 1)/(1 + 1)


i <i

which when applied to the 2P term, using equation (A.5.4), gives


3(2y2 + 6(52) = 2 + 6. Knowing y2 = 1 this gives (52 = 5/18 , £2 = 4/18 , com-
pleting the third row of the table and giving for the energy of the 2P term
Fep) = ~E p +i ED +~Es = 3FO (A.5.5c)
From the three term energies (A.5.5) we may again deduce the ratio of
excitation energies
Fep)-F(4S) = 1.67
FeD) - F(4S)

which is to be compared with the observed ratio 1.5 in nitrogen (Z = 7) with


three valence 2p-electrons.
This example illustrates the method of fractional parentage which is used for
more complex atoms and also in nuclear physics . More systematic methods for
calculating the coefficients of fractional parentage are ava ilable, making use of
other groups, notably the group U 21 -t 1 of unitary transformations in the
(2/ + 1) dimensional space of the wave functions !/JIm for a single particle and its
various subgroups, see section 18.10.

A.5.2 Recoupling coefficients (6j- and 9j-symbols)

We have seen how the Clebsch - Gordan (or vector-coupling) coefficients play
a vital role in the construction of wave functions and the calculation of matrix
elements. These coefficients are calculated entirely from the group properties of
:ft 3 and are independent of any details of the physical system other than its
symmetry. When more than two angular momenta are coupled together one
naturally encounters products of Clebsch - Gordan coefficients and certain
sums of products occur sufficiently frequently that it is worth while
introducing a symbol to denote them and tabulating numerical values for
them-see Brink a nd Satchler (1968) and Rotenberg et a/. (1959) in the
bibliography to chapter 7, for further details of the recoupling coefficients. It is
clear that unlimited complexity may be achieved by coupling more and more
angular momenta together but, in practice, it is only the first one or two steps
along this path which are of much interest. The processes which we develop in
this section are applicable to most groups but have only been developed
extensively for the group gp 3 . For the Abelian group f:1l 2 these processes are
trivial.
544 Appendix A.5 .2

The product representation ov'>xO U2) reduces simply, as given by equation


(7.44), so that there is a unique basis vector which transforms irreducibly with
givenjm, provided Ijl - j 2! ~j ~ UI + j2)' The Clebsch -Gordan coefficients
CUJ2i, m 1m 2m) are defined as the coefficients for the expansion of this
function in the product basis m lm 2. The more elaborate product repre-
sentation OU'>x OU2)X ov,)x OU.) will not reduce simply and in general a
given irreducible representation O(j) may occur several times in its reduction.
Hence the reduction is not unique and a single label jm will not uniquely
specify the basis functions of the reduced representation. The reduction may
be carried out in two parts, first reducing separately the two products
VI +j2)
ov'>x O(2) = I OUI2)
jl2 = ~, -hi
and
V, +j.)
ov,) x OU. = I O U,.) (A.5.6)
j ,. = Ij , - j. 1
and finally reducing the product

OUI2) x OU,.) = I ov) (A.S.7)


j = Ij l2- h .1

Each of these steps is unique and defines a set of basis vectors which we
denote by U12j34,jm ) . An alternative set of basis vectors may be defined by
taking the products in a different order, for example O(j,)x O(j,) and O(j2)X O(j·)
followed by OUIJ)x DU2.) would define a set IjI3j24 ,jm ) . The transformation
between these two sets is usually written as

JJ12 j34,jm ) = I [(2j12+1)(2h4+1)(2j 13+1)(~i24+1)]1


j lJiz4

(A.5.8)

which defines the 9j-symbol f ::}- Since functions with ditTerentj or mare

orthogonal it follows that the sum (A.5.8)is restricted to fixedj and m and it is
easily shown, using the raising operator J +, that the coefficients are
independent of m. From their method of construction it is clear that any of the
basis vectors on either side of equation (A.S.8) may be written as a sum of
vectors !m lm 2m3m4 ) in the simple product basis with coefficients which are
products of three Clebsch - Gordan coefficients. Hence using the orthogo-
nality of these coefficients, equation (A.5.8) leads to an expression for the 9j-
A. 5.2 Appendix 545

symbol as a sum over products of six Clebsch - Gordan coefficients. For-


tunately, there are tables of 9j-symbols! The reason for extracting the square
root factor in equation (A.5.8) is because the 9j-symbol so defined has
extensive symmetry relations. A reflection in either diagonal leaves it
unchanged, while the exchange of any two rows or columns induces a factor
( - I )P, where p is the sum of all nine j -value s in the symbol.
In an extension of the Wigner - Eckart theorem (7.53) the matrix elements of
a coupled operator in a coupled system are also given in terms of a 9j-symbol.
Consider a system with two components, like two particles, or the orbital and
spin angular momenta of a single particle. We distinguish the components by
indices I and 2 and write a coupled wave function as JJlj2jm >. We take
irreducible sets R(k,) and S(k ,) of operators (see section 7.4.6) in each
component and cduple them'':« form an irreducible set with respect to
simultaneous rotations

T(K) = L C(k 1k 2K ,QI Q2Q )R (k,)S(k, ) (A.5.9)


Q «,«, q, q,

R q(k,, Iis a un it operator so far as the second component is concerned and vice
versa . Using the Wigner - Eckart theorem and the definition (A.5.8) of the 9j-
symbol one then finds

(j l j 2j m I T( ~ ) Ij't j~j' m' >=( _1) 2K C (f Kj ,m'Qm) (jd2j IIT(K)IIj'lj~/ >


= (- I) 2K CU'Kj, m'Qm)[ (2/ + I )(2K + 1)
x (2jI + I)(2j2 + I)]!

x {~': ~~ ~'}<.i1.i21IR(k')S(k'III/lj~ >


k, k2 K (A.5.IO)
This formula may be derived by inserting all the Clebsch - Gordan coefficients
which appear on the left of equation (A.5.1O) when the various couplings are
uncoupled and the Wigner - Eckart theorem is used in each component. The
resulting sum is then seen to be identical with that appearing in equation
(A.5.8). The result should not be unexpected since the separation of the two
components in (A.5.10) has much in commo n with the transformation (A.5.8)
using the correspondence

Ut.i2h.i4jt 2h4.i1J.i24.i) -+ Ud2/tj;j/k 1 k 2 K)


The importance of equation (A.5.l0) is that the dependence of the matrix
element on K,j and/ is given explicitly in terms of known functions and does
not depend on the detailed definition of the operators. If, in the product basis
1.i1.i2ml m 2 >, the wave function separates then so does the reduced matrix
element
(A.5.II)
546 Appendix A.5 ,2

There are several special cases of these results which are in fact the ones most
frequently encountered, When one of the entries in a 9j-symbol vanishes two
other pairs of entries must be the same because of triangular conditions,
leaving only six independent j- values . In this case one writes

e}
L eA
Q b (_l)b +c+e+ f ab e
c d e = (A.5.l2)
{
f f 0 [(2,+ 1)(2! + 1)]1

and refers to {: : : } as a 6j-symbol. Again the reason for extracting the


factor in the definition (A. 5.12) is because this leads to greater symmetry in the
6j-symbol. It is unchanged by the exchange of any pair of columns or by the
exchange of any two numbers in the top row with those immediately below
them in the bottom row . Because of the symmetry of the 9j-symbol there is an
expression similar to (A.5.l2) for any 9j-symbol with a zero in any position.
By setting i, = 0, the equation (A.5.8) gives the transformation between the
two sets of irreducible basis functions for a system of three j-values

IjI2h,jm> = L (-IY12+jIJ+j,+},[(2jI2 + l)(2j1 3 + 1)]1 1

jIJ

(A.5.l3)

The result (A.5.10) may be specialised in two way. Fir st, by taking S(k,) = I,
. q,
the unit operator, so that k 2 = 0, we 0 b tam

<ilhjmIR~,')Ij'lj~ j' m' >= ( - I )j, +j, +j' - k, eU' k .t.m' ql m)oj,/"

in which the dependence on the total angular momentaj and j' is given entirely
by known functions .
As a second specialisation, set K = 0 so that k I = k 2 and the operator is n°)
now an invariant, giving

<j
' j'
I 2
j'mIT(Ollj"
° j"
1 2
jm' >= 0u..! 0mm,( _pr i', +j, +j _ k,( (2j l + l)(2j2 +
(2k I+ l)
0)1
x {j~, j'~
h h
A.5.3 Appendix 547

This formula gives thej-dependence of the matrix element of the invariant T bo1
through a 6j-symbol.
The explicit form of the invariant operator is from equation (A.5.9)

putting in the known va lues for the Clebsch - Gordan coefficients. It is fairly
common practice, however, to write

and to refer to (R1k ,).S(k ,)) as the scalar product of the two irreducible sets of
operators R(k,) and S(k ,J. In particular for the angular momentum operators
which have 1<' = 1, q ,

which agrees with the usual notion of a scalar product of two vectors. Care
must be taken to avoid errors of (-1) k,(2k 1 + 1)! due to these two ways of
expressing the inva riant.
Some of the results obtained earlier in this chapter may be obtained directly
from the formulae derived above rather than by the use of equivalent
operators. For example, equations (8.28) and (8.39) are particular cases of
equation (A.5.15) while equation (A.5.14) will be used in the next section to
deduce the ratios of transition strengths.

A.5.3 Transition strengths

We have already discussed selection rules for the emiss ion and absorption of
electric dipole radiation in sect ions 5.4 and 8.1. In an LS-coupling situation for
a many-electron atom they are given by

S' = S, J '=J,J±l, L' = L± 1 (A.5.16)

with the usual exclusion of 0 --+ 0 transitions in L or J. Here we have denoted


the initial and final states by SLJ and S'L'J', respectively. These rules follow
immediately from the fact that the dipole operator is invariant in spin space
and a vector operator in position space. The Wigner - Eckart theorem may
now be used to deduce the ratios of transition strengths between two terms for
different J and J'. In fact we need the extension of that theorem given by
equation (A.5.14). Making the correspondence
Ulj2j'lj~jj'mm' kil-+(LSL' SJJ' M M' 1)
548 Appendix A.5.3

and denoting the dipole operator by V~l) we have


= (-I)L+s+J'-IC(J' lJ,M'qM)
I}
< LSJ MIV~l)IL'SJ' M' )
J J'
X[(2L+l)(21'+I)]!lL'L
5 S <LSIIV(1)IIL'S)
The transition strength is usually defined to be the sum of sq uares of the matrix
elements ofV~l) summed over the possible projections M and M ' of initial and
final states. From normalisation of the Clebsch - Gordan coefficients

I,C 2(1'lJ ,M'qM) =1


M'

so that the strength is given by

S(J ,1') = (2J + 1)(2L + 1)(21 ' + 1){ ~, l' 1}2 <L SI IV I1
' I IL' S) 2
L S
(A.5.17)
Thus without any detailed knowledge of the wave functions we deduce that
for fixed Land L ', the dependence on J and J' is given by

S(J ,J') oc (2J+I)(2J'+I){~, ~' ~r


This result has been known as the Honl - Kronig intensity rule . A simple sum
rule may be obtained for the total intensity summed over final states J'. From
equation (A.5.13) the 6j-symbols must satisfy the normalisation condition

I, (2j12 + 1)(2j13 + 1){~1 ~12 .~2}2 = 1


i ., .I .!I 3 13
giving the J-dependence of the summed strengths as
I,S(J,J ') oc (2J+l)
J'

which is another well-known empirical rule .


For quadrupole transitions the strengths are given in just the same way by
an equation like (A.5.17) but with k 1 = 2 rather than k 1 = 1.
When an energy level with angular momentum J is split into (21 + 1) levels
by a magnetic field as described in section 8.5 for the hydrogen atom and
illustrated in figure 8.1(c) the relative st rengths of transitions between states
with d ifferent M may be related using the formulae above. As an example
consider radiation polarised in the xy-plane so that the dipole operator has
q = ± I implying that M ' = M =+= 1. Then, omitting the summations carried
out in arriving at equation (A.5.17) the strengths are given by
S(J,J',M,M') = C 2(J'lJ,M' ± 1M )(2L + I) (2J' + 1)

x{~, ~ ~r<LSIIVI1)IIL'S)2
A.5 .4 Appendix 549

Thus for example the ratio of strengths for the two transitions 151 -> 2PJ and
15r~ 2p!, shown in figure 8.1(c), both with M = -!- and M ' = is 2 -+1,
C 2(ll .1 .1 _ I _.1) x 4 x ) I ,1 I3 _i1 }2
2 2'2 2 tI 0 1

C 2(.1I.1 .1 _ I _ .1) x 2 x
I !- 1}2 - I
2 2'2 2 {
I
I 0 !-
where the values of the coefficients are taken from Rotenberg et al. (1959) in the
bibliography to chapter 7. Because of the presence of the Clebsch - Gordan
coefficients this differs from the ratio of 2 obtained from equation (A.5.l7) for
the total strengths from 1512 to 2PJ2 and 2Pb 2
respectively.

A.5 .4 The crystal field potential

The results of subsection 9.9.2 do not depend on the detailed form of the
crystal field potential but only on its symmetry which is determined by the
cry stal structure. However, this told us only the manner in which the multiplets
would split but did not tell us anything about the ordering of the split levels or
their energy spacings. If we now want to calculate such details of the states in
the crystal field we will need to evaluate matrix elements of the crystal field
potential in the states which we have found by symmetry arguments. In this
section we study the crystal field potential and in the following section
methods a re given for evaluating the matrix elements.
The potential due to the surrounding atoms or ions will have a form which is
invariant under all the point group symmetry operators and will therefore
transform according to the identity representation . For the region in which we
are interested it will also be a solution to Laplace's equation and can therefore
be expanded in spherical harmonics as

Vc = I A~k) rk flqk )( 0, ¢) (A.5.18)


kq

Since rotations can only mix spherical harmonics of the same k the
contribution to Vc from each k must al so transform according to the identity
representation of the point group. This places very strong restrictions on the
coefficients At) which we illustrate by taking the symmetry group O.
From the formula (7.42) for the character of D(k) and using table 9.5 we soon
find that for k = 1, 2, 3 and 5 the representation D(k) contains no invariant
under the group 0 and for k = 4 there is just one.
By defining the z-axis to be an axis of greatest symmetry we may also restrict
the possible values of q which occur in the expression (A.5.18). For example in
the group 0, the four-fold rotational symmetry about the z-axis implies that
the values of q must have the form q = 4n, with n an integer. Thus the only
surviving coefficients with k < 6 are AbO) , AiJ4), A~4) and A(~~ and, since there is
only one fourth-order invariant, the coefficients Ab4) , A~4) and A(~~ must be
550 Appendix A.5.5

related. To make the potential invariant under a four-fold rotation about the
x- or y-axes requires that A(~~ = A~4) = (t54)fAb4), see problem A.8. Hence the
potential for a cubic field has the form
Vc = AbO) + A~4)r4 { ylo4) + (4)t( y~4) + yi~~)}+ terms with k ~6 (A.5.19)
In this particular case it is equally easy to write down an expression for V in
powers of x, y and z which has the required symmetry. The terms must be even
in x, y and z and be symmetric with respect to permutation of x, y and z. The
serie s must therefore take the form

Vc = AbO) + IX( X 2 + y 2 + Z2) + {3(x4 + y4 + Z4) + y(x 2y 2 + y 2z2 + Z2x 2)


+ powers ~6
The potential must also satisfy Laplace's equation V 2Vc = 0 giving IX = 0 and
y = -3/3, a result which is seen to be identical with equation (A.5.19) if the
explicit formula (7.48) for the spherical harmonics is used.
It is possible to calculate numerical values for the non-zero coefficients
A~k) for a particular model, e.g. that the potential is due to point charges
located at the neighbouring lattice sites , but these models do not take into
account the polarisation of the ions and exchange effects. It is usual therefore
to regard the coefficient At) as undetermined parameters which are fixed by
comparing predictions with experiments. This procedure is meaningful since
only a few terms in the series (A. 5.18) will give non-vanishing matrix elements
for electron states of definite angular momentum, because of the restrictions
(8.3) derived from the triangle rule of vector coupling. This means that there
are just a few parameters At) to be chosen to fit consistently all the data
available from the range of experiments on specific heats, magnetic suscepti-
bilities, optical absorption and paramagnetic resonance.
For the cubic field at the vanadium ion in vanadium bromate the potential
will be given by the expression (A.5.19). No terms involving k > 4 need be
considered since we are dealing with d-elecuons and the single particle matrix
elements are of the form <dl l1;)!d >
which vanish by the triangle rule for k > 4.
Since the constant AbO) gives the same contribution in all states, the crystal field
splittings in vanadium bromate are given in terms of the single parameter Ab4 ).

A.5.5 The use of symmetry to deduce ratios of


splittings

The calculation of the crystal field splittings in terms of the parameters A~k) is
in general a complicated numerical problem involving the radial wave
functions of the electrons and details of the many-electron wave functions.
However, it is possible to deduce relations between some of the splittings by
the use of symmetry alone. This is done by using the Wigner - Eckart theorem
or, what is really the same thing, by the use of equivalent operators, see
equations (4.62), (7.53) and subsection 7.4.7.
A.5.5 Appendix 551

We illustrate the technique by calculating the ratio of the energy differences


between the three split components A 2 , T 2 and T 1 of the F-Ievel in the second
column of figure 9.9. The potential in this case is given by equation (A.5.19)
and , as shown above, the terms with k ~ 6 do not contribute. Furthermore, the
fYt 3 invariant term AbO) contributes the same amount to each component and
may therefore be ignored in calculating energy differences . Thus the differ-
ences are both proportional to Ab4 ) and the symmetry property of the operator
is defined as k = 4, with respect to fYt 3 and (invariant) A 1 with respect to the
group O. Using the notation V(k = 4, AI) for this part of Vc we need to find the
<
matrix elements L = 3, T(a) IV(k = 4, AI) IL = 3, T( a» for T( a) = A 2 J 2 and
T 1 • The form of this matrix element is very similar to that of equation (7.53)
where the Wigner - Eckart theorem for the group fYt 3 was used. The difference
is that here the group 0 is used to label the rows of the representations of fYt 3
instead of the group fYt 2 • In other words , operators and wave functions are
now labelled by Land T(a) rather than by Land M.
The required coupling coefficients for use in the Wigner - Eckart theorem in
this basis are not generally available so that it is usual to express both operators
and wave funct ions in the familiar M-basis. For the operator, this has already
been done in equation (A.5.l9) and we may write

(A.5.20)

We deliberately use the notation V~4) rather than ¥1q4 ) to emphasise that the
actual potential is summed over all electrons. Thus, whereas ¥1q4 ) must refer to a
specific electron, the expression 0q4 ) stands for the sum of such terms . It will of
course have the same rotation properties and that is all that concerns us at the
moment. Expressions for a typical wave function

IL = 3,Tf/1.) > = ~>~IL = 3, M >


M

in the M-basis are easily deduced and for practical purposes the coefficients c<u
have been tabulated. In this example we shall deduce them by setting up the
7 x 7 matrix of the operator (A.5.20) in the M-basis for L = 3 and then
+ +
diagonalising. From the reduction D(3) = A 2 T 1 T 2 we know that the
resulting eigenvalues must contain two three-fold degeneracies.
From the m-dependence of the operator (A.5.20) it is clear that the term
004 ) contributes only on the diagonal of the matrix, while the other two terms
change M by four units . Thus, we see that the 7 x 7 matrix breaks up into three
2 x 2 matrices with M = ± 2, M = + 1 and - 3 and M = - 1 and + 3,
respectively, while M = 0 is uncoupled and is therefore an eigenvector. We
thus need to calculate only a few matrix elements of the kind
<L = 3, M' IV~4) IL = 3, M > which are related , according to the Wigner -
Eckart theorem by
<L = 3,M'IV~4)1 L = 3,M >= C(343,MqM ')(L = 31IV(4) II L = 3 >
552 Appendix A.5 .5

Since we are only interested in ratios, the reduced matrix element


<L = 311 J!14) II L
= 3> may be ignored and the coupling coefficients
C(343, M qM') may be found in tables . In this way one finds the eigenvectors
given in table A.5 and the eigenvalues, with a constant factor extracted, in the
last column. (We are interested only in ratios here.) The crystal field in this
particular salt has a sign for A 1>4' which leads to the representation A 2 coming
lowest. The ratio of the splittings, deduced from the table, is shown in figure
9.9.

Table A.5

Representation Wave function Energy


of 0 shift
- -- - - - - - - - --
Az {1
L = 3, M = 2 ) - IL = 3, M = - 2 ) 12 - i - 6

T, {IL=3,M=2 )+!L=3,M=-2 )}2-!


-I
pt I L = 3, M = ±3 )-S!IL=3,M= +=1 ) IS-!
Tz {S! IL = 3, M = ± 3 ) + 3!I L = 3, M = += 1) IS- !
IL = 3, M = 0 ) +3
. __
._ - - --- - -

The use of the coupling coefficients C (343, MqM') may be avoided by


constructing equivalent operators L~4' from products L x L x L x L of the
angular momentum operators. Since there is only one such operator with
q = ± 4 we may take U1'4 r
= (L + as a fourth power of the raising or lowering
operators. The remaining operator Lb41 may now be deduced either by
successive use of equation (7.52),

L1>4' = 24(70)![ L _, [ L _, [L _, [L _ ,(L +)4]J ] ]

or by various short-cut methods, see problem A.9. The final equivalent


operator is then given by

L1>4) + (4)!(L~4) + U~~) = (154)! {14L~ - I2L 2L; + ~(U)2

(A.5.2I)

and within states with L = 3 we may write <L 2 >= 12 giving


L1>4) + (154)!(L~4) + LI~~) = lt54)! { I4L; - I34L; + 144 + L~ + L~ }
In the M-basis this operator may be evaluated directly using equation (7.40)
and <L, >= M without the need to look up values for coupling coefficients.
Of course it leads to the same result as presented in table A.5.
Finally, we remark that the formula (A.5.2I) may also be used to deduce the
splitt ing of each l-multiplet shown in the third column of figure 9.8. It is
Appendix 553

necessary only to write J2 for L 2 and J, for Lz . The relative strengths of the
splitting in different J-multiplets may be deduced with the help of equation
(A.5.l4).

Problems on appendices 4 and 5


A.I Sh ow that the choice pt e) = p(O)/(ee/ca). ~ 0 is su fficient to ens u re that equation
(A.4.14) is sat isfied .
A.2 Use the orthogonality relati on (A.4.2S) to show that the genera lised projection
operator (transfer operator) for the group Ji 3 ma y be wr itten in the form

f
P:.!.'m = (2j + I) D:.!·'m(iX. p,1') T(iX . p. l') sin{Jd :xd{Jd}'/ 8rr 2

where T(iX ,p, y) is the rotation operator.


A.3 By elementary arguments sho w that an alternative form for the sim ple projection
operator pUI is

p UI = n.{
J r- )
J 2 - j' U' + I) l!n
J r-
U U+ I) - j' U' + I)
J
l
where n denotes the infin ite pr oduct over j' = 0. 1. I. . . ., excl ud ing j' = j .
A.4 Calculate the coupling coefficients required in the deri vation of equation (A.S.2 ).
(U se the method of questi on 7.8.)
A.S In question 7.12. the ra tio s between matrix elements o f a tensor operator of
degree 2. in sta tes with I = I. were calculated from the Wigner- Eck art theorem.
Deduce the same rat ios by using the equivalent operator meth od o f sectio n 7.4.7
(Sta rt with the equivalent operator Lll ' = L + L + and co nst r uc t the other
components fro m equati on (7.52) with equat ions (7.28 ) a nd (7.30). Then use
equation (7.40) to find th e matri x elements .)
A.6 Use equation (A.S. 14 ) to sho w th at the matrix elements o f the d ipole operator
L o + gS o in LS-coupling a re given by the formul a

<L SJ ' M IL" + gSo I L SJ M ) = g M '6 J'J + ( I - g )( _ I ) I. + S + J +I

f
J 'J I }
x C (J IJ '. M OM ) [L(L +I )(2L +l)(2J+I)] 1
LL S
(Write So = J o - L o and use the result of questi on 7.13 for the reduced matrix
element of an an gular momentum operator.)
A.7 Use equation (A.S.15) to deduce the result (A.S.2). Treat P, (COS8 12 ) as a scalar
product [y(k I (I) . ylk ' (2 )] because of the addition theorem (A.4.8) and take the
reduced matrix elements of y tk l from equation (A.4.S). You will need the
following coefficients:

C (2 11, 000) = -d )l. { : 2} = {I


I
I
b' 1

A.S Sh ow that, if the potential Vc of eq ua tio n (A.S.18) is to be invariant under the


group O. then the coefficients must satisfy the relati on s A':4 = A1 = (f4)IA ~ .
(Co nsider tw o-fold and four-fold rotat ion s about the x-axis, using
YI;~ = N (x ± i.d. rt041 = N( fs)I( 3Sz4 - 30Z2, 2 + 3,4 ). )
554 Appendix

A.9 In subsection A.5.5. we derived an expression (A.5.21 )for the operator equivalent
for V(k = 4, AI)' Here, we suggest a short-cut method to achieve the same result.
The operator we require must be a fourth-order polynomial in the operators L "
Lyand L; However, we know that L~4) = L~ and that, since L~4) has q = 0, it
must have the form L~4) = C(L 4 + «L 2L; + fJL~ + yL 2 + (iL;). The coefficients a,
p, y and (i may be found by demanding that L~41 has zero matrix elements in all
states with L < 2. This is a necessary condition for a tensor operator of degree
k = 4. The choice of states L = I, M = 0; L = I, M = I; L = 1, M = 1; L = 1,
M = t is sufficient to give the four coefficients. To find the correct relative
magnitude of L~41 and L~41 we must make use of the symmetry &roup 0 and
choose the constant C so that in the complete expression L~41 + (-fs)2(L~4) + L~~)
the coefficients of L~ and L~ are the same.
A.IO Deduce the relative splitting for given J in column three of figure 9.8 by using an
equivalent operator like (A.5.21) for the cry stal field.
A.II With the method suggested in the previous question one cannot deduce the
relative splitting in different J due to the crystal field. Use equation (A.5.14),
together with the table of Rotenberg et al. (1959) in the bibliography to chapter 7,
to find this relative splitting.
Appendix 6: Solutions
to Problems in Volume 2

Chapter /4
14.4 See figure S.3 for the Brillouin zone. The little groups are C4,- at the
points rand M, C 2v at X , and C lh at ~ , Land Z . As k --+0, ~I is
compatible with r l' r 3 and r 5' and ~ 2 is compatible with r 2, r 4 and
r s . As k--+n ja, ~I is compatible with XI and X 4, while ~2 is
compatible with X 2 and X 3 ' (The irreducible representations of the
little group C4v at r are denoted by r i , where i = I to 5 corresponds to
the order in which the representations are listed in the character tables
in appendix 1. The same convention is used at the other points.)
14.5 The eigenstates are <!>I Al=!/Jl+!/J2+!/J3+!/J4' <!>IH,) = !/Jl +!/J4
-!/J3-!/J2' <!>IE)=!/JI-!/J4' <!>I E)=!/J2-!/J 3' the last two
being degenerate. The energies all contain the constant
h 2 n 2 j2Ma 2 +Vooand in addition.F, , +2 VOl for A I'vII - 2 VOl for B 2
and -VII for the E-doublet, where

Vpq = l'c-
I f
V(x ,y) exp{(px +qy)in ja}dxdy

and e , =exp{(x+y)in ja}!/J2 =exp{(x-y)in ja},


!/J3 = exp{ (-x + y)in ja} and !/J4 = exp{ (-x - y)in ja} .

555
556 Appendix 6: Solutions to Problems

""10

x
r "fa

Figure S.3

Chapter 17
17.1 (124)(36)(5)
17.4 Dimension is five: (22111), (21211), (21121), (12211), (12121).
17.5

Chapter 18
18.1 For L (3]; 111 1 ) ,1222 ) ,1333 ) and the symmetrised products formed
from 1112 ),1113 ) , 1221 ) , 12 2 3 ), 1331 ) , 13 3 2) and 11 23 ) . For
L[111] ; the antisymmetrised product formed from 1123 ) . The ortho-
gonal complement then belongs to L[21]. (Or use the Y 3 projection
operators p[21].)
I

ITIillJ
~~
18.2 (a) 1111121 1112121 1212121 1111131

0Iill] [I[lli] [ffiJ}] I 213 I 31 lJTilll


(b) There are eight as in equation (18.12), six of the kind tffD
where the first row is given in table 18.1 and finally
Append ix 6: Solu t ions to Prob lems 557

18.3

(a) (11 ) x (20) = (31)+(12 )+(20)+(01 ). (b) D l!lx D(1) = D I!l +D I!1
18.4 (300)(2 10)(120)(030)( 201)(1 1 1)(02 1)(102)(012)(003). Th us
M = 3, 2, I, 0, I, 0, - I, - I, - 2, - 3 an d so L = 3 o r L = I.
18.5 L = 3 a nd L = 1 in bo t h cases.
18.6 For [ 2] , L = 4, 2 o r O. Fo r [ 3], L = 6, 4, 3, 2 or O. For [2 1] , L = 5; 4, 3,
2 (twice) a nd 1.
3
18.7 T = 2WIt · h J = 2'
9 5 3 T
2' 1 · = 21 W1t ith J = 21311 9 7 (t .
' 2 ' 2' ) 5( t .
1 wlce,1 wlce,) 1,
3 1

Chapte r 20
20.2 See figure SA. 3a + 2{3

3a+ {3

Figure SA
A ppendices 4 and 5
A.tO See A.11.
A.II The sh ifts in the nine levels in the third column of figure 9.8 a re given by
0, - 66, 132, - 28, - 4, 36, ( - 49 ± 5 J 72 1), 196 in a rbitra ry units, wit h
the figure roughly to sca le.
Index to Volumes 1 and 2*

Abelian group, 10 Bloch sta te, 295. 299, 321


Act ion, 394 Bloch theorem, 295
Active transformation, 341 Boost, 338, 343
Adiabatic approximation, 502 Born -Oppenheimer approximati on . 281. 502
Adjoint, 34, 440 Boson , 102, 402
Ammonia molecule, 116,504 Brillou in zone, 293. 296, 300, 310, 319
Angular momentum, 97, 136- 53. 156-66
Antilinear, 202, 370
Antiparticle. 226. 259. 409 Cartan crit erion , 511
Ant i-unitary , 202, 370 Casimir ope rator, 148
Atomic structure , 102, 166-81 ,469 for the Euclidean group J') . 332
effects of crystal field in, 202-10 for the Lorentz group Y, 345
Axial distance, 448 for the Lorentz group If') , 500
Axial vector , 92 for the Poincare group ?, 356, 376
Axis for the rotation group .cjt ) , 148
bilateral, 193. 200 for the rotation group Jf 4 ' 495
n-Iold, 186 for the unitary group SU ), 247, 248, 261, 492
Causality. 408
Cayley-Klein parameters, 482
Band ,295 Cayley's theorem , 425
gap , 299 Character, 60-68
Baryon, 227 for direct product group, 81
number (charge) , 231,419 for direct product representation, 70
Basis, 25 for molecular vibration, 115
reciprocal , 292 for orthogonal group OJ, 141
* Pages 1-280 refer to volume I and pages 281-557 refer to volume 2
II Index to Volumes 1 and 2

Character (cont inued) Cubic crystal, 197. 306, 319


for point groups, 265-72 Curie point, 311
for rotation group gp2 - 131 Cycle . 426
for rotation group gp J ' 140 Cyclic boundary conditions. 300
for symmetric group !I'. , 431 Cyclic group. II
for symmetrised product, 524
for unitary group UN' 475
table. 67 Degeneracy. 90. 118. 157.485
Charge in harmonic oscill ator. 486
a nd iso spin, 222 in hydrogen atom , 167. 492 -96
conjugation. 411 splitt ing of. 100
in field theory, 410 Detailed balance. principle of. 374
independence. 214. 223. 252 Dim ension
Charm. 262 of Lie group. 126
C hro rnod yna mics, 263 of vector space . 25
Class multiplication. 529 Dipole. 88. 158
Classes. 18-22 Dirac equati on . 380-87. 405. 417
of point groups. 192-95 Direct product
of symmetric group !I'•. 428 of groups. 17
Clebsch- Gordan coefficients, 72, 79 of representations. 70-73
for rotation group :}f J. 142. 147. 160. 166. Dynkin di agram . 514
543
for symmetric group !I'. , 441
for unitary group SU 3. 245 Eigenvalue. eigenvector a nd eigenfunction. 35.
Colour. 261 86
Combination levels. 118 Electromagnetic field. 88. 162. 168. 175. 209.
Commutation relations. 128.403 247. 378. 389. 396. 400 . 419
for Lorentz group !f, 344 Electrons.
for rotation group ~ 3 ' 136 in crystals. 294 -306. 320 -23
for unitary group SU 3' 234 in mole cule s. 281-88
Commutator, 33. 89 Elementary particles. 222. 226-32, 243-48.
Commute. 10 255-62. 364. 393-423
Compatib ility relation, 323 table of. 228
Complex conjugation. 103. 373 Equivalence of representations. 52-54
Condon and Shortley phase convention. 140 Equivalent operator. 147
Conjugate elements. 18 in isopin, 220
Connectivity. 151 in unitary group SU 3. 246
Conservation law. 92 Euclidean groups s' 2' t>' 3.326-34.354.377,507
for current. 411 Eu ler angles. 519. 537
for energy, 350 Euler's theorem. 187
for fields . 400. 407 Event. 335
for momentum. 327. 350 Exchange operato rs. 172-80
Contravariant, 337 Exciton s in insulators. 313
Coordinate. 27 Expectation value. 87
normal. 109
mass-weighted. 107. 307
Coset, 22 Fermion. 102. 402
Coulomb Field. 393-423 (see also Electromagnetic field)
field. 167 Flavour. 263
force in atoms. 170 Fourier series. 127.298
force in nuclei . 215. 218. 220 Four-vector. 335-40
potential, 492 Fractional parentage. 541
Covariant. 336 Fundamental excitations. 118
Covering operation , 12
Creation and annihilation operators. 402. 486
Crystal. 196.289-323 g-factor, 166, 378
field, 202-10. 549 in a-crystal, 210
magnetic, 201 in a ferromagnet, 311
momentum. 314 in an atom, 175
Index to Volumes 1 and 2 III

G, group, 515 Index , 202, 507


G alilean invariance , 343 Indistinguishability, 102 (see also Statistics)
Gauge (see Transformation) Infinitesimal operators, 127-30,510 (see also
Glide-plane, 316 under specific group names)
Gluon, 263 Infrared spectra , 120
Group, 9-22 (see also under specific group Inner product (see Scalar product)
names , e.g, Point group) Insulator, 299
compa ct, 126, 535 exciton s in, 313
continuous, 10, 126-30 International notation, 184, 272
crystallographic, 196 Interval, 335
definition , 9 Invariant, 89
direct product, 17 Inverse, 10, 32
double , 151 , 199 Inversion (space) , 12, 96, 135, 184, 347, 360
dynamical (see non-invariance) (see also Time inversion)
element, 9 Irreducibility, 50, 63
extended , 151 Irreducible representation (see Repre sen-
factor (quotient), 507 tation)
improper, 191 Irreducible set of oper ators, 78, 91, 130, 146
integral, volume, 126, 131,476,533 Isobaric analogue, 218
Lie, 126,510-19 Isomorphism, 16, 197,517
little, 318, 330, 508 Isospin, isotopic or isoba ric spin, 215-24, 228,
magnetic (colour, Shubnikov) 201, 316 471
multiplication table , 10
non -invariance, 498-50 I
j- symbol (6j and 9j), 543
parameters, 126
Jahn - Teller effect, 502
point (see Point group)
jj-coupling, 171 , 470
rearrangement theorem, 22
representation (see Representation)
semi-direct product, 507-10 k-vector , 292
semi-simple, 510 Klein-Gordon equation, 379, 408
simply reducible, 72, 142, 243 Kramer's theorem, 373
symorphic, 316
universal covering , 151 , 481
Labelling of energy levels, 90
Lagrange
Hadron, 222, 227 equations, 108, 394, 399
Hamiltonian, 86 theorem, 23
classical, 394 Lagrangian, 394
density, 400 density , 399, 406, 417, 418
symmetry group of, 89 Lande interval rule, 175
Hamilton's principle, 394 Laplace equation, 149
Harmonic approximation, 107 Lattice , 196, 289
Harmonic oscillator, 4, 110, 485-92, 499 reciprocal , 292
Heisenberg vibrations, 306-11
interaction, 311 Lepton, 232
picture (representation), 405 charge (number), 418
Helicity, 331, 368, 390, 417, 420 Lie algebra , 511
Hermitian (see Operator) Lie group, 126,510-19
Hexagonal , 197 Light cone, 340
Homomorphism, 16,478 Linear combination of atomic orbitals
Hund's rule, 181, 541 (LCAO) approximation, 282, 304
in nuclei, 217, 255 Linear dependence , 25
Hydrogen atom, 162, 166-70,492 Little group, 318, 330, 508
Hypercharge, 227, 230, 410 Lorentz
contraction, 342
Identical particles , 102 (see also Statistics) force, 397
Identity transformation, 335--43
element, 10 Lorentz group (homogeneous) Y , 334--49, 397,
representation, 44 495
IV Index to Volumes 1 and 2

Lorentz grou p (continued) classifica tion, 113


in five dimension s, 501 coo rdina tes, 5, 109
in three dimensions, 499 d isplacement, Ill , 122
inhomoge neous (see Poinca re grou p) frequency, 5, 109, 117
LS-eoupling, 173, 470 wave-functio ns, 117
Normalisatio n (norm), 26
Nuclear force,
Magnetic field (see Electromagnetic field) strong, 214. 227, 243
Magneton wea k, 214, 231
Bohr, 162 Nuclea r shell model. 474, 492
nuclea r, 248, 379 Nucleon. 214
Magnon, 313 Nucleus, 213-22.470,492
Majo rana operato r. 180
Mapp ing, 30
Maschke's theorem, 53 Observable, 87
Mass, 364--68 Octupole, 158
-energy relation, 226, 364. 396 O mega minu s (!r ), 228, 244
number A, 214 Oper at or
splitting, 244 adjoint, 34
zero, 365, 367, 369, 373, 387-90, 417- 23 class, 450, 529
Matrix , 31 crea tion and annihilation, 402, 486
y-, 382 differential, 40
Pauli, 163 Hermitian (self-adjoi nt), 34, 87
representat ion , 44 indu ced, 36
Ma trix element, 31 infinitesimal (see Infinitesimal operat ors)
of angular momentum opera to rs ) + ' 140 inverse, 33
~ u~~~ I~ ,M5 ,~5 - irred ucible set. 78, 9 1. 130, 146
relatio ns between. 92 isospin, 215
Ma xwell eq uatio ns, 389, 401. 422 linear. 30
Mea n life. 230 number, 403
Meson projectio n, 74
decay and relativity, 341 raising an d lowering, 137
K , 226, 230, 410. 412 skew-Hermitia n. 128
n, 222, 226, 368, 407,4 10 spin, 163
Metal, 299-302 tensor , 146, 545
Metric, 336 tran sfer, 75
Michelson and Morley experiment . 342 tr an sform ed, 33, 41
Minkowski diagram , 340 un itary, 34
Mirr or rotation , 184 You ng, 446
Molecule Orb it, 509
Born -Oppenheimer method , 502-505 Orb ital (ato mic, molecular), 282
electron sta tes in, 281-88 Ord er of group, 10
vibrat ions of, 106- 23 Orthogonal, 26
Moment basis for irreducible representat ions, 68
electric dip ole, 88, 375 character s, 61, 66
magnet ic dipole , 88, 162, 247, 256, 258 compl emen t, 27
Momentum, 86, 327, 350, 394, 400, 406 (see eigenvectors, 36
also Angul ar momentum; Crysta l, mo- represent ation s, 54, 57
ment um) vector spaces, 27
Mon oclinic, 197 Orthogonal gro up, 0 3 , 18, 135, 158
Muliken notat ion , 265 Orthon orm al, 26
Mult ipole, 157 Orthorhombi c, 197
Out er produ ct, 441
Overto nes, 11 8
Near ly-free electron model , 295-99
Negative energy states, 351, 363
Neutrino, 226, 365, 368, 412, 417 Para stat istics, 103, 260 (see also Statist ics)
Neutro n, 214, 226 Par ity, 6, 96, 158, 227, 361, 368,411 ,423
Normal mode , 5, 109, 306 of a permu ta tion , 427
Index to Volumes 1 and 2 V
Partition, 427, 434, 456 equivalent, 52
Parton, 262 faithful , 44
Passive transformation, 341 identity, 44
Pauli equation, 378 irreducible , 50, 90
Pauli matrix, 163 matrix, 44
Pauli principle , 102, 171 ,255, 299, 469 (see also of direct product group, 81
Stati stics) of semi-direct product group, 508
rcr theorem. 411 reduction of, 62
Periodic potential. 294 regular, 64, 434
Periodic table . 102 symmetrised product. 72, 119, 524
Permutation, 15,39,426.525 vector, 92
group (see Symmetric group) Resonance. 230
Perturbation, 88. 99 Rest mass, 365
Phonon, 310 Rhombohedral. 197
Photon, 365, 368, 390.419.423 Root, 511
Poincare group & , 349-90, 399--423. 507 Rotation , 12. 19
Point group, 183-202 improper, 12, 191
character tables . 265-72 invariance, 156
class structure. 192-95 matrix, 47, 140. 143,519
enumeration. 186-92 mirror, 14, 184
notation, 272 Rotation group
representations, 197-201 912 , 14,96, 130-34
Polarisation, 88, 94, 97 a; 15, 134-53. 156-66, 478-82, 519-22
Positron. 411 ,tJf4 , 494
Primitive cell, 290 91.,472, 516
Primitive translation vector, 196, 290, 315 Rotc-inversion, 272
Projection operator, 74, 122, 283, 302, 307, Runge -Lentz vector , 493
309,314
Proton, 214. 228
Pseudo-vector, 92 Scalar (inner) product, 26, 333, 359
for Dirac equation, 385
in space -time, 335
Quadrupole, 158 Schmidt orthogonalisation process, 27
Quadrupole force, 492 Schoenflies notation. 184,272
Quantum mechanics . 85-104 Schrodinger equation . 30, 86. 98. 376, 405
Quark, 103, 257-62 for periodic potential, 295
Quenching, 210, 311 for spherical potential, 149
Schur 's lemmas, 55-60
Screening, 170
Raman spectra, 121 Screw-axis, 316
Rank Second quantisation, 103. 402, 486
of Lie group, 150, 511 Selection rule, 7, 91
of tensor operator, 146 angular momentum, 157
Rayleigh -Ritz method, 98 for electronic excitat ion in molecules, 287
Reciprocal lattice, 292, 318 hypercharge, 231
Recoupling coefficients. 543 in crystals , 314
Reduced matrix element , 80, 147,245,543--47 in molecular vibrations, 120
Reduction of representation . 62 isospin, 221
Reflection, 14, 184 Seniority , 473
strong, 412 Shell (atomic) , 172
Regular solid. 186-90 Simply reducible group, 72, 142,243
Relativistic fields, 396--423 Slater determinant, 172, 282
Relativity , 325-90, 396 Space
Representation. 43-84 (see also under specific Fock ,403
group names) function , 28, 36
alternating, 430 group, 315-23
complex conjugate, 153, 241, 467 Hilbert, 86
direct product. 70. 141. 242, 527 invariant, 48
double-valued, 150, 161, 166, 199, 266, 478 lanice. 196. 289
VI Index to Volumes 1 and 2

Space (continu ed) Transition, 88, 91, 92


representation , 44 forbidden, 91
vector, 24--42 strengths, 547
Space-like interv al, 335 Translation group
Space -time regions , 339 continuous, 326
Special (unim odular) unitary group (see in crystals, 289-315
Unitary group) in space -time, 349
Spectroscopic not ati on , 167 Transposition, 427
Spectrum, 87 Tri angular condition, 142
Spherical components of irreducible tensor, Tri clinic , 197
147
Spherical harmonic, 145, 149, 157,531
Spin , 102, 161-66,333, 366,469 V-spin and V-spin. 234
- o rbit interaction, 166, 168, 175 Umklapp process , 314
wave, 311 Unitary (se(' Operator)
Spin - statistics theorem, 102, 408 Unitary group
Star of k, 317, 330, 509 SV 4 , 251-55, 262, 464
Stationary sta tes, 86 SV 6 ,255-60
Statistics, 102, 260, 402, 409 SV N' 462-78,515
Stereogram, 184 V 2 , SV 2 , 215, 464, 478-83
Stoke s' lines, 121 V 3' S V 3' 232-49, 464, 488, 517
Strangeness, 231, 412 (s('(' also Hyperch arge) U N ,452-78
Structure constant, 128, 511
Subgroup, 17, 527
irreducible representation of, 73 Vacuum, 402
normal (invariant or self-conjugate), 507- Valence electron, 173, 300
10 Variational
of continuous group, 129 approximation, 97-99, 304. 322
Supermultiplet, 251-55, 470 pr inciple , 394
of elementary particles, 255-62 Vector, 25-27
Symmetric group four , 335--40
y 3' 15,21 , 254 null , 339
« ; 15, 102, 122,425-50,454-56,525 Vector coupling, 142, 159, 166, 178,543--47
Symmetrised product, 72, 119, 179, 207, coefficients, 147
524-27 Velocity of light, 338, 342
Symm etri ser and antisymmetriser, 430 Vibrations
Symmetry breaking, 99 lattice, 306-11
spontaneous, 502 molecular, 106-24
Symorphic, 316
Symplectic group, 473
Syngony, 197 Water molecule, 115, 286
Wave equation, 375-90, 408-23
Weight , 138,237,467
Tachyon, 343 function, 126, 476, 533
Tensor operator ( 5(' (' Operator) Weyl equation, 387-89,417
Term, 174 Wigner -Eckart theorem, 78, 92, 169
Tetragonal, 197 for a; 146,545
Tight-binding model , 302-306, 321 for SV 3,245
Time inversion, 362 ( 5(' (' also Time-reversal)
Time-dilation, 341
Time-like interval, 335 Yamanouchi symbol, 437
Time-reversal, 103,201,316,369-75,411 Young diagram, 434, 456
Transformation, 30--42 Young tableau, 436
gauge , 398, 410, 422, 506
induced,36
Lorentz (see Lorentz, transformation) Zeeman splitting, 162, 170, 175

Potrebbero piacerti anche