Text
                    GROUP
THEORY
IN PHYSICS
AN INTRODUCTION
ЛР
J. F. CORNWELL

This book is printed on acid-free paper Academic Press 525 В Street, Suite 1900, San Diego, California 92101-4495, USA http://www.apnet.com Academic Press Limited 24-28 Oval Road, London NW1 7DX, UK ht t p: // www. hbuk. co. uk/ap/ Copyright © 1997 by ACADEMIC PRESS All rights reserved No part of this publication may be reproduced or transmitted in any form or by any means electronic or mechanical, including photocopy, recording, or any information storage and retrieval system, without permission in writing from the publisher. A catalogue record for this book is available from the British Library Library of Congress Cataloguing-in-Publication Data ISBN 0-12-189800-8 Printed in Great Britain by the University Press, Cambridge 97 98 99 00 01 02 EB 9 8 7 6 5 4 3 2 1
Contents Preface vii 1 The Basic Framework 1 1 The concept of a group....................................... 1 2 Groups of coordinate transformations ........................ 4 (a) Rotations.............................................. 5 (b) Translations........................................... 9 3 The group of the Schrodinger equation....................... 10 (a) The Hamiltonian operator.............................. 10 (b) The invariance of the Hamiltonian operator............ 11 (c) The scalar transformation operators P(T).............. 12 4 The role of matrix representations.......................... 15 2 The Structure of Groups 19 1 Some elementary considerations.............................. 19 2 Classes..................................................... 21 3 Invariant subgroups ........................................ 23 4 Cosets...................................................... 24 5 Factor groups............................................... 26 6 Homomorphic and isomorphic mappings......................... 28 7 Direct products and semi-direct products of groups...........31 3 Lie Groups 35 1 Definition of a linear Lie group............................ 35 2 The connected components of a linear Lie group ..............40 3 The concept of compactness for linear Lie groups.............42 4 Invariant integration....................................... 44 4 Representations of Groups - Principal Ideas 47 1 Definitions................................................. 47 2 Equivalent representations.................................. 49 3 Unitary representations..................................... 52 4 Reducible and irreducible representations................... 54 5 Schur’s Lemmas and the orthogonality theorem for matrix rep- resentations ................................................... 57
iv GROUP THEORY IN PHYSICS 6 Characters................................................... 59 5 Representations of Groups - Developments 65 1 Projection operators..........................................65 2 Direct product representations .............................. 70 3 The Wigner-Eckart Theorem for groups of coordinate transfor- mations in IR3 ................................................... 73 4 The Wigner-Eckart Theorem generalized........................ 79 5 Representations of direct product groups..................... 83 6 Irreducible representations of finite Abelian groups..........85 7 Induced representations...................................... 86 6 Group Theory in Quantum Mechanical Calculations 93 1 The solution of the Schrodinger equation..................... 93 2 Transition probabilities and selection rules................. 97 3 Time-independent perturbation theory.........................100 7 Crystallographic Space Groups 103 1 The Bravais lattices ........................................103 2 The cyclic boundary conditions...............................107 3 Irreducible representations of the group T of pure primitive translations and Bloch’s Theorem..................................109 4 Brillouin zones..............................................Ill 5 Electronic energy bands......................................115 6 Survey of the crystallographic space groups..................118 7 Irreducible representations of symmorphic space groups .... 121 (a) Fundamental theorem on irreducible representations of symmorphic space groups.....................................121 (b) Irreducible representations of the cubic space groups O5h and O9h..................................................126 8 Consequences of the fundamental theorems...................129 (a) Degeneracies of eigenvalues and the symmetry of e(k) . 129 (b) Continuity and compatibility of the irreducible repre- sentations of Po(k).........................................131 (c) Origin and orientation dependence of the symmetry la- belling of electronic states................................134 8 The Role of Lie Algebras 135 1 “Local” and “global” aspects of Lie groups ..................135 2 The matrix exponential function..............................136 3 One-parameter subgroups......................................139 4 Lie algebras.................................................140 5 The real Lie algebras that correspond to general linear Lie groups 145 (a) The existence of a real Lie algebra C for every linear Lie group Q.................................................145 (b) The relationship of the real Lie algebra C to the one- parameter subgroups of Q....................................148
CONTENTS v 9 The Relationships between Lie Groups and Lie Algebras Ex- plored 153 1 Introduction......................................................153 2 Subalgebras of Lie algebras.......................................153 3 Homomorphic and isomorphic mappings of Lie algebras . . . .154 4 Representations of Lie algebras...................................160 5 The adjoint representations of Lie algebras and linear Lie groups 168 6 Direct sum of Lie algebras........................................171 10 The Three-dimensional Rotation Groups 175 1 Some properties reviewed..........................................175 2 The class structures of SU(2) and SO(3)...........................176 3 Irreducible representations of the Lie algebras su(2) and so(3) . 177 4 Representations of the Lie groups SU(2), SO(3) and 0(3) ... 183 5 Direct products of irreducible representations and the Clebsch- Gordan coefficients..............................................186 6 Applications to atomic physics ...................................189 11 The Structure of Semi-simple Lie Algebras 193 1 An outline of the presentation....................................193 2 The Killing form and Cartan’s criterion...........................193 3 Complexification..................................................198 4 The Cartan subalgebras and roots of semi-simple complex Lie algebras.........................................................200 5 Properties of roots of semi-simple complex Lie algebras .... 207 6 The remaining commutation relations...............................213 7 The simple roots................................................. 218 8 The Weyl canonical Jorm of £......................................223 9 The Weyl group of £...............................................224 10 Semi-simple real Lie algebras.....................................228 12 Representations of Semi-simple Lie Algebras 235 1 Some basic ideas..................................................235 2 The weights of a representation...................................236 3 The highest weight of a representation............................241 4 The irreducible representations of £ = A2, the complexification of £ = su(3).....................................................245 5 Casimir operators........................................251 13 Symmetry schemes for the elementary particles 255 1 Leptons and hadrons......................................255 2 The global internal symmetry group SU(2) and isotopic spin . . 256 3 The global internal symmetry group SU(3) and strangeness . . 259
vi GROUP THEORY IN PHYSICS APPENDICES 269 A Matrices 271 1 Definitions.........................................................271 2 Eigenvalues and eigenvectors .......................................275 В Vector Spaces 279 1 The concept of a vector space.......................................279 2 Inner product spaces................................................282 3 Hilbert spaces......................................................286 4 Linear operators....................................................288 5 Bilinear forms......................................................292 6 Linear functionals .................................................294 7 Direct product spaces ..........................................295 C Character Tables for the Crystallographic Point Groups 299 D Properties of the Classical Simple Complex Lie Algebras 319 1 The simple complex Lie algebra Ai, I > 1 .319 2 The simple complex Lie algebra Bi, I > 1 .320 3 The simple complex Lie algebra Ci, I > 1 .322 4 The simple complex Lie algebra I > 3 (and the semi-simple complex Lie algebra D^)............................................324 References 327 Index 335
Preface As was observed previously in the preface to my three-volume work Group Theory in Physics, thirty years or so ago group theory could have been re- garded by physicists as merely providing a very valuable tool for the eluci- dation of the symmetry aspects of physical problems. However, recent de- velopments, particularly in high-energy physics, have transformed its role, so that it now occupies a crucial and indispensable position at the centre of the stage. These developments have taken physicists increasingly deeper into the fascinating world of the pure mathematicians, and have led to an ever- growing appreciation of their achievements, the full recognition of which has been hampered to some extent by the style in which much of modern pure mathematics is presented. As with my previous three-volume treatise, one of the main objectives of the present work is to try to overcome this commu- nication barrier, and to present to theoretical physicists and others some of the important mathematical developments in a form that should be easier to comprehend and appreciate. Although my Group Theory in Physics was intended to provide a intro- duction to the subject, it also aimed to provide a thorough and self-contained account, and so its overall length may well have made it appear rather daunt- ing. The present book has accordingly been designed to provide a much more succinct introduction to the subject, suitable for advanced undergraduate and postgraduate students, and for others approaching the subject for the first time. The treatment starts with the basic concepts and is carried through to some of the most significant developments in atomic physics, electronic energy bands in solids, and the theory of elementary particles. No prior knowledge of group theory is assumed, and, for convenience, various relevant algebraic concepts are summarized in Appendices A and B. The present work is essentially an abridgement of Volumes I and II of Group Theory in Physics (which hereafter will be referred to as “Cornwell (1984)”), although some new material has been included. The intention has been to concentrate on introducing and describing in detail the most impor- tant basic ideas and the role that they play in physical problems. Inevitably restrictions on length have meant that some other important concepts and developments have had to be omitted. Nevertheless the mathematical cover- age goes outside the strict confines of group theory itself, for one soon is led to the study of Lie algebras, which, although related to Lie groups, are often
viii GROUP THEORY IN PHYSICS developed by mathematicians as a separate subject. Mathematical proofs have been included only when the direct nature of their arguments assist in the appreciation of theorems to which they refer. In other cases references have been given to works in which they may be found. In many instances these references are quoted as “Cornwell (1984)”, as interested readers may find it useful to see these proofs with the same notations, conventions, and nomenclature as in the present work. Of course, this is not intended to imply that this reference is either the original source or the only place in which a proof may be found. The same reservation naturally applies to the references to suggested further reading on topics that have been explicitly omitted here. In the text the treatments of specific cases are frequently given under the heading of “Examples”. The format is such that these are clearly distinguished from the main part of the text, the intention being that to indicate that the detailed analysis in the Example is not essential for the general understanding of the rest of that section or the succeeding sections. Nevertheless, the Exam- ples are important for two reasons. Firstly, they give concrete realizations of the concepts that have just been introduced. Secondly, they indicate how the concepts apply to certain physically important groups or algebras, thereby allowing a “parallel” treatment of a number of specific cases. For instance, many of the properties of the groups SU(2) and SU(3) are developed in a series of such Examples. For the benefit of readers who may wish to concentrate on specific appli- cations, the following list gives the relevant chapters: (i) electronic energy bands in solids: Chapters 1, 2, and 4 to 7; (ii) atomic physics: Chapters 1 to 6, and 8 to 10; (iii) elementary particles: Chapters 1 to 6, and 8 to 13. J. F. Cornwell St. Andrews January, 1997
To my wife Elizabeth and my daughters Rebecca and Jane

Chapter 1 The Basic Framework 1 The concept of a group The aim of this chapter is to introduce the idea of a group, to give some physically important examples, and then to indicate immediately how this notion arises naturally in physical problems, and how the related concept of a group representation lies at the heart of the quantum mechanical formulation. With the basic framework established, the next four chapters will explore in more detail the relevant properties of groups and their representations before the application to physical problems is taken up in earnest in Chapter 6. To mathematicians a group is an object with a very precise meaning. It is a set of elements that must obey four group axioms. On these is based a most elaborate and fascinating theory, not all of which is covered in this book. The development of the theory does not depend on the nature of the elements themselves, but in most physical applications these elements are transformations of one kind or another, which is why T will be used to denote a typical group member. Definition Group Q A set Q of elements is called a “group” if the following four “group axioms” are satisfied: (a) There exists an operation which associates with every pair of elements T and T' of Q another element T" of Q. This operation is called multipli- cation and is written as T" = TT', T" being described as the “product of T with T'”. (b) For any three elements T, T' and T" of Q (TT^T1'__T (T^ (1-1) This is known as the “associative law” for group multiplication. (The interpretation of the left-hand side of Equation (1.1) is that the product
2 GROUP THEORY IN PHYSICS TTf is to be evaluated first, and then multiplied by T" whereas on the right-hand side T is multiplied by the product T'T" (c) There exists an identity element E which is contained in Q such that ТЕ = ET = T for every element T of Q. (d) For each element T of Q there exists an inverse element T~x which is also contained in Q such that TT~r = T~rT = E . This definition covers a diverse range of possibilities, as the following ex- amples indicate. Example I The multiplicative group of real numbers The simplest example (from which the concept of a group was generalized) is the set of all real numbers (excluding zero) with ordinary multiplication as the group multiplication operation. The axioms (a) and (b) are obviously satisfied, the identity is the number 1, and each real number t 0) has its reciprocal 1/t as its inverse. Example II The additive group of real numbers To demonstrate that the group multiplication operation need not have any connection with ordinary multiplication, take Q to be the set of all real num- bers with ordinary addition as the group multiplication operation. Again axioms (a) and (b) are obviously satisfied, but in this case the identity is 0 (as a + 0 = 0 + a = a) and the inverse of a real number a is its negative —a (as a + (—a) = (—a) + a = 0). Example III A finite matrix group Many of the groups appearing in physical problems consist of matrices with matrix multiplication as the group multiplication operation. (A brief account of the terminology and properties of matrices is given in Appendix A.) As an example of such a group let Q be the set of eight matrices Ml = ' 1 0 ' -° 1 , M2 = 1 0 0 ' -1 M3 = -1 o' 0 -1 ’ -1 0 -1 о -1 0 1 ' M4 = 0 1 , M5 = 1 0 5 M6 = -1 0 J ’ о 1 ' 0 -1 M7 = 1 ° , Mg = -1 0 By explicit calculation it can be verified that the product of any two members of Q is also contained in P, so that axiom (a) is satisfied. Axiom (b) is
THE BASIC FRAMEWORK 3 automatically true for matrix multiplication, is the identity of axiom (c) as it is a unit matrix, and finally axiom (d) is satisfied as МГХ=М1, М2Х=М2, М31 = M3, М7х=М4, Mg1 = м6, Мё1 = м5, м^1 = м7, Mg1 = м8. Example IV The groups V(N) and SU(TV) U(TV) for TV > 1 is defined to be the set of all TV x TV unitary matrices u with matrix multiplication as the group multiplication operation. SU(TV) for TV > 2 is defined to be the subset of such matrices u for which detu = 1, with the same group multiplication operation. (As noted in Appendix A, if u is unitary then detu = ехр(га), where a is some real number. The “S” of SU(TV) indicates that SU(TV) is the “special” subset of U(TV) for which this a is zero.) It is easily established that these sets do form groups. Consider first the set U(TV). As (uiu2)t = u2uj and (uiu2) 1 = u2 xux \ if ui and u2 are both unitary then so is uiu2. Again axiom (b) is automatically valid for matrix multiplication and, as the unit matrix In is a member of U(TV), it provides the identity E of axiom (c). Finally, axiom (d) is satisfied, as if u is a member of U(TV) then so is u-1. For SU(TV) the same considerations apply, but in addition if ui and u2 both have determinant 1, Equation (A.4) shows that the same is true of uiu2. Moreover, In is a member of SU(TV), so it is its identity, and u-1 is a member of SU(TV) if that is the case for u. The set of groups SU(TV) is particularly important in theoretical physics. SU(2) is intimately related to angular momentum and isotopic spin, as will be shown in Chapters 10 and 13, while SU(3) is now famous for its role in the classification of elementary particles, which will also be studied in Chapter 13. Example V The groups О (TV) and SO(TV) The set of all TV x TV real orthogonal matrices R (for TV > 2) is denoted almost universally by О (TV), although О (TV, IR) would have been preferable as it indicates that only real matrices are included. The subset of such matrices R with detR = 1 is denoted by SO(TV). As will be described in Section 2, 0(3) and SO(3) are intimately related to rotations in a real three-dimensional Euclidean space, and so occur time and time again in physical applications. О (TV) and SO (TV) are both groups with matrix multiplication as the group multiplication operation, as they can be regarded as being the subsets of U(TV) and SU(TV) respectively that consist only of real matrices. (All that has to be observed to supplement the arguments given in Example IV is that the product of any two real matrices is real, that In is real, and that the inverse of a real matrix is also real.) If Т1Г2 = T2T1 for every pair of elements Ti and T2 of a group Q (that is, if all 7i and T2 of Q commute), then Q is said to be “Abelian”. It will transpire
4 GROUP THEORY IN PHYSICS Ml M2 М3 M4 M5 M6 M7 M8 Mi Ml M2 M3 M4 M5 M6 M7 M8 M2 M2 Ml M4 M3 M8 M7 M6 M5 М3 M3 M4 Ml M2 M6 M5 M8 M7 M4 M4 М3 M2 Ml M7 M8 M5 M6 M5 M5 M7 M6 M8 M3 Ml M4 M2 M6 M6 M8 M5 M7 Ml M3 M2 M4 M7 M7 M5 M8 M6 M2 M4 Ml М3 M8 M8 M6 M7 M5 M4 M2 М3 Ml Table 1.1: Multiplication table for the group of Example III. that such groups have relatively straightforward properties. However, many of the groups having physical applications are non-Abelian. Of the cases considered above the only Abelian groups are those of Examples I and II and the groups U(l) and SO(2) of Examples IV and V. (One of the non- commuting pairs of products of Example III which makes that group non- Abelian is M5M7 = M4, M7M5 = М2.) The “order” of Q is defined to be the number of elements in P, which may be finite, countably infinite, or even non-countably infinite. A group with finite order is called a “finite group”. The vast majority of groups that arise in physical situations are either finite groups or are “Lie groups”, which are a special type of group of non-countably infinite order whose precise definition will be given in Chapter 3, Section 1. Example III is a finite group of order 8, whereas Examples I, II, IV and V are all Lie groups. For a finite group the product of every element with every other element is conveniently displayed in a multiplication table, from which all information on the structure of the group can subsequently be deduced. The multiplica- tion table of Example III is given in Table 1.1. (By convention the order of elements in a product is such that the element in the left-hand column pre- cedes the element in the top row, so for example M5M8 = М2.) For groups of infinite order the construction of a multiplication table is clearly completely impractical, but fortunately for a Lie group the structure of the group is very largely determined by another finite set of relations, namely the commutation relations between the basis elements of the corresponding real Lie algebra, as will be explained in detail in Chapter 8. 2 Groups of coordinate transformations To proceed beyond an intuitive picture of the effect of symmetry operations, it is necessary to specify the operations in a precise algebraic form so that the results of successive operations can be easily deduced. Attention will be confined here to transformations in a real three-dimensional Euclidean space 1R3, as most applications in atomic, molecular and solid state physics involve only transformations of this type.
THE BASIC FRAMEWORK 5 Figure 1.1: Effect of a rotation through an angle 0 in the right-hand screw sense about Ox. (a) Rotations Let Ox, Oy, Oz be three mutually orthogonal Cartesian axes and let Ox', 0yf, 0zf be another set of mutually orthogonal Cartesian axes with the same origin О that is obtained from the first set by a rotation T about a specified axis through O. Let (ж, у, z) and (xf, y', z') be the coordinates of a fixed point P in the space with respect to these two sets of axes. Then there exists a real orthogonal 3x3 matrix R(T) which depends on the rotation T, but which is independent of the position of P, such that r' = R(T)r, (1-2) where (Hereafter position vectors will always be considered as 3 x 1 column matrices in matrix expressions unless otherwise indicated, although for typographical reasons they will often be displayed in the text as 1 x 3 row matrices.) For example, if T is a rotation through an angle в in the right-hand screw sense about the axis Ox, then, as indicated in Figures 1.1 and 1.2, x' = x, у' = у cos 0 + z sin в, z' = —?/sin$ + zcos$,
6 GROUP THEORY IN PHYSICS so that Figure 1.2: The plane containing the axes Oy, Oz, Oy' and Oz' corresponding to the rotation of Figure 1.1. R(T) = 1 0 0 0 cos в sin в 0 — sin в cos в (1-3) The matrix R(T) obeys the orthogonality condition R(T) = R(T)-1 be- cause rotations leave invariant the length of every position vector and the angle between every pair of position vectors, that is, they leave invariant the scalar product Г1.Г2 of any two position vectors. (Indeed the name “or- thogonal” stems from the involvement of such matrices in the transforma- tions being considered here between sets of orthogonal axes.) The proof that R(T) is orthogonal depends on the fact that Г1.Г2 can be expressed in ma- trix form as Г1Г2. Then, if г'г = R(T)ri and rj, = R(T)r2, it follows that Г1.Г2 = Г1Г2 = FiR(T)R(T)r2, which is equal to Г1Г2 for all iq and r2 if and only if R(T)R(T) = 1. As noted in Appendix A, the orthogonality condition implies that det R(T) can take only the values +1 or —1. If detR(T) = +1 the rotation is said to be “proper”; otherwise it is said to be “improper”. The only rotations which can be applied to a rigid body are proper rotations. The transformation of Equation (1.3) gives an example. The simplest example of an improper rotation is the spatial inversion op- eration I for which r' = — r, so that ад = -10 0 0-1 0 0 0-1 Another important example is the operation of reflection in a plane. For instance, for reflection in the plane Oyz, for which x' = —x, y' = y,z' = z,
THE BASIC FRAMEWORK 7 the transformation matrix is ' -1 0 0 0 10 0 0 1 The “product” T1T2 of two rotations Ti and may be defined to be the rotation whose transformation matrix is given by R(7iT2) = R(TX)R(T2). (1-4) (The validity of this definition is assured by the fact that the product of any two real orthogonal matrices is itself real and orthogonal.) In general R(Ti)R(T2) R(T2)R(Ti), in which case T±T2 ф T2Tr. If r' = R(T2)r and r" = R(Ti)r', then Equation (1.4) implies that r" = R(7iT2)r, so the interpretation of Equation (1.4) is that operation T2 takes place before Ti. This is an example of the general convention (which will be applied throughout this book) that in any product of operators the operator on the right acts first. With this definition (Equation (1.4)) every improper rotation can be con- sidered to be the product of the spatial inversion operator I with a proper rotation. For example, for the reflection in the Oyz plane -10 0 0 1 0 0 0 1 -10 0 0-1 0 0 0-1 1 0 0 0-1 0 0 0-1 and, as the second matrix on the right-hand side is the transformation of Equation (1.3) with 9 = 7r, it corresponds to a rotation through 7r about Ox. If a set of matrices R(T) forms a group, then the corresponding set of rotations T also forms a group in which Equation (1.4) defines the group mul- tiplication operator and for which the inverse T 1 of T is given by R(T x) = R(T)-1. As these two groups have the same structure, they are said to be “isomorphic” (a concept which will be examined in more detail in Chapter 2, Section 6). Example I The group of all rotations The set of all rotations, both proper and improper, forms a Lie group that is isomorphic to the group 0(3) that was introduced in Example V of Section 1. Example II The group of all proper rotations The set of all proper rotations forms a Lie group that is isomorphic to the group SO(3). Example III The crystallographic point group D4 A group of rotations that leave invariant a crystal lattice is called a “crys- tallographic point group”, the “point” indicating that one point, the origin (9, is left unmoved by the operations of the group. There are only 32 such
8 GROUP THEORY IN PHYSICS Figure 1.3: The rotation axes Ox, Oz, Oc and Od of the crystallographic point group D4. groups, all of which are finite. A complete description is given in Appendix C. The only possible angles of rotation are 2ir/n, where n = 2, 3, 4, or 6. (This restriction on the value of n is a consequence of the translational symmetry of a perfect crystal (cf. Chapter 7, Section 6). For a “quasicrystal”, which has no such translational symmetry, this restriction no longer applies, and so it is possible to have other values of n as well, including, in particular, the value n = 5.) It is convenient to denote a proper rotation through 2тг/п about an axis Oj by Cnj. The identity transformation may be denoted by E, so that R(F?) = 1, and improper rotations can be written in the form ICnj. As an example, consider the crystallographic point group Z)4, the notation being that of Schonfliess (1923). Z)4 consists of the eight rotations: E: the identity; Czx, Ciy, C2z- proper rotations through 7r about Ox, Oy, Oz respectively; C4y,C^yi proper rotations through тг/2 about Oy in the right-hand and left-hand screw senses respectively; C2c, C2d- proper rotations through 7r about Oc and Od respectively. Here Ox, Oy, Oz are mutually orthogonal Cartesian axes, and Oc,Od are mutually orthogonal axes in the plane Oxz with Oc making an angle of тг/4 with both Ox and Oz, as indicated in Figure 1.3. The transformation matrices are ’ 1 0 0 ’ ' i L 0 0 ’ R(£) = 0 1 0 R(C2J = t ) -1 0 _ 0 0 1 _ t ) o -1 _ -1 0 1 J -1 0 0 K(^) = 0 1 ( 1 , r(c2z) = 0-10 о о - 1 0 0 1
THE BASIC FRAMEWORK 9 ’0 0-11 Г001’ R(C4, r(c2c) ) = 0 10, R^-1) = 010, _ 1 0 0 J [-100 0 0 1 1 Г 0 0 -1 ’ = 0-10, R(C2d) = 0-10. .1 ° ° J L1 ° °. E C2x C2y C2z Ciy C^y C2c C2d E C2x C2y c2z Ciy c2c C2d E C2x C2y C2z Ciy C^1 C2c C2d C2x E C2z C2y C2d C2c Ciy C2y C2z E C2x C^1 С4у C2d C2c C2z C2y C2x E C2c C2d С4у С~) Ciy C2c C4y C2d C2y E C2z C2x Ciy1 C2d Ciy C2c E C2y C2x C2z C2c Ciy C2d Ci1 C2x C2z E C2y c2d Ci1 C2c Ciy C2z C2x C2y E Table 1.2: Multiplication table for the crystallographic point group D4. The multiplication table is given in Table 1.2. This example will be used to illustrate a number of concepts in Chapters 2, 4, 5 and 6. (b) Translations Suppose now that Ox, Oy, Oz is a set of mutually orthogonal Cartesian axes and O'xf,O'y',Ofzf is another set, obtained by first rotating the original set about some axis through О by a rotation whose transformation matrix is R(T), and then translating О to O' along a vector —t(T) without further rotation. (In IR3 any two sets of Cartesian axes can be related in this way.) Then Equation (1.2) generalizes to r' = R(T)r + t(T). (1.5) It is useful to regard the rotation and translation as being two parts of a single coordinate transformation T, and so it is convenient to rewrite Equation (1.5) as r' = {R(T)|t(T)}r, thereby defining the composite operator {R(T)|t(T’)}. Indeed, in the non- symmorphic space groups (see Chapter 7, Section 6), there exist symmetry operations in which the combined rotation and translation leave the crystal lattice invariant without this being true for the rotational and translational parts separately. The generalization of Equation (1.4) can be deduced by considering the two successive transformations r' = {R(72)|t(72)}r = R(T2)r + t(T2) and r" = {R(Ti)|t(Ti)}r' = R(Ti)r' + t(Ti), which give r" = R(T1)R(T2)r + [R(7i)t(T2) + t(T!)]. (1.6)
10 GROUP THEORY IN PHYSICS Thus the natural choice of the definition of the “product” T1T2 of two general symmetry operations Ti and T2 is {R(7i)|t(7i)} = {r(Ti)r(t2) 1 R(7i)t(T2) + t(T!)}. (1.7) This product always satisfies the group associative law of Equation (1.1). As Equation (1.5) can be inverted to give r = R(T)-1r' - R(T)-1t(T), the inverse of {R(T)|t(T)} may be defined by {R(T) It(T)}-1 = {R(T)-11 —R(T)-1t(T)}. (1.8) It is easily verified that {R(7iT2)|t(TiT2)}-1 = {R(r2)|t(r2)}-1{R(T1)|t(r1)}-1, (1.9) the order of factors being reversed on the right-hand side. It is sometimes convenient to refer to transformations for which t(T) = 0 as “pure rotations” and those for which R(T) = 1 as “pure translations”. 3 The group of the Schrodinger equation (a) The Hamiltonian operator The Hamiltonian operator H of a physical system plays two major roles in quantum mechanics (Schiff 1968). Firstly, its eigenvalues e, as given by the time-independent Schrodinger equation Нф = еф, are the only allowed values of the energy of the system. Secondly, the time de- velopment of the system is determined by a wave function which satisfies the time-dependent Schrodinger equation Нф = гТьдф/dt. Not surprisingly, a considerable amount can be learnt about the system by simply examining the set of transformations which leave the Hamiltonian invariant. Indeed the main function of group theory, as it is applied in physical problems, is to systematically extract as much information as possible from this set of transformations. In order to present the essential features as clearly as possible, it will be assumed in the first instance that the problem involves solving a “single- particle” Schrodinger equation. That is, it will be supposed that either the system contains only one particle, or, if there is more than one particle in- volved, then they do not interact or their inter-particle interactions have been
THE BASIC FRAMEWORK 11 treated in a Hartree-Fock or similar approximation in such a way that each particle experiences only the average field of all of the others. Moreover, it will be assumed that H contains no spin-dependent terms, so that the sig- nificant part of every wave function is a scalar function. For example, for an electron in this situation, each wave function can be taken to be the product of an “orbital” function, which is a scalar, with one of two possible spin func- tions, so that the only effect of the electron’s spin is to double the “orbital” degeneracy of each energy eigenvalue. (A development of a theory of spinors along similar lines that enables spin-dependent Hamiltonians to be studied is given, for example, in Chapter 6, Section 4, of Cornwell (1984).) With these assumptions a typical Hamiltonian operator for a particle of mass /1 has the form ti2 d2 d2 d2 я(г) = + v W’ (L1°) 2/z ox2 dy2 oz2 where V(r) is the potential field experienced by the particle. For example, for the electron of a hydrogen atom whose nucleus is located at (9, Я(Г) = + ду2 + ~ е2/^2 + y2 + *2}1/2- (L11) In Equations (1.10) and (1.11) the Hamiltonian is written as Я(г) to empha- size its dependence on the particular coordinate system Oxyz. (b) The invariance of the Hamiltonian operator Let ^({R(T)|t(T)}r) be the operator that is obtained from Я(г) by substi- tuting the components of r' = {R(T)|t(T’)}r in place of the corresponding components of r. For example, if H(r) is given by Equation (1.11), then й2 Я2 Я2 Я2 H((R(T)|t(T)}r) = + —+—(1.12) Hr({R(T)|t(T)}r) can then be rewritten so that it depends explicitly on r. For example, in Equation (1.12), if T is a pure translation xf — x + £i, yf = у + t^z1 = z + ^з, then /i2 Л2 г)2 r)2 #({R(T)|t(T)}r) = -J_(^ + ^L + ^L) 2/z ox2 oy2 oz2 —e2/{(ж + £i)2 + (у + ^)2 + (^ + *з)2}1/2. so that Я({Н(Т)|1(Т)}г)^Я(г), whereas if T is a pure rotation about (9, then a short algebraic calculation gives t2 02 n2 Я2 #({R(T)|t(T)}r) = -^(^2 + “ е2/{ж2 + У2 + *2}1/2>
12 GROUP THEORY IN PHYSICS and hence in this case #({R(T)|t(T)}r) = Я(г). A coordinate transformation T for which H({R(T)|t(T)}r) = tf(r) (1.13) is said to leave the Hamiltonian “invariant”. For the hydrogen atom the above analysis merely explicitly demonstrates the intuitively obvious fact that the system is invariant under pure rotations but not under pure translations. The following key theorem shows how and why group theory plays such a significant part in quantum mechanics. Theorem I The set of coordinate transformations that leave the Hamilto- nian invariant form a group. This group is usually called “the group of the Schrodinger equation”, but is sometimes referred to as “the invariance group of the Hamiltonian operator”. Proof It has only to be verified that the four group axioms are satisfied. Firstly, if the Hamiltonian is invariant under two separate coordinate trans- formations Ti and T2, then it is invariant under their product 7172. (Invari- ance under Ti implies that Я(г") = Я(г'), where r" = {R(7i)|t(7i)}r', and invariance under 7^ implies that Я(г') = Я(г), where r' = {R(72)|t(72)}r, so that Я(г") = Я(г), where, by Equation (1.7), r" = {R(7i72)|t(7i72)}r). Secondly, as noted in Section 2(b), the associative law is valid for all coor- dinate transformations. Thirdly, the identity transformation obviously leaves the Hamiltonian invariant, and finally, as Equation (1.13) can be rewritten as Я(г') = Я({Н(Т)|1(Т)}-1г'), where r' = {R(T)|t(T)}r, if T leaves the Hamiltonian invariant then so does T-1. For the case of the hydrogen atom, or any other spherically symmetric system in which V(r) is a function of |r| alone,The group of the Schrodinger equation is the group of all pure rotations in IR3. (c) The scalar transformation operators P(T) A “scalar field” is defined to be a quantity that takes a value at each point in the space IR3 (in general taking different values at different points), the value at a point being independent of the choice of coordinate system that is used to designate the point. One of the simplest examples to visualize is the density of particles. The concept is relevant to the present consideration because the “orbital” part of an electron’s wave function is a scalar field. Suppose that the scalar field is specified by a function ^(r) when the coordinates of points of IR3 are defined by a coordinate system Ox, Oy, Oz, and that the same scalar field is- specified by a function when another coordinate system O’x', O'y', O'z' is used instead. If r and r' are the position
THE BASIC FRAMEWORK 13 vectors of the same point referred to the two coordinate systems, then the definition of the scalar field implies that ^'(r') = V'(r). (1.14) Now suppose that O'x', O'y', O'z' are obtained from Ox, Oy, Oz by a coordi- nate transformation T, so that r' = {R(T)|t(T’)}r. Then Equation (1.14) can be written as V-/(r') = ^({R(T)|t(T)}-1r'), (1.15) which provides a concrete prescription for determining the function 'ф' from the function ф, namely that ?/(r') is the function obtained by replacing each component of r in ^(r) by the corresponding component of {R(T)|t(T)}-1r'. For example, if ^(r) = x2y3 and T is the pure rotation of Equation (1.3), as {R(T)|t(T)}"1r' = R(T)"1r' = R(T)r' = (xf, yf cos в — z’ sin#, y’ sin# + z’ cos#), then ф'(r') = /2(?/cos# — / sin#)3. It is very convenient in the following analysis to replace the argument r' of ф1 by r (without changing the functional form of ?/). Thus in the above example ?//(r) = x(y cos # — z sin #)3, and Equation (1.15) can be rewritten as ^(^ = ^(7)11(7)}-^). (1.16) As ^(г) is uniquely determined from ^(r) for the coordinate transfor- mation T, ф' can be regarded as being obtained from by the action of an operator P(Tfi which is therefore defined by = Р(Т)ф, or, equivalently, from Equation (1.16) by (P(TW(r) = V'({Rmit(T)}-1r). The typography can be simplified without causing confusion by removing one of the sets of brackets on the left-hand side, giving WW = ^(WXT)}-1!-). (1.17) These scalar transformation operators perform a particularly important role in the application of group theory to quantum mechanics. Their properties will now be established. Clearly P(Ti) = P(T2) only if Ti = T2. (Here P(Ti) = P(T2) means that P(7i)^(r) = Р(Т2)ф(г) for every function ^(r)-) Moreover, each operator P(T) is linear, that is P(T) W(r) + ^(r)} = aP(T)</>(r) + 6P(T)^(r) (1.18)
14 GROUP THEORY IN PHYSICS for any two functions ф(г) and ^(r) and any two complex numbers a and &, as can be verified directly from Equation (1.17); (see Appendix B, Section 4). The other major properties of the operators P(T) are most succinctly stated in the following four theorems. Theorem II Each operator P(T) is a unitary operator in the Hilbert space L2 with inner product (</>, ф) defined by ZOO POO POO / / ф*(г)'ф(г') dx dy dz, (1.19) -OO J — OO 7 — 00 where the integral is over the whole of the space IR3, that is, = (i.2o) for any two functions ф and ф of L2; (see Appendix B, Sections 3 and 4). Proof With r" defined by r" = {R(T)|t(T)} xr, from Equations (1.17) and (1.19) ZOO ЛОО POO / / </>*(r")^(r")^^^. -OO J — OO 7—00 (1-21) However, dxdydz = J dx" dy" dz", where the Jacobian J is defined by J = det dx/dx" dy/dx" dz/dx" dx/dy" dx/dz" dy/dy" dy/dz" dz/dy" dz/dz" As r = R(T)r" + t(T), it follows that dx/dx" = R^T)^ dx/dy" = R(T)i2 etc., so that J = det R(T) = ±1. In converting the right-hand side of Equa- tion (1.21) to a triple integral with respect to ж", ?/", z", there appears an odd number of interchanges of upper and lower limits for an improper rotation, whereas for a proper rotation there is an even number of such interchanges. (For example, for spatial inversion 7, x" = —ж, у" = —?/, z" = —z, so the up- per and lower limits are interchanged three times, while for a rotation through 7Г about Oz the limits are interchanged twice.) Thus in all cases Equation (1.21) can be written as ZOO /»OO /»OO / / ф*(т")ф(т")ах"ау"аг", -OO 7 — oo 7—00 from which Equation (1.20) follows immediately. Theorem III For any two coordinate transformations T\ and Т2, P(TiT2) = P(Ti)P(T2). (1.22) Proof It is required to show that for any function ф(гф P(Ti72)'^(r) = P(7i)P(T2)^(r), where in the right-hand side P(T2) acts first on ф(г) and
THE BASIC FRAMEWORK 15 P(71) acts on the resulting expression. Let <^>(r) = P(T2)0(r), 30 that 0(r) = V’({R(72)|t(72)r1r). Then P(W(r) = 0({R(Tx) |t(Ti)}_1r) = V’({R(T2)|t(T2)}-1{R(Ti)|t(T1)}-1r), the last equality being a consequence of the fact that <^({R(7i)|t(7i)}-1r) is by definition the function obtained from ф(г) by simply replacing the compo- nents of r by the components of {R(Ti)|t(7i)}-1r. Thus, on using Equation (1-9), P(Ti)P(T2)V;(r) = V’({R(TiT2)|t(TiT2)}“1r) = P^T^r). Theorem IV The set of operators P(T) that correspond to the coordinate transformations T of the group of the Schrodinger equation forms a group that is isomorphic to the group of the Schrodinger equation. Proof The product P(7i)P(72)? as defined in the proof of the previous the- orem, may be taken to specify the group multiplication operation, so that the associative law of axiom (b) is satisfied. The previous theorem then implies that group axiom (a) is fulfilled, and with P(E) being the identity operator it also implies that the inverse operator P(T)-1 may be defined by P(T)-1 = P(T-1). Finally, it also indicates that the two groups are isomor- phic. Theorem V For every coordinate transformation T of the group of the Schrodinger equation Q Р(Т)Я(г) = Я(г)Р(Т). (1.23) Proof It has to be established that for any ^(r) and any T of Q P(T){H(r)ip(r)} = Я(г){Р(ЭДг)}. (1.24) Let ф(г) = Я(г)^(г). Then, by Equation (1.17), WW = <^({R(T)|t(T)}-1r) = H({R(T)|t(T)}-1r)^({R(T)|t(T)}-1r) = H({R(T)|t(T)}-1r){P(T)^(r)}, from which Equation (1.24) follows by Equation (1.13). 4 The role of matrix representations Having shown how groups arise naturally in quantum mechanics, in this pre- liminary survey it remains only to introduce the concept of a group represen- tation and to demonstrate that it too has a fundamental role to play.
16 GROUP THEORY IN PHYSICS Definition Representation of a group If each element T of a group Q can be assigned a non-singular d x d matrix Г(Т) contained in a group of matrices having matrix multiplication as its group multiplication operation in such a way that ЦВД = Г(Г1)Г(Т2) (1.25) for every pair of elements T± and T2 of P, then this set of matrices is said to provide a d-dimensional “representation” Г of Q. Example I A representation of the crystallographic point group D4 The group D4 introduced in Example III of Section2 has the following two- dimensional representation: Г(Е) = Mb Г(С2ж) = M2, r(C2J = M3, Г(С22) = M4, Г(С4у) = М5, Г(С4-1) = Ме, Г(С2с) = М7, r(C'2d) = M8, where Mi, M2,.. .are the 2x2 matrices defined in Example III of Section 1. That Equation (1.25) is satisfied can be verified simply by comparing Tables 1.1 and 1.2. It will be shown in Chapter 4 that every group has an infinite number of different representations, but they are derivable from a smaller number of basic representations, the so-called “irreducible representations”. A finite group has only a finite number of such irreducible representations that are essentially different. The representations of the group of the Schrodinger equation are of partic- ular interest. The intimate connection between them and the eigenfunctions of the time-independent Schrodinger equation is provided by the notion of “basis functions” of the representations. Definition Basis functions of a group of coordinate transformations Q A set of d linearly independent functions ^1 (r), ^2 (r), • • •, 7/Wr) forms a basis for a d-dimensional representation Г of Q if, for every coordinate transforma- tion T of Q, d P(W) = E Г(Т)тп^т(г), n = 1, 2, . . . , d. (1.26) rn=l The function ^n(r) is then said to “transform as the nth row” of the repre- sentation Г. The definition implies that not only is each function P(T)^n(r) required to be a linear combination of ^i(r),^2(r),... ,^(r), but the coefficients are required to be equal to specified matrix elements of Г(Т). The rather unusual ordering of row and column indices on the right-hand side of Equation (1.26)
THE BASIC FRAMEWORK 17 ensures the consistency of the definition for every product T1T2, for, according to Equations (1.18), (1.22), (1.25), and (1.26), ТОЩг) = Р(Г1)Р(Т2Ж(г) d m=l d = ^Т(Т2)тпР(Т^т(г) m=l d d = mn Г(Т1) m=l p=l d = P=1 Example II Some basis functions of the crystallographic point group The functions Vh(r) = #,^2(1*) = z provide a basis for the representation Г of D4 that has been constructed in Example I above, as can be verified by inspection. (This set has been deduced by a method that will be described in detail in Chapter 5, Section 1.) Theorem I The eigenfunctions of a d-fold degenerate eigenvalue e of the time-independent Schrodinger equation H(r)V»(r) = e^(r) form a basis for a d-dimensional representation of the group of the Schrodinger equation Q. Proof Let (r), (r), • • •, V\Kr) be a set of linearly independent eigenfunc- tions of H(r) with eigenvalue e, so that Я(ГШ(Г) = e^n(r), n = l,2,...,d, and any other eigenfunction of Я(г) with eigenvalue e is a linear combina- tion of ^i(r), ^2(1*),..., V\/(r). For any transformation T of the group of the Schrodinger equation, Equation (1.23) implies that Я(г){Р(Т)^п(г)} = Р(Т){Я(г)^п(г)} = 6{P(T)^n(r)}, demonstrating that P(T)^n(r) is also an eigenfunction of Я(г) with eigen- value €, so that P(T)^n(r) may be written in the form d Pmn(r) = Г(Т)топ^го(г), n = 1,2,..., d. (1.27) m=l
18 GROUP THEORY IN PHYSICS At this stage the Г(Т)шп are merely a set of coefficients with the m,n and T dependence explicitly displayed. For each T the set Г(Т)тп can be arranged to form a d x d matrix Г(Т). It will now be shown that d Г(Т1Т2)тп = 52 (1-28) P=1 for any two transformations Ti and T% of P, thereby demonstrating that the matrices Г(Т) do actually form a representation of Q. Equation (1.27) then implies that the eigenfunctions Vh(r), ^2(1*),..., V'dW f°rm a basis for this representation. From Equation (1.27), with T replaced by T^Ti and T1T2 in turn, d Р(71Щг) = 52 Г(Т1)т^т(г), (1.29) m=l d Р(Т2)фп(т) = 52г(Т2)рптАр(г), (1.30) Р=1 d mWnW = 52 (1-31) m=l From Equations (1.29) and (1.30) d d Р^Р^Шт) = 52 52 Г(Т1)горГ(Т2)рп^го(г), (1-32) m=l p=l and, as P(Ti)P(72)^n(r) = P(7i72)V>n(r) by Equation (1.22), the right hand sides of Equations (1.31) and (1.32) must be equal. As the functions Vh(r), ^2 (r), • •., ^d(r) have been assumed to be linearly independent, Equation (1.28) follows on equating coefficients of each ^m(r). This theorem implies that each energy eigenvalue can be labelled by a representation of the group of the Schrodinger equation. In Chapter 10 it will be shown that the familiar categorization of electronic states of an atom into “s-states”, “p-states”, “d-states” etc. is actually just a special case of this type of description. More precisely, every s-state eigenfunction is a basis function of particular representation of the group of rotations in three dimensions, the p-state eigenfunctions are basis functions of another representation of that group, and so on. Having established a prima facie case that groups and their representations play a significant role in the quantum mechanical study of physical systems, the next chapters will be devoted to a detailed examination of the structure of groups and the theory of their representations. So far only a brief indication has been given of what can be achieved, but the ensuing chapters will show that the group theoretical approach is capable of dealing with a very wide range of profound and detailed questions.
Chapter 2 The Structure of Groups 1 Some elementary considerations This section will be devoted to some immediate consequences of the definition of a group that was given in Chapter 1, Section 1. As many statements will be made about the contents of various sets, it is convenient to introduce an abbreviated notation in which “71 E S” means “the element T is a member of the set 5” and “71 5” means “the element T is not a member of the set 5”. The associative law of Equation (1.1) implies that in any product of three or more elements no ambiguity arises if the brackets are removed completely. Moreover, they can be inserted freely around any chosen subset or subsets of elements in the product, provided of course that the order of elements is unchanged. The proof that (T1T2)-1 = Tf1!?1 (2.1) for any 71,72 E Q provides some examples of this, for (7Г1Т’Г1)(Т172) = Т2-1(ТГ1Т1)Т2 = Т2-1ет2 = t2-1(st2) = t2-1t2 = e, there being a similar argument for (TiT2)(T,2“1T'1_1). Definition Subgroup A subset 5 of a group Q that is itself a group with the same multiplication operation as Q is called a “subgroup” of Q. By convention, a set may be considered to be a subset of itself, so Q can be regarded as being a subgroup of itself. All other subgroups of Q are called proper subgroups. Obviously the identity E must be a member of every subgroup of Q. Indeed one subgroup of Q is the set {E} consisting only of E. It will be shown in Section 4 that if g and s are the orders of Q and 5 respectively, then gj s must be an integer.
20 GROUP THEORY JN PHYSICS A concise criterion for a subset of a group to be a subgroup is provided by the following theorem. Theorem I If 5 is a subset of a group G such that S'S-1 E S for any two elements S and S' of 5, then S is a subgroup of Q. Proof It has only to be verified that the group axioms (a), (c) and (d) are satisfied by 5, axiom (b) being automatically obeyed for any subset of Q. Putting S' = S gives S'S~1 = E, so E E S and hence axiom (c) is satisfied. Putting S' = E gives S'S-1 = S-1, so S-1 e <S, thereby fulfilling axiom (d). Finally, as S-1 e 5, S'(S-1)-1 = S'S E S, so (a) is also true. Example I Subgroups of the crystallographic point group D4 The group D4 defined in Chapter 1, Section 2 has the following subgroups: (a) s = 1 (i.e. g/s = 8): {E}; (b) 5 = 2 (i.e. <7/5 = 4): {Е,С2ж}, {E,C2y}, {E,C2z}, {E,Cic}, {E,C2d}; (c) 5 = 4 (i.e. g/s = 2): {E, C2x, C2y, C2 J, {E, C2y, C4y, Е/Д {E, C2y, C2c, C2d}; (d) s = 8 (i.e. g/s = l): {E,C2x,C2y,C2z,C4y,C4y ,C2c,C2d}- The following theorem displays an interesting property of multiplication in a group. Theorem II For any fixed element T' of a group P, the sets {T'T;T E G} and {TT'; T E G} both contain every element of G once and only once. (Here {T'T; T E 5} denotes the set of elements T'T where T varies over the whole of G- For example, in the special case in which G is a finite group of order g with elements Ti, T2,..., Tg and T' = Tn, this set consists of TnTi,TnT2,..., TnTg. The interpretation of {TT'; T E G} is similar. The theorem is often called the “Rearrangement Theorem”, as it asserts that each of the two sets {T'T;T E G} and {TT'; T E G} merely consists of the elements of G rearranged in order.) Proof An explicit proof will be given for the set {T'T; T E <y}, the proof for the other set being similar. If T" is any element of P, then with T defined by T = it follows that T'T = T". Thus {T'T;T E G} certainly contains every element of G at least once. Now suppose that {T'T; T EG} contains some element of G twice (or more), i.e. for some Ti,T2 E G, T'T± = T'T2, but Ti T2. However, these statements are inconsistent, for premultiplying the first by (T')-1 gives Ti = T2, so no element of G appears more than once in {T'T; T E G}- The Rearrangement Theorem implies that in the multiplication table of a finite group every element of the group appears once and only once in every
THE STRUCTURE OF GROUPS 21 row, and once and only once in every column. This provides a useful check on the computation of the multiplication table. Tables 1.1 and 1.2 exemplify these properties. 2 Classes Whereas in ordinary everyday language the word “class” is often synonymous with the word “set”, in the context of group theory a class is defined to be a special type of set. In fact it is a subset of a group having a certain property which causes it to play an important role in representation theory, as will be shown in Chapter 4. As a preliminary it is necessary to introduce the idea of “conjugate ele- ments” of a group. Definition Conjugate elements An element T' of a group Q is said to be “conjugate” to another element T of Q if there exists an element X of Q such that T’ = XTX-\ (2.2) If T’ is conjugate to T, then T is conjugate to T", as Equation (2.2) can be rewritten as T = X-1T'(X-1)-1. Moreover, if T, T' and T" are three elements of Q such that T' and T" are both conjugate to T, then T' is conjugate to T". This follows because there exist elements X and Y of Q such that T' = XTX-1 and T" = YTY~\ so that T' = Х(У-1Т"У)Х"1 = (ХУ“1)Т"(ХУ“1)“1 (by Equation (2.1)), which has the form of Equation (2.2) as ХУ-1 e Q. It is therefore permissible to talk of a set of mutually conjugate elements. Definition Class A class of a group Q is a set of mutually conjugate elements of Q. (For extra precision this is sometimes called a “conjugacy class”.) A class can be constructed from any T e Q by forming the set of products XTX~x for every X E Q, retaining only the distinct elements. This class contains T itself as T = ETE~T Example I Classes of the crystallographic point group D4 For the group D4 this procedure when applied to C?x gives (on using Table 1.2): ХС2хХ~у = C2xiorX = E,C2x,C2y,C2z, XC2xX-! = C2z^rX = C:y1,C4y,C2c,C2d. Thus {Cix, C2z} is one of the classes of D4. The same class would have been found if the procedure had been applied to C^z- D4 has four other classes,
22 GROUP THEORY IN PHYSICS namely {E}, {Ciy}, {64^, C\y} and {C2c,C2d}, which may be deduced in a similar way. The properties of classes are conveniently summarized in the following three theorems. Theorem I (a) Every element of a group Q is a member of some class of Q. (b) No element of Q can be a member of two different classes of Q. (c) The identity E of Q always forms a class on its own. Proof (a) As noted above, for any T e Q, ETE-1 = T, so that T is in the class constructed from itself. (b) Suppose that T E Q is a member of a class containing T' and is also a member of a class containing T". Then T is conjugate to T' and T", so T' and T" must be conjugate and must therefore be in the same class. (c) For any X e Q, XEX~r = XX~r = E,soE forms a class on its own. Theorem II If Q is an Abelian group, every element of Q forms a class on its own. Proof For any T and X of an Abelian group Q XTX-1 = XX^T = ET = T, so T forms a class on its own. Theorem III If Q is a group consisting entirely of pure rotations, no class of Q contains both proper and improper rotations. Moreover, in each class of proper rotations all the rotations are through the same angle. Similarly, in each class of improper rotations the proper parts are all through the same angle. Proof If T and T' are two pure rotations in the same class, Equations (1.4) and (2.2) imply that R(T') = R(X)R(T)R(X)-\ so that detR(T') = detR(T) (2.3) and tr R(T') = tr R(T) (2-4)
THE STRUCTURE OF GROUPS 23 (see Appendix A). Equation (2.3) shows that T and T' are either both proper or are both improper. Moreover, for any proper rotation T through an angle в (in the right- or left-hand screw sense) tr R(T) = 1 + 2 cos 9 (cf. Equation (10.4)). Equation (2.4) then implies that all proper rotations in a class are through the same angle 9. Finally, by expressing any improper rotation T as the product of the spatial inversion operator I with a proper rotation through an angle 0, it follows that trR(T) = —{1+2 cos 0}, so all proper parts involved in a class are through the same angle 9. It should be noted that the converse of the last theorem is not necessarily true, in that there is no requirement for all rotations of the same type to be in the same class. Indeed, in the above example of the point group D4, the proper rotations Czx and Czy are in different classes, even though they are rotations through the same angle 7Г. 3 Invariant subgroups The main object of this and the following section is to introduce two concepts that are involved in the construction of factor groups. Definition Invariant subgroup A subgroup 5 of a group Q is said to be an “invariant” subgroup if XSX-1 G S (2.5) for every SeS and every X e Q. Invariant subgroups are sometimes called “normal subgroups” or “normal divisors”. Because of the occurrence of the same forms in Equation (2.2) and Condition (2.5), there is a close connection between invariant subgroups and classes. Theorem I A subgroup S of a group Q is an invariant subgroup if and only if S consists entirely of complete classes of Q. Proof Suppose first that 5 is an invariant subgroup of Q. Then if S is any member of S and T is any member of the same class of Q as 5, by Equation (2.2) there exists an element X of Q such that T = XSX~x. Condition (2.5) then implies that T e 5, so the whole class of Q containing S is contained in S. Now suppose that S consists entirely of complete classes of P, and let S be any member of S. Then the set of products XSX~x for all X e G forms the class containing 5, which by assumption is contained in 5. Thus XSX~x e S for all S e S and X e G, so S is an invariant subgroup of p. This theorem provides a very easy method of determining which of the
24 GROUP THEORY IN PHYSICS subgroups of a group are invariant when the classes have been previously calculated. Example I Invariant subgroups of the crystallographic point group D4 For the crystallographic point group D4 it follows immediately from the lists of subgroups and classes given in Sections 1 and 2 that the invariant subgroups are {£}, {E, C2y}, {E, C2x, C2y, C2z}, {E, C2y, Ciy, C^}, {E, C2y, C2c, C2d}, and D4 itself. (The subgroup {E, Czx} is not an invariant subgroup as Czx is part of a class {C2X,C2z} that is not wholly contained in the subgroup. The same is true of {E, (?2C} and {E, <?2d}.) For every Q the trivial subgroups {E} and Q are both invariant subgroups. 4 Cosets Definition Coset Let S be a subgroup of a group Q. Then, for any fixed T e Q (which may or may not be a member of <S), the set of elements ST, where S varies over the whole of 5, is called the “right coset” of S with respect to T, and is denoted by ST. Similarly, the set of elements TS is called the “left coset” of S with respect to T and is denoted by TS. In particular, if S is a finite subgroup of order s with elements Si, S2, ..., Ss, then ST is the set of s elements SiT, S2T,..., SsT, and TS is the set of s elements TSi,TS2,..., TSs. In the following discussions two sets will be said to be identical if they merely contain the same elements, the ordering of the elements within the sets being immaterial. Example I Some cosets of the crystallographic point group D4 Let Q be D4 and let S = {E, C2X}- Then from Table 1.2 the right cosets are SE = SC2X = {E,C2cJ, = {C22/,C2J, SC4y = SC2d = {C4yiC2d}i SC4y = SC2c = {C4y,C2e}, and the left cosets are ES = C2xS C2vS = C2zS CiyS = C2cS C^s = c2ds {E,c2x}, {C2y,C2z}, {C*4y, C2c} 1 It should be noted that C4yS SC±y and C4yS SC4y.
THE STRUCTURE OF GROUPS 25 This example shows that the right and left cosets ST and TS formed from the same element T e G are not necessarily identical. The properties of cosets are summarized in the following two theorems. The first theorem is stated for right cosets, but every statement applies equally to left cosets. It is worth while checking that the above example of the point group D4 does satisfy all the assertions of this theorem. Theorem I (a) If T e 5, then ST = S. (b) If T qL S, then ST is not a subgroup of Q. (c) Every element of G is a member of some right coset. (d) Any two elements ST and S'T of ST are different, provided that S S'. In particular, if S is a finite subgroup of order s, ST contains s different elements. (e) Two right cosets of S are either identical or have no elements in common. (f) If T' e ST, then ST' = ST. (g) If G is a finite group of order g and S has order s, then the number of distinct right cosets is g/s. Proof (a) If T e S, the Rearrangement Theorem of Section 1 applied to S consid- ered as a group shows that ST is merely a rearrangement of S. (b) If ST is a subgroup of G, it must contain the identity E, so there must exist an element S E S such that ST = E. This implies T = 5-1, so T E S. Thus if T S, ST cannot be a subgroup of Q. (c) For any T E S, as T = ET and E e S, it follows that T E ST. (d) Suppose that ST = S'T and S S'. Post-multiplying by T-1 gives S = S', a contradiction. (e) Suppose that ST and ST' are two right cosets with a common element. It will be shown that ST = ST'. Let ST = S'T' be the common element of ST and ST'. Here 5,5' E S. Then T'T~r = (5')"15, so T'T~r E 5, and hence by (a) 5(T'T-1) = S. As 5(T'T-1) is the set of elements of the form ST'T~\ the set obtained from this by post-multiplying each member by T consists of the elements ST', that is, it is the coset ST'. Thus ST = ST'. (f) As in (с), T' E ST'. If T E ST' then ST' and ST have a common element and must therefore be identical by (e).
26 GROUP THEORY IN PHYSICS (g) Suppose that there are M distinct right cosets of S. By (d) each contains s different elements, so the collection of distinct cosets contains Ms different elements of Q. But by (c) every element of Q is in this collection of distinct cosets, so Ms = g. The property (f) is particularly important. It shows that the same coset is formed starting from any member of the coset. All members of a coset therefore appear on an equal footing, so that any member of the coset can be taken as the “coset representative” that labels the coset and from which the coset can be constructed. For example, for the right coset {C^Cid} of the point group D4, the coset representatives could equally well be chosen to be C4y or Cid- As the number of distinct right cosets is necessarily a positive integer, property (g) demonstrates that s must divide g, as was mentioned in Section 1. Theorem II The right and left cosets of a subgroup S of a group Q are identical (i.e. ST = TS for all T E Q) if and only if 5 is an invariant subgroup of Q. Proof Suppose that S is an invariant subgroup. It will be shown that if T' E ST then T' E TS. (A similar argument proves that if T' E TS then T' E ST, so, on combining the two, it follows that ST = TS.) If T' E ST there exists an element S of S such that T' = ST. Then T~rT' = T~rST, which is a member of S as S is an invariant subgroup. Thus T~rT' E S, so T' = T(T-1T') must be a member of TS. Now suppose that ST = TS for every T E Q. This implies that for any S E S and any T E Q there exists an S' E S such that TS = S'T, so TST~r = S' and hence TST~r E S. Thus S is an invariant subgroup of Q. Of course, in the above example concerning the point group D4, the sub- group 5 = {E, C2X} was carefully chosen so as not to be an invariant subgroup, in order to demonstrate that right and left cosets are not always identical. 5 Factor groups Let S be an invariant subgroup of a group Q. Each right coset of S can be considered to be an “element” of the set of distinct right cosets of S, the internal structure of each coset now being disregarded. With the following definition of the product of two right cosets, the set of cosets then forms a group called a “factor group”. Definition Product of right cosets The product of two right cosets ST± and ST2 of an invariant subgroup 5 is defined by ST1.ST2 =S(71T2). (2-6)
THE STRUCTURE OF GROUPS 27 Proof of consistency It will be shown that Equation (2.6) provides a mean- ingful definition, in that, if alternative coset representatives are chosen for the cosets on the left-hand side of the equation, then the coset on the right-hand side remains unchanged. Suppose that T{ and T% are alternative coset repre- sentatives for ST\ and ST2 respectively, so that T[ G ST± and T% G ST2. It has to be proved that Sf/T^T^) = ^(TiTla). As T[ G STi and T% G ST^ there exist S,S' G S such that T[ = STr and Tf, = S'T^ Then T[T^ = ST1S'T2. But 7i5' G TlS, so, as 5 is an invariant subgroup, TiS' G ST±. Consequently there exists an 5" e 5 such that 7\5' = 5"7\. Then ТЩ = (55")(7iT2), so that T^T^ G S(TiT2) and hence, by property (f) of the first theorem of Section 4, S(T{T£) = SlTUty. Theorem I The set of right cosets of an invariant subgroup 5 of a group Q forms a group, with Equation (2.6) defining the group multiplication opera- tion. This group is called a “factor group” and is denoted by Q/S. Proof It has only to be verified that the four group axioms are satisfied. (a) By Equation (2.6), the product of any two right cosets of 5 is itself a right coset of 5 and is therefore a member of Q/S. (b) The associative law is valid for coset multiplication because, if 5T, ST' and ST" are any three right cosets, (ST.ST').ST" = 5(7T').5T" = 5((7T')T") and ST.(ST'.ST") = ST.S(T'T") = 5(T(T'T")), where the two cosets on the right-hand sides are equal by virtue of the associative law (7T')T" = T(T'T") for Q. (c) The identity element of Q/S is SE(= 5), as for any right coset SE.ST = S(ET) =ST = S(TE) = ST.SE. (d) The inverse of ST is 5(T-1), as 5T.5(T“1) = 5(TT“1) = SE = 5(T“1T) = 5(T“1).5T. The coset 5(T-1) is a member of Q/S as T-1 e Q. If Q is a finite group of order g and S has order s, part (g) of the first theorem of Section 4 shows that there are g/s distinct right cosets. Thus Q/S is a group of order g/s with elements 57i, 57^,..., STS1 (Ti, 7^,... ,TS being a set of coset representatives). As S itself is one of the cosets, one can take Ti = E.
28 GROUP THEORY IN PHYSICS SE sc2x SC^y SC2c SE SE sc2x sc4y SC2c sc2x sc2x SE SC2c sc4y SC4y sc4y SC2c SE sc2x SC2c SC2c sc4y sc2x SE Table 2.1: Multiplication table for the factor group Q/S, where Q is the crystallographic point group D4 and S = {E, Czy}. Example I A factor group formed from the crystallographic point group D4 Let Q be D4 and let S — {E, Czy}, which is an invariant subgroup of Q (see Example I of Section 3). Then Q/S is a group of order 4 with elements SE = SCzy = {E,Czy}, SCzx = SSC2z = {C2x,C2z}, SC4y = SC^ = {C4y,C4y}, SC2c = sc2d = {c2c,c2d}, whose multiplication table is given in Table 2.1. (Here it should be noted for example that С2ХСду = (?2d, so SCix-SC^y = SCid = SC^cY 6 Homomorphic and isomorphic mappings Let Q and Q' be two groups. A “mapping” ф of Q onto P' is simply a rule by which each element T of Q is assigned to some element T' = ф(Т) of with every element of Q' being the “image” of at least one element of Q. If ф is" a one-to-one mapping, that is, if each element T' of Q' is the image of only one element T of P, then the inverse mapping ф~г of Qf onto Q may be defined by ф~\Т') = T if and only if T' = ф(Т). Definition Homomorphic mapping of a group Q onto a group Q' If ф is a mapping of a group Q onto a group P' such that Ж1Ж) = Л) (2.7) for all 71,72 G P, then ф is said to be a “homomorphic” mapping. On the right-hand side of Equation (2.7) the product of Ti with Тг is evaluated using the group multiplication operation for P, whereas on the left-hand side the product of ф(Тф) with ф^Т^) is obtained from the group multiplication operation for Qf. Although these operations may be different, there is no need to introduce any special notations to distinguish between them, because the relevant operation can always be deduced from the context and there is really no possibility of confusion. Example I A homomorphic mapping of the point group D4 Let Q be D4 and let P' be the group of order 2 with elements +1 and — 1,
THE STRUCTURE OF GROUPS 29 with ordinary multiplication as the group multiplication operation. Then ф(Е) = ф(С2у) = ф(С2х) = ф(С2г) = +1, ф(С4у) = ф(С^) = ф(С2с) = ф(С2Л) = -1, is a homomorphic mapping of Q onto P', as may be confirmed by examination of Table 1.2. For example, ф(С2х)Ф(С2С) — (+1)(—1) = —1, while Table 1.2 gives ф(С2хС2с) = ф(С±у) = -1. Clearly, if g and д' are the orders of Q and Q' respectively, then g > д'. Actually, the First Homomorphism Theorem, which will be proved shortly, implies that if g and gf are both finite, then g/g1 must be an integer. One major example of a homomorphic mapping has already been encountered in the concept of a representation of a group. Indeed, the definition in Chapter 1, Section 4 can now be rephrased as follows: Definition Representation of a group Q If there exists a homomorphic mapping of a group Q onto a group of non- singular d x d matrices Г(Т) with matrix multiplication as the group multi- plication operation, then the group of matrices Г(Т) forms a d-dimensional representation Г of Q. There is no requirement in the definition of a homomorphic mapping that the mapping should be one-to-one. However, as such mappings are particu- larly important, they are given a special name: Definition Isomorphic mapping of a group Q onto a group If ф is a one-to-one mapping of a group Q onto a group Q' of the same order such that Ж1)Ж2) = Ф{Т\Т1) TUT2 EG, then ф is said to be an “isomorphic” mapping. In the case of representations, if the homomorphic mapping is actually isomorphic, then the representation is said to be “faithful”. Clearly, if ф is an isomorphic mapping of Q onto Q', then the inverse mapping ф~г is an isomorphic mapping of gf onto g. (There is no analogous result for general homomorphic mappings, as ф~г is only well defined when ф is a one-to-one mapping.) Although two isomorphic groups may differ in the nature of their elements, they have the same structure of subgroups, cosets, classes, and so on. Most important of all, isomorphic groups necessarily have identical representations. The following theorem clarifies various aspects of homomorphic mappings. As it is the first of a series of such theorems, it is often called the “First Homomorphism Theorem”, but the others in the series will not be needed in this book.
30 GROUP THEORY IN PHYSICS Definition Kernel /С of a homomorphic mapping Let ф be a homomorphic mapping of a group Q onto a group Q'. Then the set of elements T e Q such that = E', the identity of is said to form the “kernel” /С of the mapping. Theorem I Let ф be a homomorphic mapping of Q onto p', and let /С be the kernel of this mapping. Then (а) /С is an invariant subgroup of Q\ (b) every element of the right coset /СТ maps onto the same element ф(Т) of P', and the mapping в thereby defined by 0(£T) = ф(Т) (2.8) is a one-to-one mapping of the factor group P//C onto P'; and (с) в is an isomorphic mapping of P//C onto Q’. Proof See, for example, Chapter 2, Section 6, of Cornwell (1984). One consequence of the theorem is that every element of Q' is the image of the same number of elements of Q. This has the further implication that the mapping is an isomorphism if and only if /С consists only of the identity Eof P. In the special case in which Q' is identical to Q (so that ф is a mapping of Q onto itself), an isomorphic mapping is known as an “automorphism”. For each X e Q the mapping фх of Q onto itself defined by фх(Т) = XTX-1 is an automorphism, as it is certainly one-to-one and Фх(Т.)фх(Т2) = (ХТ.Х-^ХЪХ-1) = ХЦУГ^Х-1 = фхрМ) for all 7i, 7*2 e Q. Such a mapping is called an “inner automorphism”, and any automorphism that is not of this form is known as an “outer automorphism”. The whole theory of spin for electrons and other elementary particles in non-relativistic quantum mechanics is based on the following theorem. Theorem II There exists a two-to-one homomorphic mapping of the group SU(2) onto the group SO(3). If u E SU(2) maps onto R(u) E SO(3), then R(u) = R(—u), and the mapping may be chosen so that R(u)jk = |tr {o-jUCTfeU-1} (2.9)
THE STRUCTURE OF GROUPS 31 for j, к = 1,2,3, where (2-Ю) are the Pauli spin matrices. The kernel /С of the mapping consists only of I2 and — Proof See, for example, Chapter 3, Section 5, of Cornwell (1984). 7 Direct products and semi-direct products of groups Although the abstract construction of direct product groups appears at first sight rather artificial, a number of examples of groups having this structure occur naturally in physical problems. Let Pi and P2 be any two groups, and suppose that E\ and E2 are the identities of Pi and P2 respectively. Consider the set of pairs (Ti,72), where Ti e Pi and T2 E P2, and define the product of two such pairs (Ti,^) and CM)by (Т1,Т2)(Т{,^) = (Т1Г1',ВД) (2.11) for all Ti,7i E Pi and T2, 7^ E Q2- Theorem I The set of pairs (Ti,72) (for Ti E Pi, T2 E P2) forms a group with Equation (2.11) as the group multiplication operation. This group is denoted by Pi 0 P2, and is called the “direct product of Pi with p2”. Proof All that has to be verified is that the four group axioms of Chapter 1, Section 1, are satisfied. By Equation (2.11), the product of any two pairs of Pi 0 P2 also a member of Pi 0 P2, so axiom (a) is fulfilled. Axiom (b) is observed, as {(T1,T2)(T',T')}(T1",T'') = ((Т1Т1')Т1",(Т2Г')Т'') and (T1,T2){(T[,T^(T[',T^} = the pairs on the right-hand sides being equal because the associative law applies to Pi and P2 separately. The identity of Pi 0 P2 is (Ei, E2), as for all Ti E Pi and T2 E Q2 (ТъТ2\ЕъЕ2) = (E1,E2)(T1,T2) = (T1,T2). Finally, the inverse of (Ti, T2) is (Ef1, T^-1), which is also a member of Pi0p2- If Pi and P2 are finite groups of orders #i, and g2 respectively, then Pi 0P2 has order #i#2-
32 GROUP THEORY IN PHYSICS The properties of 0 P2 are best presented in the form of a theorem (all the assertions of which have trivial proofs). Theorem II (a) Pi 0P2 contains a subgroup consisting of the elements (Ti, E2), Ti E Pi, that is isomorphic to Pi, the isomorphic mapping being ф(Т\, E2) = Ti. (b) Pi 0P2 contains a subgroup consisting of the elements (Ei, T2), T2 E P2, that is isomorphic to P2, the isomorphic mapping being ^(#1, T2) = T2. (c) The elements of these two subgroups commute with each other, that is (Ti,E2)(Ei,T2) = (Ei,T2)(Ti,E2) = (T1,T2) for all Ti E Pi and T2 E P2. (d) These two subgroups have only one element in common, namely the identity (Ei,E2). (e) Every element of Pi 0 P2 is the product of an element of the first sub- group with an element of the second subgroup. That is, for all Ti E Pi and T2 6 P2, (Ti,T2) = (Ti,E2)(Ei,T2). As isomorphic groups have identical structures, it is natural to now extend the definition of a direct product. Enlarged definition Direct product group A group P' is said to be a “direct product group” if it is isomorphic to a group Pi 0 P2 constructed as in the first theorem above. With this extension the elements of a direct product group need no longer be in the form of pairs. Such a group can be identified by the following theorem, which is essentially the converse of that immediately above. Theorem III If a group P' possesses two subgroups and P2 such that (a) the elements of P^ commute with the elements of P£, (b) Pi and P2 have only the identity element in common, and (c) every element of P' can be written as a product of an element of P^ with an element of p£, then P' is a direct product group that is isomorphic to P^ 0 p£. Proof See, for example, Chapter 2, Section 7, of Cornwell (1984).
THE STRUCTURE OF GROUPS 33 Example I The group 0(3) as a direct product group The group 0(3) is isomorphic to S0(3) ®g2, where g2 is the matrix group of order 2 consisting of the matrices I3 and —13, as the properties (a), (b) and (c) of the preceding theorem are obviously satisfied. As 0(3) is isomorphic to the group of all rotations in three dimensions, and as S0(3) is isomorphic to the subgroup of proper rotations (see Chapter 1, Section 2), this implies that the group of all rotations is isomorphic to the direct product of the group of proper rotations and the group {E, 1} consisting of the identity transformation E and the spatial inversion operator I. It should be observed that condition (a) of the last theorem can be replaced by an equivalent condition (a'), which reads: (a) and gf2 are both invariant subgroups of Qf. (Obviously (a) implies (a')- Conversely, if (a') is true, then for any T[ E g{ and Ц E gf2, T'T'(T')"1 = T£ E Sf2. Similarly (Т^Т'Щ = T" E Pf so that (T^-^'T^T')-1 = T'^T')-1 - (T^)-1^'. As T['(T')”1 e Pi and (T^)-1^' e P£, (b) implies that T[ = Tf and T2 = T2 . Thus (Т^~1Т^ = T[ for all T{ E Pi and T2 E g2, so that Pi and P£ commute.) The notion of a semi-direct product group P' is essentially a generalization of that of a direct product group in which conditions (b) and (c) of the last theorem are retained intact but condition (a') is weakened to the requirement that only Pi must be an invariant subgroup, but Pf although remaining a subgroup of P', need not be invariant. Definition Semi-direct product group A group P' is said to be a “semi-direct product group” if it possesses two subgroups Pi and P£ such that (a) Pi is an invariant subgroup of P'; (b) Pi and P2 have only the identity element in common; and (c) every element of P' can be written as a product of an element of Pi with an element of Pf P' may then be said to be isomorphic to Pi®p2- As in the special case of a direct product group, the requirement (b) always implies that the decomposition (c) is unique. Example II The Euclidean group oflR3 as a semi-direct product group The Euclidean group P' of IR3 is defined to be the group of all linear coordi- nate transformations T, with Equation (1.7) giving the group multiplication operation. Let Pi be the subgroup of pure translations and Pi the subgroup of pure rotations. Then for any T± E Pi and any T G P', from Equations (1.7) and (1.8), {R(T)|t(T)}{l I ttTOHRCOItCT)}-1 = {1 I R(T)t(7i)},
34 GROUP THEORY IN PHYSICS so that is an invariant subgroup of Q'. Moreover, for any T e S', {R(T)|t(T)} = {1 | t(T)}{R(T) | 0}, so that requirement (c) is also satisfied, while (b) is obvious. Thus Q' is isomorphic to A further important set of examples is provided by the symmorphic crys- tallographic space groups. These will be discussed in detail in Chapter 7. Although it is possible to give an abstract construction of a semidirect product of certain groups in terms of pairs of elements from the two groups, the procedure is much more elaborate than for the direct product (Lomont 1959, page 29), Fortunately, all the physically important examples of groups having a semi-direct product structure occur naturally, so this abstract construction will be omitted here.
Chapter 3 Lie Groups It is now time to formulate a definition of a Lie group and to describe some of the major properties of such groups. Readers whose interests lie only in the applications to solid state physics (where only finite groups appear) may safely omit this chapter. 1 Definition of a linear Lie group A Lie group embodies three different forms of mathematical structure. Firstly, it satisfies the group axioms of Chapter 1 and so has the group structure described in Chapter 2. Secondly, the elements of the group also form a “topological space”, so that it may be described as being a special case of a “topological group”. Finally, the elements also constitute an “analytic man- ifold”. Consequently a Lie group can be defined in several different (but equivalent) ways, depending on the degree of emphasis that is being accorded to the various aspects. In particular, it can be defined as a topological group with certain additional analytic properties (Pontrjagin 1946, 1986) or, alter- natively, as an analytic manifold with additional group properties (Chevalley 1946, Adams 1969, Varadarajan 1974, Warner 1971). Both of these formu- lations involve the introduction of a series of ancillary concepts of a rather abstract nature. Very fortunately, every Lie group that is important in physical problems is of a type, known as a “linear Lie group”, for which a relatively straightforward definition can be given. As will be seen, this definition is both precise and simple, in that it involves only familiar concrete objects such as matrices and contains no mention of topological spaces or analytic manifolds. (Readers who are interested in the general definition of a Lie group in terms of analytic manifolds may, for example, find this formulation in Appendix J of Cornwell (1984).) The basic feature of any Lie group is that it has a non-countable number of elements lying in a region “near” its identity and that the structure of this region both very largely determines the structure of the whole group and
36 GROUP THEORY IN PHYSICS is itself determined by its corresponding real Lie algebra. To ensure that this is so, the elements in this region must be parametrized in a particular analytic way. Of course, to say that certain elements are “near” the identity means that a notion of “distance” has to be composed, and it is here that the complications of the general treatment start. However, all the Lie groups of physical interest are “linear”, in the sense that they have at least one faithful finite-dimensional representation. This representation can be used to provide the necessary precise formulation of distance and to ensure that all the other topological requirements are automatically observed. Definition Linear Lie group of dimension n A group Q is a linear Lie group of dimension n if it satisfies the following conditions (А), (В), (C) and (D): (A) Q must possess at least one faithful finite-dimensional representation Г. Suppose that this representation has dimension m. Then the “distance” d(T, Tf) between two elements T and T' of Q may be defined by m m d(T, T') = +{£ £ I r(T)Jfe - v(T')jk |2}V2. (3.1) j=\ k=l (This distance function d(T,Tf) will be called the “metric”.) Then (i) d(T',T) = d(T,T'); (ii) d(T\T)=O; (iii) d(T,T')>OifT^T'; (iv) if T, Tf and T" are any three elements of P, d(T, T") < d(T, T') + d(T', T"), all of which are essential for the interpretation of d(T, Tf) as a distance. (The choice of this metric implies that the group is being endowed with the topology of the m2-dimensional complex Euclidean space (Cm (see Example II of Appendix B, Section 2).) The set of elements T of Q such that d(T,E) < b, where b is positive real number, is then said to “lie in a sphere of radius d centred on the identity E”, which will be denoted by Such a sphere will be sometimes referred to as a “small neighbourhood” of E. (B) There must exist а б > 0 such that every element T of Q lying in the sphere of radius 6 centred on the identity can be parametrized by n real parameters ^1,^2? • • • -Li (no two such sets of parameters corre- sponding to the same element T of 6), the identity E being parametrized by Xi = x2 — ... — xn = 0.
LIE GROUPS 37 Thus every element of Ms corresponds to one and only one point in an n-dimensional real Euclidean space IRn, the identity E corresponding to the origin (0,0,, 0) of IRn. Moreover, no point in IRn corresponds to more than one element T in M<$. (C) There must exist a tj > 0 such that every point in IRn for which ,2 (3-2) corresponds to some element T in M$. The set of point elements T so obtained will be denoted by Rrr Thus Rr] is a subset of and there is a one-to-one correspondence between elements of Q in Rrn and points in IRn satisfying Condition (3.2). The final set of conditions ensures that in terms of this parametrization the group multiplication operation is expressible in terms of analytic functions. Let T(rri, x^ ..., хп) denote the element of Q correspond- ing to a point satisfying Condition (3.2) and define Г(ж1, ж2, • • •, xn) by Г(ж1, ж2, • • •, xn) = Г(Т(ж1, ж2, • • •, жп)) for all (#1, ж2,..., xn) satisfy- ing Condition (3.2). (D) Each of the matrix elements of Г(ж1,ж2,... , жп) must be an analytic function of £Ei, ж2,..., for all (#i, #2, • • •, Xn) satisfying Condition (3.2). The term “analytic” here means that each of the matrix elements Tjk must be expressible as a power series in xi — х®, j;2 — x9>,..., xn — х)ъ for all (ж?, x®)..., x^f) satisfying Condition (3.2). This implies that all the derivatives dTjk/dxP, d^Tjk/dxPdx4 etc. must exist for all j, к = 1,2, ...,m at all points satisfying Condition (3.2), including in particular the point (0,0,... ,0) (Fleming 1977). In particular one can define the n m x m matrices ai, a2,..., an by (ap)jfc — (&Pjk/dxP}X1—X2='''=Xn=Q . (3-3) These conditions together imply the following very important theorem. Theorem I The matrices ai, a2,..., an defined by Equation (3.3) form the basis for a n-dimensional real vector space. Proof See, for example, Chapter 3, Section 1, of Cornwell (1984). It should be noted that, although ai,a2,... ,an form the basis of a real vector space, there is no requirement that the matrix elements of these ma- trices need be real. (This point is demonstrated explicitly in Example III.)
38 GROUP THEORY IN PHYSICS It will be shown in Chapter 8 that the matrices ai,a2,... , an actually form the basis of a “real Lie algebra”, a vital observation on which most of the subsequent theory is founded. However, the rest of the present chapter will be devoted to “group theoretical” aspects of linear Lie groups. The above definition requires a parametrization only of the group elements belonging to a small neighbourhood of the identity element. In some cases this parametrization by a single set of n parameters a?i, 372> • • •, xn is valid over a large part of the group or even over the whole group, but this is not essential. In Section 2 it will be shown that the whole of the “connected” subgroup of a linear Lie group of dimension n can be given a parametrization in terms of a single set of n real numbers which will be denoted by ?/i, y^..., yn- However, this latter parametrization is not required to satisfy all the conditions of the above definition, and so need bear little relation to the parametrization by 371, 372, • • • ч Xn • The following examples have been chosen because they illustrate all the essential points of the definition without involving any heavy algebra. Example I The multiplicative group of real numbers As in Example I of Chapter 1, Section 1, let Q be the group of real numbers t (t 0) with ordinary multiplication as the group multiplication operation, the identity E being the number 1. Q has the obvious one-dimensional faithful representation T(t) = [t], so condition (A) is satisfied and the metric d of Equation (3.1) is given by d(t,tf) = \t — t'\. In particular, d(t, 1) = \t — 1|. Let 6 = | so that | < t < 2 for all £ in M$. A convenient parametrization for t e Ms is then t = expxi. (3.4) As required in (B), the identity 1 corresponds to 371 = 0. Condition (C) is obeyed with у = log |, as x? < (log |)2 implies | < exp 371 < |. By Equation (3.4) Г(ж1) = exprri, which is certainly analytic, so that condition (D) is satisfied. Thus Q is a linear Lie group of dimension 1. It should be noted that Equation (3.3) implies that ai = [1], thereby confirming the first theorem above. It is significant that the parametrization in Equation (3.4) extends to all t > 0 (with — oo < xi < +oo) and that this set forms a subgroup of Q. Moreover, every group element t such that t < 0 can be written in the form t = (—1) exp 371 for some xi. Example II The groups 0(2) and SO(2) 0(2) is the group of all real orthogonal 2x2 matrices A, SO(2) being the subgroup for which det A = +1. If A e 0(2), Г (A) = A provides a faithful finite-dimensional representa- tion. The orthogonality conditions AA = AA = 1 require that G4n )2 + (^12)2 = G^n)2 + (^2i)2 = (^2i)2 + (>^22)2 = (^-12)2 + (A22)2 = 1 (3.5)
LIE GROUPS 39 and A11A21 + ^4-22^4-12 = АцА12 + ^4-22^4-21 = 0- (3.6) Equations (3.5) imply that (Лц)2 = (A22)2 and (A12)2 = (A21)2, so that there are only two sets of solutions of Equations (3.6), namely: (i) Ац = A22 and A12 = — A21. Equations (3.5) imply that det A = +1, i.e. A e SO(2). Moreover, from Equations (3.5), d(A, 1) = 2(1 — An)1/2. (ii) Ац = — A22 and A12 = A21. In this case det A = —1 and d(A, 1) = 2. With the choice 8 = condition (B) requires the parametrization of part of set (i) but it is not necessary to include set (ii), as it is completely outside M$. A convenient parametrization is A = Г(А) = COS£E1 sinrri — sinrri COS£E1 (3.7) Clearly xi = 0 corresponds to the group identity 1 and the dimension n is 1. Every point of IR1 such that ж2 < (тг/З)2 gives a matrix A in so con- dition (C) is satisfied. In fact the parametrization of Equation (3.7) extends to the whole of the set (i) with —7г < Xi < 7Г, that is, to the whole of SO(2). Condition (D) is obviously obeyed, so 0(2) and SO(2) are both linear Lie groups of dimension 1. Further, Equation (3.3) gives ai = 0 1 -1 0 again confirming the first theorem above. Although the parametrization of Equation (3.7) extends to the whole of SO(2), it cannot apply to the set (ii). However, every A of set (ii) can be written as A = 0 1 -1 0 COS£E1 sinrri — sinrri COS£E1 — sinrri COS£E1 COS£E1 sinrri (3.8) for some x± such that —7Г < Xi < 7Г. Example III The group SU(2) SU(2) is the group of 2 x 2 unitary matrices u with det u = 1. A faithful finite- dimensional representation is provided by Г(и) = u. The defining conditions imply that every u E SU(2) has the form u = a /? -/?* a* (3.9) where a and /3 are two complex numbers such that |a|2 + |/?|2 = 1. With a = aq + ie^ /? = /?i + ^2 (oq, being real), this latter condition becomes a2 + + (3^ + /?2 = I • An appropriate parametrization is then a2 = Ж3/2, /31 = ж2/2, /32 = Ж1/2, ai = +{1 - (l/4)(a:i + x% + x2)}1/2,
40 GROUP THEORY IN PHYSICS for then Xj = X2 = X3 = 0 corresponds to the identity 1, and d(u, 1) = 2[1 - {1 - (l/4)(^ + x% + а?з)}1/2]1/2, * so that d(u, 1) < 6 if and only if x2 + x^ + x2 < {2ё2 — ^ё4}1/2. Thus, with ё < 2\/2 and tj < 2ё2 — |64, conditions (B) and (C) are satisfied and and Rr] coincide. Condition (D) is clearly true, so SU(2) is a linear Lie group of dimension 3. Incidentally, Equation (3.3) gives (3.10) so that the first theorem above is yet again confirmed. Although this parametrization is the most convenient for establishing that SU(2) is a linear Lie group, it is not the most useful for some practical cal- culations. Indeed only the matrices u with aq > 0 can be parametrized this way, whereas it will be shown in Example III of Section 2 that there exist parametrizations of the whole of SU(2). There is no difficulty in principle in generalizing the arguments used in Examples II and III to show that for all N > 2, О (AT), SO (AT), U(AT) and SU(7V) are linear Lie groups of dimensions ^N(N — 1), ^N(N — 1), N2 and N2 — 1 respectively, but the detailed algebra is rather more lengthy. (U(l) is a special case that is very easy to treat along the lines of Example I, because u = [exp га?1], — 7Г < xi < 7Г, is a parametrization.) Finally, a Lie subgroup can be defined in the obvious way. Definition Lie subgroup of a linear Lie group A subgroup Q' of a linear Lie group Q that is itself a linear Lie group is called a “Lie subgroup” of Q. 2 The connected components of a linear Lie group Definition Connected component of a linear Lie group Q A maximal set of elements T of Q that can be obtained from each other by continuously varying one or more of the matrix elements T(T)j/c of the faithful finite-dimensional representation Г is said to form a “connected component” of g. (It can be shown that the concept of connectedness as defined for a general topological space (Simmons 1963) is equivalent, for the type of space being considered here, to that implied by the above definition.) Example I The multiplicative group of real numbers This group was considered in Example I of Section 1. The set t > 0 forms
LIE GROUPS 41 one connected component (which actually constitutes a subgroup) and the set t < 0 forms another connected component. As t = 0 is excluded from the group, one set cannot be obtained continuously from the other. Example II The groups 0(2) and SO(2) In the group 0(2) that was examined in Example II of Section 1, every matrix A of S0(2) can be parametrized by Equation (3.7) with —7Г < < я*, whereas if A is a member of the set (ii) (i.e. if det A = —1), A can be written in the form of Equation (3.8). Thus SO(2) constitutes one connected component and the set (ii) is another connected component. It is obvious that these two sets cannot be connected to each other, because in a connected component det Г(Т) must vary continuously with T (if it varies at all), but det A cannot take any values between +1 and — 1 for A e 0(2). These examples suggest the following general theorem. Theorem I The connected component of a linear Lie group Q that contains the identity E is an invariant subgroup of Q. This component is often referred to as “the connected subgroup of P”. Moreover, each connected component of a linear Lie group Q is a right coset of the connected subgroup. Proof See, for example, Chapter 3, Section 2, of Cornwell (1984). In principle Q may have a countably infinite number of connected compo- nents, but in all cases of physical interest this number is finite. The axioms imply that the connected subgroup is always a linear Lie group. Definition Connected linear Lie group A linear Lie group is said to be “connected” if it possesses only one connected component. Thus the whole of a connected linear Lie group of dimension n can be parametrized by n real numbers ?/i, y^..., yn which form a connected set in IRn in such a way that all the matrix elements Г(Т)^ are continuous functions of the parameters. There is no requirement that these functions be analytic nor that they provide a one-to-one mapping. Consequently this parametrization does not necessarily satisfy all the conditions appearing in the definition of a linear Lie group. As the sets rri, ж2? • • •, and y^ y^ ..., yn are required for different purposes, they need not be interchangeable. The parametrizations do coincide in Examples I and II above, but Example III below reflects the general situation. Example III The group SU(2) Every pair of complex numbers a and /3 of Equation (3.9) that satisfy the condition |q|2 + |/?|2 = 1 can be written as a = cos?/i ехр(г?/2), P = smy1 ехр(гт/3),
42 GROUP THEORY IN PHYSICS where о < У1 < тг/2, 0 < y2 < 2тг, 0 < Уз < 2тг. (3.11) Thus „ -гл.а-Г cosj/iexp(iy2) sinyi ехр(гуз) 1 , . [ — sini/i ехр(—гу3) cosyi ехр(-гу2) ’ whose matrix elements are obviously continuous functions of 3/1, and уз. This is therefore a parametrization of the whole of SU(2). (This parametriza- tion fails to satisfy the conditions involved in the definition of a linear Lie group because it does not provide a one-to-one mapping of the appropriate regions, for the identity corresponds to the whole set of points = 0, y<2 = 0, 0 < Уз < 27Г. Consequently дГ/дуз = 0 at yx = y2 = уз = 0.) Similar arguments show that SO (TV) and SU(TV) are connected linear Lie groups for all N > 2, as is U(TV) for all N > 1. The relationship between a connected linear Lie group and its corresponding real Lie algebra will be stud- ied in some detail in Chapter 8, where it will be shown that the Lie algebra very largely determines the structure of the group. Indeed, it is for this pur- pose that the parametrization in terms of X2, • • •, xn is required. However, the rest of this chapter is devoted to certain “global” properties of linear Lie groups, and for these it is the parametrization in terms of ?/i, y^ •. •, yn that is relevant. 3 The concept of compactness for linear Lie groups Although the concept of a “compact” set in a general topological space has a curiously elusive quality, the following theorem, often referred to as the “Heine-Borel Theorem”, provides a very straightforward characterization of such sets in finite-dimensional real and complex Euclidean spaces. As this will suffice to distinguish a compact linear Lie group from a non-compact linear Lie group, no attempt will be made to give a detailed account of compactness, nor even a definition of the notion. (A lucid account of this and other general topological ideas may be found in the book of Simmons (1963).) Theorem I A subset of points of a real or complex finite-dimensional Eu- clidean space is “compact” if and only if it is closed and bounded. As mentioned in Section 1, by introducing the faithful m-dimensional rep- 2 resentation Г, the Lie group has been endowed with the topology of (Cm . However, it is often helpful to invoke the continuous parametrization of the connected subgroup by ?/i, y^..., yn introduced in Section 2. As the continu- ous image of a compact set is always another compact set (Simmons 1963), it
LIE GROUPS 43 follows that if the linear Lie group has only a finite number of connected com- ponents and the parameters ?/i, y^..., yn range over a closed and bounded set in IRn, then the group is compact. A “bounded” set of a real or complex Euclidean space is merely a set that can be contained in a finite “sphere” of the space. The term “closed” implies something more involved, so perhaps a few words of explanation may be needed. Although the specification of a general closed set can be fairly difficult, the only subsets of IRn that are relevant here are connected, and for these the characterization is straightforward. Indeed, in IR1 every connected closed set is of the form «i < 2/1 < bi- Similarly, the set aj < yj < bj, j = 1,2,..., n, of IRn is closed, but if any of the end points aj and bj are not attained the set is not closed. This set is bounded if and only if all the aj and bj are finite. These considerations imply the following identification. Characterization Compact linear Lie group of dimension n A linear Lie group of dimension n with a finite number of connected compo- nents is compact if the parameters 2/1, 1/2, • • •, Уп range over the closed finite intervals aj < yj < bj, j = 1,2,..., n. The Lie groups of physical interest that are non-compact usually fail to 2 be compact by virtue of giving an unbounded set of matrix elements in (Cm . As the sets of matrix elements Г(Т)^ of a linear Lie group Q are bounded if and only if there exists a finite real number M such that d(T, E) < M for all T E P, such groups are very easy to recognize in practice. If a Lie group Q is compact, then every Lie subgroup S of Q must also be compact (except in the rare case when S has a “non-closed” parametrization). On the other hand, if Q is non-compact, 5 may easily possess compact Lie subgroups. For semi-simple Lie groups there exists a criterion for compactness that is expressed purely in Lie algebraic terms, as will be shown in Chapter 11, Section 10. The real importance of the distinction between compact and non-compact groups lies in the fact that the representation theory of compact Lie groups is very largely the same as that for finite groups, whereas for non-compact groups the theory is entirely different. Example I The multiplicative group of real numbers As noted in Example I of Section 1, a faithful one-dimensional representation of this group is provided by T(t) = [t\. Obviously this set is unbounded in (C1, so the group is non-compact. Example II The groups О (TV) and SO (TV) For 0(2) and SO(2), Examples II of Sections 1 and 2 imply that the range of the only parameter 2/1 (= #i) is —7Г < y\ < тг. Similar statements are true for О (TV) and SO (TV) for TV > 3, and consequently О (TV) and SO (TV) are compact
44 GROUP THEORY IN PHYSICS for all N > 2. Example III The groups U(AT) and SU(7V) As all the intervals in Conditions (3.11) are closed and finite, SU(2) is compact. The same is true of SU(7V) for all N > 2, and of U(AT) for all N > 1. 4 Invariant integration If to each element T of a group Q a complex number f(T) is assigned, then f(T) is said to be a “complex-valued function defined on P”. One example that has been met already is the set of matrix elements T(T)jk (for J, к fixed) of a matrix representation Г of Q. For a finite group sums of the form f(T) are frequently encountered, particularly in representation theory. Because the Rearrangement Theorem shows that the set {T'T; T E G} has exactly the same members as P, it follows that for any T' E Q E /(t't) = £ /(T), тед тед and the sum is said to be “left-invariant”. Similarly £ Z(tt') = £ /(T), тед тед so such sums are also “right-invariant”. Moreover, with f(T) = 1 for all T E Q, the sum is finite in the sense that 1 = g, the order of Q. In generalizing to a connected linear Lie group, it is natural to make the hypothesis that the sum can be replaced by an integral with respect to the parameters ?/i, y^... , yn- However, questions immediately arise about the left-invariance, right-invariance and finiteness of such integrals. For general topological groups these become problems in measure theory. Using this the- ory Haar (1933) showed that for a very large class of topological groups, which includes the linear Lie groups, there always exists a left-invariant in- tegral and there always exists a right-invariant integral. (Accounts of these developments, including proofs of the theorems that follow, may be found in the books of Halmos (1950), Loomis (1953) and Hewitt and Ross (1963).) Let r rbi rbn / f(T)diT= / dyi... dynf(T(yi,...,yn))ai{yi,...,yn) (3.13) J g J di «/ dn and [ f(T) drT = I dy1... f dynf(T(y1,...,yn)')ffr(y1,...,yn) (3.14) g Jai J an be the left- and right-invariant integrals of a linear Lie group P, so that [ f(T'T)dlT = [ f(T)dtT JG JG (3.15)
LIE GROUPS 45 [ f(TTf)drT = [ f(T)drT (3.16) Jg Jg for any Tr e Q and any function /(T) for which the integrals are well defined. Here ..., ?/n) and <тг(?/1, • • •, ?/n) are left- and right-invariant “weight functions”, which are each unique up to multiplication by arbitrary constants. The left- and right-invariant integrals may be said to be finite if /• pbi rbn diT = dyi... dync?i(y-L,...,yn) J g J <21 «/ d-n, and /• rbi rbn / drT = / dyr... dynar(y!,...,yn) J g Jai J an are finite. If the multiplicative constants can be chosen so that ..., Уп) and crr(?/i,...,?/n) are equal, so that the integrals are both left- and right- invariant, then Q is said to be “unimodular”, and one may write d[T = drT = dT and If Q has more than one connected component, the integrals in Equations (3.13) and (3.14) can be generalized in the obvious way to include a sum over the components. The significance of the distinction between compact and non-compact Lie groups lies in the first two of the following theorems, the first of which was originally proved by Peter and Weyl (1927). They imply that compact Lie groups have many of the properties of finite groups, summation over a finite group merely being replaced by an invariant integral over the compact Lie groups, whereas for non-compact groups the situation is completely different. Theorem I If Q is a compact Lie group, then Q is unimodular and the invariant integral r rb± rbn f(T)dT = dyi... I dynf(T(y1,...,yn))a(y1,...,yn) J g J ai J dn exists and is finite for every continuous function f(Tfi Thus <t(s/i, ..., yn) can be chosen so that r rbi rbn I dT = dyi...l dyna(y1,...,yn) = 1. Jg J ai J an (A function /(T) is continuous if and only if /(T(?/i,..., yn)) is a continuous function of ?/i,...,?/n.) Theorem II If Q is a non-compact Lie group then the left- and right- invariant integrals are both infinite.
46 GROUP THEORY IN PHYSICS For non-compact groups the question of when Q is unimodular is partially- answered by the following theorem. Theorem III If Q is Abelian or semi-simple then Q is unimodular. The definition of a semi-simple Lie group is given in Chapter 11, Section 2. The other non-compact linear Lie groups have to be investigated individu- ally. In practice, explicit expressions for weight functions are seldom needed. Indeed, in dealing with the compact Lie groups all that is usually required is the knowledge (embodied in the first theorem above) that finite left- and right-invariant integrals always exist.
Chapter 4 Representations of Groups - Principal Ideas 1 Definitions The concept of the representation of a group was introduced in Chapter 1, Section 4, where it was shown that representations occur in a natural and significant way in quantum mechanics. It is worth while starting the detailed study of representations by repeating the definition as rephrased in Chapter 2, Section 6: Definition Representation of a group Q If there exists a homomorphic mapping of a group Q onto a group of non- singular d x d matrices Г(Т), with matrix multiplication as the group multi- plication operation, then the group of matrices Г(Т) forms a d-dimensional representation Г of Q. It will be recalled that the representation is described as being “faithful” if the mapping is one-to-one. Theorem I If Г is a d-dimensional representation of a group P, and E is the identity of P, then Г(Е*) = 1^. Proof As E2 = E then Г(Е'){Г(Е') — 1} = 0. Suppose first that this is the minimal equation for Г(Е*) (see Appendix A, Section 2). This implies that Г(Е') is diagonalizable and has at least one eigenvalue equal to zero, which in turn implies that det Г(Е*) = 0. As this is not permitted, the minimal equation must be of degree less than two and so must be of the form Г(Е*) — 7I = 0, and clearly the only allowed value of 7 is 1. It follows that Г(Т-1) = Г(Т)-1 for all T e Q. Every group Q possesses an “identity” representation, which is a one-dimensional representation for
48 GROUP THEORY IN PHYSICS which Г(Т) = [1] for all T e Q. Although mathematically extremely trivial, physically this representation can be very important. Example I Some representations of the crystallographic point group D4 Several representations of the group D4 have already been encountered either explicitly or implicitly, and it is worth while gathering them together for future reference. As the subsequent developments will show, this list is far from being exhaustive. (i) Equation (1.4) implies that the matrices Г (Г) listed in Example III of Chapter 1, Section 2, form a faithful three-dimensional representation of D4. (ii) A faithful two-dimensional representation of D4 was explicitly noted in Example I of Chapter 1, Section 4. (iii) A non-faithful but non-trivial one-dimensional representation of D4 is given implicitly in Example I of Chapter 2, Section 6. In this represen- tation = Г(С2у) = Г(С2ж) = Г(С2г) = [1], Г(с4у) = Г(С4;1) = Г(С2С) = Г(С2Й) = [-1]. (iv) Finally there is the identity representation for which Г(Т) = [1] for all Teg. For a Lie group it is necessary to supplement the definition by the require- ment that the homomorphic mapping must be continuous. For a connected linear Lie group this implies that the matrix elements of the representation must be continuous functions of the parameters 3/1, 2/2, • • •, Уп of Chapter 2, Section 2. (The extension to analytic representations and the relationship between the two concepts will be considered in Chapter 9, Section 4.) For groups of coordinate transformations in three-dimensional Euclidean space IR3 it has already been demonstrated how useful are the operators P(T) and the basis functions ^n(r) that were defined in Chapter 1, Sections 2 and 4 respectively. It is profitable to partially generalize these concepts to make them available for any group Q. To this end, consider a d-dimensional repre- sentation Г of P, let Vh, ^2, • • •, be the basis of a d-dimensional abstract complex inner product space (see Appendix B, Section 2) called the “carrier space” V, and for each T e Q define the operator Ф(Т) acting on the basis by d Ф(Т)фп = £ (4.1) m=l for n = 1,2,..., d. With the further definition that d d адЕ{М1} = ЕМФ(т-} (4.2) 5=1 5=1
REPRESENTATIONS - PRINCIPAL IDEAS 49 for any set of complex numbers 6i,62,... ,bd, such an operator is a linear operator. Moreover, Equation (4.1) implies the operator equalities Ф(ТхТ2) = Ф(Тх)Ф(Т2) (4.3) for all Ti,T2 e G, so that the operators form a group and there is a homo- morphic mapping of G onto this group. The operators Ф(Т) and the carrier space V are sometimes said to collectively form a “module”. However, there is no guarantee that the operators are unitary for a given representation, that is, in general (ф(тмф(эд / (<m (See Section 3 for further discussion of this point). Finally, if the basis is chosen to be an ortho-normal set, then Equation (4.1) implies that Г(Т)тп = (^т,Ф(ТШ (4.4) for any T e G- Conversely, any set of operators acting on a d-dimensional inner product space and satisfying Equation (4.3) will produce, by Equation (4.4), a d-dimensional matrix representation. (This provides the best way of introducing infinite-dimensional representations, the finite-dimensional inner product space merely being replaced by an infinite-dimensional Hilbert space, but these will not be discussed in this book.) It is entirely a matter of taste and convenience whether one works with an explicit matrix representation or with the corresponding module consisting of the operators Ф(Т) and the carrier space V on which they act. Theoretical physicists normally prefer to deal with the more concrete matrix representa- tions, whereas pure mathematicians tend to prefer the module formulation. It should be noted that for groups of coordinate transformations in IR3, for which both the operators Ф(Т) and P(T) are defined, there are two major differences between these sets of operators. Firstly, the Ф(Т) depend on the representation Г under consideration, whereas the P(T) are independent of the representation. Secondly, the operators Ф(Т) act in a finite-dimensional space, whereas the P(T) act in the infinite-dimensional Hilbert space L2. As the theory is developed in this chapter it will become apparent that every group has an infinite number of different representations, but these can be formed out of certain basic representations, the so-called “irreducible representations”. For a finite group there is essentially only a finite number of these. It will have become evident already that vector spaces and inner product spaces play an important part in representation theory. Readers who are not very familiar with them are advised to study Appendix В before proceeding further. 2 Equivalent representations Theorem I Let Г be a d-dimensional representation of a group P, and let S be any d x d non-singular matrix. Define for each T e G a d x d matrix
50 GROUP THEORY IN PHYSICS Г'(Т) by r'(T) = S-1r(T)S. (4.5) Then this set of matrices also forms a (/-dimensional representation of Q. The representations Г and Г' are said to be “equivalent”, and the transformation in Equation (4.5) is called a “similarity transformation”. Proof For any 71,72 E G, by Equations (1.25) and (4.5), Г'(Т1)Г'(Т2) = S-1r(Ti)SS“1r(T2)S = S-1r(Ti)r(T2)S = S-1r(TiT2)S = T'tTYT^. In Section 6 there will be given a simple direct test for the equivalence of two representations which does not require actually finding the matrix S that induces the similarity transformation. As all 1 x 1 matrices commute, if d = 1 then Г'(Т) = Г(Т) for all T E G and for every lxl non-singular matrix S. Thus two one-dimensional representations of G are either identical or are not equivalent. For d > 2 the situation is not so simple. In general a similarity trans- formation will produce an equivalent representation whose matrices Г'(Т) are different from those of Г(Т). However, these differences are in a sense superficial, for it will become clear that to a very large extent equivalent representations have essentially the same content. The following theorem on basis functions provides the first indication of this. Theorem II Let Г be a d-dimensional representation of a group of coor- dinate transformations in IR3, let Vh(r), ^2(1*), • • •, V^(r) be a set of basis functions of Г and let S be any d x d non-singular matrix. Then the set of d linearly independent functions ^1(г), > ^d(r) defined by d C(r) = E (4.6) m=l for n = 1,2,..., d form a set of basis functions for the equivalent representa- tion Г', where, for all T E G, Г'(Т) = S-1r(T)S. (4.7) Proof For any T E G, from Equations (1.18), (1.26) and (4.6), d d F(T)C(r) = Smn{P(T)V>m(r)} = E Smnr(T)pm^p(r). m=l m,p=l However, inverting Equation (4.6) gives d <7=1
REPRESENTATIONS - PRINCIPAL IDEAS 51 Thus, from Equation (4.7), d d Pm'n(r)= £ (S-1)?pr(T)pmSmn^(r)=£r/(T)</n^(r)- m,p,Q=l q=l The functions (r), ^2 (r), • • • , 'Фа (r) form a basis for a complex d- dimensional inner product space. With the definition in Equation (4.6), the functions ^i(r), > ^d(r) form an alternative basis for the same space. Thus the effect of a similarity transformation is merely to rearrange the basis of this space without changing the space itself This result has particular significance for the solutions of the time- independent Schrodinger equation. It was shown in Chapter 1, Section 4, that the eigenfunctions of a d-fold degenerate energy eigenvalue form a basis for a d-dimensional representation Г of the group of the Schrodinger equation. However, any d linearly independent linear combinations of these eigenfunc- tions also form a set of eigenfunctions belonging to the same eigenvalue, and there is no reason to prefer the original set to this new set, or vice versa. As the new set forms a basis for a representation equivalent to Г, the represen- tation of the group of the Schrodinger equation that corresponds to an energy eigenvalue is determined only up to equivalence. This section will be concluded by stating the analogous theorem which is valid for the carrier space of any representation of any group. Theorem III Let Г be a d-dimensional representation of a group P, let 'Фхч ^2, • • •, 'Фа be a basis of its carrier space and define the operators Ф(Т) for all T e Q by Equation (4.1) and its extension (4.2). Let S be any d x d non- singular matrix. Then the set of d linearly independent vectors ^1, ^2? • • •, defined by а 'Фп = m=l (for n = 1,2,..., d) forms a basis for the equivalent representation Г', where, for all T e Q, r'(T) = S-1r(T)S, in the sense that а Ф(т)< = £ г'(Т)топ< тп=1 for all T e G and n = 1,2,..., d. Proof This is essentially identical in content to that given above. Again and ^1, • • • > V>d are merely two different bases for the same carrier space.
52 GROUP THEORY IN PHYSICS 3 Unitary representations Definition Unitary representation of a group A “unitary” representation of a group Q is a representation Г in which the matrices Г(Т) are unitary for every T e Q. The following theorems show the profound difference between compact and non-compact Lie groups and the affinity between compact Lie groups and finite groups. Theorem I If Q is a finite group or a compact Lie group then every repre- sentation of Q is equivalent to a unitary representation. Proof See, for example, Appendix C of Cornwell (1984). It will be recalled that all the point groups and space groups of solid state physics are finite. Likewise, the rotation groups in three dimensions and the internal symmetry groups of elementary particles are compact Lie groups. Thus, in all these situations, advantage may be taken of the considerable simplifications that result from using representations that are unitary. Although the technical definition of “simple” and “semi-simple” Lie groups must be deferred until Chapter 11, Section 2, this is the appropriate place to mention some relevant properties of their representations. Theorem II If Q is a non-compact simple Lie group then Q possesses no finite-dimensional unitary representations apart from the trivial representa- tions in which Г(Т) = 1 for all T e Q. Proof This will be given in Chapter 12, Section 2. A non-compact Lie group that is not simple may possess both unitary representations and representations that are not equivalent to unitary repre- sentations, as the following example shows. Example I The multiplicative group of positive real numbers This group was considered previously in Examples I of Chapter 3, Sections 1, 2 and 3. A typical element is ехртд, —сю < y\ < сю. It has a set of one-dimensional unitary representations defined by г(exp 3/1) = [ехр(шух)], where a is any fixed real number. It has also a set of one-dimensional non- unitary representations given by Г (exp 3/1) = [exp(Z?3/i)],
REPRESENTATIONS - PRINCIPAL IDEAS 53 where /3 is any fixed real number. These latter representations, being one- dimensional, cannot be transformed by any similarity transformation into unitary representations. Theorem III If Q is a group of coordinate transformations in IR3 and if the representation Г of Q possesses a set of basis functions, then Г is unitary if the basis functions form an ortho-normal set. Proof Suppose that the basis functions ^i(r), ^2(1*), • . ., V^(r) °f Г f°rm an ortho-normal set, i.e. ('фгт'Фп) = ^mn for m, n = 1,2,..., d. As the operators P(T) are unitary, it follows from Equations (1.19), (1.20) and (1.26) that for each T E Q &т.п = (Фт,Фп) = (Р(Т)фт,Р(Т)фп) d = £ Г(Т)*тГ(Т)дп(фр,фд) p,q=l d = £ r(T);TOr(T)pn, p,Q=l so that Г(Т)^Г(Т) = 1 and hence Г(Т) is unitary. From a set of basis functions ^i(r), ^2(1*), • • •, V^(r) °f a non-unitary rep- resentation Г an ortho-normal set (r), ^2 (r), • • • > V^(r) can alwaYs be con- structed by the Schmidt orthogonalization process (see Appendix B, Section 2). As each ?/>'• (r) is a linear combination of the (r), the set ^1 (r), ^2 (r) > • • • > ^(r) must be basis functions for a unitary representation Г' that is equivalent to Г. Indeed, on defining the coefficients 5mn by V4(r) = 22m=i Smn^Tn(y), the matrix S having these coefficients as elements is precisely the matrix that induces the similarity transformation from Г to Г'. However, there exist groups of coordinate transformations in IR3 that have at least some representations that do not possess basis functions, so this ar- gument does not imply that every representation of every group of coordinate transformations is equivalent to a unitary representation. For any abstract group Q there exists a generalization of the last theorem. If an ortho-normal basis is used in the construction of the operators Ф(Т) of Equations (4.1) and (4.3), it follows by an argument similar to that given in the above proof that, for each T E Q, Ф(Т) is a unitary operator if and only if Г(Т) is a unitary matrix. The amount of attention that has just been devoted to non-unitary rep- resentations should not be allowed to obscure the main point, which is that in most cases of physical interest all the representations can be chosen to be unitary. This section will be concluded with an important theorem that demon- strates the special role played in similarity transformations by matrices that
54 GROUP THEORY IN PHYSICS are unitary. (As noted in Appendix B, Section 2, such transformations trans- form ortho-normal bases into ortho-normal bases.) Theorem IV If Г and Г' are two equivalent representations of a group G related by the similarity transformation r'(T) = S-1r(T)S for all T e G> and if Г is a unitary representation and S is a unitary matrix, then Г' is also a unitary representation. Conversely, if Г and Г' are equivalent representations that are both unitary, then the matrix S in the similarity transformation relating them can always be chosen to be unitary. Proof The first proposition is almost obvious, but the converse requires a rather lengthy proof, which may found, for example, in Appendix C of Corn- well (1984). 4 Reducible and irreducible representations Suppose that the d-dimensional representation Г of a group Q can be parti- tioned so that it has the form V(T\- [ Г11(Т) Г12(Т) 1 /до\ Г(т) - [ 0 r22(T) J (4-8) for every T e Q, where Гц(Т), Г12(Т), Г22(Т) and the zero matrix 0 have dimensions si x si, si x s2, s2 x s2 and s2 x si respectively. (Here si +s2 = d, si > 1, s2 > 1 and si and s2 are the same for all T e G-) Then (cf. Equation (A.7)) for any Ti,T2 eG _ Г Г11(Т1)Г11(Т2) Г11(Т1)Г12(Т2) +Г12(Т1)Г22(Т2) 1 Г(ТХ)Г(Т2)-^ 0 1^)1^) ]’ so that, as the matrices Г(Т) form a representation of Q, Гц(Т1Т2) = Г11(71)Г11(72) (4.9) Г22(Т1Т2) = (4.10) Equations (4.9) and (4.10) imply that the matrices Гц(71) and the matri- ces Г22(Т) both form representations of p. Thus the representation Г of G is made up of two other representations of smaller dimensions, so it is natural to describe such a representation as being “reducible”. In order that this description should apply equally to all equivalent representations, the formal definition can be stated as follows: Definition Reducible representation of a group G A representation of a group G is said to be “reducible” if it is equivalent to a representation Г of G that has the form of Equation (4.8) for all T e G-
REPRESENTATIONS - PRINCIPAL IDEAS 55 It follows from Equations (4.1) and (4.8) that Si Ф(Т)^п = E rn(T) тп'Фпч m=l for n = 1,2,. ..,$i and all T e Q. Thus the $i-dimensional subspace of the carrier space V having basis , V'si invariant under all the operations of Q in the sense that if is any vector of this subspace then Р(ТУф is also a member of this subspace for all T e Q. (It should be noted that in general the ^-dimensional subspace with basis ., 'фа is not invariant.) The following definition is the most important in the whole of the theory of representations. Definition Irreducible representation of a group Q A representation of a group Q is said to be “irreducible” if it is not reducible. This definition implies that an irreducible representation cannot be trans- formed by a similarity transformation to the form of Equation (4.8). Conse- quently the carrier space V of an irreducible representation has no invariant subspace of smaller dimension. Some simple tests for irreducibility will be developed in Sections 5 and 6. Returning to the reducible representation Г of Equation (4.8), the ques- tion arises as to whether Гц(Т) and Г22(Т) are also reducible or not. If Гц(Т) is reducible, then by a similarity transformation it too can be put in the form of Equation (4.8) with submatrices of some dimensions. The same is true of Г22(Т). Obviously this process can be continued until all the repre- sentations involved are irreducible. Thus every reducible representation Г by an appropriate similarity transformation S can be put into the form " r'n(T) Г'12(Т) Г'13(Т) • •• Г'1Г(Т) 0 П2(т) г23СО . • • ПДТ) r'(T) = S-1r(T)S = 0 0 •• Г'Г(Т) 0 0 0 r;r(T) _ where all the matrices ГС(Т) form irreducible representations, for j = 1, 2, ...,r. (Here Tjk(T) is an s' x sk, matrix, ^j=isj = sj 1 for eac^ j = 1,2, ...,r, and s'1? ^2,..., s'r are the same for all T E Q.) It is now apparent that the irreducible representations are the basic building blocks from which all reducible representations can be constructed. The final question is whether all the upper off-diagonal submatrices Г'k(T) (k > j) can be transformed into zero matrices by a further similarity transfor- mation, leaving only the diagonal submatrices non-zero. If so, Г is equivalent
56 GROUP THEORY IN PHYSICS to a representation of the form - глт 0 0 0 0 Г'2'2(Т) 0 0 Г"(Т) = 0 0 ПзСП . 0 (4-11) 0 0 0 .. г"г(т) _ in which the T'T are all irreducible representations of Q. Definition Completely reducible representations of a group Q A representation Г of a group Q is said to be “completely reducible” if it is equivalent to a representation Г" that has the form in Equation (4.11) for all TeP. A completely reducible representation is sometimes referred to as a “de- composable” representation. Theorem I If Q is a finite group or a compact Lie group then every reducible representation of Q is completely reducible. The same is true of every reducible representation of a connected, non-compact, semi-simple Lie group and of any unitary reducible representation of any other group. Proof See, for example, Chapter 4, Section 4, of Cornwell (1984). Suppose that </>i, </>2,..., </>d form a basis for the carrier space of the com- pletely reducible representation Г" of Equation (4.11) and that the irreducible representation T'T has dimension d7, j = 1,2, ...,r, so that dj = d. Then it follows from Equations (4.1) and (4.11) that ф\, ..., form a basis for the carrier space of Г'^, that <^1+i,(^1+2,... ^ф^+а^ form a basis for the carrier space of Г22, and so on. The carrier space of Г" is therefore a direct sum of carrier spaces belonging to each of the irreducible represen- tations Г'^,Г22, • • •,Г"г (see Appendix B, Section 1). Correspondingly, the completely reducible representation Г" is said to be the “direct sum” of the irreducible representations Г'^, Г22, • • •, Г"г, this statement being expressed concisely by Г" — Г" m Г" m m Г" — x 11 “ x 22 “ • • • “ x rr' (The symbol ф here indicates that the sum involved is not that of ordinary matrix addition.) Similarly, the equivalence of a representation Г to a direct sum of irreducible representations Г'^,, Г"г can be written as Г ~ Г" m Г" m m Г" In the case in which Г is equivalent to a unitary representation, all the irre- ducible representations in the direct sum are themselves equivalent to unitary representations.
REPRESENTATIONS - PRINCIPAL IDEAS 57 5 Schur’s Lemmas and the orthogonality the- orem for matrix representations The name “Schur’s Lemma” is often attached to one or other (or sometimes both) of the following two theorems. Theorem I Let Г and Г' be two irreducible representations of a group p, of dimensions d and df respectively, and suppose that there exists a d x d' matrix A such that Г(Т)А = АГ'(Т) for all T e Q. Then either A = 0, or d = d' and det A 0. Proof See, for example, Appendix C, Section 3, of Cornwell (1984). Theorem II If Г is a d-dimensional irreducible representation of a group Q and В is a d x d matrix such that Г(Т)В = ВГ(Т) for every T e G, then В must be a multiple of the unit matrix. Proof Let A = В — /31, where the complex number (3 is chosen so that det A = 0. Then Г(Т)А = АГ(Т) for all Te(7, so by the previous theorem the only alternative not excluded is A = 0, that is, В = /31. The following corollary shows how very straightforward are the irreducible representations of Abelian groups. Theorem III Every irreducible representation of an Abelian group is one- dimensional. Proof Let Г be an irreducible representation of an Abelian group G- As Г(Т’)Г(Т") = Г(Т')Г(Т) for all T and Tf of P, it follows from the preceding theorem that, for each Tf e G, Г(Т") = 7(7") 1, where 7(7') is some complex number that depends on T'. Clearly, such a representation is irreducible if and only if it is one-dimensional. The “orthogonality theorem for matrix representations” is a second corol- lary which will be used time and time again. As will be seen, it applies both to finite groups and compact Lie groups. Theorem IV Suppose that Гр and Г9 are two unitary irreducible represen- tations of a finite group G which are not equivalent if p q (but which are identical if p = q). Then (i/ff) £ r₽(T);fcr’(T)st = (i/dpy>pq6js6kt, T&3 where g is the order of G and dp is the dimension of Гр. Similarly, if G is a com- pact Lie group, the summation can be replaced by an invariant integration,
58 GROUP THEORY IN PHYSICS giving [ r”(T)*fcn(T)st dT = (l/dp)6pg6js6kt. J Q Proof See, for example, Appendix C, Section 3, of Cornwell (1984). It is in the application of this theorem that the main practical advantage of working with unitary representations lies. For example, one immediate consequence is the following partial converse to Theorem III of Section 3. Theorem V If </>i(r), </>£(r),..., aud (rL ^2 (r) > • • • > are respectively basis functions for the unitary irreducible representations Гр and Г9 of a group of coordinate transformations Q that is either a finite group or a compact Lie group, and Гр and Г9 are not equivalent if p q (but are identical if p = q), then (C,C) = ° unless p = q and m = n. If p = q and m = n, then (^,^m) is a constant independent of m. Proof From Equations (1.20) and (1.26), for any Te(7, (C,C) = (F(T)C,P(T)C) dp dq j = 1 k=l dp and dq being the dimensions of Гр and Г9 respectively. Summing or in- tegrating over all the transformations T e G, the orthogonality theorem for matrix representations gives dp 3 = 1 Thus when p q, or when p = q but m n, it follows immediately that (</>^,^Я) = When p = q and m = n the right-hand side of this last equation is independent of m, so (^,^m) must be independent of m. One immediate implication of this theorem is that for all m,n = 1,2,...,dp. Thus, if (r) is normalized, then so too are ^2 (r)> Фз (r), • • •• Henceforth it will usually be assumed that every set of basis functions of an irreducible representation is a mutually ortho-normal set, that is, for all m, n = 1,2,..., dp.
REPRESENTATIONS - PRINCIPAL IDEAS 59 6 Characters Although equivalent representations have essentially the same content, there is a large degree of arbitrariness in the explicit forms of their matrices. How- ever, the characters provide a set of quantities which are the same for all equivalent representations. Indeed, for finite groups and compact Lie groups the characters uniquely determine the representations up to equivalence. The characters have a number of other very useful properties which, for the most part, are valid for finite groups or compact Lie groups but not for non-compact Lie groups. Definition Characters of a representation Suppose that Г is a d-dimensional representation of a group Q. Then d x(T) = trr(T) (=£г(Т)^-) is defined to be the “character” of the group element T in this representation. The set of characters corresponding to a representation is called the “character system” of the representation. As Г(£7) = 1^ for the identity E of P, then x(E) = d. Theorem I A necessary condition for two representations of a group to be equivalent is that they must have identical character systems. Proof Let Г and Г' be two equivalent representations of a group P, both of dimension d, so that there exists a d x d non-singular matrix S such that Г'(Т) = S-1r(T)S for all T e Q. Then, as noted in Appendix A, tr Г'(Т) = tr Г(Т). Thus, if x(T) and yf (T) are the characters of T in Г and Г' respectively, then yf (T) = x(T) for all T e Q. The characters therefore provide a set of quantities that are unchanged by similarity transformations. The converse proposition will be considered shortly. The invariance property of the trace also provides another simple result: Theorem II In a given representation of a group P, all the elements in the same class have the same character. Proof Suppose that the elements T' and T of Q are in the same class. Then (see Chapter 2, Section 2) there exists a group element X such that T' = XTX~\ so that Г(Т') = Г(Х)Г(Т)Г(Х)“1. Consequently tr Г(Т) = tr Г(Т), and hence x'(T) = x(T). There are two orthogonality theorems for characters. The first is as follows:
60 GROUP THEORY IN PHYSICS Theorem III Let %P(T) and %q(T) be the characters of two irreducible representations of a finite group Q of order g, these representations being assumed to be inequivalent if p q. Then (i/ff) E х^тУхЧт) = 6pq. TeQ Similarly, if Q is a compact Lie group, the summation can be replaced by an invariant integration, giving [ Хр(ТУхд(Т) dT = 6pq. Jq Proof Theorem I of Section 3 shows that for the groups under consideration similarity transformations may be applied to the two irreducible representa- tions to produce unitary representations. The result then follows immediately from the orthogonality theorem for matrix representations (Theorem IV of Section 5) on putting j = к and s = t and summing over j and s. The converse theorem referred to previously can now be proved fairly easily. Theorem IV If Q is a finite group or a compact Lie group then a sufficient condition for two representations to be equivalent is provided by the equality of their character systems. Proof See, for example, Appendix C, Section 4, of Cornwell (1984). It should be noted that this sufficient condition does not extend to non- compact Lie groups. The characters provide a complete specification (up to equivalence) of the irreducible representations that appear in a reducible representation. This knowledge can prove very useful, as will be seen later. The details are as given by the following theorem. Theorem V The number of times np that an irreducible representation Гр (or a representation equivalent to Гр) appears in a reducible representation Г is given for a finite group Q by np = (Ш E x(T)xp(ry, TeQ where xp(T) and x(^) are the characters of Гр and Г respectively and g is the order of Q. For a compact Lie group this generalizes to np = j х(Т)хр(ТГ dT.
REPRESENTATIONS - PRINCIPAL IDEAS 61 Proof See, for example, Chapter 4, Section 6, of Cornwell (1984). The following theorem gives a convenient criterion for irreducibility ex- pressed solely in terms of characters, and so provides a very simple test for irreducibility, particularly for finite groups. Theorem VI A necessary and sufficient condition for a representation Г of a finite group Q to be irreducible is that (i/p)£|x(t)|2 = i, T&3 where x(T) is the character of the group element T in Г and g is the order of Q. The corresponding condition for a compact Lie group is [ |x(T)|2dT=l. Jg Proof See, for example, Chapter 4, Section 6, of Cornwell (1984). Characters may also be used to prove a theorem on the number of inequiv- alent irreducible representations of a finite group P, as well as a useful result on their dimensions. Theorem VII For a finite group p, the sum of the squares of the dimensions of the inequivalent irreducible representations is equal to the order of Q. Proof See, for example, Appendix C, Section 4, of Cornwell (1984). Theorem VIII For a finite group P, the number of inequivalent irreducible representations is equal to the number of classes of Q. Proof See, for example, Appendix C, Section 4, of Cornwell (1984). These two theorems taken together are often sufficient to uniquely specify the dimensions of the inequivalent irreducible representations. Example I Dimensions of the inequivalent irreducible representations of the crystallographic point group D^ As noted in Chapter 2, Section 2, D^ is of order 8 and has five classes. Thus it has five inequivalent irreducible representations. Let dj, j = 1,2,..., 5, be their dimensions, so that ^.=1 dj2 = 8, which has the solution di = d2 = d3 = d^ = 1 and (/5 = 2. This solution is unique up to a relabelling of representations. The second orthogonality theorem for characters is as follows.
62 GROUP THEORY IN PHYSICS Theorem IX If Xp(fij) is the character of the class Cj of a finite group Q for the irreducible representation Гр of P, then £xp(G)*xp(c^ = ^fe, p where the sum is over all the inequivalent irreducible representations of <y, g is the order of Q and Nj is the number of elements in the class Cj. Proof See, for example, Appendix C, Section 4, of Cornwell (1984). The character systems of the irreducible representations of a finite group are conveniently displayed in the form of a “character table”. The classes of the group are usually listed along the top of the table and the inequivalent ir- reducible representations are listed down the left-hand side. As a consequence of the Theorem VIII above, this table is always square. For groups of low order it is quite easy to completely determine the char- acter table directly from the theorems that have just been stated, without first obtaining explicit forms for the matrices, as the following example will show. Example II Character table for the crystallographic point group D4 The classes of D4 (see Chapter 2, Section 2) are Ci = {E}, C2 = {C2X,C2Z}, C3 = {C2y}, c4 = {С'^с'гД C5 = {C2c,C2d}. Consider first the four one-dimensional representations Г1, Г2, Г3 and Г4. As Clx = Clc = E then хр(С2ж)2 = Xp(C2c)2 = 1, p = 1,2,3,4. Moreover, from Table 1.2, Ciy = C2cC2X, so that xp(C4) = Хр(С2)хр(Сб)« Finally, C2y = C%y, so xp(C2y) = 1 for p = 1,2,3,4. Thus the four one-dimensional irreducible representations of D4 may be chosen to be such that: = 1, /(C5) = 1; X2(C2) = 1, X2(C5) = -1; x3(C2) = -1, X3(C5) = -1; y4(C2) = -1, %4(C5) = 1. From the first orthogonality theorem for characters (Theorem III) the two- dimensional representation Г5 must satisfy the conditions: Х5(С1) + 2х5(С2)+х5(Сз) + 2х5(С4) + 2х5(С5) = 0, ) Х5(С1) + 2Х5(С2)+х5(Сз)-2х5(С4)-2х5(С5) = 0, Х5(С1)-2х5(С2) + х5(Сз) + 2х5(С4)-2х5(С5) = 0, Х5(С1)-2х5(С2)+х5(Сз)-2х5(С4) + 2х5(С5) = 0. > Adding these equations gives х5(^1)+х5(Сз) = 0 and, as x5(Ci) = №(£') = 2, this implies х5(Сз) — ~2- Moreover, Theorem VI above gives lx5(T)|2 = 8, while |x5(£)|2 + \хъ(С2у)|2 = 8, so that x5(C2) = X5(C4) = ХЧ^б) = 0. The complete character table for D4 is given in Table 4.1. It is interesting to relate these irreducible representations to the representations of D4 dis- cussed in Example I of Section 1. Г1 is clearly the “identity” representation (iv), Г2 is the one-dimensional representation (iii), Г5 is the two-dimensional representation (ii), and the three-dimensional representation (i) is reducible, being given by the direct sum Г3фГ5.
REPRESENTATIONS - PRINCIPAL IDEAS 63 E ClxiCiz C2y C20 C^d Г1 1 1 1 1 1 Г2 1 1 1 -1 -1 Г3 1 -1 1 1 -1 Г4 1 -1 1 -1 1 Г5 2 0 —2 0 0 Table 4.1: Character table for the crystallographic point group D4. Although a number of results of physical significance follow immediately from a knowledge of the characters, it is often necessary to obtain explicit expressions for the matrices of the representations. A method for constructing such explicit expressions from the characters is described in Chapter 5, Section 1. Of course, for one-dimensional representations the characters themselves are the matrix elements. Hitherto all results on finite groups have had an immediate generalization for compact Lie groups. For Theorems VII and VIII above this generalization is more far-reaching and is embodied in the following theorem due to Peter and Weyl (1927). Theorem X For a compact Lie group P, the number of inequivalent irre- ducible representations is infinite but countable. This theorem implies that the irreducible representations of a compact Lie group can be specified by a parameter that only takes integral values (or, if more convenient, by a set of parameters taking integral values). This result has been anticipated in some of the notations already employed (but not in any of the proofs).
г кжу д гАа: 7 <Алt н НС'’'’ Ха1аЬь а./1 X. ' ah аааа а» HAG * нааа . а’ ; аан ,-, U:4aaaUX 4?ХЦА\ АГА ь‘ ’A '-VAA -.',’;ht -g r. aA'g., .A / ,;Ь A.g Ah^VAAA. AOfrf | A*M H f Л 1& * -I, 4 л \ ' f ' !’ । •, i • '"J i/i i . j *ч ’,’ J’i’ 1 • ’ 'i ’v n > j 1 *, F - i f ’,, ’ i И 1 l') i 1 i j”, t । j-/ V»’ A | Г jA' / ^д}Ж, -Л A'A^ I dh L ;J(i >' A >;’AAAA Up’b нМ А аЬааАА A^ А/A A A*F ;A’A гр‘.ДАЦ Ы.‘ -р '«(’.и;г’г A . .A A A’ : A,A A' <'••• !,“iv., ! , ;A - r* , ’Ю 5 -<j m'>4VV 1 . Г ' > A \ -A -> u ! ,i,-. !< -, 5 ( , ; (\ i’ j ч: ’/ * 1L ) \ 4 > ’-d h Д . V *. " /1 1« ь? г '.ни ,U j ;J '• ,4 j. <гьп,. 1 цг -i il * C ’* i :; - , . ‘ A; ; 1 ьf t 4 r<tU ;Л ’ 4 ( " V \_or ,-; 4lh\ 5 / // bf- JLx^u -'- г|%йН ’’'''^ ’54 - '‘d^Ai -A .< j, VA ’ AHA X \ - > --A-; >kAp (" !•; ^'A5 ' 1 , . ' " A , , J g ' v -lh ,/ < ” -g<. . s r !<;<'. j, A A X ."'d'/HIA^'A)niA 'A’T , 4a);vX АНГ,а‘'. •' - ,-': A’ r .., ?.A AAA ' ''• b/A’i!A'A‘^ h Лйн<лА AlO ‘ 4a ‘AA'’A А-AA -3 >’ I’HlA b ; d A 4A-A?Ff a/A#AA '/A ? ’ d > j-Aa-A'' , A -’ 'A . / ! f Ы|Ы A fAr a( A;
Chapter 5 Representations of Groups - Developments Having laid the foundations of the theory of group representations in the previous chapter, attention will now be concentrated on certain developments that are particularly significant in the applications to quantum mechanics. 1 Projection operators For any finite group Q of coordinate transformations in IR3, in particular for any crystallographic point group or space group, the basis functions of uni- tary irreducible representations are easily determined by a purely automatic process involving certain “projection operators”. Before defining these it is necessary to state a theorem which has many applications. Theorem I Any function ф(т) of L2 can be written as a linear combination of basis functions of the unitary irreducible representations of a group Q of coordinate transformations in IR3. That is dp t5-1) P .7 = 1 where ^(r) is a normalized basis function transforming as the J th row of the dp-dimensional unitary irreducible representation Гр of P, ap are a set of complex numbers and the sum over p is over all the inequivalent unitary irreducible representations of Q. Here L2 is the space of square-integrable functions, as defined in Ap- pendix B, Section 3. The basis functions фр(г) and coefficients ap depend on ф(г) and some of the coefficients ap- may be zero. For example, for the crystallographic point group D4 it will be demonstrated in Example I below
66 GROUP THEORY IN PHYSICS that with the choice ф(г) = (x + z) exp(—r) (where r = {x2 + y2 + z2}1/2), ф(т) = A-1{^(r)+^f (r)} with = Arexp(—r) and^r) = Azexp(—r), whereas with ф(т) = y(x + z)exp(—г), ф(т) = B-1{^(r) — ^(r)} with </>i(r) = Byzexp(—r) and </>2(1*) = -Bxyexp(-r). There is no suggestion in the theorem that the functions фР (r) on the right- hand side of Equation (5.1) form a fixed basis for the space L2. Indeed this would be impossible for a finite group as L2 is infinite-dimensional, whereas there is only a finite number of functions on the right-hand side of Equation (5.1) when Q is a finite group. On the other hand, the theorem can be applied to every member of a complete basis for the space L2 in turn, thereby producing a basis for L2, all of whose members are basis functions of unitary irreducible representations of Q. A situation where this proves very useful is examined in Chapter 6, Section 1. Proof See, for example, Appendix C, Section 5, of Cornwell (1984). Definition Projection operators Let Гр be a unitary irreducible representation of dimension dp of a finite group of coordinate transformations Q of order g. Then the projection operators are defined by Vpmn = (dp/g) £ Гр(Т)*тпР(Г), (5.2) T&3 for m, n = 1, 2,..., dp. If the group of coordinate transformations is a compact Lie group, the definition may be generalized to n = dp f rp(3XnP(T) dT. (5.3) Jq Theorem II The projection operators Ppin have the following properties: (a) For any two functions ф(т) and ^(r) of L2 (Р^ф) = (ф,Рртф). (5.4) In particular (Ррпф,ф) = (ф,Ррпф), (5.5) so that Ppn is a self-adjoint operator. (b) If the projection operators Pfnn and P^k belong to two unitary irre- ducible representations Гр and Г9 of Q that are not equivalent if p q (but are identical if p = q), then = 6PqSnjPqmk. (5.6) In particular (5-7)
REPRESENTATIONS - DEVELOPMENTS 67 (c) If ^1(г)> V,2(r)^ • • • ’ are basis functions transforming as the unitary irre- ducible representation Г9 of Q, then = Mn#m(r). (5-8) (d) For any function ф(т) of L2 (5.9) where and ^p(r) are the coefficients and basis functions of the ex- pansion of ф(г) (Equation (5.1)) that relate to the nth row of Гр. Proof See, for example, Appendix C, Section 5, of Cornwell (1984). The properties in Equations (5.5) and (5.7) are characteristic of any pro- jection operators. The nature of the projection associated with the operator 7?pin is apparent from Equation (5.9), which shows that Ppn projects into the subspace of L2 consisting of functions transforming as the nth row of Гр. As operators with the property in Equation (5.7) are known as “idem- potent” operators, the projection operator technique is sometimes called the “idempotent method”. For a finite group this theorem provides a simple automatic method for the construction of basis functions. (For a compact Lie group it is preferable to use other methods, as for example in Chapter 10, Section 4.) A set of ortho-normal basis functions transforming as the rows of Гр can be found by first selecting a function ф(т) such that Ppn^(r) is not identically zero for some arbitrarily chosen n = 1,2, ...,dp. With cp = (Р^пФ-» Р^пФУ^2) the function ^p(r) defined by C(r) = (iM)^^(r) is normalized and transforms as the nth row of Гр. Should its ortho-normal partners ^m(r) (ш = 1, 2,..., dp; m 7^ n) be required, they can be found by operating on ^p(r) with 'Pmn- It will be seen later (Chapter 6, Section 1) that in physical problems it is usually only necessary to work with basis functions belonging to one arbitrarily chosen row of each irreducible representation. Example I Construction of basis functions of irreducible representations of the crystallographic point group First let ф(т) — (x + z)exp(—r), where r = {x2 + y2 + z2}1/2. Then, from Equations (1.8) and (1.17), for any pure rotation T, F(T)</>(r) = = </>(R(T)r), the orthogonal matrices R(T) for being given in Example III of Chapter 1, Section 2. For example, for T = (?4P,
68 GROUP THEORY IN PHYSICS and as </>(R(C42/)r) is defined to be the function in which the ж, у and z in ф(т) are replaced by the 11, 21 and 31 components of R(C42/)r respectively, and here ф(г) = (x + z) exp(—r), then P(C42/)</>(r) = (z - x) exp(-r). The following is a complete list of functions Р(Т)ф(т) obtained in this way: F(E)<^(r) Р(С2х)ф(г) Р(С2у)ф(г) Р(С2г)0(г) Р(С2с)Ф(г) Р(С^)ф(г) Р(С2(1)ф(г) Р(С4у)ф(г) (x + z) exp(—r), (x — z) exp(—r), (—x — z) exp(—r), (—x + z) exp(—r). (5.10) Being one-dimensional, the matrices of the irreducible representations Г1, Г2, Г3 and Г4 are given directly in terms of the characters of Table 4.1 by rj(T) = [xJ(D] f°r T of Z)4 and j = 1,2,3,4. As noted in Example II of Chapter 4, Section 6, the matrices of the two-dimensional irreducible representation Г5 may be taken to be: Г5(Е) = r5(c2z) = 1 0 ' 0 1 ’ -1 0 5 0 ' 1 r5(C2a:) = , r5(C4p) = ’ 1 0 " 0 1 0 ' -1 -1 ’ 0 , r5(C2v) = , Г5(С47) = ’ -1 o’ 0 -1 0 1 ' -1 0 Г5(С2с) = ’ 0 1 1 0 5 r5(C2d) = 0 -1 -1 0 Then, by Equation (5.2), Pfi^(r) = 0 for p = 1,2,3,4, whereas Pii</>(r) = жехр(—r). Thus the function Ажехр(—r), where A = (1/cf) -- (жехр(—г),жехр(—r))-1/2, is a normalized basis function transforming as the first row of Г5, its partner transforming as the second row of Г5 is PfiM37 exp(—r)}, which is equal to Az exp(—r). It will be seen that, as P(T) exp(—r) = exp(—r) for all T e Q, the factor exp(—r) plays no role in the construction apart from ensuring that the basis functions can be normalized. Consequently, if exp(—r) is replaced by any function F(r) such that (xF(r), xF(r)) is finite, then A'xF(r) and A'zF(r) (where A' = (xFfr), rF(r))-1/2) are ortho-normal basis functions of Г5 trans- forming as the first and second rows respectively. Clearly no harm comes from temporarily being less precise than usual and saying that “ж and z transform as the first and second rows of Г5”. Such statements about basis functions of irreducible representations of groups of pure rotations appear quite commonly in the literature. A similar analysis applied to ф(г) = (xy + yz) exp(—r) shows that Pfi^(r) = 0 for p = 1,2,3,4, but Pii</>(r) = yz exp(-r). Thus Byz exp(-r) (where В = (l/ci) = (yz exp(-r),^ exp(-r))-1/2) is a normalized basis function
REPRESENTATIONS - DEVELOPMENTS 69 transforming as the first row of Г5. Its partner transforming as the second row of Г5 is p2i{B?/z exp(—r)}, which is equal to —Bxyexp(-r). Again, loosely one could say that “yz and —xy transform as the first and second rows of Г5”. The procedure for constructing basis functions that has just been described requires an explicit knowledge of the matrix elements of the representations, and not merely a knowledge of the character system alone, which is usually the only information which is given in the published literature. Of course, for one-dimensional representations the characters give the matrix elements immediately, but for the other representations some further analysis is needed. A method involving “character projection operators”, which can be used in such cases, will now be described. Definition Character projection operator Let Гр be an irreducible representation of dimension dp of a finite group of coordinate transformations Q of order g, xp(T) being the character of T e Q in Гр. Then the character projection operator for Pp is defined by = (dp/g) £ xp(T)*P(T). (5.11) TeG Obviously Pp can be constructed from the character table alone and dp pp — V4 pp ' / J ' nn’ n=l so that Pp has the property of projecting out of a function ф(т) the sum of all the parts transforming according to the rows of Гр. This implies that if Ррф(т) is not identically zero, it is a linear combination of basis functions of Гр (which are as yet undetermined). However, as noted in Chapter 4, Section 2, linear combinations of basis functions are themselves basis functions in an equivalent representation, so Ррф(т) may be taken to transform as the first row of some form of the pth irreducible representation. This particular form will henceforth be denoted by Гр. (Up to this stage Гр was only specified up to a similarity transformation.) The procedure to be described then generates explicit matrix elements for this form of Гр, which is, of course, as good as any other equivalent form. Having chosen a normalizable ф(т) such that Ррф(т) is not identically zero, construct Р(Т){Ррф(т)} for each T e Q. (Each of these must be linear combinations of the dp basis functions of Гр.) From these functions abstract dp linearly independent functions, taking one of these to be Ррф(т) itself. Apply the Schmidt orthogonalization process (see Appendix B, Section 2) to these functions to produce dp orthonormal functions ^p(r), n = 1, 2,..., dp, ^f(r) being a multiple of Ррф(т). These functions can be taken as the basis
70 GROUP THEORY IN PHYSICS functions of a unitary representation of Гр. The matrix elements can then be found from Equation (1-26), that is, from dp F(T)C(r) = £ Г(Тр)тп<(г), (5.12) m=l as P(T)^p(r) can be found for each T e Q using Equation (1.17). This method will be illustrated by using it to obtain a matrix for the irreducible representation Г5 of D4, which is an academic exercise here, as a set is already known. Example II Determination of matrix elements of the two-dimensional ir- reducible representation Г5 of the crystallographic point group D4 from its character system Take ф(г) = zF(r), where F(r) is any function of r such that ф(т) is nor- malized. Then, from Table 4.1 and Equation (5.11), Ррф(т) = zF(r). As P(C^/1{Pp^(r)}) = xF(r\ and as d5 = 2, zF(r) and xF(r) together give the totality of linearly independent functions Р(Т){Ррф(г)}. It happens here that zF(r) and xF(r) are orthogonal, so the Schmidt process is not needed. Then, as xF(r) is also normalized, one may take V>i(r) = zF(r) and 7^2 (r) = ж-^(р)- Then, for example, F(C'4!/)V'i(r) = —xF(r) = 1 = zF(r) = V’i(r), J which, in comparison with Equation (5.12), gives as the matrix representing (>4^ 0 1 ' _ -i 0 ' The matrices representing the other elements of D4 may be found in the same way. They are not identical to those quoted in Example I above, but could be obtained from them by a similarity transformation (Equation (4.5)) with 0 1 2 Direct product representations In Appendix A, Section 1, the definition is given of the direct product A®B of an m x m matrix A and an n x n matrix В in which A 0 В is an mn x mn matrix whose rows and columns are each labelled by a pair of indices in such a way that (cf. Equation (A.8)) (A 0 B)jS ^^ — AjkBst (1 < J, к < m; 1 < s, t < n). Certain properties of direct product matrices are also deduced in that section.
REPRESENTATIONS - DEVELOPMENTS 71 The following theorem shows that this definition allows the construction of “direct product representations”, which, through the Wigner-Eckart The- orem, play a major role in the applications of group theory in quantum me- chanical problems. (They are sometimes called “Kronecker product represen- tations” or “tensor product representations”.) Theorem I If Гр and Г9 are two unitary irreducible representations of a group Q of dimensions dp and dq respectively, then the set of matrices defined by Г(Т) =Гр(Т)®Г«(Т) (5.13) for all T e Q form a unitary representation of Q of dimension dpdq. The character x(T) of T e Q in this representation is given by x(T) = xp(T)x«(T). (5.14) Proof For any Ti and T^ of P, by Equation (5.13), Г(Т1)Г(Т2) = {ГР(Т1)0П(Т1)}{ГР(Т2)0П(Т2)} = {ГР(Т!)ГР(Т2)} 0 {Г^ТОГ^Т,)} (on using Equation (A.9)), so Г(Т1)Г(Т2) = Гр(Т1Т2) 0 Г9(Т1Т2) (as Гр and Г9 are themselves representations), and hence Г(Т1)Г(Т2)=Г(Т1Т2). Thus the matrices Г(Т) of Equation (5.13) certainly form a representation of Q and its dimension is obviously dpdq. As the direct product of any two unitary matrices is itself unitary (see Appendix A, Section 1), each matrix Г(Т) of Equation (5.13) must be unitary. Finally, as the diagonal elements of ГР(Т) 0 Г9(Т) are labelled by the pairs (j, s) and (fc, t) with j = к and s = for any T e Q dp dq х(П = ££(rp(T)®r*(T))JSJS j = l s = l dp dq = ££rp(T)^r«(T)ss j = l S = 1 = xp(T)xq(T)- The direct product representation defined by Equation (5.13) will be de- noted by Гр 0Г9. Although its definition in terms of matrix elements of Гр
72 GROUP THEORY IN PHYSICS and Г9 may appear complicated, in terms of bases the definition is completely natural, as will be demonstrated in the next two sections. In general the representation Гр 0 Г9 is reducible, even if Гр and Г9 are themselves irreducible. For example, for the crystallographic point group jD4, as Г5 is two-dimensional, Г5 0 Г5 must be four-dimensional. However, Z)4 has no irreducible representations of dimension greater than two, so Г5 0 Г5 must be reducible. Henceforth in this Section it will be assumed that Q is a finite group or a compact Lie group. Then all the irreducible representations of Q may be assumed to be unitary and every direct product Гр 0 Г9 is either irreducible or is completely reducible. Suppose that a similarity transformation with a dpdq x dpdq non-singular matrix C is applied to Гр 0 Г9 to give an equivalent representation that is a direct sum of unitary irreducible representations, and the unitary irreducible representation Гг of Q appears times in this sum. This can be written formally as ГЮë (5.15) or more precisely as cr1^ ® r9)c = ®4«rr’ (5.16) where the right-hand side is called the “Clebsch-Gordan series for Гр 0 Г9”. For the case in which Q is a finite group of order g, Theorem V of Chapter 4, Section 6 gives nrpq = (1/p) £ xp WCOxW, (5.17) T&3 the corresponding expression when Q is a compact Lie group being = [ хчттттту dT. (5.18) Jg Thus in these cases the Clebsch-Gordan series is determined solely by the characters. Obviously, as Гр 0 Г9 is of dimension dpdq, dpdq = nT)Qdr, where dr is the dimension of the irreducible representation Гг. Example I Clebsch-Gordan series for the crystallographic point group D4 Table 4.1 and Equation (5.17) together imply that Г5 0 Г4 « Г5 (i.e. n|4 = n|4 = n|4 = n^4 = 0, n|4 = 1), Г5 ®Г3 « Г5,
REPRESENTATIONS - DEVELOPMENTS 73 and Г5 ® Г5 » Г1 ф Г2 ф Г3 ф Г4. These particular Clebsch-Gordan series will be needed in later examples. The other series may be found in the same way. Two useful symmetry properties follow immediately from Equations (5.17) and (5.18), namely ^qP ^pq (5.19) and nP*r Прд-) (5.20) where in Equation (5.20) Гр* is the irreducible representation of Q defined by ГР*(Т) = {ГР(Т)}* for all T e P, the identity (Equation (5.20)) being a consequence of the fact that ri^ must be an integer and so must be real. 3 The Wigner-Eckart Theorem for groups of coordinate transformations in IR3 Theorem I Suppose that Q is a group of coordinate transformations in IR3 having irreducible representations Гр and Г9 of dimensions dp and dq respectively, and ^(r), j = 1, 2,..., dp, and ^(r), s — 1,2,... ,dq, are basis functions of Гр and Г9 respectively. Then the set of dpdq functions ^(r)^j(r) (where j = 1, 2,..., dp and s = 1, 2,..., dq) form a basis for the direct product representation Гр 0Г9, provided that they form a linearly independent set. Proof For any T G P, from Equation (1.17), P(T){<^(r)^(r)} = ^({R(T)|t(T)}-1r)^({R(T)|t(T)}-1r) = {P(T)^(r)}{P(T)^(r)} dp dq = {£r₽(T)fcj^(r)}{£r(T)ts^(r)} k=l t=l dp dq - EE(r k=l t=l which is of the form of Equation (1.26). Now suppose that Q is a finite group or a compact Lie group and is the number of times that the irreducible representation Гг appears in Гр 0Г9, where Гр, Г9 and Гг are all assumed to be unitary. If 0 there must be n?q linearly independent sets of basis functions for Гг formed from linear combinations of the products </>^(r)^f(r). Let these be denoted by $pQ(r),
74 GROUP THEORY IN PHYSICS for all T e I = 1,2,..., dr, and a the form d'p dt u= = 1 / P The coefficients \ J are They can be regarded as forming a dpdq Q к a Q к where a = 1, 2,..., np , and I = 1, 2,..., dr, dr being the dimension of Гг, so that dr Р(Т)^(г) = £г(Т)ы^«(г) (5.21) 1 , 2,..., npq. These may be written in 7 Q)^’(r). (5.22) called “Clebsch-Gordan coefficients”. x dpdq non-singular matrix, the rows being labelled by the pairs (j, fc), where j = 1, 2,..., dp and к = 1,2,..., dq and the columns being labelled by the triples (r,a, Z), where r appears only if npq / 0, in which case a = 1, 2,..., npq and I = 1, 2,..., dr. In fact this is the matrix C of Equation (5.16). The inverse of Equation (5.22) can be written as npq dr z r a=l1=1 a (5.23) for all j = 1, 2,..., dp and к = 1, 2,..., dq, where the sum over r is over all those irreducible representations Гг for which npq 0. The coefficients ( т Q P Q \ ’ . , I again form a dpdq x dpdq non-singular matrix, but this \ J к / time the rows are labelled by the triples (r, a, Z) and the columns by the pairs (j, k). Clearly this is the matrix C-1 of Equation (5.16). As ГР®Г9 is unitary and the direct sum on the right-hand side of Equation (5.16) is also unitary, Theorem IV of Chapter 4, Section 3 shows that C may be chosen to be unitary, which implies that P Q \ j к J r, I (5.24) a P Q j к Even if the matrix of Clebsch-Gordan coefficients is chosen to be unitary, there is still a degree of arbitrariness in the Clebsch-Gordan coefficients. (See, for example, Chapter 5, Section 3, of Cornwell (1984) for a detailed discussion of this point). If, for a given pair p and q, npq < 1 for every r, then it is easy to construct a simple formula for the Clebsch-Gordan coefficients that corresponds to this choice. For a finite group Q this is p q r, a\ = (dr/g)^ ГР(Т)^(Т)^Гта s t u ) ЕтебГР(Т)^^(Т)^Г’'(Т)й where the set of indices J, к,1 are chosen so that the denominator is non-zero. (See Chapter 5, Section 3, of Cornwell (1984) for a proof of this result.)
REPRESENTATIONS - DEVELOPMENTS 75 Although this formula generalizes in the obvious way when Q is a compact Lie group, for such a group it is much easier to use Lie algebraic methods to calculate the Clebsch-Gordan coefficients. For example, for SO(3) and SU(2) a description may be found in Chapter 12, Section 5, of Cornwell (1984), the extension for other simple compact Lie groups being given Chapter 16, Section 6, of Cornwell (1984). When there exists an n^q such that n^q > 1 the construction of a unitary matrix of Clebsch-Gordan coefficients is more difficult and for finite groups there is little advantage in making such a choice. (A detailed discussion may be found in the work of van den Broek and Cornwell (1978).) The case in which Г9 = Г1, where Г1 is the one-dimensional identity representation defined by Г1(Т) = [1] for all T E G, provides a simple but important example. In this situation Г 0Г1 = rp (5.26) for every irreducible representation Гр of Q. Moreover Equation (5.25) and its generalization for compact Lie groups, when taken with the orthogonality theorem for matrix representations (Theorem IV of Chapter 4, Section 5), show that the corresponding Clebsch-Gordan coefficients are given by 1 1 и P s (5-27) Example I Clebsch-Gordan coefficients for the crystallographic point group A Using the matrices of the two-dimensional irreducible representation Г5 spec- ified in Example I of Section 1, the Clebsch-Gordan coefficients corresponding to the series Г5 0 Г4 « Г5 are given by Equation (5.25) (with j = 1, к = 1, Z = 2) as 5 4 1 1 5 4 1 1 5, 1 2 5, 1 1 5 4 2 1 5 4 2 1 5, 1 1 5, 1 2 1, 0. Similarly for Г5 0 Г3 ~ Г5, Equation (5.25) (with j = 1, к = 1, I = 2) gives /53 5, 1 V 1 1 2 /53 5, 1 \ 1 1 1 5 3 2 1 5 3 2 1 5, 1 1, 0. 5, 2 1 1 Likewise, for Г5 0 Г5 « Г1 ф Г2 ф Г3 ф Г4, Equation (5.25) implies that all
76 GROUP THEORY IN PHYSICS the Clebsch-Gordan coefficients are zero except for the following: 5 5 1 1 1, 1 1 ) / 5 \ 2 5 2 1, 1 A 1 J = 2-1/2, 5 5 2, 1 / 5 5 2, 1 \ = 2-V2, 1 1 1 \ 2 2 1 ) 5 5 3, 1 f 5 5 з, i A = 2-V2, 2 1 1 1 2 1 J 5 5 4, 1 \ / 5 5 4, 1 A = 2-V2. 2 1 1 \ 1 2 1 / The Wigner-Eckart Theorem depends on one further concept, that of “ir- reducible tensor operators”. Definition Irreducible tensor operators for a group of coordinate transfor- mations in IR3 Let С??, • • • be a set of dq linear operators that act on functions belonging to the Hilbert space L2 and which satisfy the equations dq P(T)Q'p>(T)-1 = £r’(7%Q’ (5.28) k=l for every j = 1, 2,..., dq and every T of a group of coordinate transforma- tions P, where Г9 is an irreducible representation of Q of dimension dq. Then QiiQh... are said to be a set of “irreducible tensor operators” of the irre- ducible representation Г9 of Q. Equations (5.28) are to be interpreted as operator equations, that is, both sides must produce the same result when acting on any function of the com- mon domain in L2 of the operators Qq-. Moreover, on the left-hand side of Equations (5.28), each operator acts on everything to its right. Example II The Hamiltonian operator as an irreducible tensor operator Let Q be the group of the Schrodinger equation for some system. Then, from Equation (1.23), Р(Т)Я(г)Р(Т)-1 = Я(г) for all T e Q. Comparison with Equations (5.28) shows that Я(г) is an irreducible tensor operator for the one-dimensional identity representation of the group of the Schrodinger equation. Example III Differential operators as irreducible tensor operators of the crystallographic point group For any rotation T, PtT^PtT)-1 PtT^PtT)-1 ад11£+ад21^ + ад31£,' ад12^ + ад22^ + вд32^, > Вд1з^+ад2з<+вд33<,, (5.29)
REPRESENTATIONS - DEVELOPMENTS 77 where R(T) is the 3x3 orthogonal matrix specifying T. The first of the Equations (5.29) will now be proved in some detail to illustrate the type of manipulation that is usually involved. For any differential function /(r) of L2, Equation (1.17) gives F(T)-1/(r) = P(T-1)/(r) = /(?), where r' = R(T)r. Thus — lP(T~Af(PA =-----------+ — - — H------------- dxX V dx dx' dx By' dx dzf - й(Т)117^+ВД21^ + ВДз1~д^’ (5.30) on using Equation (1.2). Now define h(r) = df(r)/dx and put g(r) = where r' = R(T)r.. Then, by Equation (1-17), P(T){df(r')/dx'} = P(T)h(r') = F(T)ff(r) = ff(R(T)-1r) = h(r) = df(r)/dx. A similar argument applied to the second and third terms of Equation (5.30) then gives the first of Equations (5.29) immediately. Inspection of the matrices R(T) for D4 (see Example III of Chapter 1, Section 2) shows that R(T) 12 = ^(T)2i = #(Т)2з = Р(Т)з2 = 0 for all T of D4 and R(T)22 = r3(T)n (= x3(7")), where Г3 is the one-dimensional irreducible representation of D4 given in the character table, Table 4.1. Thus, from Equations (5.29), рт^р(т)-1 = г3(т)и-^, so that d/dy is an irreducible tensor operator transforming as Г3. Inspection also shows that R(T)^ R(T)13 _ Я(Т)з1 Я(Г)зз . for all T of Z?4, where Г5 is the two-dimensional irreducible representation of D4 (see Example I of Section 1). Thus Equations (5.29) show that Equation (5.28) is satisfied with q = 5 and = d/dx, = d/dz, so d/dx and d/dz constitute a set of irreducible tensor operators for Г5. Example IV Multiplication by a basis function as an irreducible tensor op- erator Let ^J(r), Э = 1,2, ...,dg, be a set of basis functions for the irreducible representation Г9, and define Qj by Qj/(r)=V-J(r)/(r) r5(T) =
78 GROUP THEORY IN PHYSICS for J = 1,2,... ,dq, i.e. QJ is the operation of multiplication by V,J(r)- Then Qi, Q2, • •. form a set of irreducible tensor operators of Г9, for F(T)QjP(T)-1/(r) = P(T)[^(r){F(T)-1/(r)}] = {P(T)V>J(r)}{P(T)P(T)-1/(r)} = {£r’(T)fcj<(r)}/(r) k=l = £r’(T)fcJQ’/(r). k=l a p q j к (5.31) Theorem II The Wigner-Eckart Theorem for a group of coordinate trans- formations in IR3 Let Q be a group of coordinate transformations that is either a finite group or a compact Lie group. Let Гр, Г9 and Гг be unitary irreducible representations of Q of dimensions dp, dq and dr respectively, and suppose that (r), j = 1,2,..., dp, and (r), I = 1, 2,..., dr, are sets of basis functions for Гр and Гг respectively. Finally, let Qqk, к = 1, 2,... ,dq, be a set of irreducible tensor operators of Г9. Then пРЯ .QIW = E a=l for all j = 1, 2,..., dp, к = 1,2,..., dq, and I = 1, 2,..., dr, where (r\Qq|p)a form a set of n^q “reduced matrix elements” that are independent of 7, к and I. Proof See, for example, Appendix C, Section 6, of Cornwell (1984). It should be noted that it is not required that the matrix of Clebsch-Gordan coefficients must be unitary. The Wigner-Eckart Theorem provides both the most succinct and the most powerful expression in the whole field of application of group theory in physical problems. Indeed, most physical applications depend directly on it. It shows that the 7, k,l dependence of the quantities is given completely by the Clebsch-Gordan coefficients. Moreover, the whole set of dpdqdr elements (^[, depend only on n^q reduced matrix elements. The theorem has been stated here for the case in which Q is a group of coordinate transformations in IR3 that is either a finite group or a compact Lie group. However, it may be generalized quite easily to any non-compact, semi-simple Lie group, both for the case in which the representations are finite- dimensional (Klimyk 1975) and the case in which they are unitary but infinite- dimensional (Klimyk 1971). Further generalization to unitary representations of non-semi-simple, non-compact Lie groups has also been achieved (Klimyk 1972). See also Agrawala (1980).
REPRESENTATIONS - DEVELOPMENTS 79 The actual definition of the reduced matrix elements is dp dq ^pq z (ri№)« = (w£EEEP qt S = 1 t=l tt=l Q=1 2 ay^,Qqt<K\ (5.32) but in practice the simplest way of determining them is to find nrpq non-zero elements (V>[, (either by direct evaluation or by fitting to experimental data) and then regard the nrpq equations (Equations (5.31)) in which these elements appear on the left-hand side as a set of simultaneous equations in (’’IQ9|p)a, OC= The application of the Wigner-Eckart Theorem to a number of physical problems is described in detail in Chapter 6, particularly in Sections 2 and 3. Frequent use is also made of the following special case. Theorem III If </>? (r) (for j = 1,2,..., dp) and ^(r) (for A; = 1,2,..., dg) are respectively basis functions for the unitary irreducible representations Гр and Г9 of the group of the Schrodinger equation Q that is either a finite group or a compact Lie group, and Гр and Г9 are not equivalent if p q (but are identical if p = q), and if H(r) is the Hamiltonian operator, then (С,я<) = о unless p = q and m = n. Moreover, if p = q and m = n, then (c/)^, Нф1^) is a constant independent of m. Proof As noted in Example II above, H(r) is an irreducible tensor operator of the one-dimensional identity representation Г1 of Q. The required result then follows immediately from the Wigner-Eckart Theorem on using Equations (5.26) and (5.27). Alternatively, this theorem may be proved by a simple generalization of the proof of Theorem V of Chapter 4, Section 5, of which it is an obvious extension. 4 The Wigner-Eckart Theorem generalized It will now be shown how the developments of the previous section can be expressed in terms of the linear operators and carrier spaces first introduced in Chapter 4, Section 1, thereby enabling the theory to apply to any group Q and not merely to groups of coordinate transformations in IR3. Suppose that the irreducible representations Гр and Г9 of Q have dimen- sions dp and dq and that (for j = 1,2,..., dp) and (for 5 = 1,2,..., dq) are ortho-normal bases for the two corresponding abstract inner product spaces Vp and Vq. A dpdg-dimensional “direct product space” Vp 0 Vq may be defined as the set of all quantities ф of the form dp dq Ф = а^Рз 0 j = \ s = l
80 GROUP THEORY IN PHYSICS where aJS are a set of complex numbers. (This concept is developed in more detail in Appendix B, Section 7.) With an inner product in Vp 0 Vg defined by dp dq (Ф’Ф) = ^2^ajsbjS, j = l s=l where dp dq Ф = ®Фз, j=l S=1 the products 0 'Фв f°r 7 = 1,2,..., dp, and s = 1, 2,..., dq, form an ortho- normal basis for Vp 0 Vq. Now define the linear operators ФР(Т) and Ф9(Т) for all T e Q acting on the bases of Vp and Vq respectively by dp k=l for j = 1, 2,..., dp, and dq = '£rg(T)ts^ t=l for s = 1,2, ...,dg. These are essentially just Equations (4.1) embellished with extra indices, so ФР(Т) and $q(T) may be extended to the whole of Vp and Vq respectively. For each T e Q, a further linear operator Ф(Т) acting on Vp 0 Vq may be defined by Ф(Т){^р ® Ф1} = {Фр(Т)<} ® {Ф9(ЭД} (5.33) and again extended to the whole of Vp and V9, so that dp dq Ф(Т){^ ® = £ E(r₽(T) ® r«(T))fctJsM ® V’t} (5.34) k=l t=l for all 7 = 1,2,..., dp, and s = 1,2,..., dq. Thus the operators Ф(Т) are the linear operators corresponding to the direct product representation Гр 0 Г9 of Q. / n As the Clebsch-Gordan coefficients , \ J к ements of a matrix C that completely reduces Гр 0Г9 (see Equation (5.16)), a ) are the matrix el- it follows that for 0 the elements of Vp 0 Vq defined by i / P I Q (5.35)
REPRESENTATIONS - DEVELOPMENTS 81 satisfy dr Ф(Т)^’“ = £Г(Т)Ы^ (5.36) U=1 for all T E G, I = 1,2,..., dr and a = 1,2,..., nrpq. That is, again the Clebsch- Gordan coefficients give the appropriate linear combinations that form bases for the various irreducible representations of Гр 0P, the similarities between Equations (5.35) and (5.22) and between Equations (5.36) and (5.21) being particularly significant. (In comparing the developments of this section with those of Section 3, it must be observed that the products </>p(r)V>f (r) of the basis functions </>p(r) of Гр and Vd?(r) °f Г9 form a basis for a dpdg-dimensional subspace of L2 only if they are linearly independent, and even then these products do not necessarily form an ortho-normal set with respect to the usual inner product of L2. By contrast, the products фр- 0 of basis vectors and of Vp and Vq always form an ortho-normal basis of Vp 0 Vq with the inner product defined as above. Thus for basis functions in general one cannot identify фр(г) 0 ^f(r) whh ^(r)^j(r) and at the same time take the inner product of Vp 0 Vq to be that of L2.) To proceed further it is necessary to redefine the concept of a set of irre- ducible tensor operators. To this end let Q be a linear mapping of Vp into Vr (Vp and Vr being carrier spaces for the irreducible representations Гр and Гг of G) so that Qi'p E Vr for all Ф E Vp. Defining the sum (Qi + Q2) of two such operators Q± and Q2 by (Qi + Qz№ = Qi^ + (for all £ Vp\ the product aQ for any complex number a by (aQ)V> = «(QVO (for all 'Ф £ V”p), and the “zero” mapping 0 by (Уф = 0 (for all G Vp, where the 0 on the right-hand side here is the “zero” element of Vr), it follows that the set of all linear mappings Q from Vp to Vr form a vector space, which will be denoted by L(VP, Vr). If Vp and Vr are of dimensions dp and dr respectively, then L(VP, Vr) is of dimension dpdr (Shephard 1966). Now define for each T E G an operator Ф'(Т) acting on L(VP, Vr) by $\T)Q = $r(T)Q$p(T)~1 for all Q E L(VP, Vr), ФР(Т) and ФГ(Т) being the operators acting in Vp and Vr belonging (in the manner described above) to the irreducible representa- tions Гр and Гг. Then Ф'(Т) is a linear operator, and for any Ti,T2 EG Ф^Т^Ф'^) = Ф'СГ^), so that the set of operators Ф'(Т) correspond to a representation of G for which the carrier space is L(VP, Vr). (The proof of this statement is as follows. For any Ti, T2 E G and any Q E L(VP, Vr), Ф'(Т1)Ф'(т2)е = ФГ(Т1){ФГ(Т2)<2ФР(Т2)-1}ФР(Т1)-1 = Ф'(Т1Т2)<Э.)
82 GROUP THEORY IN PHYSICS Let Г' be the representation of Q for which the operators Ф'(Т) and the carrier space L(VP, Vr) form a module. That is, if Qi, Q2, • • • are a basis of the vector space L(VP, Vr), the matrix elements Г'(Т)шп are defined by dp d-p &(T)Qn = r'(T)mnQm m=l for all T e Q. In general Г' is reducible. Suppose that Г' is completely reducible and that Г9 is an irreducible representation that appears in its reduction, and let Qq^Q^... be a basis for the corresponding subspace of L(Vp,Vr)- Then dq &(T)Q«n = £ r«(T)mnQ^ m=l for n = 1,2,..., dq and all T E Q. That is, by the definition of Ф'(Т), dq ФГ(Т) Qqn ФР(Т)-1 = 52 r’(T)mnQ^ (5.37) 771=1 for n = 1,2, ...,dg, and all T e Q. This set of operators will be called “irreducible tensor operators of the irreducible representation Г9 of P”. Theorem I The generalized Wigner-Eckart Theorem Let Q be a finite group or a compact Lie group. Let Гр, Г9 and Гг be unitary irreducible representations of Q of dimensions dp, dq and dr respectively, and suppose that фр- (J = 1,2,..., dp) and (Z = 1,2,..., dr) are basis vectors of ortho- normal bases of the carrier spaces Vp and Vr of Гр and Гг respectively. Finally, let Qqk (k = 1,2,..., dq) be a set of irreducible tensor operators of Г9, defined as above. Then пря / = E • I CE=1 r, a I (r\Qq\P)a (5.38) for all j = 1,2,..., dp, к = 1, 2,..., dq and I = 1,2,..., dr, where (r|Q9|p)a are a set of n^q “reduced matrix elements” that are independent of j, к and I. Proof See, for example, Chapter 5, Section 4, and Appendix C, Section 6, of Cornwell (1984). It should be noted that the appropriate inner product on the left-hand side of Equation (5.38) is that of Vr, as and Qk(fp are both members of Vr. The remarks made in Section 3 about the Wigner-Eckart Theorem for a group of transformations in IR3 apply equally to the theorem as generalized above. In particular, although the theorem is stated and proved here for the case in which Q is a finite group or compact Lie group, the conclusion is valid
REPRESENTATIONS - DEVELOPMENTS 83 much more generally. A detailed discussion of the range of validity has been given by Agrawala (1980). In a minor extension of this formalism, one could introduce an inner prod- uct space V that is a direct sum of carrier spaces of certain unitary irreducible representations of Q and which contains at least Vp ф Vr (and which, in the extreme case, may contain one carrier space for every inequivalent irreducible representation of P). Then, for each T e Q an operator Ф(Т) can be defined which maps elements of V into V, and which acts as ФР(Т) on Vp, as ФГ(Т) on Vr, and so on. The irreducible tensor operators are then required to each map V into V and to be such that Ф(Т) Ф(Т)-1 = Г«(Т)тп(2Яг for all T E Q and all n = 1,2,..., dq. In this case the Wigner-Eckart theorem deals with inner products defined on V, but is otherwise the same as above. 5 Representations of direct product groups The concept of direct product groups was discussed in some detail in Chapter 2, Section 7. In studies of their representations it is most convenient (and quite sufficient) to revert to the original formulation in terms of pairs. Theorem I Let Г1 and Г2 be representations of Pi and P2 respectively. Then the set of matrices Г((Т1,Т2)) defined for all Ti e Pi and T2 e P2 by Г((Т1,Т2)) = Г1(Т1) 0Г2(Т2) (5.39) provides a representation of Pi 0P2. This representation of Pi 0P2 is unitary if Г1 and Г2 are unitary representations and is faithful if Г1 and Г2 are faithful representations. Proof For any T^T^ E Pi and any T^Tj E P2, from Equation (5.39), Г^Т^Г^Т'.Т')) = {Г1(Т1)0Г2(Т2)}{Г1(Т1')0Г2(Т')} (on using Equation (A. 9)), so Г^Т^Г^Т')) = Г1(Т1Т') 0Г2(Т2Т'), (as Г1 and Г2 are assumed representations of Pi and P2 respectively), and hence Г^Т^Г^Т')) = Г^Т.Т^Т')) (on using Equation (2.11)). Consequently the matrices Г of Equation (5.39) form a representation of Pi 0 Qz- The unitary property follows from the fact that the direct product of two unitary matrices is itself unitary (see Appendix A, Section 1), while the faithful property is obvious.
84 GROUP THEORY IN PHYSICS This theorem allows the nature of Pi 0 P2 to be investigated when Pi and P2 are finite groups or linear Lie groups. There are essentially three distinct cases: (i) Pi and P2 are both finite groups. In this case clearly Pi 0 P2 is a finite group whose order is the product of those of Pi and P2 separately. (ii) Pi is a finite group and P2 is a linear Lie group. Suppose that Pi has order g± and has a faithful finite-dimensional rep- resentation Г1. Suppose that P2 has a faithful finite-dimensional repre- sentation Г2, that the elements of P2 near the identity are specified by n real parameters #i, #2, • • •, and that P2 has N connected compo- nents. Then the faithful finite-dimensional representation of Equation (5.39) can be used to show that Pi 0 P2 is a linear Lie group with Ng1 connected components, whose connected subgroup is isomorphic to the set of matrices Fi(E'i) 0 Г2(Т2) for all T2 of the connected subgroup of p2. Moreover, the elements of Pi 0 P2 near the identity of Pi 0 P2 may be specified by the same n real parameters as for P2. As the “in- variant integral” of Pi 0 P2 involves an integral over n variables with the same weight function as for P2 and a sum over the Ng± connected components, it is obvious that Pi 0 P2 is compact if and only if P2 is compact. (iii) P2 is a finite group and Pi is a linear Lie group. This is just the same as the previous case with the roles of Pi and P2 interchanged. (iv) Pi and Q2 are both linear Lie groups. Suppose that Tj is a faithful finite-dimensional representation of pj, that the elements of P7 near the identity of Py are specified by n3 real parameters, and that Qj has Nj components (J = 1,2). Then the faithful finite-dimensional representation (Equation (5.39)) of Pi 0 P2 can be employed to prove that P10P2 is a linear Lie group with TV] A2 connected components and that the elements of Pi 0P2 near the identity of Pi 0P2 are specified by (ni + n2) real parameters. The “invariant integral” of Pi 0 P2 therefore involves an integral over (ni + n^) variables (whose weight function is the product of those of Pi and P2 separately) and a sum over the N1N2 components, so that Pi 0 P2 is compact if and only if Pi and P2 are both compact. Theorem II If Pi 0 P2 is a finite group or a compact linear Lie group and Г1 and Г2 are irreducible representations of Pi and P2 respectively, then the representation Г defined by Equation (5.39) is an irreducible representation of Pi 0 P2. Moreover, every irreducible representation of Pi 0 P2 is equivalent to a representation constructed in this way. Proof See, for example, Appendix C, Section 7, of Cornwell (1984).
REPRESENTATIONS - DEVELOPMENTS 85 6 Irreducible representations of finite Abelian groups The irreducible representations of every finite Abelian group Q may now be found very easily. It should be recalled that Theorem III of Chapter 4, Sec- tion 5 shows that these representations must be one-dimensional, and as ev- ery representation of a finite group is equivalent to a unitary representation (equivalence implying identity for one-dimensional representations), all these irreducible representations are automatically unitary. Moreover, Theorem VIII of Chapter 4, Section 6 and Theorem II of Chapter 2, Section 2 together imply that the number of inequivalent irreducible representations of Q is equal to the order of Q. The first stage is to consider a special type of Abelian group. Definition Cyclic group A group is said to be “cyclic” if every element can be expressed as a power of a single element. The most general form of a finite cyclic group of order g is therefore {E, T, T2,..., Tp-1}, with T9 = E, the element T being called the “gen- erator” of p. Obviously every cyclic group is Abelian. It is easily shown that all cyclic groups of the same order are isomorphic. Theorem I The set of all unitary irreducible representations of a cyclic group of order g is given by Гр(Тт) = [ехр{2тггт(р - 1)/#}] (5.40) for m = 1,2,... ,g. Here p takes values p = 1,2,..., #, and T is the generator of Q. Proof Suppose that Г is an irreducible representation and Г(Т) = [7], where 7 is some complex number. Then у9 = 1 as T9 = E, so 7 can take any of the g possible values 7 = ехр{2тп(р — 1)/#}, where p = 1,2,... ,^. These g values of 7 then give the g inequivalent irreducible representations, which may therefore be labelled by p. As ГР(ТШ) = {Гр(Т)}т = [7m], Equation (5.40) follows immediately. The factor ехр{2тггт(р — 1)/#} has been introduced in Equation (5.40) instead of the factor ехр{2тггтр/^} simply to ensure conformity with the usual convention that Г1 is the identity representation. The following theorem shows that any finite Abelian group is made up of cyclic groups and so enables all its irreducible representations to be calculated immediately. Theorem II Every finite Abelian group is either a finite cyclic group or is isomorphic to a direct product of a set of finite cyclic groups.
86 GROUP THEORY IN PHYSICS Proof See, for example, Rotman (1965) (pages 58 to 62). Example I Irreducible representations of groups isomorphic to Cr®Cs, Cr and Cs being cyclic groups of order r and s respectively Let Tf be an irreducible representation of Cr, so that, by Equation (5.40), rf(Tim) = [ехр{2тггт(р — l)/r}], where T± is the generator of (7r. Simi- larly, let Г2 be an irreducible representation of Cs, with generator T2, so that Г2(Т2П) = [ехр{27гш(д — l)/s}]. Then, by the theorems of the previous sec- tion, the irreducible representations of every group isomorphic to Cr 0 Cs may be labelled by a pair (p, q), where p = 1,2,..., r, and q = 1,2,..., s, and, from Equation (5.39), ^((T^, T2n)) = [exp{27rz({m(p - l)/r} + {n(q - l)/5})}] for all m = 1, 2,..., r and n = 1,2,..., s. The crystallographic point group Z)2 (see Appendix C) is an example hav- ing this structure, as it is isomorphic to C2 0 C2. 7 Induced representations The method of “induction” provides a very powerful technique for construct- ing representations of a group from representations of its subgroups. It will be described here for the case in which the group is finite and the results ob- tained will be applied in Chapter 7 to the symmorphic crystallographic space groups. However, the technique is not restricted to finite groups. Indeed, one of the most significant developments of the last few years has been the generalization to arbitrary, locally compact topological groups, including par- ticularly Lie groups. This development has been largely pioneered by the work of Mackey (1963, 1968, 1976). It has proved extremely valuable in the con- struction of the infinite-dimensional unitary representations of noncompact, semi-simple Lie groups (Stein 1965, Lipsman 1974, Barut and Raczka 1977), thereby putting into a general context the original work on the homogeneous Lorentz group (Gel’fand et al. 1963, Naimark 1957, 1964). Other physi- cally important non-compact Lie groups that are particularly well suited to treatment by the induced representation method include the Poincare group (Wigner 1939, Bertrand 1966, Halpern 1968, Niederer and O’Raifeartaigh 1974a,b), the Galilei group (Inonii and Wigner 1952, Voisin 1965a,b, 1966, Brennich 1970, Niederer and O’Raifeartaigh 1974a) and the Euclidean group of IR3 (Miller 1964, Niederer and O’Raifeartaigh 1974a). Most of the results to be derived in this section for finite groups carry over to the general case of locally compact topological groups with their group theoretical content essentially unchanged. The complications of the general case lie in the measure theoretic questions involved, together with the fact that nearly all the representations that appear are infinite-dimensional. Coleman (1968) has given a very readable introduction to these matters. The basic theorem on induced representation is easily stated and proved:
REPRESENTATIONS - DEVELOPMENTS 87 Theorem I Let S be a subgroup, of order s, of a group Q of order g, and let 71,72,... be a set of M(= д/s') coset representatives for the decomposition of Q into right cosets with respect to 5. Let A be a d-dimensional unitary- representation of 5. Then the set of Md x Md matrices Г(Т), defined for all TePby ( А(тктт~^г, if TkTT-'eS, r(TW = |0; HTkTT-^s, (5-41) provides an Md-dimensional unitary representation of Q. If ty(S) are the characters of the representation A of 5, then the characters x(T) of the representation Г of Q are given by x(T) = Y^mTT-1), (5.42) j where the sum is over all coset representatives Tj such that TjTTfi1 e S. Proof See, for example, Appendix C, Section 8, of Cornwell (1984). This representation Г of Q is said to be “induced” from the representation A of the subgroup 5, this being indicated by writing Г = A(5) T Q. In Equations (5.41) the rows and the columns of Г(Т) are each separately labelled by a pair of indices, exactly as in the theory of direct product rep- resentations (see Section 2 and Appendix A, Section 1). The theorem (and proof) is also valid when Q and S are compact Lie groups such that the de- composition of Q into right cosets with respect to 5 contains only a finite number M of distinct cosets. For one physically important type of group the induced representation method not only produces irreducible representations of the group, but it generates the whole set of such representations. This satisfactory situation occurs when Q has the semi-direct product structure A® В and the invari- ant subgroup A is Abelian (see Chapter 2, Section 7). Physically important groups with this structure include the Euclidean group of IR3 (see Example II of Chapter 2, Section 7), the Poincare group, and the symmorphic crystal- lographic space groups (see Chapter 7). Of these only the latter are finite but all the results to be described can be generalized easily to the other groups. The construction of the unitary irreducible representations of Q involves a number of stages which will now be described in detail. It will be assumed that the orders of P, A and В are g, a and b respectively, so that g = ab. (a) As A is Abelian it possesses a inequivalent irreducible representations, all of which are one-dimensional and therefore completely specified by their characters. Let these characters be denoted by Хд(А), q = 1, 2, ..., a, for all A e A.
88 GROUP THEORY IN PHYSICS (b) Let B(q) be the subset of elements В of В such that x\(BAB~^ = x\(A) (5.43) for all A E A. Then B(q) is a subgroup of B. B(q) is called a “little group”. (As A is an invariant subgroup of P, BAB-1 E A for all A E A and all В E B, so В(q) is well defined. The subgroup property follows because, if В and Bf are members of B(q), then for any A E A, from Equation (5.43), x^B'B-^B'B-1)-1) = X^B-'AB) = х^(Л), so that B'B-1 is also a member of B(q).) Let b(q) be the order of B(q). (c) Let be the set of M(q)(= b/b(q)) coset representatives for the decomposition of В into right cosets with respect to B(q). (d) For each В E В define the quantities (A) for all A E A by Х^д\А) = x^BAB-1). (5.44) Then, for each fixed B, the set Хд^(-4) is a set of characters of a one-dimensional irreducible representation of A, so that B(q) is an in- teger in the set 1, 2,..., a. That Хд^(А) are such characters can be demonstrated as follows. Let A and A' be any two elements of A. Then = x^BAB-^x^BA'B-1) = x^BAB-^BA'B-1)) = х^(В(ЛЛ')В-1) (e) Obviously Equations (5.43) and (5.44) imply that B(q) = q for all В E B(q). More generally, B(q) = Bj(q) for every В E В belonging to the right coset B(q)Bj. This follows because if В E B(q)Bj then there exists an element B' of B(q) such that В = B'Bj. Then for any A E A Хл^Ча) = x^((B'JBi)A(B'JBj)-1) (by Equation (5.44)) = 1) (by Equation (5.43)) = Хд^9\А) (by Equation (5.44)). (f) The set of M(q)(= 6/6(q)) integers {Bi(q)(= q), B2(q), B3(q), ...} is known as the “orbit” of q. (g) The groups B(Bj(q)) are all isomorphic to B(q) for J = 1, 2,..., M(q). That is, all members of the orbit of q are associated with essentially
REPRESENTATIONS - DEVELOPMENTS 89 the same group B(q). (Equation (5.43) implies that B(Bj(g)) is the subgroup of В consisting of all В G В such that ХУЧ\ВАВ-^ = for all A e A. By Equation (5.44) this can be rewritten as X^BAITBN = xA(B:)AB~}') (5.45) for all A e A. Now consider the automorphic mapping of Q onto itself defined by фд(Т) = BjTB^1. As A is an invariant subgroup of P, this provides a one-to-one mapping of A onto itself. Consequently, let A' be any element of A and let A = B^A'Bj. Then Equation (5.45) can be further rewritten as XA((BjBBj~1)A,(BjBB~1')~1) = xqA(A') for all Af e A. Thus фд(В) = BjBB~r maps B(Bj(q)) onto B(q) and, as фj is an isomorphic mapping, B(q) is isomorphic to B(Bj(q)).) (h) Let S(q) be the set of all products AB, where A e A and В e B{q). Then S(q) is a subgroup of Q with the semi-direct product structure A@B(q). (If A, A' e A and B,B' e B(q), then, as A is an invari- ant subgroup of P, there exists an A" e A such that (B'B-1)A-1 = A"(B'B-1). Consequently (A'B')(AB)“1 = A'B'B^A-1 = AA"B'B-1, ' which is a member of <S(q), as AA" e A and B'B-1 e B(g). The semi-direct product structure of 5(q) follows directly from that of Q.) (i) Let be a unitary irreducible representation of B(q) of dimension dp. Then the set of dp x dp matrices A9,P(AB) defined by ^(АВ)=хдл(.А)Трвм(В) (5.46) for all A G A and В G В(g) form a dp-dimensional unitary representation of S(q). (That is a representation S(q) can be proved as follows. Let A, A' be any two elements of A and В, B' any two elements of B(q). Then there exists an A" G A such that BA'B-1 = A", so, from Equation (5.43), xp(^-z) — Xg(A")« Thus, from Equation (5.46), A9,P((AB)(A'B')) = A9,P(AA"BB') = X\{AA!'}T^BB'} = Д«’г’(АВ)Д«’р(Л'В/). The unitary property is obvious.)
90 GROUP THEORY IN PHYSICS (j) The set of M(q)(— b/b(q)) coset representatives Bi,B2,... for the de- composition of В into right cosets with respect to B(q) also serve as coset representatives for the decomposition of Q into right cosets with respect to <S(q). (The numbers of distinct right cosets in the two de- compositions are equal, as the number in the latter decomposition is g/s(q) = (ab)/(ab(q)) = M(q) (s(q) being the order of <S(q)). Moreover, S(q)Bj and S(q)Bk are distinct if and only if B(q)Bj and B(q)Bk are distinct. (To verify this, first suppose that S(q)Bj and S(q)Bk possess a common element. Then there exist A, A' E A and В, B' E B(q) such that ABBj = A'B'Bk- However, as S(q) is a semi-direct product of A and B(q), A = A' and BBj = B'Bk, so that B(q)Bj and B(q)Bk possess a common element. The demonstration of the converse proposition is then obvious.) (k) Unitary representations Tq,p of Q of dimensions M(q)dp may be induced from the unitary representations A9,p of S(q) by applying the previous theorem with S = S(q) and A = A9,p, That is, symbolically, Г9,р = A^p(5(q)) T Q. Let T be any element of Q and suppose that T = AB, where A E A and В E B. By (j) the coset representatives Ti, T2,... of the theorem may be identified with Bi,B2,.... Then B/C(AB)BJ“1 = (B/cAB^"1)(B/cBBj“1), where B^AB^1 E A, so Bk(AB)B~1 E <S(q) if and only if B&BB”1 E B(q). When B/-BB71 e B(q), Equations (5.44) and (5.46) give ^BkBB-1). Thus, from Equations (5.41), f HBkBBT1 ( ’ ’Jr 1 0, ifBkBB-1 (5-47) Similarly, Equation (5.42) implies that the characters xq'p(AB} of Tq,p are given by Xg,p(AB) = ^Xa^Wb^BB-1), (5.48) j where the sum is over all coset representatives Bj such that BjBB”1 E B(q), and where X#(g)(B) are the characters of The remarkable properties of these representations Г9,р of Q are summa- rized in the following theorem. Theorem II Let Tq,p be the unitary representation of the semi-direct prod- uct group P(= A®B) defined by Equations (5.47). Then
REPRESENTATIONS - DEVELOPMENTS 91 (a) Tq,p is an irreducible representation of and (b) the complete set of unitary irreducible representations of Q may be de- termined (up to equivalence) by choosing one q in each orbit and then constructing Г9,р for each inequivalent of 13 (q). Proof See, for example, Appendix C, Section 8, of Cornwell (1984). This construction will be used in Chapter 7 in the discussion of irreducible representations of symmorphic crystallographic space groups.

Chapter 6 Group Theory in Quantum Mechanical Calculations 1 The solution of the Schrodinger equation One of the most valuable applications of group theory is to the solution of the Schrodinger equation. Only for a small number of very simple systems, such as the hydrogen atom, is it possible to obtain an exact analytic solution. For all other systems it is necessary to resort to numerical calculations, but the work involved can be shortened considerably by the application of group representation theory. This is particularly true in electronic energy band calculations in solid state physics, where accurate calculations are only feasible when group theoretical arguments are used to exploit the symmetry of the system to the full. For simplicity consider the “single-particle” time-independent Schrodinger equation described in Chapter 1, Section 3(a), namely Я(г)^(г) — c^(r), (6.1) H(r) being the Hamiltonian operator (see Equation (1.10)). It is required to find the low-lying energy eigenvalues e and their corresponding eigenfunctions ^(r). The unknown function ^(r) can be expanded in terms of a complete set of known functions (r), fa (r),... that form a basis for L2 (see Appendix B, Section 3), that is, oo V'(r) = (6.2) .7=1 where ai, «2, • • • form a set of complex numbers whose values are unknown at this stage. The assumption is now made that the series (Equation (6.2))
94 GROUP THEORY IN PHYSICS converges sufficiently rapidly that only the first N terms need be retained. Then it can be replaced by the approximation N v>(r) = (6-3) Some judgement is required as to the best choice of the set ^q(r), ^(r), • • • that will ensure the validity of this approximation. Indeed, the different types of energy band calculation, described for instance in the article of Reitz (1955), essentially differ merely in this choice. For example, in solid state problems where the valence electrons are expected to be tightly bound to the ions, it is natural to take the Vb‘(r) be atomic orbitals, thereby giving the so- called “method of linear combinations of atomic orbitals”, often described more briefly as the “L.C.A.O. method”. At the other extreme, when the valence electrons are nearly free, it is natural to form the Vb’(r) from plane waves (orthogonalized to the ionic electronic energy eigenfunctions to prevent the expansion giving ionic electron energy eigenfunctions), thereby giving the so-called “orthogonalized plane wave method”, or “O.P.W. method” for short. Substituting Equation (6.3) into Equation (6.1) and forming the inner product of Equation (B.18) with V>fc(r) gives N Н-ф^ - е(^,^)} = 0. (6.4) j=i This is a matrix eigenvalue equation of the form of Equation (A. 10), in which the matrix elements (^., are known but the eigenvalues 6 and the el- ements aj of the eigenvector are to be determined. Equation (6.4) has a non-trivial solution if and only if det{(V>k, Wj) - = 0, (6.5) in which the matrix involved is of dimension N x N (cf. Equation (A. 11)). The left-hand side of the scalar equation derived from Equation (6.5) is a polynomial of degree AT, the roots of which are the eigenvalues 6. Both the explicit determination of the polynomial and the calculation of its roots are very lengthy processes if N is large. With the roots obtained it is possible to go back to Equation (6.4), regarded now as a system of N linear algebraic equa- tions, and for each root e obtain the corresponding set of complex numbers aj thereby giving by Equation (6.3) an approximation to the corresponding eigenfunction V>(r)- The number N is here quite arbitrary, but clearly, as N is increased, two effects follow. Firstly, the accuracy of the approximations to the lower energy eigenvalues is improved, which is very desirable. Secondly, more eigenvalues at higher energies appear, although these are usually less important. However, as noted above, the numerical work involved increases rapidly as N increases, this work being roughly proportional to N\. By invoking group representation theory this numerical work can be cut tremendously without any loss of accuracy. All that has to be done is to
QUANTUM MECHANICAL CALCULATIONS 95 arrange that the members of the complete set of known functions Vb(r) Equation (6.3) are each basis functions of the various irreducible representa- tions of the group of the Schrodinger equation, Q. In practice this is achieved by applying the projection operators of Chapter 5, Section 1 to the atomic orbitals, orthogonalized plane waves, or other given functions that are judged appropriate to the particular system under consideration. An extra pair of indices m and p has now to be included in the designation of the functions, so that ^Jm(r) transforms as the mth row of the irreducible representation Гр of P, the index j distinguishing linearly independent basis functions having this particular symmetry. Equation (6.3) is then rewritten as dp ^(Г) ='E'E'EaPjm^m^’ j p m=l (6.6) and Equation (6.5) becomes det{«, H^m) - e^kn, ^m)} = 0. (6.7) If Q is a finite group or a compact Lie group, each irreducible representation Гр may be taken to be unitary. Theorem V of Chapter 4, Section 5 then shows that W’L,tfm) = S9PSnm№km, tfm)- (6-8) Similarly, Theorem III of Chapter 5, Section 3, implies that Wjm) = Wjm). (6.9) The rows and columns of the determinant of Equation (6.7) may be rear- ranged so that all the terms corresponding to a particular row of a particular irreducible representation of Q are grouped together. (This can be achieved by successively interchanging pairs of rows of the determinant and then pairs of columns. Such interchanges in general change the sign of a determinant. However, here the value of the determinant is zero so its value is unchanged by such a rearrangement.) Equations (6.8) and (6.9) then imply that the determinant of Equation (6.7) takes the “block form” " D(l, 1) 0 0 D(l,2) 0 0 0 0 0 0 det 0 0 .. D(l,d!) 0 0 = 0, 0 0 0 D(2,l) 0 0 0 0 0 D(2,2) ... (6.10) where D(p, m) is a submatrix defined by D(p, m)kj = (^km, H^m) - e^km, ^m).
96 GROUP THEORY IN PHYSICS Thus D(p, m) involves only basis functions corresponding to the mth row of Гр. The matrices 0 consist entirely of zero elements. Equation (6.10) can be factorized to give П П detD(p,m) = 0. p m=l The complete set of eigenvalues of Equation (6.10) are then obtained by taking detD(p, m) = 0 (6.11) for every p and every m = 1, 2,..., dp. The energy eigenvalues corresponding to the mth row of Гр are therefore given by the secular equation, Equation (6.11), which only involves basis functions corresponding to the mth row of Гр. As the dimensions of D(p, m) are usually very much smaller than those of the determinant of Equation (6.7), very much less numerical work is now needed to find the energy eigenvalues and eigenfunctions for the same degree of accuracy. A further valuable saving of effort is provided by noting that Equations (6.8) and (6.9) also imply that D(p, 1) = D(p, 2) = ... = D(p, dp) (6.12) for each irreducible representation Гр. Thus only one secular equation (Equa- tion (6.11)) has to be solved for each irreducible representation Гр and each of the resulting energy eigenvalues can be taken to be dp-fold degenerate. It is very interesting to relate these results to the general conclusion drawn in Chapter 1, Section 4, that every d-fold degenerate energy eigenvalue is asso- ciated with a d-dimensional representation Г of the group of the Schtodinger equation, the corresponding d linearly independent eigenfunctions being basis functions of this representation . Suppose first that this representation Г is irreducible and is identical to Гр. Then the d-fold degeneracy of the energy eigenvalue follows automatically from the identities in Equations (6.12). It is not unexpected that the energy eigenfunction ^(r) transforming as the mth row of Гр involves only basis functions transforming the same way. That is, af,n = О for q p and n m in the expansion in Equation (6.6). The situation when Г is reducible is more complicated. Suppose that Г is the direct sum of two inequivalent irreducible representations Гр and Г9 of di- mensions dp and dq, so that d = dp + dq. Then dp of the energy eigenfunctions can be taken as basis functions of Гр, the remaining dq eigenfunctions being basis functions of Г9. The identities in Equations (6.12) produce a dp-fold de- generacy in the energy eigenvalue. Similarly, the corresponding identities with p replaced by q give rise to a dg-fold degeneracy. The overall d(= dp + dg)-fold degeneracy must be a consequence of the secular equations detD(p, m) = 0 and detD(g, n) = 0 possessing a common eigenvalue, but no reason for this can be attributed to the symmetry of the system. Consequently the extra degeneracy associated with this common eigenvalue is called an “accidental”
QUANTUM MECHANICAL CALCULATIONS 97 degeneracy, and the energy levels corresponding to Гр and Г9 are said to “stick together”. In general, an arbitrarily small change in the potential that preserves its symmetry will break the accidental degeneracy. It is to be expected that accidental degeneracies occur only very rarely, the normal situation being that in which the representation Г is irreducible. However, in some exceptional systems, such as the hydrogen atom, these ac- cidental degeneracies occur so extensively and in such a regular fashion that they cannot be truly coincidental. Their origin lies in a “hidden” symmetry which gives rise to an invariance group that is larger than the obvious invari- ance group. For the hydrogen atom the situation is described in detail in, for example, Chapter 12, Section 8, of Cornwell (1984). 2 Transition probabilities and selection rules Suppose that a small time-dependent perturbation H'(t) is applied to a system whose time-independent “unperturbed” Hamiltonian is Hq, so that the total Hamiltonian becomes я(*) = я0 + я'(г). Suppose that before the perturbation is applied (that is, at time t = —сю) the system is in an eigenstate фi of Hq with energy eigenvalue бг. Then, according to first-order perturbation theory (Schiff 1968), the probability of finding the system at time t in another eigenstate ф/ of Яо (whose energy eigenvalue is бу) is given by Zoo (Ф/, ехр{г(е/ - ei)t'/h} dt'\2. -oo (Here it is assumed that фi and ф/ have been appropriately normalized.) The significant part of this expression is the inner product (</>/, Я'^')^)- The analysis of Chapter 1, Section 4 shows that фi and фf must be basis func- tions of some representations of the invariance group Po of the unperturbed Hamiltonian Яд. With the perturbation Hf(t') expressed in terms of irre- ducible tensor operators of Pg, such inner products can be studied using the Wigner-Eckart Theorem of Chapter 5, Section 3. In particular, it is possible to deduce when symmetry requires that (</>y, Я'(^')^) = 0- When this is so, transitions from фi to ф/ are forbidden (at least in first-order perturbation theory). This basic idea will now be developed for the very important case in which the system interacts with an external electromagnetic field, resulting in ab- sorption or induced emission of radiation. It will be assumed that the unper- turbed system is described by a “single-particle” Schrodinger equation (in the sense of Chapter 1, Section 3(a)), so that (cf. Equation (1.10)) h2 d2 d2 d2 Яд(г) = ~(-^“^ + + ^(rh 2p dx2 oy2 oz2
98 GROUP THEORY IN PHYSICS where /z is the mass of the particle and V (r) the potential that it experiences. The perturbing operator H'(r,t) may be taken to be Я'(г,£) = (ze/i//zc)A(r,t).grad, (6.13) where c is the speed of light and the vector potential A(r,t) for a plane electromagnetic wave with wave vector к and frequency wfTn is of the form A(r, t) = 2Ao cos(k.r — a>t + a). Here Aq is a constant vector, with real components, that specifies the polar- ization of the radiation, a gives its phase, Ao.к = 0, and ca = |k|c. Then the transition probability for absorption of a photon from the radiation field causing a transition from ф{(т) to </>/(r) is given by 4iv2e2h2I((ef — еЛ/К). ...,2 /-i ---- 2 2z _ x2 1Ш A0.grad<fc)| , (6.14) the angular frequency of the absorbed radiation being (б/ — ei)/h. Similarly, the transition probability for induced emission of a photon associated with a transition from to </>/(r) is given by (6.15) the angular frequency of the emitted radiation being (б* — 6y)//z. In Expres- sions (6.14) and (6.15) I(cT) is defined to be such that the intensity of the radiation field in the angular frequency range w to w + dw is I(w)dw. In de- riving Expressions (6.14) and (6.15) it has been assumed that the wavelength of the radiation is much greater than the dimensions of the regions in which the eigenfunctions фгфс) and </>/(r) are significantly different from zero. As the inner products of Expressions (6.14) and (6.15) can be rewritten in the form (0/, A0.grad</>i) = —{/z(ey - ei)/h2}^f, А0.г</>Д (6.16) the transitions are often called “electric dipole transitions”. (The detailed derivations of Expressions (6.13), (6.14), (6.15) and (6.16) may be found in the book by Schiff (1968).) It is interesting to note (see Schiff (1968)) that the transition probability for spontaneous emission of radiation of frequency (б* — Cf)/h polarized in the direction of the unit vector n in a transition from ф^ (r) to ф/ (r) is given by | (</>/, n.grad</>j)|2. ОС pL Equation (6.16) allows this too to be expressed in terms of “dipole” inner products. Let Q = Ад.grad for absorption or induced emission and Q = n.grad for spontaneous emission. Then Example III of Chapter 5, Section 3 shows that
QUANTUM MECHANICAL CALCULATIONS 99 Q is easily expressed as a linear combination of irreducible tensor operators. There is no intrinsic difficulty in carrying out the analysis for the most general case, but for simplicity it will be assumed that Q is an irreducible tensor operator transforming as the fcth row of the unitary irreducible representation Гд of Qq. Suppose also that ф{(г) = <^(r) and </>/(r) = </>[(r), where </>j(r) and </>[(r) are respectively basis functions transforming as the jth and Zth rows of the unitary irreducible representations Гд and Гд of Po • If Go is a finite group or a compact Lie group, the Wigner-Eckart Theorem shows that = о if npq — 0, that is, if Гд does not appear in the reduction of Гд 0 Гд. Thus a list of forbidden transitions can be found from the Clebsch-Gordan series alone. However, a complete list requires a knowledge of the Clebsch-Gordan coefficients, for if npq 0 the Wigner-Eckart Theorem implies that 7 r, a I (6.17) and it can happen that P Q r, a \ _ J к I J for all a = 1,2, ...,n£g, thereby giving again a zero transition probability. (This situation occurs in the example that will be given shortly.) The Wigner-Eckart Theorem can also be employed in the analysis of the magnitudes of the transition probabilities of the allowed transitions. If npq 0, Equation (6.17) shows that (</>/, Q</>i) depends on nrpq reduced ma- trix elements. The transition probabilities for all other transitions involving initial states that are partners of ф^т) in the same basis, final states that are partners of ф/(т) in the same basis, and other directions of polarization whose operators form part of the same set of irreducible tensor operators as Q also depend on the Clebsch-Gordan coefficients and these nrpq reduced matrix elements. Thus a complete description of dpdqdr possible transitions depends only on npq reduced matrix elements, where npq can be considerably smaller than dpdqdr. The most important case is that in which Qq is the group of all rotations in IR3. This will be considered in detail in Chapter 10, Section 6, after the irreducible representations of this group have been derived. The following simple example will serve until then to illustrate the power of the technique. Example I Optical selection rules associated with the crystallographic point group Let Po = ^4 and Го denote the irreducible representations of previously referred to as Гр, p = 1,2,3,4,5. Consider absorption from an initial state
100 GROUP THEORY IN PHYSICS that transforms as the first row of the two-dimensional irreducible rep- resentation Гд(= Г5) (the explicit matrices being as in Example I of Chapter 5, Section 1). For polarization vector Ao in the ж-direction, Ag.grad = Aq^B/Bx^ which (by Example III of Chapter 5, Section 3) is an irreducible tensor operator transforming as the first row of Гд. As Гд 0 Гд = Гд ф Гд ф Гд ф Гд (see Example I of Chapter 5, Section 2), the Clebsch-Gordan series implies that the final state </>/(r) cannot transform as either row of Гд. However, the Clebsch-Gordan coefficients (see Example I of Chapter 5, Section 3) show that </>/(r) can only transform as Гд or Гд. Similarly, for Ад in the ^/-direction, Ад.grad = AQzB/By, which is an irreducible tensor operator transforming as Гд. As Гд 0 Гд Гд transitions can only occur to final states </>/(r) transforming as Гд. Examination of the Clebsch-Gordan coefficients shows that </>/(r) must transform as the second row of Гд . Finally, for Ад in the ^-direction, Ад.grad = AqqB/Bz^ an irreducible tensor operator transforming as the second row of Гд. Again the Clebsch- Gordan series indicates that </>/(r) cannot transform as either row of Гд, while the Clebsch-Gordan coefficients show that </>/(r) can only transform as Гд or Г4 1 o- 3 Time-independent perturbation theory Suppose that the Hamiltonian H of a system is time-independent and is made up of two time-independent parts Hq and H', where Hq is sufficiently simple that its eigenfunctions and eigenvalues are known, and Н' is sufficiently small that its effect can be considered to be a perturbation on Яд. The problem is then to find the eigenfunctions and eigenvalues of H = Hq + Hf in terms of those of Hq. It is assumed that the eigenfunctions of H and Hq can be put into a one-to-one correspondence, in the sense that they are the limits as A —> 0 and A —> 1 respectively of the eigenfunctions of the operator Hq + XH\ 0 < A < 1. If </>i, </>2, • • •, </>d are a set of eigenfunctions of Hq corresponding to the d-fold degenerate energy eigenvalue бд and the associated eigenfunctions 1 фч , • • •, Фа of H correspond to eigenvalues that are not equal to each other, then the perturbation Н' may be said to “split” 6q. It will be shown that a considerable amount of information about such splittings can be found very simply using group representation theory. For simplicity it will be assumed that H and Hq give “single- particle” Schrodinger equations (in the sense of Chapter 1, Section 3(a)). Let Qq and Q be the invariance groups of Hq and H respectively. Usually there exists a coordinate transformation T of Hq such that #'({R(T)|t(T)}r) / Я'(г), so that Equation (1.13) implies that Q is a proper subgroup of Qq. Suppose that the unperturbed energy eigenvalue 6q corresponds to a representation Гд
QUANTUM MECHANICAL CALCULATIONS 101 of Po- Then Го provides a representation of the subgroup Q. However, even if Го is an irreducible representation of Po, as a representation of Q Го may- be reducible. For convenience it will be assumed that Q is a finite group or a compact Lie group, so that Го is then completely reducible on Q (see Chapter 4, Section 4). Thus suppose that as a representation of Q r0«J2®nrrr, (6.18) where the sum is over all inequivalent unitary irreducible representations Гг of P, nr being the number of times that Гг appears in the reduction of Tq. Here nr may be zero for some Гг. These integers are easily evaluated from a knowledge of the characters of Qo and Q alone, for Theorem V of Chapter 4, Section 6 gives nr = (i/<?) £ Mir? тед for the case in which Q is finite and of order g. Here Xo(T) and xp(T) denote the characters of Го and Гг respectively. Similarly, if Q is a compact Lie group, nr= [ X^T)xr(TYdT. Jg Suppose that = n, so that Equation (6.18) contains n irreducible representations of Q. Then the d perturbed energy eigenfunctions ^2, .. .^d in general belong to n different eigenvalues of H. Thus c0 is split by the perturbation Hf into n different values. If dr is the dimension of Гг, then the perturbed energy eigenvalue corresponding to Гг is dr-fold degenerate. Naturally ^2rdr = d. As dr = хг(Е), both the number of perturbed energy eigenvalues and their degeneracies can be predicted using the characters of Qq and P alone. In the special case in which Го provides an irreducible representation of P, the unperturbed eigenvalue co is not split. (In particular, this happens when Q coincides with p0 and Го is an irreducible representation of Po-) As all expressions for the perturbed eigenvalues and eigenfunctions involve only the unperturbed energy eigenvalues and matrix elements of Hf between unperturbed energy eigenvalues (Schiff 1968), the Wigner-Eckart Theorem of Chapter 5, Section 3, can be brought into use again. For example, sup- pose that Го is equal to Гд, a unitary irreducible representation of Po, and «/^(r), </>£(r),..., <^d(r) are a se^ °f linearly independent eigenfunctions of Hq, with eigenvalue 6q, that are basis functions for Гд. Suppose also that H' is an irreducible tensor operator transforming as the fcth row of a unitary irreducible representation Гд of Po- Then, if d = 1 (that is, if 6q is “non- degenerate” ), the corresponding perturbed eigenvalue e is given to first order by б = б0 + (^,Я'^),
102 GROUP THEORY IN PHYSICS while if d > 1 the corresponding first-order perturbed eigenvalues are the eigenvalues of the d x d matrix A whose elements are given by Ay =e0^+ (<#,#'$), (6.19) j, I = 1, 2,..., d. In both cases there is a first-order effect only if Гд 0 Гд contains Гд (that is, when ri^q 0). When this is so the Wigner-Eckart Theorem shows that n₽<? / \ * = £ p q p’ a (р\н'\р) q=1 к J / (6.20) so that the matrix elements depend on the Clebsch-Gordan coefficients and n^q reduced matrix elements. For a further analysis of the case d > 1, see, for example, Chapter 6, Section 3, of Cornwell (1984).
Chapter 7 Crystallographic Space Groups 1 The Bravais lattices An infinite three-dimensional lattice may be defined in terms of three linearly independent real “basic lattice vectors” ai, a2 and аз. The set of all lattice vectors of the lattice is then given by tn = niai + n2a2 + n3a3, where n = (ni,n2,n3), and ni, n2 and n3 are integers that take all possible values, positive, negative and zero. Points in IR3 having lattice vectors as their position vectors are called “lattice points” and a pure translation through a lattice vector tn, {1 |tn}, is called a “primitive” translation. Suppose that in a crystalline solid there are S nuclei per lattice point, and that the equilibrium positions of the nuclei associated with the lattice point r = 0 have position vectors ri, t2, ... , 75. Then the equilibrium positions of the whole set of nuclei are given by ГП7=^п+Т7? (7-1) where 7 = 1, 2,..., S and tn is any lattice vector. In the special case when S = 1, n may be taken to be 0 and the index 7 may be omitted, so that ГП = ^п- The set of all primitive translations of a lattice form a group which will be denoted by T°°. T°° is Abelian but of infinite order. In Section 2 the Born cyclic boundary conditions will be introduced. They have the effect of replacing this infinite group by a similar group of large but finite order, so that all the theorems on finite groups of the previous chapters apply. The “maximal point group” Q^iax of a crystal lattice may be defined as the set of all pure rotations {R(T)|O} such that, for every lattice vector tn,
104 GROUP THEORY IN PHYSICS Figure 7.1: Basic lattice vectors of the simple cubic lattice, Гс. the quantity R(T)tn is also a lattice vector. Clearly R(T) e Qqicix if and only ifR(T)a,- is a lattice vector for J = 1, 2,3. There are essentially 14 different types of crystal lattice. They are known as the “Bravais lattices”. These will be described briefly, but no attempt will be made to give a logical derivation or to show that there are no others. (In this context two types of lattice are regarded as being different if they have different maximal point groups, even though one type is a special case of the other. For example, as may be seen from Table 7.1, the simple cubic lattice Гс is a special case of the simple tetragonal lattice Гд with a = 6, but Qqicix = Oh for Гс, whereas g^iax = D^h for Гд.) Lattices with the same maximal point group are said to belong to the same “symmetry system”, there being only seven different symmetry systems. Complete details are given in Table 7.1, in which the notation for point groups is that of Schonfliess (1923). (A full specification of these and the other crystallographic point groups may be found in Appendix C.) The cubic system is probably the most significant, the body-centred and face-centred lattices occurring for a large number of important solids. The basic lattice vectors of the cubic lattices are shown in Figures 7.1, 7.2 and 7.3. The lattice points of the simple cubic lattice Гс merely form a repeated cubic array, and the basic lattice vectors lie along three edges of a cube. For the body-centred cubic lattice the basic lattice vectors join a point at the centre of a cube to three of the vertices of the cube, so that the lattice points form a repeated cubic array with lattice points also occurring at every cube centre. For the face-centred lattice the lattice points again form a repeated cubic array with additional points also occurring at the midpoints of every cube face, the basic lattice vectors then joining a cube vertex to the midpoints of the three adjacent cube faces.
CRYSTALLOGRAPHIC SPACE GROUPS 105 Figure 7.2: Basic lattice vectors of the body-centred cubic lattice, Г£. A symmetry system a may be regarded as being “subordinate” to a sym- metry system (3 if Qq1(1x for a is a subgroup of д$ъах for (3 and at least one lattice of (3 is a special case of a lattice of a. The complete subordination scheme is then: triclinic < monoclinic < orthorhombic < tetragonal < cubic; monoclinic < rhombohedral; orthorhombic < hexagonal; (Here a < /3 indicates that a is subordinate to /3.) For a perfect crystalline solid the group of the Schrodinger equation is a crystallographic space group, which contains rotations as well as pure prim- itive translations. The crystallographic space groups will be investigated in detail in Section 6. However, it is very enlightening, as a first stage in their study, to limit attention to the subgroup T of pure primitive translations of the relevant lattice. Only the translational symmetry is then being taken into account. In particular, the energy eigenfunctions must transform according to the irreducible representations of this subgroup, which is equivalent to saying that they satisfy Bloch’s Theorem, as will be demonstrated in Section 3. Bloch’s Theorem has now become so much an essential part of the theory of solids that it is sometimes forgotten that it is basically a group theoretical result. The elementary energy band theory based upon Bloch’s Theorem itself requires no knowledge of group theory and so is presented in most textbooks on solid state theory. However, the neglect of rotational symmetries in this elementary theory does mean that some phenomena are overlooked, and, in particular, it cannot predict the extra degeneracies which can occur in electronic energy levels. Moreover, it is only by taking into account the rotational symmetries that it is possible to reduce the numerical work in energy band calculations to a manageable amount and still produce accurate results.
106 GROUP THEORY IN PHYSICS (1) Triclinic symmetry system ((7™аж = Ci): (i) simple triclinic lattice, I\: ai,a2 and аз arbitrary. (2) Monoclinic symmetry system (ffTax = C^h): (i) simple monoclinic lattice, Tm: аз perpendicular to both ai and аг; (ii) base-centred monoclinic lattice, Г^: ai = (a, 6,0), a2 = (a, —6,0), a3 = (c, 0, d). (3) Orthorhombic symmetry system (G™ax = D^h): (i) simple orthorhombic lattice, Го: ai = (a, 0,0), a2 = (0,6,0), a3 = (0,0, c); (ii) base-centred orthorhombic lattice, Г^: ai = (a, 6,0), a2 = (a,—6,0), аз = (0,0, c); (iii) body-centred orthorhombic lattice, Г^: ai = (a, 6, c), a2 = (a, 6, —с), аз = (a, —6, —c); (iv) face-centred orthorhombic lattice, r£: ai = (a, 6,0), a2 = (0, b, с), аз = (a, 0, c). (4) Tetragonal symmetry system (Q™ax = D^h): (i) simple tetragonal lattice, Tq: ai = (a, 0,0), a2 = (0, a, 0), a3 = (0,0,6); (ii) body-centred tetragonal lattice, Г^: ai = (a, a, 6), a2 = (a, a, —6), аз = (a, —a, 6). (5) Cubic symmetry system ((7™аж = Oh): (i) simple cubic lattice, Гс: ai = (a, 0,0), a2 = (0, a, 0), a3 = (0,0, a); (ii) body-centred cubic lattice, Г^: ai = |a(l,l,l), a2 = ja(l,l,-l), a3 = |a(l,-l,-l); (iii) face-centred cubic lattice, r£: ai = |a(l, 1,0), a2 = |a(0,1,1), a3 = |a(l,0,1). (6) Rhombohedral (or trigonal) symmetry system (Q™ax = B3cz): (i) simple rhombohedral lattice, Ггн: ai = (a, 0,6), a2 = (|а\/3, —6), аз = (—|a\/3, —6). (7) Hexagonal symmetry system (Q™ax = Dqh): (i) simple hexagonal lattice, I\: ai = (0,0, c), a2 = (a, 0,0), аз = (—— |a\/3,0). Table 7.1: The Bravais lattices. (The real parameters a^b^c and d are arbi- trary.)
CRYSTALLOGRAPHIC SPACE GROUPS 107 Figure 7.3: Basic lattice vectors of the face-centred cubic lattice, Г£. A proof of Bloch’s Theorem that involves only an elementary application of the ideas of the previous chapters is given is Section 3. Sections 4 and 5 are then devoted to a brief account of the elementary electronic energy band theory that is based on this theorem. Section 8 then describes, for the case of symmorphic space groups, how this theory is modified when the full space group is introduced in place of its translational subgroup T. It will be seen there that the concepts introduced in Sections 4 and 5 still play a fundamental role. 2 The cyclic boundary conditions Strictly speaking, a real crystalline solid cannot possess any translational sym- metry because it is necessarily finite in extent. Consequently any translation will shift some electron or nucleus from just inside some surface to the outside of the body, that is, to a completely different environment. On the other hand, for a normal sample the inter-nuclear spacing is so much smaller than the dimensions of the sample and the interactions that di- rectly affect each electron or nucleus are of such short range, that for electrons and nuclei well inside the body the situation is almost exactly as if the solid were infinite in extent. Moreover, the evidence of X-ray crystallography is that the nuclei within a solid can be ordered as if they were based on an infinite lattice, except near the surfaces. As most of the properties of a solid depend only on the behaviour of the vast majority of electrons or nuclei that lie in the interior, it is a very reasonable approximation to idealize the situation by working with models based on infinite lattices. The translational symmetry possessed by such models then permits a considerable simplification of the analysis.
108 GROUP THEORY IN PHYSICS However, the symmetry groups based on infinite lattices are necessarily of infinite order, and it is easier to work with groups of finite order. This can be achieved by imposing “cyclic” boundary conditions on the infinite lattice. For the electrons it may be assumed that for every energy eigenfunction </>(r) ф(г) = ф(г + Mai) = ф(г + 7V2a2) = ф(г + 7V3a3), (7.2) where M, A2 and A3 are very large positive integers and ai, a2 and аз are the basic lattice vectors of the lattice. This implies that the infinite crystal is considered to consist of a set of basic blocks in the form of parallelepipeds having edges Mai, Ma2 and Маз, and that the physical situation is identical in corresponding points of different blocks. These boundary conditions cannot affect the behaviour of electrons well inside each basic block to any significant extent, so the bulk properties are again unchanged. The integers M, A2 and A3 may be taken to be as large as desired. The integration involved in the inner product (</>, defined in Equation (1.19) must now be taken as being over just one basic block of the crystal, B. For any pure primitive translation T the operators P(T) retain the unitary property of Equation (1.20), provided all functions involved satisfy Equation (7.2). This follows because (P(T)^, P(T)^) is equal to Ф(? — t(T))*?/j(r — t(T)) dxdydz = j j j ф(т}'к'ф(т') dxdydz, where B' is obtained from В by a translation — t(T). As every part of Bf can be mapped into a part of В by an appropriate combination of translations through Mai, M^2 and Маз, by Equations (7.2) the last integral becomes The conditions in Equations (7.2) are often referred to as the “Born cyclic boundary conditions”, as they are the analogues for electronic states of the vi- brational boundary conditions first proposed by Born and von Karman (1912). They imply that P({l|MaJ) = P({l|0}) (7.3) for every function of interest and for j = 1, 2,3. (Of course P({l|0}) is merely the identity operator.) Consequently P({l|tn + ZiMai + /гМаг + /зМаз}) = P({l|tn}) for any lattice vector tn and any set of integers Zi, Z2 and Z3, so that only N = N1N2N^ of these operators are distinct. The set of distinct operators may be taken to be P({l|niai + + п3аз}) with 0 < nj < M, j = 1,2,3. (7.4) Moreover, as P({l|Maj}) = Р({1|а7})^, it follows from Equation (7.3) that P({l|aJ)^ = P({l|0}), j = 1,2,3. (7.5)
CRYSTALLOGRAPHIC SPACE GROUPS 109 Thus this set of distinct operators forms a finite group T of order N = Henceforth this group T will be used in place of the infinite group of pure primitive translations T°°. Incidentally, as it remains true that F({l|tn}) F({l|tn4) = F({l|tn}{l|tn4) for any two lattice vectors tn and tn' of a lattice, the mapping </>({1 |tn}) = P({l|tn}) is a homomorphic mapping of T°° onto T. The kernel /С of this mapping is the infinite set of pure primitive translations of the form {1 |Zj.7Viai + I2N2&2 + /з^Узаз}, where Zi, Z2 and Z3 are any set of integers. 3 Irreducible representations of the group T of pure primitive translations and Bloch’s Theorem As the group T is a finite Abelian group of order N = 7V1-/V2AZ3, it possesses N inequivalent irreducible representations, all of which are one-dimensional (see Chapter 5, Section 6). These are easily found, for T is isomorphic to the direct product of three cyclic groups. Consider a particular one-dimensional irreducible representation Г of T and suppose that Г({1|ау}) = [су], for j = 1,2,3. Then, from Equation (7.5), it follows that = 1, (7.6) so that Cj = exp(—2i{ipj/Nfifi j = 1,2,3, where pj is an integer. As exp(—2ivi(pj + Nfi)/Nfi) = exp(—2mpj/Njfi there are only Nj different values of Cj allowed by Equation (7.6) and each of these, by convention, may be taken to correspond to a pj having one of the values 0,1,2,..., Ny - 1. Then r({1l^aj}) = [exp(-27rij>j^/^)] and hence r({l|tn) = [ехр(-2тгг{(р1П1/М) + (ргпг/Л^) + (рз^з/^з)})], (7.7) where tn = mai + П2а2 + пзаз. There are N — N1N2N3 sets of integers (Р1,Р2,Рз) allowed by the above convention which can be used to label the N different irreducible representations of T. Equation (7.7) can be simplified and given a simple geometric interpreta- tion by introducing the following notation. Define the “basic lattice vectors of the reciprocal lattice” bi, b2 and Ьз by a^.bfc = 2Tr6jk, j, к = 1, 2,3, (7.8)
по GROUP THEORY IN PHYSICS so that, explicitly, bi = 2тга2 A a3/{ai.(a2 A a3)}, (7.9) with similar expressions for b2 and Ьз. Then define the so-called “allowed к-vectors” by к = Aqbi + ^2^2 + &зЬз, (7.10) where kj —pj/Nj. Thus k.tn = 2m{(p1n1/N1) + (p2«2/№) + (рз'«з/№)}, so that Equation (7.7) becomes rk({l|tn}) = [exp(-ik.tn)], (7.11) where the N irreducible representations are now labelled by the allowed k- vectors. Suppose that </>k(r) is a basis function transforming as the first (and only) row of Гк. Then, by Equations (1.26) and (7.11), P({l|tn})</>k(r) = rk({l|tn})</>k(r) = exp(-zk.tn)^(r). (7.12) However, by Equation (1.17), P({l|tnM(r) = ^({1|tn}-ir) = ^k(r _ tn)) so that </>k(r - tn) = exp(-ik.tn)</>k(r). Thus </>k(r) = exp(ik.r)z<k(r), (7-13) where ttk(r) is a function that has the periodicity of the lattice, that is, ^(r — tn) = ^k(r) for any lattice vector tn. Equation (7.13) is the statement of the theorem of Bloch (1928) in its usual form, for electronic energy eigenfunctions must be basis functions of the irreducible representations Гк of T. A function of the form in Equation (7.13) is therefore called a “Bloch function”. The corresponding energy eigenvalue may be denoted by e(k), so that Я(г)^к(г) = б(к)^(г). (7.14) The notation for basis functions here follows the standard practice in which the irreducible representation is specified by a superscript (or set of super- scripts) and the rows by a subscript (or set of subscripts). In particular, the wave vector к appears as a superscript with this convention. However, it should be pointed out that in most of the solid state literature к is written as a subscript, so that (r) would be written as </>k(r) and Equation (8.20) would become Я(г)</>к(г) = б(к)</>к(г).
CRYSTALLOGRAPHIC SPACE GROUPS 111 Figure 7.4: The basic parallelepiped of k-space. 4 Brillouin zones The set of lattice vectors of the reciprocal lattice is defined by Km = mibi + m2b2 + m3b3, (7-15) where m = mi, m2 and m3 are integers, and bi, b2 and Ьз are the basic lattice vectors of the reciprocal lattice defined by Equation (7.8). They have the property that exp(iKm.tn) = 1 (7-16) for any Km and tn. It is useful to note that ^2exp(ik.tn) = N, 0, ifk = Km, ifk^Km, (7-17) where the sum is over all the lattice vectors of one basic block of Section 2, this result being a consequence of the fact that the left-hand side is a product of three simple geometric series. Similarly, / 1 + \ _ f -N, if tn = 0, XexP( *k.tn)- I Oj iftn^o, к the sum being over all allowed k-vectors. In Section 3, N irreducible representations of T were found and described by the allowed к-vectors (Equation (7.10)). These к-vectors can be imagined as being plotted in the so-called “к-space” or “reciprocal space” defined by the reciprocal lattice vectors. The allowed k-vectors lie on a very fine lattice (defined by Equation (7.10)) within and upon three faces of the parallelepiped having edges bi, b2 and Ьз that is shown in Figure 7.4.
112 GROUP THEORY IN PHYSICS Figure 7.5: Construction of a Brillouin zone boundary. It is, however, more convenient to replot the allowed k-vectors into a more symmetrical region of к-space surrounding the point к = 0. To do this consider the equation , k' = k + Km, (7.18) where Km is a reciprocal lattice vector. Two vectors к and k' satisfying Equa- tion (7.18) are said to be “equivalent”, because exp(—zk'.tn) = exp(—zk.tn) by Equation (7.16), and hence rk({l|tn}) = rk({l|tn}) for every {l|tn} of T. Thus the irreducible representation described by к could equally be described by k'. The more symmetrical region of к-space is called the “Brillouin zone” (or sometimes the “first Brillouin zone”), and it is defined to consist of all those points of к-space that lie closer to к = 0 than to any other reciprocal lattice points. Its boundaries are therefore the planes that are the perpendicular bisectors of the lines joining the point к = 0 to the nearer reciprocal lattice points, the plane bisecting the line from к = 0 to к = Km having the equation k.Km = ||Km|2, as is clear from Figure 7.5. For some lattices, such as the body-centred cubic lattice Г£, only nearest neighbour reciprocal lattice points are involved in the construction of the Brillouin zone, but for others, such as the face-centred cubic lattice r£, next-nearest neighbours are involved as well. The irreducible representations of T then correspond to a very fine lattice of points inside the Brillouin zone and on one half of its surface. The mapping of the parallelepiped of Figure 7.4 into the Brillouin zone can be quite complicated because different regions of the parallelepiped are mapped using different reciprocal lattice vectors. The following two- dimensional example shown in Figure 7.6 of a square lattice demonstrates
CRYSTALLOGRAPHIC SPACE GROUPS 113 2' bi О L4' 3' Figure 7.6: Construction of a two-dimensional Brillouin zone. this clearly. In this example the analogue of the three-dimensional paral- lelepiped of Figure 7.5 is the square with sides bi and b2, which consists of four regions 1, 2, 3 and 4, and the analogue of the Brillouin zone is the square having к = 0 at its centre, which consists of the four regions 1', 2', 3' and 4'. The region 1 is mapped into 1' by K(0?0 0) = 0, 2 is mapped into 2' by K(-i,до) = —bi, 3 is mapped into 3' by K(O,-i,o) = — b2, and 4 is mapped into 4' by K(_i _lj0) = -bi - b2. By construction, the volume of the Brillouin zone is the same as that of the parallelepiped from which it is formed, namely bi.(b2 Л Ьз). It follows from Equation (7.9) that this is equal to (27r)3/{ai.(a2 Л аз)}, where ai.(a2 Л аз) is the volume of the parallelepiped whose sides are ai, a2 and аз. For the simple cubic lattice Гс, the basic lattice vectors of the reciprocal lattice obtained from Table 7.1 and Equation (7.9) are bi = (2тг/а)(1,0,0), b2 = (2тг/а)(0,1,0), b3 = (2тг/а)(0,0,1). The Brillouin zone is given in Figure 7.7. The position vectors of the “sym- metry points” are as follows: for Г, к = (0, 0, 0); for X, к = (тг/а)(0,0,1); for M, к = (тг/а)(0,1,1); and for R, к = (тг/а)(1,1,1). The significance of the term “symmetry point” will be explained in Section 7. The notation is that of Bouckaert et al. (1936). Similarly, for the body-centred cubic lattice the basic lattice vectors of the reciprocal lattice are bi = (27r/n)(l,0,l), b2 = (2тг/п)(0,1,-1), b3 = (2тг/п)(1,-1,0), the Brillouin zone being shown in Figure 7.8. The position vectors of the symmetry points are as follows: for Г, к = (0,0,0); for Я, к = (тг/а)(0,0,2);
114 GROUP THEORY IN PHYSICS Figure 7.7: Brillouin zone corresponding to the simple cubic lattice Гс. Figure 7.8: Brillouin zone corresponding to the body-centred cubic lattice Г".
CRYSTALLOGRAPHIC SPACE GROUPS 115 Figure 7.9: Brillouin zone corresponding to the face-centred cubic lattice Г{. for JV, к = (тг/а)(0,1,1); and for P, к = (тг/а)(1,1,1), the notation being that of Bouckaert et al. (1936). Finally, for the face-centred cubic lattice the basic lattice vectors of the reciprocal lattice are bi = (2тг/а)(1,1, -1), b2 = (2тг/а)(—1,1,1), b3 = (2тг/а)(1, -1,1). The Brillouin zone is given in Figure 7.9, the position vectors of the points indicated (in the notation of Bouckaert et al. (1936)) being: for Г, к = (0,0,0); for К, к = (тг/а)(0, j, j); for L, к = (тг/а)(1,1,1); for P, к = (тг/а)(|, |, 2); for W, к = (тг/а)(0,1,2); and for X, к = (тг/а)(0,0, 2). The Brillouin zones corresponding to the other eleven Bravais lattices may be found in the review article by Koster (1957). 5 Electronic energy bands The set of energy eigenvalues corresponding to an allowed к-vector may be denoted by 6i(k), 62(k),..., with the convention that en(k) < £n+i(k) (7-19)
116 GROUP THEORY IN PHYSICS Figure 7.10: Part of the electronic energy band structure of iron. for all n = 1,2,.... The set of energy eigenvalues en(k) corresponding to a particular n are said to form the “nth energy band” and the set of energy bands is said to constitute the electronic energy band “structure”. To visualize the energy band structure, it is convenient to consider, one at a time, the axes of the Brillouin zone that join the symmetry points, and for each allowed k-vector on each axis to plot the energy eigenvalues 6n(k).' Two typical examples are shown in Figures 7.10 and 7.11, which give the energy levels along the axis Д for iron (as calculated by Wood (1962)) and silicon (as calculated by Chelikowsky and Cohen (1976)), the lattices being the body-centred cubic and face-centred cubic respectively. The number in curved brackets gives the band index n, as defined in Expression (7.19) and the number in square brackets indicates the degeneracy of the corresponding eigenvalue. (The occurrence of degenerate eigenvalues is a consequence of the rotational symmetry, which is being neglected in this section but which will be investigated in Sections 7 and 8, where the other symbols will also be explained.) The positive integers №, Nz and Л3 introduced in Equations (7.2) are arbitrarily large, and it is frequently convenient to consider the limiting case in which they tend to infinity. The allowed k-vectors can then take all values inside the Brillouin zone and on half of its surface, and the cn(k) are con- tinuous functions of к for each n. Moreover, gradk6n(k) are also continuous functions of k, except possibly at points where two bands touch. The plots in Figures 7.10 and 7.11 are made for this limiting case.
CRYSTALLOGRAPHIC SPACE GROUPS 117 Figure 7.11: Part of the electronic energy band structure of silicon. In the “single-particle” approximation (see Chapter 1, Section 3(a)), the Pauli exclusion principle implies that no two electrons can “occupy” the same one-electron state. With the present neglect of spin-dependent terms, such a state is specified by an allowed k-vector, a band index n, and a spin quantum number that can take one of two possible values. It follows that each energy level cn(k) can “hold” two electrons and hence each energy band can hold 2N electrons. If there are V valence or conduction electrons per atom and S atoms per lattice point of the crystal, there will be NVS valence or conduction electrons in each large basic block of the crystal (in the sense of Section 2), which will therefore require the equivalent of bands to hold them. In the ground state of the system all the energy levels will be doubly occupied up to a certain energy бр, the “Fermi energy”, and all levels above this energy will be unoccupied. The surface in к-space defined by en(k) = £f is called the “Fermi surface”. The distribution of energy levels near the Fermi energy largely determines the electronic properties of a solid. If one and only one band contains the Fermi energy, and all others are entirely above it or below it, then the Fermi surface merely consists of one sheet. If no band
118 GROUP THEORY IN PHYSICS contains the Fermi energy, as happens for insulators and semiconductors, there is no Fermi surface. In all other cases the Fermi surface consists of several sheets, to visualize which one considers a number of identical Brillouin zones, with one zone for each band. A full band corresponds to a full Brillouin zone, but a partially occupied band corresponds to a partially occupied Brillouin zone and hence to a sheet of the Fermi surface in that zone. As an example, consider body-centred cubic iron for which 5=1 and V = 8, so that the equivalent of four complete bands are needed to hold the valence electrons. The Fermi energy 6p is shown in Figure 7.10 and is clearly consistent with this. Bands 3, 4, 5 and 6 are partially occupied, giving rise to four sheets in the Fermi surface. For silicon V = 4 and 5 = 2, so that again four bands (or their equivalent) are required to hold the valence electrons. However, in this case there is no overlap between the 4th and 5th bands, so that the Fermi energy lies between the Xi and Г25' levels of Figure 7.11. There are therefore four completely filled bands and no Fermi surface. 6 Survey of the crystallographic space groups Consider an infinite crystalline solid for which the equilibrium positions of the nuclei are given by Equation (7.1). The set of all coordinate transformations that map the set of equilibrium positions into itself forms an infinite group Q°° that is known as a “crystallographic space group”. Clearly Q°° contains as a subgroup the infinite group of pure primitive translations T°° of the relevant lattice, but contains no other pure translations. If {R|t} is a member of the space group Q°° and tn is any lattice vector of its lattice, then Rtn must also be a lattice vector. (This follows because (by Equations (1.7) and (1.8)) {R|t}{l|tn}{R|t}-1 = {l|Rtn} , so that {l|Rt„} must be a member of Q°° and, being a pure translation, must be a primi- tive translation. Thus the set of all rotational parts R of the space group operations {R|t} form a subgroup Q$ of the maximal point group Q^iax of its crystal lattice (though Po need not be a proper subgroup of д™ах). Qq is known as the “point group of the space group”. Detailed investigations show that the only possible proper rotations of Po are through multiples of 2тг/6 or 2тг/4, and the only possible improper rotations are products of these proper rotations with the spatial inversion op- erator. Consequently Qq is always a finite group, and must be one of the 32 crystallographic point groups that are specified in detail in Appendix C. (Of course these restrictions disappear if the requirement of translational sym- metry is abandoned. Consequently, for a “quasicrystal”, which has no such translational symmetry, it is possible to have other rotations as well, includ- ing, in particular, the proper rotations through 2тг/5.) Space groups having the same point group Q$ are said to belong to the same “crystal class”, so there are 32 different crystal classes. In the classification of space groups by Schonfliess (1923), each space group is denoted by the
CRYSTALLOGRAPHIC SPACE GROUPS 119 Schonfliess symbol for its point group Po, together with a superscript. Thus, for example, the space groups for crystals possessing the cubic Bravais lattices Гс, and Г{, and having only nuclei at the lattice points, all have Oh as point group, and are denoted by O^, Of and O^ respectively. Similarly, the space group of the diamond structure, which also has the Bravais lattice and point group Oh but which has two nuclei per lattice site, is denoted by O^. As will be seen, the assignment of superscripts by Schonfliess is rather arbitrary. An alternative that is more explicit but more complicated is the “international notation” (see Henry and Lonsdale 1965, Shubnikov and Koptsik 1974, Burns and Glazer 1978). To every rotation {R|0} of Qq there exists a vector tr which is such that {R|tr + tn} is a member of Q°° for every lattice vector tn of the lattice. Moreover, tr is unique (up to a lattice vector), as if {R|tr} and {R|t^} are both members of Q°° then so must be {R|TR}{R|rJEt}-1 = {1|tr — t^J, which, being a pure translation, is bound to be a primitive translation, so that = tr + tn for some tn. For definiteness it will be assumed henceforth that tr is always chosen so that tr = Qiai + Q2&2 + qs&3 with 0 < qj < 1, j = 1,2,3. If tr = 0 for every {R|0} E Qo, then Q°° is said to be a “symmorphic” space group. That is, for a symmorphic space group every transformation is of the form {R|tn}, the translational part always being a lattice vector. Obviously, if Q°° is symmorphic, then Q$ is a subgroup of P°°. Only 73 of the 230 crystallographic space groups in IR3 are symmorphic. Important examples include the cubic space groups O^ and O^- However, if Q°° is non-symmorphic, then for some {R|0} E Go there exists a non-zero tr. Thus, if Q°° is non-symmorphic, Go is not a subgroup of P°°. If a crystal has nuclei of only one species and their equilibrium positions lie only at the lattice points of a Bravais lattice, then the corresponding space group is symmorphic. The same is true if the crystal has more than one species of nuclei and if the arrays of nuclei of each species each form a Bravais lattice. (This is only possible if there is only one nucleus of each species per lattice site.) By contrast, non-symmorphic space groups are associated with crystals in which there are more than one nuclei of a given species per lattice site. Only the representation theory of symmorphic space groups will be con- sidered in Section 7. For a description of the corresponding theory for non- symmorphic space groups see, for example, Chapter 9, Section 3, of Cornwell (1984). A complete description of all 230 space groups may be found in the Inter- national Tables for X-ray Crystallography (Henry and Lonsdale 1965), which employ both the Schonfliess and international notations. (A simple pre- scription for determining the symmetry elements of a space group from the “general position” listed in the International Tables has been given by Won- dratschek and Neubuser (1967).) Another complete specification that is par- ticularly clear and thorough has been given by Shubnikov and Koptsik (1974). A further clear and comprehensive description of the crystallographic space
120 GROUP THEORY IN PHYSICS groups has been given by Burns and Glazer (1978). Lists of the space groups to which elements and compounds belong have been compiled by Wyckoff (1963,1964,1965) and by Donnay and Nowacki (1954). The effect of imposing Born cyclic boundary conditions is to replace the infinite space group Q°° by a finite group Q. As a preliminary, let M = A^2 = A3, where are as defined in Equations (7.2). (For every {R|0} e Qo this ensures that ЩЛ^-а?) = ZiMai + I2N2&2 + for some integers Zi, I2 and Z3 and each j = 1,2,3.) Thus, in the symmorphic case, Q may be defined to be the group of op- erators P({R|tn}) for all {R|0} e Po and all lattice vectors tn of the finite group T with Equation (7.5) applied. Then Q has order goN, where go and N are the orders of Q$ and T respectively. As it remains true that P({R|t})P({R'|t'}) = P({R|t}{R'|t'}) for every {R|t} and {R'|t'} of P, the mapping <!>({R|t}) = P({R|t}) is a homomorphic mapping of onto Q. The kernel /С of this mapping is again the infinite set of pure primitive translations of the form {l|ZiMai + /2^2^ + £зАзаз}, where Zi, I2 and Z3 are any set of integers. With the inner product defined as in Section 2, i.e. so that it involves an integral over just one basic block of the crystal, B, the operators P({R|t}) retain the unitary property of Equation (1.20), provided all the functions on which they act satisfy the Born cyclic boundary conditions (Equations (7.2)). The following sections will be devoted to the study of the representations of the finite symmorphic space groups Q and of their consequences. In this context no confusion will be caused if the Schonfliess or international notations are applied to the finite space groups as well as to the corresponding infinite groups. An identity that is worth noting is t.(Rk) = (R-1t).k, (7.20) which is valid for any rotation R and any vectors t and k. This follows because R is a 3 x 3 orthogonal matrix (see Chapter 1, Section 2((a)), so that 333 t.(Rk) = tiRijkj = (JR-1)jitikj = (R-1t).k. г=1 i,j=l i,j=l This implies that the reciprocal lattice (as defined in Section 4) has the same symmetry as the crystal lattice to which it belongs. This is shown by the following theorem. Theorem I If {R|0} of Q$ and Km is a lattice vector of the reciprocal lattice, then RKm is also a lattice vector of the reciprocal lattice. Proof Suppose that RKm is not a lattice vector of the reciprocal lattice. Then there exists a lattice vector tn of the crystal lattice that is such that exp{ztn.(RKm)} 7^ 1. Thus, by Equation (7.20), exp{z(R-1tn).Km} 1.
CRYSTALLOGRAPHIC SPACE GROUPS 121 However, {R 1|0} is a member of Po, so R 1tn must be a lattice vector of the crystal lattice. Equation (7.16) then provides a contradiction. 7 Irreducible representations of symmorphic space groups (a) Fundamental theorem on irreducible representations of symmorphic space groups The first stage in the analysis of symmorphic space groups is to observe that they possess a particularly straightforward structure. Theorem I If Q is a symmorphic space group, then Q is isomorphic to the semi-direct product T@£7o- Proof All that has to be verified is that the three requirements of the defi- nition of Chapter 2, Section 7, are satisfied. Firstly, T is clearly an invariant subgroup of Q. Moreover T and Q$ have only the identity in common. Fi- nally, if Q is symmorphic, every element of Q is the product of a pure primitive translation of T with a pure rotation of Qq. As T is Abelian, the theory of induced representations given in Chapter 5, Section 7, can be applied to produce all the irreducible representations of Q. Moreover, all the “little groups” from which the irreducible representations of Q are induced are subgroups of Po, and hence every “little group” is a crystallographic point group. As all the irreducible representations of the crystallographic point groups are known (and are listed in Appendix C), the irreducible representations of the space group Q follow immediately. In applying the results of the induced representation theory it is very helpful to re-cast some of the concepts in terms of the geometric picture of the Brillouin zone and its allowed k-vectors developed in Section 4. The quantities that will now be introduced will be identified in the proof of the fundamental theorem with certain of the entities of the induced representation theory. Definition Po(k)> point group of the allowed wave vector к The point group (7o(k) is the subgroup of the point group Qq of the space group Q that consists of all the rotations {R|0} of <y0 that rotate к into itself or an “equivalent” vector (in the sense of Equation (7.18)). That is, {R|0} of Po is a member of <7o(k) if there exists a lattice vector Km of the reciprocal lattice (as defined in Equation (7.15)), which may be zero, such that , Rk = k + Km. (7.21) Let g§ and #o(k) be the orders of and <70(k) respectively, and let M(k)
122 GROUP THEORY IN PHYSICS be defined by M(k) = g0/g0(k). (7.22) Then (see Chapter 2, Section 4) M(k) is always an integer. Definition General points, symmetry points, symmetry axes and symmetry planes of the Brillouin zone If Po(k) is the trivial group consisting only of the identity transformation {l|0}, then к is said to be a “general point” of the Brillouin zone. If к is such that Qq (k) is a larger group than those corresponding to all neighbouring points of the Brillouin zone, then к is known as a “symmetry point”. If all the points on a line or plane have the same non-trivial (7o(k), then this line or plane is said to be a “symmetry axis” or “symmetry plane”. It is easily shown that this list of definitions exhausts all the possible situations. Example I Symmetry points, axes and planes for the space group Of The Brillouin zone of the body-centred cubic lattice Г'" is shown in Figure 7.8. For the space group Of with this lattice the symmetry points are Г, H, N and P, and the symmetry axes are Д, Л, S, D, F and G, and the symmetry planes are the planes containing two symmetry axes. Example II Symmetry points, axes and planes for the space group Of The Brillouin zone of the face-centred cubic lattice is given in Figure 7.9. For the space group Of with this lattice the symmetry points are Г, L, W and X, the symmetry axes are Д, Л, S, Q, S and Z, and the symmetry planes are IKWX. TKL, TLUX and UWX. The point К is not a symmetry point because its <7o(k) is the same as that for the axis S that ends at K. For the same reason U is not a symmetry point. It should be noted also that KLUW is not a symmetry plane. Finally, the lines LK, KW, LU and UW have only the symmetry of the symmetry planes to which they belong, so they are not regarded as symmetry axes. Definition The “star” of к Let {RjO}, j = 1,2,... ,M(k), be a set of coset representatives for the de- composition of Po into left cosets with respect to (7o(k) (see Chapter 2, Section 4). Then the set of M(k) vectors k7 defined by k7 = R7k (j = 1,2,..., M(k)) is called the “star” of k. Of course any member of a left coset can be chosen to be the coset rep- resentative, but once a choice is made it should be adhered to. If {R'|0} E {RJO}^o(k), then there exists an element {R|0} of (7o(k) such that R' = RjR. Then R'k = Rj(Rk) = R^k + RjKm (by Equation (7.21)), so that R'k is equivalent to R7k, as R7Km is a reciprocal lattice vector. Thus a different choice of the coset representatives merely results in a set of vectors that are equivalent to those of the set ki, кг,.... Consequently the star of к
CRYSTALLOGRAPHIC SPACE GROUPS 123 is unique up to equivalence. It is always convenient to choose Ri = 1, so that ki = k. The fundamental theorem on the irreducible representations of a symmor- phic space group can now be presented. Theorem II Let к be any allowed к-vector of the Brillouin zone and let {Rj|O} (J = 1,2,..., M(k)) be a set of coset representatives for the decom- position of Po into left cosets with respect to Po(k). Let Г^0(к) be a unitary irreducible representation of (7o(k), assumed to be of dimension dp. Then there exists a corresponding unitary irreducible representation of the space group Q of dimension dpM(k), which may be denoted by Гкр, such that ( exP{—^(Rfck).tn} Г^о(к)({Rfc RRjO})*r, rkp({R|tn})H,jr = < if R^RRj e 0o(k), (7.23) ( 0, otherwise, for j, к = 1,2,..., M(k), and r, t = 1, 2,..., dp. (Here each row and column is specified by a pair of indices.) Moreover, all the inequivalent irreducible representations of Q may be obtained in this way by working through all the inequivalent irreducible representations of Qo (k) for all allowed k-vectors that are in different stars. Proof All that is required is to identify the concepts introduced above with those developed for induced representations in Chapter 5, Section 7. The required result is then an immediate consequence of Theorem II of Chapter 5, Section 7. Theorem I above showed that Q is isomorphic to so the groups A and В can be identified with T and Po respectively. As the characters of T are specified by the allowed k-vectors and are given by Equation (7.11), the label q may be identified with k. Thus with A = {1 |tn} Хд(А) = exp(-zk.tn). (7.24) With В = {R|0}, BAB-1 = {1 |Rtn}, so Хд(ВАВ-1) = exp{—ik.(Rtn)} = exp{—z(R-1k).tn} (by Equation (7.20)). Thus Equation (5.43) implies that the subgroup B(q) of В is merely (7o(k), that is, the (7o(k) are the “little groups”. Clearly b = go, b(<?) = #o(kL and M(q) = If {RJO} (J = 1,2,..., M(k)) are the coset representatives for the de- composition of Qq into left cosets with respect to (7o(k), then {Rj|0}- (j = 1,2,... ,M(k)) may be taken as the coset representatives for the decompo- sition of Po into right cosets with respect to (7o(k). (This follows because, if {Rj|0}-1 and {R^O}-1 belong to the same right coset, there exists an {RIO}’1 of Qo such that {Rfc|0} 1 = {R|0}{Rj|0} \ Then {R&|0} = {Rj|0}{R|0}-1, which implies that {Rj|0} and {Rfc|0} belong to the same
124 GROUP THEORY IN PHYSICS left coset.) Thus the coset representatives Bj (j = 1,2,..., M(qf) for the decomposition of В into right cosets with respect to B(q) may be identified with {Rj|O}-1 (J = l,2,...,M(k). Then, by Equations (5.44), (7.20) and (7.24), with A = {1 |tn}, %^(9)(Л)=ехр{-г(К,к)Лп}, (7.25) and Equations (5.47) reduce to Equations (7.23). Finally, the orbit of q is obviously merely the star of k. This theorem removes the need for an explicit display of the character table for P, which would in any case be very difficult because of the vast number of irreducible representations that Q possesses. All the information required about Q on such things as basis functions and degeneracies is imme- diately provided by the theorem in terms of the corresponding quantities for crystallographic point groups, which are very easily obtained. The character ykp({R|tn}) of {R|tn} in the irreducible representation Гкр of Q is given by Xk”({R|tn}) = ^exp{-»(R,k).tn} ^({R-'RRj-lO}), (7.26) 3 where the sum is over all the coset representatives R7 such that {R“1RRJ|0} e Po(k) and where denotes the character of the irreducible represen- tation Г£о(к) of Po(k). (This follows immediately from Equation (5.48) using Equation (7.25).) In applications the following properties of the basis functions of Гкр of Po(k) are useful. Theorem III Let (r) (J = 1,2,..., M(k), and r = 1,2,..., dp) be a set of basis functions of the unitary irreducible representation Гкр of the space group Q defined by Equations (7.23). Then (a) V’jT (r) a Bloch function with wave vector R7k, (b) the functions Vhr (r) (r = 1? 2,..., dp) form a basis for the unitary irre- ducible representation Гкр of Po(k), anc^ (c) V-^(r) = F({Rj|0})^(r), for j = 1,2,... ,M(k). Proof In the double subscript notation and the present context, Equation (1.26) becomes M(k) dp P({R|tn})^(r) = £ £rk₽({R|tn})fctjr<r(r). (7.27) k=l j = l
CRYSTALLOGRAPHIC SPACE GROUPS 125 (a) It follows from Equations (7.23) that rkp({l|tn})fct,jr = <5ifc^rexp{-i(R>k).tn}, so, by Equation (7.27), Л{!|ММкг (r) = exp{-«(Rjk).tn}V’]T(r)- Comparison with Equations (7.12) shows that (r) is a Bloch function with wave vector R7k. (b) If {R|0} e £o(k), Equations (7.23) imply that rkp{R|OhMr = 6kl rPo(k)({R|O})tr . Then, by Equation (7.27), for {R|0} E (7o(k) cZp F({R|O})<(r) = £rPo(k)({R|O})tr<(r). t=l (c) From Equations (7.23) Гкр(^|0})ы,1г = 6jk6tr , so that, by Equation (7.27), F({Rj|O})^kp(r) = V’jT(r)- This latter theorem has a converse. Theorem IV Suppose that <^kp(r), r = 1,2, ...,dp, are Bloch functions, with wave vector k, that are also basis functions of the unitary irreducible representation of <7o(k). Let <(r) = F({RJ|O})<p(r), j = l,2,...,M(k). (7.28) Then the set of dpM(k) functions (r) f°rms a basis for the irreducible representation Гкр of Q defined by Equations (7.23). Proof All that has to be shown is that the functions ^j^(r), as defined in Equation (7.28), satisfy Equation (7.27). This follows immediately from the identity P({R|tn})P({Rj|O}) = P({Rfc|O})P({l|R;1tn})P({R;1RRJ|O}), in which R/c can be chosen so that {R^1RRJ|O} E (7o(k). The analysis of Chapter 6, Section 1, demonstrates that considerable sim- plification of the calculation of electronic energy eigenvalues can be achieved
126 GROUP THEORY IN PHYSICS if the energy eigenfunctions corresponding to an irreducible representation of the group of the Schrodinger equation are expanded in terms of basis func- tions of that representation. Moreover, it is sufficient to restrict attention just to functions transforming as one row of the irreducible representation. In the present context, the group of the Schrodinger equation is a symmorphic space group P, and so for an energy eigenvalue corresponding to Гкр it is sufficient to restrict attention to functions transforming as the j = 1, r = 1 row of Гкр, that is, to Bloch functions with wave vector к that transform under the rotations of Po(k) as the first row of Such “symmetrized wave functions” can be constructed using the pro- jection operator technique described in Chapter 5, Section 1. Applied to spherical harmonics, plane waves or atomic orbitals this technique produces the “symmetrized spherical harmonics”, “symmetrized plane waves” or “sym- metrized atomic orbitals” that are needed for the various methods of electronic energy band calculation. (Good accounts of these methods exist in the re- views of Callaway (1958,1964), Fletcher (1971), Pincherle (1960,1971) and Reitz (1955).) Details of these applications of the projection operator tech- nique may be found elsewhere (e.g. Cornwell (1969)). Tabulations of the symmetrized spherical harmonics for the space groups O^ O^ and O^ the so-called “kubic harmonics”, have been given by von der Lage and Bethe (1947), Howarth and Jones (1952), Bell (1954), Altmann (1957) and Altmann and Cracknell (1965). The construction of symmetrized plane waves for all 73 symmorphic space groups may be simplified by the use of tables given by Luehrmann (1968), which supplement those given previously for O^ Of and Tj by Schlosser (1962). (b) Irreducible representations of the cubic space groups Oh, 01 and O9h The simple cubic space group O^ the face-centred cubic space group O^ and the body-centred cubic space group O^ provide examples of symmorphic space groups of great importance. The following description of them will serve to introduce the standard notations and conventions that are used for all symmorphic space groups. As shown by Theorem II of subsection (a), an irreducible representation of the space group Q is specified by an allowed к-vector and a label p of the irreducible representation of <70(k) to which it corresponds. The convention is that the symmetry points and axes of the Brillouin zone are denoted by capital letters, as for example in Figures 7.7, 7.8 and 7.9. The irreducible representations of the corresponding point groups (7o(k) are then labelled by assigning a subscript or set of subscripts to the appropriate capital letter. For example, the centre of the Brillouin zone for the space groups Oj^ Of and Ofr is the point Г, so that the ten irreducible representations of the point group Po(k) for Г (which is actually Oh) are called Г1, Г2, Г12, Г15, Г25, Г1/,Г2', Г12', Г15/ and Г25' in the most commonly used notation of Bouckaert et al. (1936). (This assignment of subscripts is almost entirely arbitrary and
CRYSTALLOGRAPHIC SPACE GROUPS 127 Point Coordinates Go (к) Г (0,0,0) oh M (тг/а) (0,1,1) D^h R (тг/а)(1,1,1) oh X (тг/а)(0,0,1) Dih Axis Coordinates, 0 < к < 1 0o(k) A (тг/а)(0,0, к) Civ A к, к) c3v S (тг/а)(0, ft, ft) C2v s (?r/a)(ft, ft, 1) C2v T (тг/а)(«, 1,1) Civ z 0/0(0, к, 1) c2v Plane Equation (k) ГМХ Г RM TRX MRX kx = 0 ky — kz > kx kx — ky <C kz kz = 7Г/a Cs(IC2x) Cs(IC2f) Cs(IC2b) Cs(IC2z) Table 7.2: The point groups (7o(k) for the symmetry points, axes and planes of the simple cubic space group O^ unfortunately conveys no direct information about the nature of the corre- sponding irreducible representation. More informative notations have been proposed by Howarth and Jones (1952) and by Bell (1954), but the notation of Bouckaert et al. (1936) is so widely used that to employ any other notation now would just cause confusion.) The point groups (7o(k) for the symmetry points, axes and planes of the space groups O^ Of and O^ are given in Tables 7.2, 7.3 and 7.4 respectively. The notation for the point groups is that of Schonfliess (1923). This notation is used again in Appendix C, where the character tables for all 32 crystal- lographic point groups are listed, together with explicit sets of matrices for all irreducible representations of dimension greater than one. The tables of Appendix C also give, when appropriate, the notation of Bouckaert et al. (1936) for the irreducible representations of P0(k). (The rotations of each Po(k) listed in Appendix C are for the value of к specified in Tables 7.2, 7.3 and 7.4.) It is possible for two or more points in different stars to have point groups Po(k) that are isomorphic. In such a situation the actual group elements of the Po(k) may be different, the isomorphic groups merely differing in the orientation of their defining axes. For example, for the body-centred cubic space group O^ the points on the axes S and D correspond to a (7o(k) that is
128 GROUP THEORY IN PHYSICS Point Coordinates 0о(к) Г (0,0,0) oh L (тг/а)(1,1,1) Dsd W (тг/а)(0,1,2) D2d X (тг/а)(0,0,2) D^h Axis Coordinates So (k) A (тг/а)(0,0, 2k), 0 < к < 1 Civ A (тг/а)(к, к, к), 0 < к < 1 c3v S (тг/а)(0, |к, |к), 0 < к < 1 C2v Q (тг/а)(1 — к, 1,1 + к), 0 < к < ;1 C2 s (тг/а)(|к, |к,2), 0 < к < 1 c2v z (тг/а)(0, к, 2), 0 < к < 1 C2v Plane Equation 0o(k) TKWX кх = 0 TKL ку — kz кх Cs(IC2f) FLUX кх — ку <С kz Cs(IC2b) UWX kz = 2ir/a Cs(IC2z) Table 7.3: The point groups Po(k) for the symmetry points, axes and planes of the face-centred cubic space group O^ C2v- However, with the coordinates of S and D given in Table 7.4, the group elements of (7o(k) for S (arranged in classes) are Ci = E,C2 = C2e, C3 = IC2x, C4 = IC2f, whereas the group elements of (7o(k) for D are Ci = E, C2 = C2x, C3 = IC2e, C4 = IC2f. The irreducible representations of Po(k) for S are denoted by Si, S2> S3 and S4, while those for D are denoted by Eh, D2, D3 and Z)4. The different sets of group elements occurring in this way are all listed in the description of the corresponding point group in Appendix C. For every point к on a symmetry plane, the group (7o(k) is Cs, which contains just the identity transformation and a reflection. The appropriate reflection for each plane is indicated in parentheses in the third column of Tables 7.2, 7.3 and 7.4. The irreducible representation of Cs for which the character of the reflection is +1 is described as being “even” and is denoted by a +, and the other irreducible representation is described as being “odd” and is denoted by a —. Figures 7.10 and 7.11 are examples of energy bands employing the notation described above.
CRYSTALLOGRAPHIC SPACE GROUPS 129 Point Coordinates <7o(k) Г (0,0,0) oh H (?r/a)(0,0,2) oh N (тг/а) (0,1,1) Dih P (тг/а)(1,1,1) Td Axis Coordinates, 0 < к < 1 Go (k) A (тг/а)(0,0,2к) A к, к) c3v S (тг/а)(0, ft, ft) c2v D (?r/a)(ft, ft, 1) c2v F (тг/а)(1 — ft, 1 — ft, 1 + ft) c3v G (тг/а)(0,1 — ft, 1 + ft) c2v Plane Equation <7o(k) IHN kx = 0 C3(IC2x) LNP ky — kz kx Cs(IC2f) VHP kx — ky kz C.s(IC2b) HNP ky + kz = Tn ja C3(IC2e) Table 7.4: The point groups (7o(k) for the symmetry points, axes and planes of the body-centred cubic space group O^ 8 Consequences of the fundamental theorems (a) Degeneracies of eigenvalues and the symmetry of e(k) Suppose first that к is a general point of the Brillouin zone. Then Po(k) consists only of the identity transformation {l|0} and so has only one irre- ducible representation, namely the one-dimensional representation for which Гх({1|0}) = [1]. As #o(k) = 1, it follows from Equation (7.22) that M(k) = go. There is therefore only one irreducible representation of Q corresponding to this k, and its dimension is M(k)di = #o-l = so that the corresponding energy eigenvalue is #o-f°ld degenerate. Any Bloch function exp(zk.r)ttk(r) with this wave vector к is a basis function for the irreducible representation of this (7o(k) and the set of ba- sis functions of the corresponding irreducible representation of Q formed from this function are exp(zk.r)ttk(r), exp(zk2.r)ttk(R2 lr)> • • •• Now suppose that these Bloch functions are energy eigenfunctions. As they correspond to wave vectors k(= ki),k2,..., they correspond, according to Equation (7.14), to energy eigenvalues e(k),e(k2),.... As they are degenerate, being a basis for an irreducible representation of P, it follows that e(R,k) ее e(k,) = e(k) (7.29) for J = 1,2,..., #0, the Ry being the set of rotations of Qq. Thus the go wave
130 GROUP THEORY IN PHYSICS vectors in the star of к have the same energies, and c(k) has the symmetry of the point group Go of the space group Q. This means that if the band structure is known in one basic section of the Brillouin zone containing only (l/<?o) of the volume of the Brillouin zone and no two wave vectors in the same star, then it can be obtained immediately throughout the whole Brillouin zone. For example, for the body-centred cubic space group O^ whose Brillouin zone is shown in Figure 7.8, go = 48 and the basic section is the wedge-shaped region THNP (or, more precisely, the region bounded by the planes containing three of the four points Г, Я, N and P). The symmetry of c(k) is widely employed in calculations of energy band structures and in determinations of the Fermi surface from experimental mea- surements, such as those of the de Haas-van Alphen effect. For a general point of the Brillouin zone the inclusion of the rotational parts of G in addition to the translational parts of T already taken into account in Bloch’s Theorem is of no assistance in simplifying the numerical task of actually finding the energy eigenvalues by the technique described in Chapter 6, Section 1. For this reason, relatively few accurate calculations of c(k) have been performed for general points of the Brillouin zone. The second case that will be considered is at the other extreme. For the point к = 0, Rk = 0 for every {R|0} of Po(k), so that (7o(O) = Po- The star of к then consists only of к = 0 itself and so M(0) = 1. The basis functions of Q corresponding to к = 0 are merely the periodic basis functions of Go, and the corresponding degeneracies of energy eigenvalues are those of the dimensions of the irreducible representations of <y0. In this case the technique of Chapter 6, Section 1, allows an appreciable simplification of the numerical work involved in finding the energy eigenvalues, even beyond that already brought about by the consideration of T alone. Some space groups possess other symmetry points for which (7o(k) = Go- These do not require further examination, as the comments made about the point к = 0 also apply to these points. The point H of the Brillouin zone of the body-centred cubic lattice is an example. The third and final case is that of the intermediate situation in which Po(k) is not trivial but is a proper subgroup of Go- Included in this case are all symmetry points other than those of the second case and all points on symmetry axes and planes. For such a point (7o(k) will have more than one irreducible representation, some of these possibly being of more than one dimension. As #o(k) < go, it follows from Equation (7.22) that M(k) > 1. Consider the energy eigenvalue corresponding to a dp-dimensional irre- ducible representation of Po(k). This will be dpM (k)-fold degenerate by the theorem, and this degeneracy is made up as follows: (a) By a similar argument to that used in the case of a general point, it follows that б(к,) = е(к), J = 1,2, ...,M(k), so that again c(k) exhibits the symmetry of the point group Go-
CRYSTALLOGRAPHIC SPACE GROUPS 131 (b) In addition, each e(kj) is “dp-fold degenerate”, in the sense that there are dp linearly independent Bloch eigenfunctions of H(r) corresponding to this eigenvalue and to this particular wave vector k7. The degree of simplification of numerical work depends on the order #o(k)- These arguments show that, although the concept of a star appears in the fundamental theorems for symmorphic space groups, it is in some contexts possible and convenient to revert to the description that appeared in Section 4 with Bloch’s Theorem, in which there corresponds a set of energy levels to every allowed к-vector of the Brillouin zone and not merely to those lying in different stars. In this description, a dp-fold degeneracy of e(k) means, as above, that there are dp linearly independent Bloch eigenfunctions of Я(г) corresponding to this eigenvalue and to the particular wave vector k. A dp-fold degeneracy of e(k) then corresponds to a dp-dimensional irreducible representation of Po(k). This degeneracy was indicated in Figures 7.10 and 7.11 by [dp\. Furthermore each energy band then has the symmetry of Qq. This is the description that is commonly used in the solid state literature. It is worth while pointing out here that even if does not contain the spatial inversion operator 7, the symmetry e(k) = e(—k) always remains be- cause of time-reversal symmetry. (For details, see, for example, Chapter 8, Section 5, of Cornwell (1984)). (b) Continuity and compatibility of the irreducible rep- resentations of Qo (k) A typical section of an energy band diagram for a symmetry axis is shown in Figure 7.12. The axis displayed there is the Д axis of the cubic space groups Ofr, Of and Ofr. The numbers in parentheses are the band labels, as defined in Expression (7.19), so that the bands 1 and 2 “touch” at one point kg. This figure exhibits two general characteristics of energy bands. The first feature is that, along any part of a band which does not touch another band, the corresponding irreducible representation of (7o(k) remains the same. Thus, for example, the whole of the left-hand part of band 1 cor- responds to Д2 and the whole of the right-hand part to ДХ/. It is therefore possible to talk of the symmetry of a band, or part of a band, along an axis. The reason for this behaviour is essentially that the energy eigenvalues corre- sponding to a particular irreducible representation of <7o(k) can be obtained from a secular equation involving only basis functions of that representation, as was noted in Chapter 6, Section 1. A small change in к will then only produce a small change in the energy eigenvalues emerging from this secular equation. The second general characteristic is that the symmetries of the bands are interchanged at a point where the bands touch. For example, in Figure 7.12 band 1 changes from Д2 to ДХ/ on moving from left to right, while band 2 changes from ДХ/ to Д2. The reason for this is that the secular equation corresponding to an irreducible representation of Po(k) produces energy eigen- values that are analytic functions of k, the degeneracy corresponding to the
132 GROUP THEORY IN PHYSICS Figure 7.12: “Touching” of energy bands along the Д axis of the cubic space groups. touching of two bands having no effect on this as it is “accidental”. This also implies that gradkc(k) is continuous for bands, or parts of bands, belonging to the same irreducible representation, and this continuity is not affected by the interchange of band labels when bands touch. Thus, for example, in Figure 7.12 the limit as к —> k0 from the left of gradkc(k) for band 1 is equal to the limit as к —> k0 from the right of gradkc(k) for band 2. The “touching” of energy bands has been investigated in detail by Herring (1937). The concept of “compatibility” is best described by an example. Consider therefore а Г12 energy level at the point Г of the Brillouin zone for the body- centred cubic space group O^. This level, being two-fold degenerate, belongs to two energy bands. What then are the symmetries of these bands near Г along the symmetry axis Д? That is, what irreducible representations of the group Po(k) for Д are “compatible” with Г12? The investigation proceeds as follows. Let (7о(Г) and (7о(Д) be the point groups Po(k) for к at Г and on Д respectively. Then (7о(Д) is a subgroup of Ро(Г) and so the irreducible representations of Ро(Г) are representations of Р0(Д) that are, in general, reducible. The actual reduction can be determined immediately from the characters, for if Г and Гр are irreducible representa- tions of Ро(Г) and Qq (Д) respectively, with characters denoted by x and yp, then the number of times np that Гр appears in the reduction of Г is given by np = (Ш(Д)) £ x({R|O})^({R|O})*, {R|O}e^o(A) where <?o(A) is the order of (7о(Д). (This is just an immediate application of Theorem V of Chapter 4, Section 6.) Thus, for example, the reduction of Г12
CRYSTALLOGRAPHIC SPACE GROUPS 133 Figure 7.13: “Compatibility” of the Г12 energy level with the Д1 and Д2 energy levels along the Д axis. in this context is given by Г12 = Д1 Ф Д2- The point Г could be considered as an ordinary point of Д by ignoring the elements of (7о(Г) that are not in (7о(Д), and then the energy levels at Г could be classified in terms of the irreducible representations of (7о(Д). The Г12 level would then be regarded as a non-degenerate Д1 level and a non-degenerate Д2 level that happen to have the same value. Because of the continuity of irreducible representations along axes that was mentioned above, the two bands that touch at Г in the Г12 level will along the Д axis near Г have symmetries Д1 and Д2. The degeneracy that exists at Г is split on moving away from Г. This situation is shown in Figure 7.13. Although this argument shows that the Д1 and Д2 irreducible represen- tations are compatible with the Г12 level, it cannot predict whether the band having Д1 symmetry lies lower or higher than the band having Д2 symmetry. This can only be determined by direct calculation. The same analysis can be applied to all irreducible representations at every symmetry point for all the symmetry axes going through that point. A similar analysis can also be used to determine the compatibility of the irreducible representations corresponding to a symmetry axis with those of the symmetry planes containing the axis. These results can be expressed in “compatibility tables”. A typical example is exhibited in Table 7.5. A comprehensive set of such tables for the space groups O^ Oft and O^ was given by Bouckaert et al. (1936) (see also Cornwell (1969)). It is possible to develop this type of analysis to investigate certain finer fea- tures of the electronic energy bands, particularly the vanishing of components of gradkc(k) and the form of the intersection of constant energy contours with symmetry axes. This allows the location of the “critical points” (van Hove
134 GROUP THEORY IN PHYSICS Г1 Г2 Г12 Г15 Г25 Ai Д2 Д1Д2 Д1Д5 Д2Д5 Si S4 E1S4 S1S3S4 S2S3S4 Ai Л2 A3 А1ЛЗ А2Л3 Гг r2z Г12' Г15' Г25' Ar Д2' Д1/ Д2' Д1/Д5 Д2'Дб s2 S3 S2S3 S2S3S4 S1S2S3 Л2 Ai A3 А2Л3 Л1Л3 Table 7.5: Compatibility relations between the symmetry point Г and the symmetry axes Д, S and Л for the cubic space groups O^ Of and O^. 1953, Phillips 1956) to be located. For details see Cornwell (1969), and also Rashba (1959), Sheka (1960) and Kudryavtseva (1967). (c) Origin and orientation dependence of the symmetry labelling of electronic states Consider, as an example, the NaCl structure, in which the Na nuclei occupy the lattice sites of one r£ lattice and the Cl nuclei occupy the lattice sites of another Г£ lattice, one lattice being obtained from the other by a pure translation through to = |n(l, 0, 0). (Here a is the quantity appearing in the definition of the basic lattice vectors of in Table 7.1). With the origin of the coordinate system taken at one of the Na nuclei, Equation (1.13) is satisfied for every coordinate transformation T of the space group O^. Similarly, with the origin at a Cl nucleus, Equation (1.13) is again satisfied for every T of O^. Let H(r) and H'(r) be the Hamiltonian operators corresponding to these two choices of origin. Although Я(г) and H'(r) are related, they are obviously not identical, that is, Я(г) /= H'(r). It can be shown (Cornwell 1971) in a situation such as this that, while the actual values of the electronic energy levels are naturally independent of the choice of origin of the coordinate system, the labelling of the states in terms of the irreducible representations of the space group can be origin dependent. The symmetry labels of states also depend on the choice of the orientation of the coordinate axes, as has been discussed in detail by Cornwell (1972).
Chapter 8 The Role of Lie Algebras 1 “Local” and “global” aspects of Lie groups It is now time to begin the systematic study of Lie groups. They were in- troduced at an early stage in Chapter 3 so that the general features of their representation theory could be presented at the same time as the represen- tation theory of finite groups. Although the definition of a linear Lie group given in Section 1 of Chapter 3 necessarily involved the “local” coordinates (#i, Х2ч •. • , xn) which parametrize elements near the identity, the emphasis in the subsequent sections of that chapter was on the “global” properties (that is, the properties of the whole group), particularly the concept of compactness and integration on the group. In the closer study of Lie groups both the “local” and the “global” aspects are important, but it is fair to say that most of the information concerning the structure of a Lie group comes from the investigation of its “local” properties. It is the main purpose of this chapter to show how these “local” properties are themselves determined by the corresponding “real Lie algebra”. The link is provided for linear Lie groups by the matrix exponential function, which is described in Section 2, and which leads in turn to the idea of a “one-parameter subgroup” of Section 3. The concept of a real Lie algebra is introduced first for the group of proper rotations in IR3, for which the argument (in Section 4) is helped by the geometrical nature of the elements under consideration. At the same time the very useful and closely related notion of a “complex Lie algebra” is defined. For general linear Lie groups a slightly different line of argument is needed, and this is provided in Section 5. Chapter 9 will be mainly concerned with introducing for Lie algebras many of the ideas previously discussed for groups, and relating these to the analo- gous properties of linear Lie groups. Again the emphasis will be largely on the “local” aspects of the groups. Chapter 10 is devoted to the rotation groups in IR3, to the related groups SO(3), 0(3) and SU(2), and to their Lie algebras. Not only are they important for their applications in atomic and nuclear physics, but their representations
136 GROUP THEORY IN PHYSICS lie at the heart of much of the representation theory that follows in later chapters. With Chapter 11 attention begins to be concentrated on the so-called “simple” and “semi-simple” Lie algebras, which have very important physical applications, and the structure theory of the semi-simple complex Lie algebras is investigated in detail. The representation theory of semi-simple Lie algebras is described in Chapter 12. It may be helpful to anticipate some of the discussion by alerting the reader to the fact that, although to every Lie group there is a real Lie algebra which is unique (up to isomorphism), in general several non-isomorphic Lie groups can correspond to the same real Lie algebra. Also, although to every real Lie algebra the complexification is unique (up to isomorphism), in general several non-isomorphic real Lie algebras correspond to the same complex Lie algebra. In all the arguments that follow involving matrices, the “commutator” [A.B] of any two m x m matrices A and В is defined by [A, B] = AB - BA. 2 The matrix exponential function The matrix exponential function provides the link between a linear Lie group and its corresponding real Lie algebra. Its definition and certain of its prop- erties are simple generalizations of those of the familiar exponential function of a real or complex number. Definition The matrix exponential function If a is an m x m matrix, then exp a is the m x m matrix defined by exp a = 1 + ^a^/j! . (8.1) 3 = 1 Theorem I The series for exp a in Equation (8.1) converges for any m x m matrix a. Proof See, for example, Chapter 10, Section 2, of Cornwell (1984). The following example will prove to be very significant. Example I The proper rotation matrices R(T) of SO(3) as matrix exponen- tial functions Consider the 3x3 matrix ai defined by " 0 0 0 ’ ai = 0 0 1 (8-2) 0-10
THE ROLE OF LIE ALGEBRAS 137 Then, with a = 0ai, a-7 = (—l/-7 1^2^-7ai for j odd, and 0 0 0 0 1 0 0 0 1 for j even, so that exp(0ai) = 1 0 0 0 cos 0 sin 0 0 — sin 0 cos 0 (8-3) As noted in Chapter 1, Section 1, the right-hand side of Equation (8.3) speci- fies a proper rotation through an angle в in the right-hand screw sense about the axis Ox (cf. Equation (1.3)). It will be demonstrated in Section 4 that every matrix of SO(3) can be expressed in matrix exponential form with a suitable choice of exponent. The multiplication properties of matrix exponential functions are more complicated than those of exponential functions of real or complex numbers, as the following theorem shows. Theorem II (a) If a and b are any m x m matrices that commute, (exp a) (exp b) = exp(a + b) = (exp b) (exp a). (8.4) (b) If a and b are m x m matrices whose entries are sufficiently small (exp a) (exp b) = exp c, where 1 1 c = a + b + - [a, b] + — {[a, [a, b]] + [b, [b, a]]} + ..., (8.5) where the infinite series in Equation (8.5) contains commutators of in- creasingly higher order. Thus, in general, (expa)(expb) (expb)(expa). Equation (8.5) is known as the “Campbell-Baker-Hausdorff formula”. Proof (a) If [a,b] = 0 then (a + b)-7 = _ k)l)}a.kbJ~k, so that Equations (8.4) follow in exactly the same way as the corresponding results for real or complex numbers.
138 GROUP THEORY IN PHYSICS (b) The precise conditions on the smallness of the elements of a and b re- quired to ensure the convergence of the series in Equation (8.5), together with the complete expression for c, may be found in the original papers (Campbell 1897a,b, Baker 1905, Hausdorff 1906). All that will be done here is to demonstrate the correctness of Equation (8.5) to second order. From Equation (8.1) (to second order) (expa)(expb) = {1 + a+ |a2 + .. .}{1 + b + |b2 + ...} = 1 + (a + b) + ^2^ ’ (8-6) However, to second order, expc = exp{a + b + |[a,b] + ...} = l + {(a + b) + |[a,b] + ---} + -{(a + b) + -[a,b] + .. .}2 + ... Z z = 1 + {a + b + |(ab - ba) + ...} +—{a2 + ba + ab + b2 + •> (8.7) from which Equation (8.5) follows to second order, on equating the right-hand sides of Equations (8.6) and (8.7). Theorem III The matrix exponential function formed from an m x m ma- trix a possesses the following properties: (a) (expa)* = exp(a*). (b) The transpose of (exp a) is exp(a). (c) (expa)t = exp (at). (d) For any m x m non-singular matrix S exp(SaS-1) = S(expa)S-1. (e) If Ai, A2,..., Am are the eigenvalues of a, then eA1, eA2,..., eAm are the eigenvalues of exp a. (f) det(expa) = exp(tra). (g) exp a is always non-singular and (exp a)-1 = exp(—a).
THE ROLE OF LIE ALGEBRAS 139 (h) The mapping </>(a) = exp a is a one-to-one continuous mapping of a small neighbourhood of the m x m zero matrix 0 onto a small neighbourhood of the m x m unit matrix 1. Proof See, for example, Appendix E, Section 1, of Cornwell (1984). 3 One-parameter subgroups Definition One-parameter subgroup of a linear Lie group A “one-parameter subgroup” of a linear Lie group Q is a Lie subgroup of Q consisting of elements TfL) which depend on a real parameter t that takes all values from —ею to +сю such that T(s)T(t) =T(s + t) (8.8) for all s and —сю < s, t < +сю. In particular, if Q is a group of m x m matrices then a one-parameter subgroup of Q is a Lie subgroup of matrices A(t) such that A(s)A(t) = A(s + t) (8.9) for all s and —сю < s, t < +сю. Clearly T(s)T(f) = T(t)T(s) for all s and so every one-parameter sub- group is Abelian. Moreover, Equation (8.8) with s = 0 implies that T(0) — E, the identity of Q. Obviously a one-parameter subgroup is a Lie group of di- mension 1, so that in the matrix case dA/dt for t = 0 exists and is not identically zero. Example I A one-parameter subgroup o/SO(3) The 3x3 matrices A(t) defined by A(i) = 1 0 0 0 cos t sin t 0 — sin t cos t satisfy Equation (8.9) and form a subgroup of SO(3) that is (by Equation (3.7)) isomorphic to the Lie group SO(2). Thus these matrices form a one- parameter subgroup of SO(3). As shown by Example I of Section 2, A(f) = exp(tai), where ai is specified by Equation (8.2). The property exhibited in this example is completely general, as the fol- lowing theorem shows. Theorem I Every one-parameter subgroup of a linear Lie group Q of m x m matrices is formed by exponentiation of m x m matrices. Indeed, if the
140 GROUP THEORY IN PHYSICS matrices A(t) form a one-parameter subgroup of P, then A(t) = exp{ta}, (8.10) where a = dA/dt evaluated at t = 0. Proof For brevity write A(t) = dA/dt, so a = A(0). Let B(t) = A(t) exp{—tA(0)}, so that B(t) = {A(£) — A(£)A(0)} exp{—£A(0)}. However, from Equation (8.9), for any t, A(t) = lim[A(t + s) — A(t)]/s = lim A(t)[A(s) — A(0)]/s, 8—>0 8—>0 so that A(£) = A(t)A(0). (8.11) Thus B(t) = 0 for all t and consequently B(t) = B(0) = 1, from which Equation (8.10) follows immediately. 4 Lie algebras For the linear Lie group SO(3) it will now be shown that it is possible to introduce the corresponding real Lie algebra in a very direct way by a combi- nation of algebraic and geometric arguments. For the other linear Lie groups the essential results are similar, but the arguments are rather longer and less direct. It is essential to bear in mind that, in this context, there are three mutually isomorphic groups, namely (a) the groups of all proper rotations T in IR3, (b) the group SO(3) of rotation matrices R(T), and (c) the corresponding group of linear operators P(T) (as defined in Equation (1-17)). Consider first any proper rotation T in IR3. Suppose this is a rotation through an angle 0q about a certain axis. Then the set of all rotations about that axis form a one-parameter subgroup. Consequently every proper rota- tion lies in some one-parameter subgroup of the group of proper rotations in IR3. Correspondingly, every matrix of SO(3) must lie in some one-parameter subgroup of SO(3). By the theorem of Section 3, if Aft) are the elements of such a one-parameter subgroup of SO(3), there exists a non-zero 3x3 ma- trix a such that A(t) = exp(ta). As A(t) is real, Theorem III of Section 2 implies that a also can be taken to be real. This theorem and the condition A(t) = A(£)-1 also imply that a = —a. Conversely, if a is real and a = —a,
THE ROLE OF LIE ALGEBRAS 141 then A(t) = exp(ta) is a member of S0(3). Thus every element o/SO(3) is obtained by exponentiation from some 3x3 real antisymmetric matrix. The set of all 3 x 3 real antisymmetric matrices forms a three-dimensional real vector space (see Appendix B, Section 1). (It forms a real vector space because, if a and b are any two such matrices, then so too is aa + /?b, for any real numbers a and (3. The dimension is three because, for any such matrix a, <2ц = <222 = <233 = 0 and &2i = “<212, <231 = — <213, and «23 = —<232.) Consequently a can be specified by three real parameters, such as its «12, «13 and «23 elements. A convenient basis for this vector space is formed by the matrices ai = 0 0 0 " 0 0 1 0-10 , a2 = 0 0 -1 ' 0 0 0 1 0 0 > a3 — ’010' -10 0 0 0 0 (8-12) These generate one-parameter subgroups of matrices R(T) corresponding to rotations about Ox, Oy and Oz respectively (cf. the Examples of Sections 2 and 3). As the commutator [a, b] (— ab — ba) of two real 3x3 antisymmetric matrices is also a real 3x3 antisymmetric matrix, [a, b] is a member of the vector space whenever a and b are members. Thus the set of all 3 x 3 real antisymmetric matrices satisfy the conditions in the following definition of a “real Lie algebra”. Definition Real Lie algebra C A “real Lie algebra” C of dimension n (> 1) is a real vector space of dimension n equipped with a “Lie product” or “commutator” [a, b] defined for every a and b of C such that (i) [«, b] e C for all a, b G £; (ii) for all «, 6, с e C and all real numbers a and (3 [aa + /3b, c] = a[a, c] + /3[b, c]; (8.13) (iii) [«, b] = — [6, a] for all «, b e £; and (iv) for all a, b, с e C [a, [b, c]] + [6, [c,«]] + [c, [«, 6]] = 0. (8-14) (This is known as “Jacobi’s identity”.) In the particular case of a Lie algebra of matrices the commutator [a, b] will always be defined by [a, b] = ab — ba, (8.15)
142 GROUP THEORY IN PHYSICS and then conditions (ii), (iii) and (iv) are automatically satisfied. Similarly, for a Lie algebra of linear operators the commutator [a, b] will always be defined by [а, Ь]ф = а(Ьф) — Ь(аф). (8.16) for any pair of linear operators a and b and for any element ф of the vector space in which they act for which the right-hand side of Equation (8.16) is meaningful. It is not absolutely essential to have an explicit definition of [a, b] in terms of other possibly simpler products such as those of Equation (8.15) or (8.16). Indeed, it is an interesting intellectual exercise to proceed solely on the basis of the properties (i), (ii), (iii) and (iv) without making any other assump- tions about the nature of [a, b], The resulting development is the theory of “abstract” Lie algebras. This does not lead to any new structures, for there exists a theorem by Ado (1947) which states that every abstract Lie algebra is isomorphic to a Lie algebra of matrices with the commutator defined as in Equation (8.15). Nevertheless, much of the development of abstract Lie algebras is no more complicated than the corresponding theory for matrices, and indeed has the additional advantage that it applies equally to matrices and linear operators as special cases. Consequently the formulation will be given in general terms whenever it is convenient to do so. This is certainly the case for certain immediate consequences of the defi- nition. For example, (ii) and (iii) imply that [a, fib + 7c] = fi[a, b] + 7[a, c] (8-17) for all a, 6, с e C and all real numbers fi and 7. Further, let «1, ..., an be a basis of the real vector space of C. As [ap, aq] e C for all p, q = 1,2,..., n, there exists a set of n3 real numbers cpq known as the “structure constants of C with respect to the basis a1? a2,..., an \ that are defined by n [ap,aQ] = ^2cpqar, p,q=l,2,...,n. (8.18) r=l (Conditions (iii) and (iv) of the definition of C imply that crpq = — cqp (for p,q = 1,2,...,n), and £"=i{<$,4 + + с*рф} = 0 (for p,q,r,t = 1, 2,..., n), so these constants are not independent.) Then, if a = ]Cp=i apap and b = Y^q=i Pqaq are апУ ^w0 elements of C (so that 07, ce2, ..., an and /?i,/?2, • • • ifin are all real), by Equations (8.13), (8.17) and (8.18), n [«>4= Y apfiqCpqar- (8.19) p, q,r=i Thus every commutator can be evaluated from a knowledge of the structure constants. In particular, in the real Lie algebra C = so(3) associated with SO(3), with the basis elements ai, a2 and аз defined by Equations (8.12), [ai,a2] = —a3, [a2,a3] = -ax, [a3,ai] = -a2. (8.20)
THE ROLE OF LIE ALGEBRAS 143 Consequently the structure constants with respect to ai, a2 and аз are given by f 1, if (p, q, r) = (1,2,3), (2, 3,1), (3,1,2), ~crpq = epqr=l -1, if (p,Q,r) = (2,1,3), (1,3, 2), (3,2,1), (8.21) [ 0, for all other values of (p, q, r). The Campbell-Baker-Hausdorff formula (see Section 2) then indicates that the product of any two elements of the group SO(3) lying close to the identity can be determined (at least in principle) from the structure constants. That is, the structure of the group SO(3) close to its identity is specified by the structure of its corresponding real Lie algebra C (=so(3)). The commutation relations (Equations (8.20)) take a very familiar form when the real Lie algebra associated with the group of linear operators P(T) corresponding to the group of proper rotations in IR3 is considered. Let T be the rotation corresponding to the matrix exp(ta) of SO(3), so that, by Equation (1.17), = /({exp(ta)}-1r) = /({exp(-ta)}r). As a = lim^o{exp(ta) — l}/t, it is natural to define a corresponding linear operator P(a) by the same limiting process. That is, let P(a) = lim{P(exp(ta)) - P(l)}/t. Thus, for any function /(r) in the domain of P(a), P(a)/(r) = lim[/({exp(—ta)}r) - f(r)]/t. However, for small t, /({exp(—ta)}r) ~ /({1 — ta + ...}r) = /(r - tar + ...) - /(r)-tragrad/(r), so that P(a) = — fagrad. (8.22) (Here r and grad are to be interpreted as 3 x 1 column matrices with entries x,y,z and д/dx., d/dy, d/dz respectively.) Thus, from Equations (8.12), P(ai) = yd/dz — zd/dy, 1 Р(аг) = zdfdx — xdjdz. > (8.23) P(a3) = хд/ду — уд/дх. J Equation (8.22) implies that [P(a),P(b)]=P([a,b]). (8.24)
144 GROUP THEORY IN PHYSICS (8.25) In particular, by Equations (8.20), (or directly from Equations (8.23)), [P(ai),P(a2)] = -P(a3), [P(a2),P(a3)] = -P(ai), > [P(a3),P(ai)] = -P(a2). The key observation is that the quantum mechanical orbital angular mo- mentum operators Lx, Ly and Lz are just multiples of P(ai), P(a2) and P(a3). In fact Lx = (fi/i)P(aiL Ly = (П/г)Р(а2), Lz = (8.26) (see Schiff (1968)), so that Equations (8.25) imply the familiar angular mo- mentum commutation relations [Lx,Ly] — ihLz, [LyyLz] — itiLX') ^LZ1LX^ — ihLy. (8.27) There is therefore an intimate connection between the quantum theory of an- gular momentum and the group of proper rotations in IR3. In particular, it will be shown in Chapter 10, Section 4, that the determi- nation of the basis functions of the irreducible representations of the group of proper rotations in IR3 can be reduced to the construction of the simultaneous eigenfunctions of the orbital angular momentum operators Lz and L2, where L2 is defined by p2 = V + V + V. This latter problem is treated in nearly every book on elementary quantum mechanics (e.g. Schiff (1968)). These eigenfunctions and their corresponding eigenvalues can be found by solving certain differential equations (see Chap- ter 10, Section 4), but there exists a well-known, purely algebraic method of determining the eigenvalues (see Chapter 10, Section 3). It will become ap- parent that it is merely the prototype of a method that is applicable to the representations of a large class of Lie algebras. The details of this development will be given in Chapter 10, Section 4, but it is convenient to note here that the argument for angular momentum oper- ators involves the “ladder operators” L+ and L_, defined by L± = Lx ±iLy, that is, it involves the operators P(ai) ± гР(а2). The significant point is the appearance here of the imaginary number г, indicating that it is useful to extend the definition of a Lie algebra to embrace such complex linear combi- nations. The resulting structure is called a “complex Lie algebra”. Definition Complex Lie algebra C A “complex Lie algebra” C of dimension n (> 1) is a complex vector space of dimension n equipped with a “Lie product” or “commutator” possessing the properties (i), (ii), (iii) and (iv) listed in the definition of a real Lie algebra, except that in (ii) a and /3 are now any complex numbers. Equations (8.17), (8.18) and (8.19) apply also to complex Lie algebras (although now the a and /3 of (8.17), the structure constants crpq of (8.18), and the ai, q2, . • •, otn and /?i, /?2,..., /?n of (8.19) may be complex numbers).
THE ROLE OF LIE ALGEBRAS 145 In the case of a real Lie algebra of matrices or of linear operators whose basis elements are linearly independent over the complex field there is no difficulty in “complexifying” the real Lie algebra to produce a unique complex Lie algebra of the same dimension. In these situations the complex vector space may be taken to have the same basis elements as the real vector space, but, in the complex space, complex linear combinations of these basis elements are allowed. In fact this is the process already encountered in connection with the angular momentum ladder operators. (For the more general case the process is rather more elaborate. See, for example, the detailed discussion of Chapter 13, Section 3, of Cornwell (1984).) Somewhat paradoxically, the study of complex Lie algebras is more straightforward than that of real Lie algebras. Consequently it is convenient to investigate the properties of a linear Lie group by first introducing the cor- responding real Lie algebra, and then proceeding almost immediately to the associated complex Lie algebra. This section will be concluded with a definition that applies equally to real and complex Lie algebras: Definition Abelian Lie algebra A Lie algebra C is said to be “Abelian” if [a, b] = 0 for all a, b E C. Thus in an Abelian Lie algebra all the structure constants are zero. Such a Lie algebra may alternatively be called “commutative”. 5 The real Lie algebras that correspond to general linear Lie groups For Q = SO(3) it was elementary to demonstrate the occurrence of one- parameter subgroups, the existence of the corresponding real Lie algebra following from these. However, for a general linear Lie group Q it is nec- essary to reverse the order of the argument. First (in subsection (a)) it will be shown that for every such Q a corresponding real Lie algebra of matrices exists, and only then (in subsection (b)) will the existence and properties of the one-parameter subgroups of Q be deduced. (a) The existence of a real Lie algebra £ for every linear Lie group Q As a preliminary, the essential points of the definition of a linear Lie group Q of dimension n given in Chapter 3, Section 1, will be re-cast in the special case in which Q actually consists oimxm matrices A (so that T = A and Г(Т) = A). There is a one-to-one correspondence between these matrices lying close to the identity and the points in IRn satisfying Condition (3.2) which define the matrix function A(2q, ж2, • • •, Xn) (e P), for all (aq, ж2, • • •, %n) satisfying Condition (3.2). By assumption the elements of A(#i, ж2, • • •, xn) are analytic
146 GROUP THEORY IN PHYSICS functions of Ж1, ^2, • • •, xn. The n m x m matrices ai, a2,..., an defined by — (dAjk/dxp)X1=X2—^,=Xn=Q (8.28) (for j, к = 1, 2,..., m; p = 1, 2,..., n) (cf. Equation (3.3)) then form the basis for an n-dimensional real vector space V. Definition Analytic curve in Q Let rri(t), ^2^),..., xn(t) be a set of real analytic functions of t defined in some interval [0, to), where to > 0, such that rrj(O) = 0 for j = 1,2,..., n, and the point (rri(t), ^(t),...,rrn(t)) satisfies Condition (3.2) for all t in [0, to)- Then the corresponding set of m x m matrices A(t) of P, defined by A(t) = A(;ri(t),^(t),... ,rrn(t)), is said to form an “analytic curve” in Q. As A(0) = 1, every analytic curve starts from the identity of Q. There is no requirement at this stage that an analytic curve must form part of a one-parameter subgroup of Q. Definition Tangent vector of an analytic curve in Q The “tangent vector” of an analytic curve A(t) in Q is defined to be the m x m matrix a, where a = dA(t)/dt evaluated at t = 0. (More precisely, this is the tangent vector “at the identity”, but this extra phase will be omitted as no other tangent vectors will be considered here.) Theorem I The tangent vector of any analytic curve in Q is a member of the real vector space V having the matrices ai? a2,..., an of Equation (8.28) as its basis. Conversely, every member of V is the tangent vector of some analytic curve in Q. Proof As dA(i)/dt = ^p=1(dA/dxp)(dxp/dt)) it follows then that a = 52p=i £p(0)ap, where ip(0) = (dxp/dt)t=o. Thus a E V. Conversely, suppose a = 52p=1 Apap is any member of V. Then xj(t) = Xjt, j = 1,2,..., n, defines an analytic curve that has a as its tangent vector. Theorem II If a and b are the tangent vectors of the analytic curves A(t) and B(t) in P, then [a, b] (= ab — ba) is the tangent vector of the analytic curve C(t) in P, where C(t) = A(Vt)B(Vt)A(Vt)-1B(Vt)-1. (8.29) Proof Theorem II of Chapter 3, Section 1, implies that the curve C(t) defined by Equation (8.29) is an analytic curve in p. With s = д/t, A(s) = 1 + sa + |s2a' + ... and B(s) = 1 + sb + |s2b' + ..., where a' = (d2A/dt2)t=o and b' = (d2B/dt2)t=o. Then, to second order, A(s)-1 = 1 — sa+s2(a2 — |a')+... and B(s)-1 = 1 — sb + s2(b2 — |b') + .... Thus, after some algebra, C(f) = 1 + s2[a, b] + ..., so that (dC/dt)t=o = [a, b].
THE ROLE OF LIE ALGEBRAS 147 This leads immediately to the fundamental theorem: Theorem III For every linear Lie group Q there exists a corresponding real Lie algebra C of the same dimension. More precisely, if Q has dimension n then the m x m matrices ai,a2,... , an defined by Equation (8.28) form a basis for £. Proof All that has to be shown is that if a and b are any two members of V, the real vector space with basis ai,a2,...,an of Equation (8.28), then so is [a, b]. However, by Theorem I above, a and b are tangent vectors to some analytic curves A(t) and B(t) in Q. Then, by Theorem II, [a, b] is the tangent vector of the analytic curve C(t) of Equation (8.29), so [a, b] must be a member of V. Having shown that the vector space V with basis ai, a2,..., an is actually a real Lie algebra, henceforth V will be denoted by C (as in the statement of Theorem III above). In the mathematical physics literature ai, a2,..., an are often referred to as the “generators” of the Lie algebra C. While the above construction of C depends explicitly on the parametrization of P, it can be shown that a different parametrization merely produces a Lie algebra that is isomorphic to C (in the sense of Chapter 9, Section 3). That is, the real Lie algebra corresponding to a linear Lie group is essentially unique. This will become very clear in many cases of interest after the role of the one-parameter subgroups is developed in subsection (b). Henceforth the convention will be adopted that for the linear Lie groups SU(AT), U(AT), SO (AT) and so on, the corresponding real Lie algebras are denoted by su(AT), u(AT), so (AT) and so on. Example I The real Lie algebra C = su(2) of the linear Lie group Q = SU(2). It follows from Example III of Chapter 3, Section 1, and from Equation (8.28) that the generators of C = su(2) are 1 Г 0 i ai - 2 г 0 1 Г oi a2 “ 2 -10 _1 [ i O’ ai - 2 0 —i ' (8.30) so that, by direct calculation, the basic commutation relations are [ai,a2] = —a3, [a2,a3] = -аъ [a3,ax] = -a2. (8.31) It will be observed that ai, a2 and аз are all traceless anti-Hermitian matrices, so C is the set of all 2 x 2 traceless anti-Hermitian matrices. (This result will be derived more directly in Example II below.) This example also demonstrates that the matrices of a real Lie algebra need not themselves be real, for clearly ai and аз are not real. The reality condition of a real Lie algebra C requires only that the elements of C be real linear combinations of ai, a2,..., an.
148 GROUP THEORY IN PHYSICS Theorem III has the following converse: Theorem IV Every real Lie algebra is isomorphic to the real Lie algebra of some linear Lie group. Proof See Freudenthal and de Vries (1969). (b) The relationship of the real Lie algebra £ to the one- parameter subgroups of Q Theorem V Every element a of the real Lie algebra £ of a linear Lie group Q is associated with a one-parameter subgroup of Q defined by A(t) = exp(ta) for —oo < t < oo. Proof See, for example, Appendix E, Section 2, of Cornwell (1984). Clearly all elements of C of the form Aa, where A ranges over all real values but a is fixed, give the same one-parameter subgroup of Q. Theorem VI Every element of a linear Lie group Q in some small neigh- bourhood of its identity belongs to some one-parameter subgroup of p. That is, every such element of Q can be obtained by exponentiating some element of the corresponding real Lie algebra. Proof For any set of n real numbers (я^, x2,..., a^), define the m x m matrix by . -,£„)= exp-f^iai + Ж2а2 + • • • +4an}, (8.32) ai,a2,...,an being the basis of £. That А(ж'1,is guaranteed by Theorem V. Consequently the original parameters aq, , xn can be expressed as analytic functions of x^, x2,..., x’n1 with xi = X2 = ... = xn=Q corresponding to x^ = x2 = ... = x'n = 0. As the Jacobian (dxj/dx'jf) is non-zero at x± — x^ = ... = xn = 0, this is a one-to-one mapping between small neighbourhoods of the two origins. It follows that every element of Q in some small neighbourhood can be expressed in the form of Equation (8.32). It is worth noting that as the set of coordinates x2l... ,xfn) of Equa- tion (8.32) satisfies all the conditions of Chapter 3, Section 1, it provides an alternative to the original set x<^ ..., xn, and because it is necessarily simply related to C it is called a set of “canonical coordinates” for Q. There remains the question of whether this result extends to the whole of the connected subgroup of Q. The next theorem shows that this is so if Q is
THE ROLE OF LIE ALGEBRAS 149 compact, but it is possible to construct examples that demonstrate that this need not be so if Q is non-compact. (See, for example, Example III of Chapter 10, Section 5, of Cornwell (1984)). Theorem VII If Q is a compact linear Lie group, every element of the connected subgroup of Q can be expressed in the form exp a for some element a of the corresponding real Lie algebra £. In particular, if Q is connected and compact, every element of Q has the form exp a for some a e C. Proof See Price (1977) or Dynkin and Oniscik (1955). Even for compact connected Lie groups this mapping need not be one- to-one, for it is possible that exp a = expb with a b. For example, for Q — SO(3) Equation (8.3) shows that exp(^ai) = exp{($ + 27rn)ai} for n = ±1,±2,.... The exponential mapping provides a direct way of determining the real Lie algebras corresponding to a number of important linear Lie groups that does not require an explicit parametrization. The following example illustrates the method. Example II The real Lie algebra C = for Q = SU(7V) for N >2. Let exp(ta) be any one-parameter subgroup of Q = SU(7V), so that a is some N x N matrix. As exp(ta) is required to be unitary, parts (c) and (g) of Theorem III of Section 2 imply that a+ = -a. (8.33) Moreover, as it is required that det (exp(ta)) = 1 for all real t, part (f) of that theorem shows that exp(tr (ta)) = 1 for all real t. Clearly this is only possible if tra = 0. (8.34) Thus C = su(JV) is the set of all traceless anti-Hermitian N x N matrices. The dimension n of C (and hence of Q) can be calculated as follows. Equa- tion (8.33) implies that the diagonal elements of a must all be purely imag- inary. Taking Equation (8.34) into account, the set of diagonal elements is specified by N — 1 real parameters. (For example, ia11)ia221 • • • ,iajv-i,2V-i may be taken to have arbitrary real values, but ад/w = — ajj-) Simi- larly, Equation (8.33) implies that the “lower” off-diagonal elements of a (i.e. the ajk with j > k) are completely specified by the corresponding “upper” off-diagonal elements, as a^j = —ajk*- There are |(V2 — N) upper off- diagonal elements, and as each has an independent real and imaginary part, the set of all off-diagonal elements is specified by 2.| (TV2 — N)(= (V2 — N) real parameters. Thus a (e C) requires (V2 — N) + (N — 1)(= (V2 — 1)) real parameters, or, put another way, there exist N2 — 1 linearly independent traceless anti-Hermitian N x N matrices. Hence n = N2 — 1. In particular, n = 3 for Q = SU(2), and n = 8 for Q = SU(3).
150 GROUP THEORY IN PHYSICS Q Conditions on A 6 Q £, Conditions on a 6 C n GL(A, <D) - gl(A, <D) - 2N2 GL(A, IR) A real gl(A,IR) a real N2 SL(A,C) det A = 1 sl(A, C) tr a = 0 2№-2 SL(A, JR) ( A real, [ det A = 1 sl(A,IR) ( a real, [ tr a = 0 N2 - 1 U(AQ At = A"1 u(A) a'*' = —a N2 SU(JV) ( At = A-1, [ det A = 1 su(A) f — —a, [ tr a = 0 N2 - 1 U(p, q) Atg = gA-1 u(p, Q) a^g = -ga N2 SU(p,<z) f Atg = gA-1, f det A = 1 su(p, q) ( a+g = -ga, [ tr a = 0 N2 - 1 O(N, C) A = A"1 so(A, C) a = —a N2-N SO(A, <D) ( A = A-1, [ det A = 1 so(A, C) a = —a N2 -N O(7V) Г A = A-1, [ A real Г A = A-1, so(A) f a — —a, { a real f a = —a, { a real ±(N2-N) SO(JV) < A real, ( det A = 1 so(A) ±(N2-N) O(p, q) f Ag = gA~1, [ A real so(p, g) ( ag = -ga, { a real ±(N2-N) ( Ag = gA, f ag = -ga, 1 a real SO(p.q) < A-1 real, 1 det A = 1 so(p, q) |(№-a) SO* (A) 7 и и so* (A) f a = —a, t at J = —Ja |(№-A) Sp(f.C) AJA = J sP(f,C) aJ = —Ja N2 +N Sp(f,IR) ( AJA = J, [ A real sp(f,IR) ( a J = —Ja, { a real |(№ + a) sp(f) J AJA = J, | At = A'1 sp(f) ( aJ = —Ja, I a^ = —a |(№+A) Sp(r, s) f AJA = J, | At GA = G sp(r, fi) ( a J = —Ja, t atG = -Ga |(№ + A) SU*(A) Г JA* = AJ, [ det A = 1 su*(A) ( Ja* = aJ, [ tr a = 0 № -1 Table 8.1: The real Lie algebras C of some important linear Lie groups Q. A and a are N x N matrices, which are complex unless otherwise stated; g is an TV x TV diagonal matrix with p diagonal elements +1 and q N — p) diagonal elements — 1, p > q > 1. In the last six entries N is even, and J and G are the TV x TV matrices defined in Equations (8.35) and (8.36).
THE ROLE OF LIE ALGEBRAS 151 Table 8.1 lists the details of the real Lie algebras belonging to a number of important linear Lie groups that can be obtained this way. In Table 8.1 J and G are the N x N matrices defined by J = о I TV/2 -ljV/2 0 (8.35) and -lr 0 0 0 /-1 _ 0 1S 0 0 Z V G “ 0 0 -lr 0 ’ 36) 0 0 0 ls where 1 < r < and s = — r. That the exponential mapping remains invaluable even for non-compact linear Lie groups is demonstrated by the following theorem. Theorem VIII Every element of the connected subgroup of any linear Lie group Q can be expressed as a finite product of exponentials of its real Lie algebra C. Proof See, for example, Appendix E, Section 2, of Cornwell (1984). These results may be summarized by the statement that the matrix ex- ponential function always provides a mapping of C into Q. This is onto if Q connected and compact, and even when Q is connected but non-compact every element of Q is expressible as a finite product of exponentials of members of £.

Chapter 9 The Relationships between Lie Groups and Lie Algebras Explored 1 Introduction This chapter is concerned primarily with introducing for Lie algebras a num- ber of concepts that were defined in previous chapters for groups, and then investigating in detail the relationships between these concepts for a linear Lie group and its corresponding real Lie algebra. For the most part these relation- ships are very straightforward, but in some instances there are complications in which the global aspects of the Lie groups make themselves apparent. Section 2 introduces the idea of a subalgebra and Section 3 examines isomorphic and homomorphic mappings for Lie algebras. This enables the basic ideas of representation theory for Lie algebras to be developed in Section 4. Section 5 is devoted to the study of the so-called “adjoint” representations, which are defined for both Lie algebras and Lie groups and which play an important role in later chapters. The chapter is concluded in Section 6 with an investigation of direct sums of Lie algebras and their representations. 2 Subalgebras of Lie algebras The definitions that follow apply equally to real or complex Lie algebras. Definition Subalgebra of a Lie algebra A “subalgebra” CJ of a Lie algebra £ is a subset of elements of C that them- selves form a Lie algebra with the same commutator and field as that of C. This implies that CJ is real if C is real and CJ is complex if C is complex.
154 GROUP THEORY IN PHYSICS £J is said to be a “proper” subalgebra of C if at least one element of C is not contained in Cf. In this case the dimension of is smaller than that of C. (Throughout this book the convention will be adopted that every Lie algebra and subalgebra has dimension greater than zero.) Definition Invariant subalgebra of a Lie algebra A subalgebra £' of a Lie algebra C is said to be “invariant” if [a, b] E CJ for all a E CJ and b E C. An alternative name for an invariant subalgebra is an “ideal”. In the special case of a real Lie algebra of matrices the following theorems show that there is an intimate connection between these concepts and the corresponding concepts for linear Lie groups. Theorem I If Q and Q' are linear Lie groups, £ and CJ are their corre- sponding real Lie algebras and P' is a subgroup of P, then £! is a subalgebra of £. Moreover, if P' is an invariant subgroup of P, then CJ is an invariant subalgebra of Proof See, for example, Appendix E, Section 3, of Cornwell (1984). Theorem II Let C be the real Lie algebra corresponding to a linear Lie group Q. Then each subalgebra of C is the Lie algebra of exactly one connected Lie subgroup of Q. Proof See Helgason (1962, 1978). This section will be concluded with a useful result on invariant subalgebras. Let £' and £" be two subspaces of a Lie algebra £ and let [£', £"] denote the subset of £ that consists of all linear combinations (with coefficients in the same field as that of £) of elements of the form [a', a"], where a' E £! and a" E £". Theorem III If £' and £" are invariant subalgebras of £, then [£',£"] is also an invariant subalgebra of £ (or it is the “trivial” set consisting only of the zero element 0 of £). Proof See, for example, Appendix E, Section 3, of Cornwell (1984). 3 Homomorphic and isomorphic mappings of Lie algebras The following definitions apply equally to real and complex Lie algebras, the “field” being the set of all real numbers in the first case and the set of all complex numbers in the second. The concepts are in essence the same for
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 155 Lie algebras as for groups, that is, homomorphisms are structure-preserving mappings and isomorphisms are homomorphisms that are one-to-one. Definition Homomorphic mapping of a Lie algebra C onto a Lie algebra CJ Let J) be a mapping of a Lie algebra C onto a Lie algebra CJ having the same field such that (i) for all a, b e C and all a, (3 of the field ф(аа + (3b) = + (Зф(Ь); (9.1) and (ii) for all a, b E C i>]) = [V’(a), V’(i’)]- (9-2) Then ф is said to be a “homomorphic” mapping of C onto Cf. Definition Isomorphic mapping of a Lie algebra C onto a Lie algebra CJ A mapping J) of a Lie algebra C onto a Lie algebra Cr with the same field is said to be “isomorphic” if it is both homomorphic and one-to-one. In the special case in which Cf is identical to C (so that is a mapping of C onto itself) an isomorphic mapping is called an “automorphism”. With the product of two automorphisms ф and ф of C defined by (фф(а) = ф(ф(а)) (for all a E C), the set of all automorphisms of C forms a group, which will be denoted by Aut(£). Theorem I Suppose that there exists a homomorphic mapping ф of a Lie algebra C onto a Lie algebra CJ with the same field. Then C and CJ have the same dimension if and only if ф is an isomorphic mapping. Proof Suppose that «i,a2,... ,an is a basis of C and consider the equa- tion а7^(а7') = 0? where ai, a2> • • •, are members of the field of C (and CJ). By virtue of Equation (9.1), this equation can be rewritten as V’EjLi ajaj) — 0- If Ф is one-to-one, this implies that Y^j=i ajaj = 0? which, because of the linear independence of a±, a2,..., an, has only the so- lution on = a2 = ... = an = 0. Consequently ф(аф), ^(n2), • •., Ф(^п) must also be linearly independent. Conversely, if ф(аф), ^(n2), • • •, Ф(ап) are linearly independent then the only solution of ф(а) = 0 is a = 0. (This follows as any such a can be written as a = ajaL so ^(a)(= = 0 implies Qi = a2 = ... = an = 0 and hence a = 0). Thus ф is one-to-one (for if ф(Ь) = ф(Ь'), then ф(Ь — bf) = 0 and hence b — bf = 0, implying that b = bf). In order to directly connect these notions with those for linear Lie groups, it is necessary to modify the definition of homomorphic mappings for linear Lie groups.
156 GROUP THEORY IN PHYSICS Definition Analytic homomorphism of a linear Lie group Q onto a linear Lie group Let ф be a mapping of a linear Lie group Q onto a linear Lie group P' such that (i) ф is a homomorphic mapping in the “abstract” group sense of the defi- nition of Chapter 2, Section 6; and, in addition, (ii) for every A(#i, #2, • • •, яп) £ Q in some small neighbourhood of the iden- tity of Q, ф(А(х1,Х2,... , rrn)) is a matrix whose elements are analytic functions of Ж1, X2,..., xn (in the sense of Chapter 3, Section 1). Then ф is said to be an “analytic” homomorphism of Q onto Qf. (Strictly speaking, the above definition requires only that ф be analytic at the identity of Q. However, it can be shown (Sagle and Walde 1973) that such a homomorphism is necessarily analytic on the whole of p. Consequently nothing is lost in calling such a homomorphism “analytic” without further qualification.) In a similar way it is possible to define a “continuous homomorphism” of a connected linear Lie group Q onto a connected linear Lie group Q’ as a mapping that is a homomorphism in the abstract group sense and is continuous in the sense that the elements of the matrix <^(A(?/i, y2,..., ynf) are continuous functions of the parameters y\, y2,..., yn of the connected subgroup of Q (see Chapter 3, Section 2). It is a very remarkable fact that if a homomorphic mapping is continuous then it is certainly analytic. (For a proof see, for example, Sagle and Walde (1973).) Definition Analytic isomorphism of a linear Lie group Q onto a linear Lie group Q' This is simply an analytic homomorphism that is also one-to-one. Theorem II If Q and P' are two linear Lie groups and £ and CJ are their corresponding real Lie algebras, and if ф is an analytic homomorphic mapping of Q onto P', then the mapping ф of C into CJ defined for each a e C by V’(a) = (9-3) is a homomorphic mapping of C onto . Moreover, for all a e C and — oo < t < oo ехр{й/>(а)} = </>(exp(fa)). (9.4) Proof See, for example, Appendix E, Section 4, of Cornwell (1984). This theorem has an obvious generalization to isomorphic mappings. Theorem III If Q and Q' are two linear Lie groups and C and Cf are their corresponding real Lie algebras, and if ф is an analytic isomorphic mapping
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 157 of Q onto P', then the mapping ф oi C onto £,' defined by Equation (9.3) is an isomorphism. Proof All that has to be shown is that ф is one-to-one. Suppose that ^(a) = ^(b). Then, by Equation (9.4), </>(exp(fa)) = </>(exp(fb)) for all real t. If ф is an isomorphism this implies that exp(^a) = exp(^b) for all real t, so that a = b. Example I The analytic isomorphic mapping o/U(l) onto SO(2) and the associated isomorphic mapping of their real Lie algebras Consider the mapping ф of Q = U(l) onto P' = SO(2), defined for all real aq by 0([exp(ia:i)]) = cosrri sinaq — sinrri cosaq It is easily shown that this is an analytic isomorphism. With the basis of C = u(l) taken to be the 1x1 matrix ai = [г], by Equation (9.3) = (s cos£ — sin£ sin£ cost t=o 0 1 -1 0 (9-5) As C = u(l) and Cf = so(2) are one-dimensional Lie algebras, they are nec- essarily Abelian and hence they must be isomorphic. Much more remarkable is the next theorem, for which a preliminary defi- nition is required. Definition Discrete subgroup of a linear Lie group A subgroup /С of a linear Lie group Q is said to be “discrete” if either (i) /С is a finite group; or (ii) /С has a countable infinity of elements, but there exists a small neigh- bourhood of the identity of Q that contains no element of /С (apart from the identity of Q itself). Theorem IV If the kernel /С of an analytic homomorphic mapping ф of a linear Lie group Q onto a linear Lie group Q’ is discrete, then the corresponding mapping ф (defined in Equation (9.3)) of the real Lie algebra C of Q onto the real Lie algebra CJ of Qf is an isomorphic mapping. Proof If /С is discrete there exists a small sphere S centred on the identity of Q such that the only element in S that maps under ф into the identity of Q' is the identity of Q. Consequently ф provides a one-to-one mapping of S onto a small neighbourhood of the identity of Q'. Thus Q and Q' must have the same dimension, which implies that C and CJ have the same dimension. Theorems I and II of this section combine to show then that ф must be an isomorphic mapping.
158 GROUP THEORY IN PHYSICS This result is of the greatest significance, as it shows that in general the structure of a linear Lie group is not completely determined by its corre- sponding real Lie algebra, for two (or more) linear Lie groups that are not isomorphic can have isomorphic real Lie algebras. The terminology used to describe this situation is that the real Lie algebra does not determine the structure of its corresponding Lie groups “globally”, but only “locally”. (In fact it can be shown that all the connected linear Lie groups having isomorphic real Lie algebras can be determined (at least in principle) from the “universal covering group” of the Lie algebra. Moreover, all semi-simple connected lin- ear Lie groups with isomorphic Lie algebras can be constructed quite easily from the “universal linear group” of the Lie algebra, this group being itself a semi-simple connected linear Lie group. For details, see, for example, Chapter 11, Section 7, of Cornwell (1984)). The following examples demonstrate explicitly the conclusions of the above theorem. The second is particularly important in the quantum theory of angular momentum. Example II The analytic homomorphic mapping of the multiplicative group of positive real numbers IR+ onto the group SO(2) and the associated isomor- phic mappings of their real Lie algebras. Let Q be the multiplicative group of positive real numbers IR+ and Q’ the group SO(2). Each element of Q can be considered as the component of a 1x1 matrix. As shown in Examples I and II of Chapter 3, Section 1, every element of Q then has the form [exp^i], where x± can take any real value, and every element of P' can be written in the form of Equation (3.7). It is easily verified that the mapping ф defined by </>([expa?i]) = COS£E1 — sinrri sinrri COS£E1 (9-6) is an analytic homomorphism of Q onto Q1. Clearly its kernel /С is the set [exprri] with x± = 2n:k, к = 0, ±1,±2,.... Although countably infinite, /С satisfies the condition of Theorem IV, for in any neighbourhood — e < x± < e with 0 < 6 < 2тг the only point corresponding to an element of /С is x± = 0, which corresponds to the identity. It is obvious in this case that Q and Q' are both one-dimensional, so that £ and £J are also both one-dimensional and hence necessarily isomorphic. With the basis of £ taken to be the 1x1 matrix ai = [1], by Equations (9.3) and (9.6), V'(ai) = d cost dt - sin t sin£ cos£ t=o 0 1 -1 0 It should be noted that, although Q and P' have isomorphic real Lie algebras, Q' is compact but Q is non-compact (see Examples I and II of Chapter 3, Section 3).
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 159 Example III The analytic homomorphic mapping o/SU(2) onto S0(3) and the associated isomorphic mapping of their real Lie algebras Let Q = SU(2) and Q' = S0(3). It was shown in Chapter 2, Section 6, that there exists a homomorphic mapping ф (in the abstract group sense) of SU(2) onto SO(3) defined by <Ku)jk = (l/2)tr{crjucrfcu_1} (9.7) for all u e SU(2) and j,k = 1,2,3 (see Equations (2.9) and (2.10)). It is easily verified that ф is also analytic. Moreover, the kernel /С is finite (consisting only of I2 and —12), so that the mapping ф of Equation (9.3) must be an isomorphic mapping. Indeed both SU(2) and SO(3) are of dimension 3. Moreover, Equations (9.3) and (9.7) imply that for any a e £(=su(2)) = (l/2)tr{o-y [a, er*]}. Consequently, with the basis а1,а2,аз of £ (=su(2)) defined by Equations (8.30), as a? = (7 = 1,2,3) and 3 y 2iepqr<Tr (where epqr is defined by Equations (8.21)), it follows that ^(ap)j/c = epjk- Thus " 0 0 0 0 0 -1 0 1 0 V>(ai) = 0 0 1 , v>(a2) = 0 0 0 , V’(as) = -1 0 0 0 -1 0 1 0 0 0 0 0 These are precisely the basis elements of £' = so(3) chosen in Equations (8.12). The isomorphic nature of ф is confirmed by the fact that the commu- tation relations of so(3) (Equations (8.20)) and of su(2) (Equations (8.31)) are identical. In this case both Q = SU(2) and Q' = SO(3) are compact. This point will be taken up again in Chapter 11, Section 10. The only remaining case of interest is that in which /С itself is a linear Lie group. Theorem V Let /С be the kernel of an analytic homomorphic mapping ф of a linear Lie group Q of dimension n onto a linear Lie group Q1 of dimension n', and suppose that /С is also a linear Lie group. Then /С has dimension (n — n') and its corresponding real Lie algebra is an invariant subalgebra of the real Lie algebra £ of Q. Proof Let £" be the real Lie algebra of /С. Theorem I of Section 2 implies that £" is an invariant subalgebra of £ because /С is an invariant subgroup
160 GROUP THEORY IN PHYSICS of Q. As ф maps the whole of /С onto the identity of P', Equations (9.3) and (9.4) imply that a matrix a of £ is a member of £" if and only if ^(a) = 0. Let ai, a2,..., an be a basis for C chosen so that ai, a2,..., ап„ is a basis for £," (so that £," has dimension n"). Then ^(aj) = 0 for J = 1,2,..., n", while ^(aj) for j = nH + 1, n" + 2,..., n are linearly independent and so form a basis for £', the real Lie algebra of Q1. (For if an"+i,..., an are real numbers such that аз^(aj) = °, then ^(E7=n"+i aJaj) = °’ implying that 52j=n//+1 £ C" and hence an//+i = ... = an = 0.) Consequently nf = n — n". Hitherto the term “Lie subgroup” Q' of a linear Lie group Q has been used to describe a subset of Q that is itself a linear Lie group (see Section 2 and Chapter 3, Section 1). Now suppose that Q" is a linear Lie group that is analytically isomorphic to P'. Although, strictly, Q” is not a subgroup of Q (as the elements of Q” need not be members of P), in practice very little confusion can arise if such a Q" is described as being a Lie subgroup of Q. Indeed this looser usage is frequently encountered in the literature. Example IV SU(2) regarded as a Lie subgroup o/SU(3) Suppose u e SU(2) and let 0(u) = u 0 0 1 where the matrix on the right-hand side is a member of SU(3). Let Q' denote the subset of elements of SU(3) that have this form. It is easily established that P' is a Lie subgroup of Q in the strict sense and that ф is an analytic isomorphism of SU(2) onto Qf. Thus SU(2) can be regarded as being a Lie subgroup of SU(3). 4 Representations of Lie algebras The definition of a representation and much of the subsequent discussion apply equally to real and complex Lie algebras. Definition Representation of a Lie algebra C Suppose that to every a E C there exists a d x d matrix Г (a) such that (i) for all a, b E C and a, fl of the field of C T(m + flb) = aT(a) + /?Г(0, (9.8) and (ii) for all a, b E C Г([а,Ь]) = [Г(а),Г(Ь)].
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 161 Then these matrices are said to form a “d-dimensional representation” of £. Clearly the set of matrices Г (a) form a Lie algebra with the same field as C and there is a homomorphic mapping of C onto this Lie algebra. In specifying a representation of £, it is obviously sufficient to specify only the matrices T(aj) for every aj of the basis of Many of the ideas on representations discussed in Chapter 4 for groups apply equally to Lie algebras. In particular, the duality between matrix rep- resentations and modules, the existence of similarity transformations and the concepts of reducible, completely reducible and irreducible representations re-appear. Consider first the duality with modules (see Chapter 4, Section 1). If Г is a d-dimensional representation of a Lie algebra £, and ^2? • • • ? form the basis of a d-dimensional abstract complex inner product space V, the “carrier space”, then for each a E C an operator Ф(а) may be defined (cf. Equation (4-1)) by d = Уj = l,2,...,d. (9.9) k=i With the further definition that d d i=i J=1 for any set of complex numbers iq, 62, • • •, it follows immediately that the set of operators Ф(а) form a Lie algebra and there is a homomorphic mapping of C onto this Lie algebra. Moreover, if the basis ^1, ^2, • • •, °f is chosen to be an ortho-normal set, then (cf. Equation (4.4)) T(a)kj = (^,Ф(а)^) for any a E C. Again the set of operators Ф(а) and the carrier space V are said to collectively form a “module”. Similarly, if Г is a d-dimensional representation of C and S is any d x d non-singular matrix, then the set of matrices Г'(а), defined for all a E C by Г'(а) = S^r^S, also form a d-dimensional representation of £,. Again Г and Г' are said to be “equivalent” representations (see Chapter 4, Section 2). Likewise, a d-dimensional representation of C is said to be “reducible” if it is equivalent to a representation Г of £ that can be partitioned in the form г<“) = [ Г,о(“> Йм] <9Л“> for every a E £, the dimensions of the submatrices being as in Equation (4.8). A representation of C is then defined to be “irreducible” if it is not
162 GROUP THEORY IN PHYSICS reducible. Finally, a representation of C is said to be “completely reducible” it is equivalent to a representation Г" of C that has the form Г Г'ЛМ 0 0 0 0 0 0 Г"(а) = 0 0 0 (9.U) 0 0 0 .. Г'» _ for every a E £,, where Г'^, Г^, • •. are all irreducible representations of C. Because of the linear property of Equation (9.8), it is sufficient to consider only the matrices representing the basis elements of C when checking for reducibility or complete reducibility. The name “Schur’s Lemma” is also attached to one (or both) of the two following theorems. Theorem I Let Г and Г' be two irreducible representations of a Lie algebra C of dimensions d and df respectively, and suppose that there exists a d x df matrix A such that Г(а)А = АГ'(a) for all a e C. Then either A = 0, or d = d' and det A 0. Proof The proof given, for example, in Appendix C, Section 3, of Cornwell (1984) applies with the group Q replaced by the Lie algebra £. (Indeed the proof is applicable to any set of linear operators.) Theorem II If Г is a d-dimensional irreducible representation of a Lie al- gebra £, and В is a d x d matrix such that Г(а)В = ВГ(а) for every a e £, then В must be a multiple of the unit matrix. Proof This is exactly as for Theorem II of Chapter 4, Section 5, but with Q replaced by As for groups, there is an immediate corollary for the Abelian situation. Theorem III Every irreducible representation of an Abelian Lie algebra is one-dimensional. Proof This is exactly as for Theorem III of Chapter 4, Section 5, but with Q replaced by £. Of particular importance is the connection between the representations of a linear Lie group and those of its corresponding real Lie algebra. The transition from group to algebra is completely straightforward, as the following theorem shows, but the procedure does not necessarily work in the reverse direction. Definition Analytic representation of a linear Lie group Q Let be a representation of Q in the abstract group sense (see Chapter 1,
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 163 Section 4) such that, for every А(ж1, , xn) e Q in some small neighbour- hood of the identity of P, the elements of Г^(А(ж1, x?^..., xn)) are analytic functions of xi, X2, • • •, xn (in the sense of Chapter 3, Section 1). Then is said to be an “analytic representation” of Q. Theorem IV Let be a d-dimensional analytic representation of a linear Lie group P, whose corresponding real Lie algebra is £. (a) Then there exists a d-dimensional representation Г*£ of C defined for each a E C by Г£(а) = [£re(exp(ta))]f=0 (9.12) (b) For all a E C and all real t exp{tr£(a)} = r^(exp(ta)). (9.13) (c) If and Гд are two d-dimensional analytic representations of Q and Г*£ and Г2 are the associated representations of C defined as in Equation (9.12), then 1?£ is equivalent to Г'с if is equivalent to Гд . The converse is also true if Q is connected. (d) 1?£ is reducible if is reducible, and Г*£ is completely reducible if is completely reducible. The converses are also true if Q is connected. (e) If Q is connected then Г*£ is irreducible if is irreducible. Conversely, is irreducible if Г*£ is irreducible. (f) If is a unitary representation, then Г^(а) is anti-Hermitian for all a E C. The converse is also true if Q is connected. Proof (a) and (b) The proofs are essentially the same as those of the corre- sponding parts of Theorem II of Section 3. (c) Suppose that Г^(А) = S-1r^(A)S for all A E Q. Then from Equations (9.12) and (9.13) and part (d) of Theorem III of Chapter 8, Section 2 rr(a) = [^{S-1r6(exp(ia))S}]t=0 = [i{s“lexP{ir-c(a)}s}]t=o = [^{exp(iS'-1r£(a)S)}]t=o = S-1r£(a)S. Conversely, if Г^(а) = S-1r£(a)S, then Equation (9.13) implies that r^(exp(ta)) = S-1r^(exp(ta))S. Theorem VIII of Chapter 8, Section 5, then extends this result to the whole of Q if Q is connected. (d) Equation (9.12) implies that Г*£ is reducible if is reducible, and Г*£
164 GROUP THEORY IN PHYSICS completely reducible if is completely reducible. Conversely, as every power of a matrix of the form in Equation (9.10) also has this form, so too has the exponential of such a matrix. The same is true of a matrix of the form in Equation (9.11). The converse results then follow from Equation (9.13) and Theorem VIII of Chapter 8, Section 5, provided that Q is connected. (e) This is an immediate consequence of (d). (f) This follows from Equation (9.13), with Theorem VIII of Chapter 8, Sec- tion 5 being invoked in the converse proposition. It is most important to realize that this theorem (and in particular Equa- tion (9.13)) does not imply that every representation Г*£ of C gives a represen- tation of Q by exponentiation. Rather, Equation (9.13) merely shows that if is an analytic representation of Q then can be obtained from by ex- ponentiation. The essential point is that, although the matrices ехр{£Г£(а)} are well defined for all a E C and all real t, they do not necessarily form a representation of Q. The following examples will demonstrate this explicitly. Example I Connection between the representations of C = so(2) and Q = SO(2) C = so(2) is one-dimensional. Let ' 0 1 ’ ai - [ -1 0 be its basis element (see Example II of Chapter 3, Section 1). As the only relevant commutation relation is the trivial one [ai,ai] = 0, it follows that Tr(ai) = [p] provides a one-dimensional representation of C for any complex number p. Then exp{tF£(ai)} = [exp(tp)], while the elements of Q have the form ч Г cos t sin t exp (tai) = . . . • 7 — sin t cos t However, exp((t + 2?r)ai) = exp(tai), but exp{(t + 2тг)Г£(а1)} = ехр(2тгр) exp{tr£(ai)}. Consequently this representation Г*£ of C gives a representation of Q = SO(2) by exponentiation if and only if p = iq, where q is some integer. Example II Connection between the representations of C = so(3) and Q = SO(3) Let the matrices ai, a2 and аз of Equations (8.12) provide a basis for C = so(3). As demonstrated in Example II, so(3) is isomorphic to the real Lie algebra su(2). Thus a two-dimensional representation of C = so(3) is provided by inverting the isomorphic mapping of Example II, giving, by Equations (8.30), г \ 1 Г 0 « Г£( 2 i 0 Гг(а2) = | 0 1 -1 0 г£(а1) = I i 0 0 —i ' (9-14)
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 165 Then while ехр{£Г£(а3)} = exp(|zt) 0 0 exp(—|zt) exp(ta3) = cos t sin t — sin t cos t 0 0 0 0 1 Thus ехр{(£+2тг)а3} = exp(ta3), but ехр{(£+2тг)Г£(а3)} = - exp{d?r(a3)}. Consequently the representation Г*£ of C = so(3) defined by Equation (9.13) does not on exponentiation give a representation of Q = SO(3). The question of which irreducible representations of C = so(3) do provide representations of Q = SO(3) will be examined further in Chapter 10, Section 4. Finally, there remains the question of the relationship between the analytic representations of a linear Lie group Q defined above and the continuous representations of Q introduced in Chapter 4, Section 1. Theorem V If Q is a compact linear Lie group then every continuous rep- resentation of Q is analytic, and vice versa. Proof See Naimark (1964). (The compactness of Q comes in because invari- ant integration of Q is necessary.) It is worth noting that the continuity assumption was used in Chapter 4 only in the proofs of theorems for compact Lie groups, that is, precisely for those groups for which the concepts of continuity and analyticity are equiva- lent. It is very useful to examine the connection between the representation of Q and C in terms of modules. Denoting quantities associated with Q and C by the appropriate subscripts, Equation (4.1) becomes (with T = exp(ta)) d Фб(ехр(£а))^ = ^^(exp^a))^*, (9.15) fe=i and Equation (9.9) becomes d Фг(а)^ = (9.16) fe=i Clearly Equations (9.12) and (9.13) imply that, for all a G £, Фг(а) = [^Ф<?(ехр(«а))]4=0 , (9.17) ехр{^Фс(а)} = Фе(ехр(<а)) (9.18) and
166 GROUP THEORY IN PHYSICS for all a e C and all real t. The linear operators Ф^(ехр(£а)) and Фг(а) all operate in the same vector space V, and Vh, ^2, • • •, V'd form a basis both of and of Г*£. Of particular significance is the representation of C that corresponds to the direct product Р0Г9 of Q. Adding subscripts Q to Equations (5.33) and (5.34) gives (for T = exp(ta)) Фе(ехр(4а)){^ ® f*} = {$g(exp(ia))V>J} 0 {Ф£(ехр(<а))^’} (9.19) and dp dq Фр(ехр(*а)){^ = ^^(rP(exp(ta))0(r’(exp(Za)))fc4JS{^0 0t<'}, k=lt=l (9.20) where dp $pg(exp(tay)rf = ^2rg(exp(ia))fcj^ (9.21) k=l and dq Фр(ехр(*а))^ = ^r^expfta))^-'. (9.22) t=l Thus, if and Г^. are the irreducible representations of C related to Гд and Tqg by Equations (9.12) and (9.13) (with superscripts p and q added), and Ф£(а) and Ф^(а) are the corresponding linear operators (related to Фд and ФЯд by Equations (9.18) with superscripts p and q added), so that dp Ф£(а)^р = £г£(а)^< (9.23) k=l and dq Ф’(а)^=£П(а)^, (9.24) t=l and if ф£ (a) are the linear operators corresponding to the Фд, then Equation (9.19) gives Фг(а){< ® = {Ф£(а)<} ® № + < ®(9-25) for all a e C. Then, from Equations (9.23) and (9.24), dp dq Ф£(а){^®^} = ££(Гр£(а)®Ц +4 ®r«(a))fcMs{<®V-?} (9.26) k=lt=l for all a e j = 1,2, ...,dp, and s = 1,2, ...,dg. Equation (9.26) shows that the representation of C corresponding to the direct product representation Гр 0 Гq of Q is given by the set of matrices r£(a)01dg+ldp®r’(a). (9.27)
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 167 Because the basis vectors of this representation are still the direct products of those of the two constituent representations, this representation of C will be called the “direct product” (or “Kronecker product”) of the representations Tpc and Г^, and will be denoted symbolically by Tpc 0 Г^, even though its explicit form is not that of a straight direct product. Further, as Equation (5.16) can be rewritten as с-1{Г<Иехр(£а)) ® T^(exp(ia))}C = ^фп^Г^ехр^а)), where C is the matrix of Clebsch-Gordan coefficients, it follows that С-Ч1ЭД ® ldg + 4 ® Г’(a)}C = ^©п^ГИа). Thus the irreducible representation Trc of C appears nrpq times in the direct product of Tpc and Tqc ifTg appears npq times in the direct product Г^0Г^ of P, and the reduction is performed using the same matrix of Clebsch-Gordan coefficients for both G and Consequently, with defined exactly as in Equation (5.35) by (9.28) Q к are the Clebsch-Gordan coefficients, then dr Ф£(а)^’“ = £Г2(а)и;^ U=1 (9.29) for all a e I = 1, 2,..., dr and a = 1, 2,..., nrpq, this being the Lie algebraic analogue of Equation (5.36). It is important to note that the Clebsch-Gordan coefficients for the Lie group Q and the real Lie algebra C are identical. These considerations can be generalized to any abstract Lie algebra £, real or complex. As C need not be directly related to a linear Lie group, all the relevant concepts will be defined without reference to any group, and so the subscripts C of the previous analysis may be omitted. If Фр(а) and Ф9(а) are the linear operators for a e C of the irreducible representations Гр and Г9 of £, defined by **(<< = ^(«,1 f93(n then the operators Ф(а) corresponding to the “direct product” representation Гр 0 Гд of C will be defined (by analogy with Equation (9.25)) by Ф(«){< 0 = {Фр(«)<} 0 < + < 0 {$*(«)<}, (9.31)
168 GROUP THEORY IN PHYSICS for all a e j = 1,2,..., dp, and s = 1,2,..., dq. It is easily verified that Ф([а, a']) = [Ф(а),Ф(а')] for all a, a' E C. Then, by Equations (9.30) and (9.31), dp dq ф(а){^₽ ® = £ £(Гр(а) 0 4 + 4 ® ® tf}, (9.32) k=lt=l so the matrices Г (a) defined by Г(а) = Гр(а) 0 ldq + ldp 0 П(а) (9.33) form a dpdg-dimensional representation of C that will be called the “direct product” representation and denoted by Гр 0P. If the irreducible represen- tation Гг of C appears nrpq times in Гр 0 Г9 then the basis vectors 9r^a can again be expressed in the form of Equation (9.28), thereby defining Clebsch- Gordan coefficients of C. As before, the matrix C of such coefficients reduces Гр 0 Г9 of C into its irreducible constituents. It is useful to express the definitions of irreducible tensor operators in Lie algebraic terms. For irreducible tensor operators of a linear Lie group Q of coordinate transformations in IR3, letting the transformation T of Equation (5.28) correspond to exp(ta), where a is an element of £, the real Lie algebra associated with P, and considering the limit as t —> 0 gives dq (P(a),Qj] = Er*(a)fcjQ’, (9.34) k=l which is valid for all a E C and j = 1,2,..., dq. Here the operators P(a) are as defined as in Equation (8.22). Similarly, for a general linear Lie group P, Equation (5.37) gives dq $r(a)QJ - Q^(a) = £r’(a)^ (9.35) k=l for all a of the corresponding real Lie algebra C and j = 1,2,..., dq. (Equation (9.35) can be taken as the definition of irreducible tensor operators for a general Lie algebra, modified if desired along the lines indicated in the last paragraph of Chapter 5, Section 4.) 5 The adjoint representations of Lie algebras and linear Lie groups First the adjoint representation ad will be defined for a Lie algebra. (In Chapter 11, it will be shown that this representation plays a key role in the analysis of semi-simple Lie algebras.) Then the adjoint representation Ad of
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 169 a linear Lie group will be introduced and the connection between ad and Ad discussed. Theorem I Let £ be a real or complex Lie algebra of dimension n and let «1, «2,..., an be a basis for £. For any a G £, let ad (a) be the n x n matrix defined by n [a,ay] = ^{ad(a)}fcjafc (9.36) k=i for j = 1,2,... ,n. Then the set of matrices ad(a) forms an n-dimensional representation of C called the “adjoint representation” of £. Proof See, for example, Chapter 11, Section 5, of Cornwell (1984). Equations (9.36) and (8.18) together imply that {ad(np)}/cj = cpj (9.37) for J,k,p = 1,2,... ,n, where cpj are the structure constants of This in turn implies that, if £ is a real Lie algebra, all the elements of ad (a) are real for all a e C. It is occasionally convenient to use the corresponding operators ad(n), defined for each a e C by ad(n)6 = [a, b] for all b e C. Then, by Equations (9.36), n ad(a)aj = ^2{ad(a)}fcjafc, k=i from which it follows by the preceding theorem that ad(An + pb) = Aad(n) + /zad(6) for all a, b e C and all real or complex numbers A, p (as appropriate) and ad([n, b]) = [ad(n), ad(6)] for all a, b e C. Clearly the vector space on which the operators ad (a) act is C itself The effect of taking a different basis .., a'n instead of eq, «2,..., an is merely to induce a similarity transformation on the matrices ad(a). Indeed, with ap = Sqpaq, ad (a) is replaced by S-1ad(n)S for each a e C. Both the definition of the adjoint representation Ad of a linear Lie group Q and certain results concerning the automorphisms of its corresponding real Lie algebra C depend on the following theorem.
170 GROUP THEORY IN PHYSICS Theorem II If Q is a linear Lie group and C is its corresponding real Lie algebra, then, for any A e Q and any b G £, Ab A-1 is a member of £. Moreover, with A = exp(ta), where ae£, {exp(ta)} b {exp(ta)} 1 b + t[a, b] + |t2[a, [a, b]] + |i*3[a> [a,b]]] + ... (9.38) Proof Consider the set of elements A{exp(sb)}A-1 of P, where A G P, b E £, and —oo < s < oo. They form a one-parameter subgroup of Q whose generator is AbA-1, so that Ab A-1 e C. Let F(t) = {exp(ta)} b {exp(ta)}-1. Then dF/dt = [a, F(t)], d2F/dt2 = [a, [a, F(t)]] and so on. As the elements of F(t) are obviously analytic functions of t, F(t) = F(O) + i(dF/dt)t=o + it2(d2F/dt2)t=o + ... = b + t[a, b] + it2[a, [a,b]] + ... Theorem III Let Q be a linear Lie group of dimension n and let ai, a2, ..., an be a basis of its corresponding real Lie algebra C. For each A E Q let Ad (A) be the n x n matrix defined by Aa7A 1 = J2{Ad(A)}fcjafc fe=i (9.39) for j = 1,2,..., n. Then (a) the set of matrices Ad (A) forms an n-dimensional analytic representa- tion of Q called the “adjoint representation” of P; and (b) the associated representation of C defined by Equation (9.12) is the adjoint representation of £; that is, ad(a) = Ad(exp(te))]t=0 for any a E £, so that for any a G C and all real t exp{tad(a)} = Ad(exp(ta)). (9.40) Proof See, for example, Chapter 11, Section 5, of Cornwell (1984). This is a convenient point to derive a useful result concerning the auto- morphisms of a real Lie algebra.
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 171 Theorem IV Let Q be a connected linear Lie group and £ its corresponding real Lie algebra. Then, for any A £ (7, the mapping ^a of £ onto itself, defined by V-A(b) = AbA1 (ii) (9.41) for all b e £, is an automorphism of £ and is called an “inner automorphism” of £. Proof Theorem II of this section shows that ^a is definitely a mapping of £ into itself. Moreover, for any b' E £ there exists a b E £ (namely b = A-1b'A) such that ^а(Ь) = b', and the mapping is obviously one-to-one. The linear requirement of Equation (9.1) is trivially satisfied, while for any b,b' e£ ^A([b, b']) = A(bb' - b'b) A-1 = AbA-1 Ab'A-1 - Ab'A-1 AbA-1 = [V'A(b), V’A(b')], so that Equation (9.2) is also satisfied. The set of all inner automorphisms of £ will be denoted by Int(£). As Int(£) is the connected component of Aut(£), the group of all automorphisms of £, it follows that Int(£) must be an invariant subgroup of Aut(£). 6 Direct sum of Lie algebras Again the basic definition applies both to real and complex Lie algebras. Definition Direct sum of two Lie algebras A Lie algebra £ is said to be the “direct sum” of two Lie algebras £± and £2 (all with the same field) if (i) the vector space of £ is the direct sum of the vector spaces of £1 and £2 (see Appendix B, Section 1), and (ii) for all a! E £\ and o" E £2, [«', a"] = 0. This is expressed by writing £ = £1 ©£2- The concept can be clarified by expressing it in terms of the basis elements of £1, £2 and £. Let ai, «2,..., an be a basis for £ constructed in such a way that ai, «2,..., ani is a basis for £± and ani+i, ащ+2, • ••,«п is a basis for £2- If [dp, aq\ = 0 for p = 1,2,..., ni and q = n± + 1, n± + 2,..., n, then £ = £1 Ф £2- Obviously the dimension of £ is the sum of the dimensions of £1 and £2. It is clear that £1 ф £2 can be constructed for any two Lie algebras £1 and £2 having the same field. The relationship of this concept for real Lie algebras to that of direct products of linear Lie groups is succinctly expressed in the following theorem.
172 GROUP THEORY IN PHYSICS Theorem I If and P2 are two linear Lie groups of dimensions ni and П2 respectively, then Pi 0 P2 is a linear Lie group of dimension (n± + П2). Moreover, if £1 and £2 are the real Lie algebras of Pi and P2, then the real Lie algebra of Pi 0 P2 is isomorphic to £1 ф £2. Proof See, for example, Chapter 11, Section 6, of Cornwell (1984). This theorem shows that there always exists a direct product group whose real Lie algebra is isomorphic to £1 ф £2. However, not every linear Lie group with real Lie algebra isomorphic to £1 ф £2 is a direct product group, as the following example shows. Example I P — U(2) and its direct sum real Lie algebra As noted in Table 8.1, the real Lie algebra £ = u(2) of P = U(2) is the set of all 2 x 2 anti-Hermitian matrices (which need not be traceless). Thus a convenient basis of u(2) is provided by the matrices а1,а2,аз of Equations (8.30), together with a4 = И2. Clearly [ap,a4] = 0 for p = 1,2,3, so £ = u(2) is the direct sum of a real Lie algebra £1, with basis ai, a2 and аз, and a real Lie algebra £2 with basis a4. Let Pi and P2 be the linear Lie groups obtained by exponentiating the matrices of £1 and £2. As £1 = su(2), it follows that Pi = SU(2), while P2 is clearly the set of matrices exp(zt)12 for all real t. Although Pi and P2 are both subgroups of P = U(2) and the elements of Pi commute with those of P2, nevertheless P is not isomorphic to Pi 0 P2. The reason is that Pi and P2 possess two common elements, namely I2 and —12, so that the fourth necessary condition of Theorem II of Chapter 2, Section 7, is not satisfied. A very similar argument applied to U(AT) for N = 3,4,5,... shows that U(AT) is not isomorphic to SU(7V) 0 U(l), even though u(N)= ф u(l). The close connection with direct product groups means that all the results on representations of direct product groups discussed in Chapter 5, Section 5, have Lie algebraic analogues. The fact that representations of the direct product Pi 0 P2 of two linear Lie groups Pi and P2 are given by the direct products of the representations of Pi and P2, that £1 Ф £2 is the real Lie algebra of Pi 0 P2, and that the Lie algebraic version of a direct product representation is given by Expression (9.27) all combine to suggest the following theorems. Theorem II Let Г1 and Г2 be representations of dimensions di and (/2 of two Lie algebras £1 and £2 respectively (where £1 and £2 are either both real or both complex). Then the set of matrices defined Г(а' + a") = Г1(«') 0 ld2 + ldl 0 Г2(а") (9.42) for all a' e and a" E £2 provides a di ^-dimensional representation of £1 Ф £2-
RELATIONSHIPS BETWEEN LIE GROUPS AND LIE ALGEBRAS 173 Proof If a', b' e £1 and a", b" € £2> then Equation (9.42) gives [Г(а' + а"),Г(6' + Ь")] = [{Г^а') ® ld2 + ldl ® Г2(а")}, {ГЛУ) ® ld2 + ldl ® Г2(Ь")}] = [Г1(а'),Г1(У)] ® ld2 + ldl ® [Г2(а"),Г2(6")] = Г1([а',6']) ® ld2 + ldl ® [Г2([а",6"])] = Г([(а' + а"),(У + 6")]) Theorem III If £1 and £2 are the real Lie algebras of two compact lin- ear Lie groups and Г1 and Г2 are irreducible representations of £1 and £2 respectively, then the representation Г defined by Equation (9.42) is an irre- ducible representation of £1 ©£2. Moreover, every irreducible representation of £1 ф £2 is equivalent to a representation constructed in this way. Proof This follows immediately from the corresponding result for direct prod- ucts of compact Lie groups (given as Theorem II of Chapter 5, Section 5) and the results of Section 4.

Chapter 10 The Three-dimensional Rotation Groups 1 Some properties reviewed When the rotation groups in IR3 were first introduced in Chapter 1, Section 2(a), it was noted immediately that the group of all rotations in IR3 is isomor- phic to 0(3) and its subgroup of all proper rotations is isomorphic to SO(3). Example I of Chapter 2, Section 7, demonstrated that 0(3) is just the direct product of S0(3) with the group of order 2 consisting of the identity and the spatial inversion matrix. It was also noted (in Chapter 2, Section 6) that there is a homomorphic mapping of SU(2) onto SO(3), which is responsible for the isomorphism that exists between their corresponding real Lie algebras su(2) and so(3) (as was demonstrated in detail in Example III of Chapter 9, Section 3). The properties of SO(3), 0(3), SU(2) and the rotation groups of IR3 will now be investigated in detail. The conjugacy classes of these groups will be studied in Section 2. Then, in Section 3, the irreducible representations of the isomorphic Lie algebras su(2) and so(3) will be derived. These are of great importance not merely for their immediate physical applications but also be- cause they play an essential part in the analysis of the representations of any semi-simple Lie algebra. When the irreducible representations of the corre- sponding groups are examined in the next section, attention will be drawn to the fact that every irreducible representation of su(2) exponentiates to give an irreducible representation of SU(2), but this representation of SU(2) may or may not provide a representation of S0(3). The irreducible representa- tions of both S0(3) and 0(3) will be identified. The Clebsch-Gordan series and coefficients are discussed briefly in Section 5. Finally some applications in atomic physics are described in Section 6. The close connection with the theory of angular momentum in quantum mechanics appears as a constantly recurring theme.
176 GROUP THEORY IN PHYSICS 2 The class structures of SU(2) and SO(3) It is useful to have another parametrization of S0(3) and SU(2). Let T be the proper rotation through an angle cu in the right-hand screw sense about an axis that lies in the direction of the unit vector n = (ni, n2, П3). Then the transformation matrix R(T) is given by 3 R(T)jk = 6jk coscj + rijrik(l — coscj) + since ^2 ejki^i (10.1) 1=1 (for j, к = 1, 2, 3), where 6jki is the permutation symbol defined in Equations (8.21). It is convenient to adopt the convention that ce is allowed to take any value in the interval — 7г < се < 7Г, which implies that if n is an allowed unit vector then — n is not permitted. (For example, the possible values of the components of n could be restricted so that П3 > 0, or if П3 = 0 then n2 > 0, or if n3 = 0 and n2 = 0 then ni = 1. With this convention a negative value of ш corresponds to a rotation in the left-hand screw sense.) The corresponding element u of SU(2) is then given by u = 1 cos(cj/2) + г{п1<71 + n2cr2 + П3СГ3} sin(cu/2), (10.2) where <7i, <t2 and (T3 are the Pauli spin matrices of Equations (2.10). Here n is restricted as above, but cj will be allowed to take any value in the interval —2tt < cj < 2tt. (The two-valued nature of the homomorphic mapping of SU(2) onto SO(3) is reflected in the fact that cj and cj + 2тг correspond to the same proper rotation R(T) while giving matrices u with opposite signs.) The matrix u defined in Equation (10.2) will be said to “correspond to the rotation cj about the axis n (—2тг < cj < 2тг)”. (The proofs of Equations (10.1) and (10.2) may be found, for example, in Chapter 12, Section 2, of Cornwell (1984)). Rewriting Equation (10.1) in full, R(T) = cos о? + П|(1 — cos a?) nl n2 (1 — cos w) ~ n3 sin ш п1пз(1 — cos a?) + sina; Thus, as £^=1 «j = 1, nl n2 (1 — cos + n3 sin ш cos a? + 712(1 — cos a?) п2пз(1 — cos a?) — ni since? П1773(1 — COSCe?) — 772 since? 772773(1 — COSCe?) + 771 since? COSCe? + 77^(1 — COSCe?) (10.3) tr R(T) = 1 + 2coscj. (10-4) These results lead to the main conclusion of this section: Theorem I Two elements of SU(2) belong to the same class of SU(2) if and only if they correspond to rotations having the same value of |cu|. Similarly, two proper rotations in IR3 belong to the same class of the group of all proper rotations in IR3 if and only if they have the same value of |cu|. Finally, two rotations of the group of all rotations in IR3 are in the same class of that
THREE-DIMENSIONAL ROTATION GROUPS 177 group if and only if both rotations are proper or both are improper and the proper parts have the same value of |cu|. Proof For details, see, for example, Chapter 12, Section 2, of Cornwell (1984). It should be pointed out that this theorem applies to the group of all proper rotations in IR3 and to the group of all rotations in IR3. For proper subgroups of these groups it is possible to have two rotations of the same type (that is, both corresponding to the same value of |cu| and both proper or both improper) lying in different classes, as was discussed in detail in Chapter 2, Section 2. It follows from this theorem that, in all representations of SU(2) and of SO(3), the characters depend only on the value of |cu| and not on the direction n that specifies the axis of rotation. For n fixed the set of rotations T whose transformation matrices R(T) are given by Equation (10.3) clearly form a one-parameter subgroup, the parameter being cj. This subgroup is generated by the 3x3 matrix a, such that a = lim{R(T) — 1}/cj. Thus, from Equation (10.3), 0 n3 -n2 a = -n3 0 n2 -П1 so that a = mai + n2a2 + n2a2, where ai,a2,a3 are the basis elements of the Lie algebra so(3) defined by Equations (8.12). Thus niai+n2a2+n2a2 generates a one-parameter subgroup of proper rotations about the axis specified by the unit vector n = (ni,n2,n3). 3 Irreducible representations of the Lie alge- bras su(2) and so(3) The intimate connection between the isomorphic Lie algebras su(2) and so(3) and the algebra of quantum mechanical angular momentum operators was noted in Chapter 8, Section 4. It should come as no surprise that the argument that will now be given for determining the irreducible representations of su(2) and so(3) is essentially that which gives the eigenvalues and eigenvectors of the angular momentum operators and which appears in many books on quantum mechanics (e.g. Schiff 1968). As su(2) and so(3) are isomorphic they have the same representations. The following arguments will be given for su(2), but naturally they apply equally to so(3). As the basis elements in Equations (8.30) of su(2) are lin- early independent over the field of complex numbers, the complexification C
178 GROUP THEORY IN PHYSICS of su(2) can be taken to be the complex vector space consisting of complex linear combinations of the basis elements ai, a2 and аз of su(2), their basic commutation relations remaining as in Equations (8.31). (£ is denoted by Ai in the Cartan classification of simple complex Lie algebras given in Chapter 11.) Essentially the following argument produces the irreducible representa- tions of £, from which those of su(2) follow immediately as C and su(2) have the same basis elements. As the Lie group SU(2) is compact, each irreducible representation of SU(2) can be taken to be unitary. Part (f) of Theorem IV of Chapter 9, Section 4, shows that the corresponding representations of su(2) consist of anti-Hermitian matrices. Thus if 'fe • • •, 'Фа is an ortho-normal basis of such a d-dimensional representation Г of su(2) in an inner product space V and the linear operators Ф(а) are defined for all a of su(2) by d Ф(а)^р = '^уг(а)др‘фд, p = l,2,...,d, (10.5) <7=1 then, as (?/^, Ф(а)^р) = Г(а)др = —Г(а)*д = — (Ф(а)^д, ^P), it follows that (ф, Ф(а)^) = -(Ф(а)</>, ф) (10.6) for all ф^ф е V and all а е su(2). It is therefore convenient to define three linear operators Ai, A2, A3 by Ap = —г'Ф(ар), p = 1,2,3 (10.7) so that, by Equation (10.6), (^,АрФ) = (Ар^,^) (10.8) for all ф,ф eV and p = 1, 2,3. That is, Ai, A2, A3 are self-adjoint operators. Also, as [Ф(ар), Ф(ад)] = Ф([ар,ад]), Equations (8.31) give 3 [Ap, Ag] = i €pqrAr<) P) q — 1,2,3, (10.9) r=l where epqr is the permutation symbol of Equations (8.21). Now define the “ladder operators” A+ and A- by A+ = Ai + iA2, A_ = Л1 - iA2, (10.10) so that Ai = (A+ + A_)/2, A2 = -i(A+ - A_)/2. (10.11) Then, by Equation (10.9), [A3,A+]=A+, (10.12) [A3,A_] = —A_, (10.13) [A+,A_]=2A3. (10.14)
THREE-DIMENSIONAL ROTATION GROUPS 179 Moreover, Equation (10.8) implies that (A_^,^)-(^,A+^) (10.15) for all ф^ф eV. Finally, define the operator A2 by A2 = A2 + A^ +A2. (10.16) Then, by Equation (10.9), [A2,Ap]=0, p= 1,2,3, (10.17) which in turn implies that [A2, A+] = [A2, A_] = 0. (10.18) Also, as A_A+ = (Ai — iA^A^ + zA2) = A2 + A^ + z[Ai, A2], Equations (10.9) and (10.16) give A_A+ = A2 - A2 - A3. (10.19) Similarly, A+A_ = A2 - A2 + A3. (10.20) In the quantum mechanical theory of angular momentum, the operators Jx^Jy, Jz corresponding to the components about Ox, Oy and Oz satisfy the commutation relations Jy\ — itllz-! ['А/, Jz\ — ^^Jxi [J*> itlJy. (10.21) Jx, Jy and Jz are general angular momentum operators, in the sense that they may correspond to intrinsic spin, orbital angular momentum, or a combination of both. The symbols Lx, Ly and Lz will be reserved for the special case of orbital angular momentum (as in Chapter 8, Section 4). Apart from a factor П, Equations (10.9) and (10.21) are identical. Therefore the identifications Jx — IA]., I у — TlA^ Jz J+ = ЙА+, J_ = ЙА_, J2 = П2А2, /zA3, > (10.22) provide the connection between the representation theory of su(2) and the theory of angular momentum. ~ Returning to su(2), in any representation Г of £ the matrix representing A2 (defined by Г(А2) = ]Cp=1 {Г(АР)}2) commutes (by Equation (10.17)) with Г (a) for all a e C. Schur’s Lemma (see Chapter 9, Section 4) then implies that if Г is irreducible then Г (A2) must be a multiple of the unit matrix, so all the basis vectors ^2,..., фа of V are eigenvectors of A2 with the same eigenvalue.
180 GROUP THEORY IN PHYSICS It is convenient to summarize the main conclusions concerning the irre- ducible representations of su(2) in the form of a theorem. The results have immediate application in the theory of angular momentum and the theory of isotopic spin, as well as forming the basis of the representation theory of semi-simple Lie algebras that will be developed in Chapter 12. Theorem I To every non-negative integer or half-integer j (i.e. for j = 0, |, 1, |,...) there exists an irreducible representation of su(2) (and its com- plexification Ai) of dimension d = 2j + 1. The ortho-normal basis vec- tors of the irreducible representation specified by j may be denoted by where m takes all values from j down to —j in integral steps (i.e. m = 3,3 — 1, j — 2,..., — j + 1, —j). Each basis vector may be chosen to be a simultaneous eigenvector of A2 and A3 with eigenvalues j(j + 1) and m respectively, that is (10-23) Л3< = (10.24) (for m = — 1,..., — j). Moreover, the relative phases of the basis vectors may be chosen so that A+^m = {U + m + (10.25) А-'Фт = Ш+ ™)G'-™ + 1)}1/2Сг-1, (10.26) for m = j)j — 1)... )—j. Up to equivalence these are the only irreducible representations of su(2) (and its complexification Ai). The matrices of this irreducible representation may be denoted by D7 (a), with elements ZP (a)m/m, where mf and m take the values j, J — 1,..., — 3 + 1, —j (these all being integers if 3 is an integer and all being half-integers if 3 is a half-integer). Then (э.1)т,т Z)-7 (a2 )т/тп IN (^з)т,т |г[<5т',т+1{0' - m)(j + m + I)}1/2 +<5m',m-l{(J + rn)(j -m+ 1)}1/2], i[<5m',m+l{(j -m)(j + m + l)}1/2 > + rn)(j - m + 1)}1/2], (10.27) Proof By an appropriate similarity transformation, each irreducible repre- sentation Г of su(2) may be chosen so that the matrix representing аз is diagonal. This implies that the basis vectors may all be chosen to be eigen- vectors of A3, as well as being eigenvectors of A2. As A3 is self-adjoint, all its eigenvalues are real. Let denote an eigenvector of A3 with eigenvalue m (so that A3^ = т'ф). Then, by Equation (10.12), Аз(А+^) = (m + 1)(A+^), so A+^ is an eigenvector of A3 with eigenvalue m +1 unless А^'ф = 0. Similarly, by Equation (10.13), A3(A_'0) = (m - 1)(А_^), so A_^ is an eigenvector of A3 with eigenvalue m — 1, provided that А_ф> 0. (One can picture the
THREE-DIMENSIONAL ROTATION GROUPS 181 eigenvalues of A3 as providing the rungs of a ladder with unit spacing between the rungs. A+ and A_ are called “ladder operators”, as they correspond to taking steps up and down the ladder.) As all basis vectors of V are eigenvec- tors of A2 with the same eigenvalue, A+^ and А_ф are eigenvectors of A2 with the same eigenvalue as ф. Let j be the maximum eigenvalue of A3 in the set belonging to a particular irreducible representation Г, assume that it is non-degenerate, and let be the corresponding eigenvector. Then A^j = 0, as otherwise (7 + 1) would be an eigenvalue of A3. Thus A_A+^ = 0 and so, from Equation (10.19), (A2 — A2 — A3)i/jj =0. Consequently A2^- = j(J + l)^j, so the eigenvalue of A2 is j(j + 1). Thus all the basis vectors of this representation must have 7(7 + 1) as their eigenvalue with A2. Now consider the set of vectors A_^, (A_)2/0j, (A_)3^-,..., which cor- respond to eigenvalues j — l,j — 2,j — 3,..., with A3. As the irreducible representation under consideration is finite-dimensional, the vectors in this sequence must become zero after a finite number of steps. Suppose, therefore, that A; is a non-negative integer such that (A_ )fc^- 0 but (A_)fc+1^ = 0, so the minimum eigenvalue of A3 is (j — k). As A_(A_)/c^ = 0, it follows that A+A_{(A_)/c'^J} = 0 and consequently, by Equation (10.20), (A2 - A2 + A3){(A_)^} = 0. But A3{(A_)^} = (7 - k){(A_)%} and A2{(A_)fc^} = J(J + l){(A_)fcVij}, so this implies that {j(j + 1) - (J - kf + (j - k)}{(A-)%} = 0. By assumption {(A_)fc^} 0, so {j(j+l) — (J — k)2 + (J — k)} = 0, which gives j = jA;. Thus the only possible values of j are j = 0,|,l,j,.... Moreover, the minimum eigenvalue of A3 is j — к = j — 2j = —j. Thus the eigenvalues of A3 are J, j — 1, j — 2,..., — j + 1, — j. As there are (2j + 1) such values, the dimension d = 2j + 1. Let denote the simultaneous eigenvector of A2 and A3 with eigenvalues 7(7 + 1) and m respectively (so that is the eigenvector previously denoted by фф). Then, for m = 7, j — 1,..., — j + 1, as A_^^ is an eigenvector of A3 with eigenvalue (m — 1), (Ю.28) where is a complex number. As ф3т and are assumed to be normal- ized. l/^ml2 = (MmV’m-uMmV’m-l) = (on using Equation (10.15)). However, it follows from Equation (10.20) that А+А-'Фт = (A2 -Al + А3)ф3т = O(J + 1) — m2 + т}ф3т = {(J+"*)(> ~m + 1)Ж, SO iMml2 = 0'+m)(j - m + 1).
182 GROUP THEORY IN PHYSICS However, Equation (10.28) also implies that л+с_х = (1М)Л+А_< = {|<I7<}<- Following the phase convention established by Condon and Shortley (1935), /z^ will be chosen to be real and positive, so that (4n = {(j +ni)(j- m + 1)}1/2. As = 0 for j = —m, Equation (10.26) follows for m = 7,7 — 1,..., — j + Z, and —j. Moreover, with this convention {|/^|2//z^} = /z^, so ^+^772—1 РтФгп for m = 7, j — 1,..., — j + 1. Thus A+^rn = /4+1V4+1 = {O' + m + 1)0 - ™)}1/2<Ц-1 for m = j — 1,7 — 2,..., — j. As /z^+1 = 0 for m = 7, Equation (10.25) holds for m = j as well. Clearly this representation is irreducible. As Equation (10.5) can be rewritten as Ф(аЖ = Y DJ m' = —j for all a of su(2) (and of C = Ai) and m = 7, j — 1,..., —7, it follows that ^(а)т,т = (С„Ф(а)<). Consequently ^(aiU = (С„(г72)(А++Л_)О, D4a2)m'm = ^т„Ш(А+-А_Ш, which immediately give Equations (10.27) on using Equations (10.25), (10.26) and (10.24). This proofJmade use of the properties of the operator A2, which is not itself a member of For proofs that do not involve A2 see, for example, Samelson (1969) and Varadarajan (1974). It should be noted that, as the irreducible representations of su(2) are specified by 7, which takes values 0, |, 1, j,..., there is a countable infinity of such representations, in agreement with the theorem of Peter and Weyl (1927) that is stated as Theorem X of Chapter 4, Section 6. It is interesting to examine in more detail the special cases j = 0, | and 1. For J = 0 the irreducible representation is one-dimensional and D°(a) = [0]
THREE-DIMENSIONAL ROTATION GROUPS 183 for all a e su(2). For j: = | the first and second rows are labelled by m' = | and — | respectively, while the first and second columns are similarly labelled by m = | and — Then Equations (10.27) give D1/2(ai) = a>, the aj being as in Equations (8.30), that is, D1/2 is identical to the defining two-dimensional representation of su(2). Similarly for j = 1 the rows are labelled by mf = 1, 0 and —1, and the columns are labelled by m = 1, 0 and —1, and Equations (10.27) give Dx(a3) = i 0 0 0 0 0 0 0-2 This representation is equivalent (but not identical) to that given by the 3x3 matrices of Equations (8.12), that is, to the defining representation of so(3). 4 Representations of the Lie groups SU(2), SO(3) and 0(3) It is not difficult to show explicitly that every irreducible representation of su(2) exponentiates to give an irreducible representation of SU(2). (See, for example, Chapter 12, Section 4, of Cornwell (1984) for details.) Henceforth the (2j + l)-dimensional irreducible representation of su(2) and the corre- sponding irreducible representation of SU(2) will both be denoted by the same symbol D7. Let X,7(u) be the character of D7 for an element u of SU(2) spec- ified by a rotation through an angle cj about an axis n, as in Equation (10.2). As all such u corresponding to the same value of cj lie in the same class (ir- respective of the direction of the unit vector n), X'7(u) can be calculated by taking n = (0, 0,1). Then, by Equation (10.2), cos(ttcj) + i sin( t;cj) 0 11 — x Z 7 x Z 7 [ 0 cos(|cj) — zsin(|cj) _ exp(jzcj) 0 0 exp(—|zcj) = exp(cjaa), (10.29) where аз is defined in Equations (8.30). Thus D^u) = exp(cjDJ (a3)), which, by Equations (10.27), is a diagonal matrix with diagonal elements exp(zmcj), m taking the values j, j — 1,..., — j. Thus з xJ(u) = E exp(imw)>
184 GROUP THEORY IN PHYSICS which can be rewritten as XJ(u) = sin{(2j + 1)cj/2}/sin(cu/2). (10.30) Incidentally, as x1(u) = 1 + 2coscj, Equation (10.4) shows that the three- dimensional representations D1(u) and R(u) are equivalent. Turning now to the representations of the Lie group SO(3), it is clear that the representation D7 of SU(2) provides a representation of SO(3) if and only if DJ (u) = I2J+1 for every u belonging to the kernel /С of the homomorphic mapping of SU(2) onto SO(3). But /С consists only of I2 and —12 (see Chapter 2, Section 6), so this condition is simply that DJ (—12) = 12J+1- However, by Equation (10.2), —12 corresponds to cj = 2тг and any direction of n. Thus, taking n = (0,0,1), Equation (10.29) shows that —12 = ехр(2тгаз), so that D-7(—12) = ехр(2тгВ-7(аз)). But Equations (10.27) indicate that this latter matrix is diagonal with diagonal entries taking the values ехр(2тггт) with m = j,j — 1,..., — J, which are all equal to +1 if j is an integer, but are all equal to —1 if j is a half-integer. Consequently the irreducible representation D7 o/su(2) (and so(3)) exponentiates to give an irreducible representation 1У 0/80(3) if and only if j is a non-negative integer. The basis functions of these representations D7 (7 = 0,1,2,...) of SO(3) are quite easily determined. As they are basis functions of the corresponding representation D7 of the Lie algebra so(3) they must satisfy the equations P(a)<(r) = £ ^(a)m,mC,(r) for all a4 e so(3) and m = 7,7 — 1,..., —j. By virtue of Equations (10.7), (10.10), (10.16), (10.23), (10.24) and (10.26), this implies that -{P(a1)2 + P(a2)2 + P(a3)2}<(r) = j(j + l)<(r), (10.31) -iP(a3)V4(r) = гщ^(г), (10.32) and - i{P(ai) - гР(а2)}<1(г) = {(j + m)(j - m + l)}x/2V4-i(r)- (Ю.ЗЗ) Expressing Equations (8.23) in spherical polar coordinates (r, #, ф) gives P(ai) = — sinфд/дв — cot# cosфд/дф, ) Р(аг) = cos ф д/дв — cot # sin^ д/дф, > P(a3) = д/дф, ) so that Equations (10.31), (10.32) and (10.33) become + =i0' + iX(r), (Ю-34) d (10.35)
THREE-DIMENSIONAL ROTATION GROUPS 185 and +icote ^№(r) = {(> + rn)(j -m + l)}1/2^_i(r). (10.36) Clearly the solution of Equation (10.35) is ^m(r) = e2m<^f(r,0). Substi- tuting into Equation (10.34) gives = Ж + 1)Ж«)- This is the associated Legendre equation, which has as its solution /(r,0)=P7(cos0)F(r), where F(r) is any function of r and P™^) is the associated Legendre function, which may be defined in terms of the Legendre polynomials Pj (£) by p™(e) = (i - for m = 0,1,2,..., J, and by РГ^ = + m)!} PT(e) for negative values of m (see Bateman 1932, Edmonds 1957). With the “spher- ical harmonics” Yjm(0, ф) defined by = (-l)ro{(2j + l)(j-m)!/47r(j+m)!}1/2e^p™(cos0), (10.37) Equations (10.34), (10.35) and (10.36) are satisfied with the ortho-normal basis functions <(r) = yjTO(MW). Here R(r) is any function of r alone (that is the same for all m = J, J — 1,..., —j) such that /* |P(r)|2r2 dr = 1. Jo This depends on the ortho-normal property of the spherical harmonics: [ [ У*т(0,ф)УГт,(0,ф') sine <№(1ф = 6^6тт,. Jo Jo Clearly only the angular parts of the basis functions are significant, and this analysis shows that they are given by the spherical harmonics YjmfO, ф). (In the literature there appear several different definitions of PJ77, (£) for m negative and also different choices of the numerical factor on the right-hand side of Equation (10.37). The definitions given here are chosen so that the factor {(J + m)(j — m + l)}1/2 appears on the right-hand side of Equation (10.36) in order to correspond to the phase convention of Condon and Shortley (1935). See Edmonds (1957) for a thorough discussion of these points.)
186 GROUP THEORY IN PHYSICS Finally, consider the group of all rotations in IR3, both proper and im- proper, which is isomorphic to 0(3). As noted in Example I of Chapter 2, Section 7, 0(3) is isomorphic to the direct product of S0(3) with the group of order 2 consisting of matrices I3 and —13. As this latter group has only two irreducible representations, both one-dimensional, which may be labelled Г1 and Г-1 in such a way that ГЧ13) = [1], ^(-13) = [1], Г-1(13) = [1], Г-Ч-13) = [-1], Theorem I of Chapter 5, Section 5, shows that for j = 0,1,2,... the group 0(3) has two inequivalent irreducible representations of dimension (2j + 1), which may be denoted by rpj (p = 1 or — 1) and which are defined by r^R) = DJ(R), Г1,17(—R) = DJ(R), 1 r-h^R) = D-?(R), T-h^-R) = —DJ(R), J for each R E S0(3), D7 being the irreducible representation of S0(3) dis- cussed above. The quantity p may be called “parity”. As for the spatial inversion operator /, as P(I)Y)m(0, ф) = (—1)-7У7т(0, </>), the basis functions Yjm(9, $)F(r) of D7 are basis functions of Гр’-7 only when p = (—l)-7. Consequently the irreducible representations Гр’-7 of 0(3) with p = — (—l)-7 possess no basis functions. This does not mean that the irreducible representations Гр’-7 of 0(3) with p = — (—l)-7 have no physical significance, for it is possible to have irreducible tensor operators transforming according to such a representation. The most important example where this is so is that of the orbital angular momentum operators Lx, Ly and Lz of Equations (8.26). Indeed, Equations (8.26) and (9.34) show that LT, Ly and Lz transform as irreducible tensor operators of the three-dimensional adjoint representation of S0(3), which is easily shown to be equivalent to D1. Moreover, for the spatial inversion operator I, Equations (5.29) show that РЩд/дхРЩ-1 = -д/дх, РЩд/дуРЩ-1 = -д/ду, and PfjLfd/dzPfjL)-1 = —d/dz, while Example IV of Chapter 5, Section 3, shows that P{I)xP{I)-x = -x, РЩуРЩ-1 = -у, and P^zP(I)-1 = —z. Consequently Equations (8.23) and (8.26) imply that Р(/)РЖР(/)-1 = Lx, P^LyP^I)-1 = Ly, and P^L^P^)"1 = Lz. Thus Lx, Ly and Lz transform as irreducible tensor operators of a representation of 0(3) that is equivalent to Г1,1. In fact, as [Р(аз),Р2] = 0, the operator Lz transforms as the row labelled by m = 0 of Г1,1. 5 Direct products of irreducible representa- tions and the Clebsch-Gordan coefficients The analysis of Chapter 9, Section 4, shows that it is merely a matter of conve- nience whether the Clebsch-Gordan series and Clebsch-Gordan coefficients are derived in terms of the Lie algebra su(2) (or so(3)) or the Lie group SU(2) (or S0(3)). In this particular situation, the Clebsch-Gordan series is found most
THREE-DIMENSIONAL ROTATION GROUPS 187 directly by a group theoretical argument, but the Clebsch-Gordan coefficients are most easily deduced by algebraic considerations. (In other cases it is usu- ally much easier to use the Lie algebraic arguments for both Clebsch-Gordan series and coefficients.) Theorem I The Clebsch-Gordan series for the direct product of two irre- ducible representations D-71 and D-7'2 of SU(2) is D?1 ® D>2 « ГР^'2 ф DJ'1+J'2-1 Ф ... Ф DlJ1-J'2l+1 Ф (10.38) where D7 appears on the right-hand side once and only once for j taking values from (д + j2) down to |д — j2| in integral steps. (This is also the Clebsch-Gordan series for su(2) and so(3), and for SO(3) provided Ji and j2 take integral values.) Proof By Equations (5.14) and (10.30), the character x71072^) of D-71 0D7'2 for an element u of SU(2) specified, as in Equation (10.2), by a rotation through an angle co about an axis in the direction n is X31®32 (u) = [sin{(2ji + l)w/2} sin{(2j2 + l)w/2}]/sin2(w/2). But 71+72 £ xJ(u) = Im[£-’L^_j2|exp{(2.?+ l>’/2}]/sin(cj/2) 3=\ji-h I = Im[i{exp(i|ji - j2|w) - exp(«(Ji + j2 + l)w)}]/sin2(w/2) = [cos{(ji - j2)w} - cos{(ji + j2 + l)w}]/sin2(w/2) = (u) for all u e SU(2), from which Equation (10.38) follows immediately. As Equation (10.38) implies that j if 7 = 71 +72,71 +72 - 1, • • •, bi ~ J2I, /inoQ\ -71’-72 [ 0, for all other values of J, ‘ ' the index a that appeared in the general expressions in Equations (5.35), (5.36), (9.28) and (9.29) is redundant and can, therefore, be omitted from Equations (5.36) and (9.29). At the same time, the Clebsch-Gordan coeffi- cients will be rewritten in a slightly different notation that is widely used in the literature on angular momentum, as follows: 71 mi 72 m2 ) = (ji,‘m,1,j2,m2 I (Edmonds 1957). In this context Equations (5.35) and (9.28) become 7i 32 £ £ U^m1,j2,m2 \ji,j2,j,m) ГП2--32 (10.40)
188 GROUP THEORY IN PHYSICS where 0:in transforms as the row labelled by m (= —J, — j + 1,... , J) of the irreducible representation D7 appearing in D-71 0 D-7'2. Of course the double appearance of Ji and 7'2 in these coefficients is un- necessary, but is accepted because of the natural way in which this notation is suggested in the theory of addition of angular momentum (cf. Chapter 12, Section 5, of Cornwell (1984)). These quantities are sometimes referred to alternatively as “vector-coupling” or “Wigner” coefficients. The book by Edmonds (1957) contains a description of the various other notations that have been used for these coefficients. The Lie algebraic considerations that lead to explicit expressions for the Clebsch-Gordan coefficients will be omitted here (but a detailed introductory account appears, for example, in Chapter 12, Section 5, of Cornwell (1984)). Wigner (1959), Racah (1942), Schwinger (1952) and Edmonds (1957) have used these ideas to produce general formulae for the Clebsch-Gordan coeffi- cients, the expression derived by Edmonds (1957) being (ii,mlt j2,m2 = {(2J + l)0i + h - - h + + b + i)!/G'i +b + j + 1)!}1/2 x{(ji +mi)!(ji -mi)!(j2 + m2)!(j2 - m2)l(j + - m)!}1/2 x ^2(-l)z{d(ji +j2 - j - -rrn- z)l(j2+m2 - z)\ z x(j - h + mi + г)!(; - ji - m2 + г)!}-1 6mi+m2,m. (10.41) (Here the sum is over those non-negative integer values of z for which all of the factors (ji+>2 - .7 - 2)!, (ji-^i-г)!, 0'2 + т2- {j-hPmxPz}\ and (7 — ji — m2 + z)l are non-negative.) Simpler formulae may be obtained for special cases, and are discussed in the references just given and in the books by Rose (1957) and Biedenharn and Louck (1979a,b). These works also contain explicit tabulations of Clebsch-Gordan coefficients, further extensive tables being given by Condon and Shortley (1935) and Cohen (1974). See also Butler (1981). All these Clebsch-Gordan coefficients may be taken to be real It is easy to extend these results to the group 0(3). With notation for irre- ducible representations of 0(3) introduced in Section 4, the Clebsch-Gordan series for 0(3) is ppi Jl 0 pP2,j2 ~ pPlP2,jl+j2 ф pPlP2,jl+j2-l ф _ ф pPlP2,b’l-j2| (Ю.42) (pi and P2 taking the values 1 and —1). Note that the parity p of every ir- reducible representation Г77’-7 in the Clebsch-Gordan series is the product of the parities pi and P2 of the irreducible representations in the direct prod- uct. (This follows because with Фр^(/)^р>> = , Equation (5.33) gives ® = (Р1Р2){С;71 ® Фт22}^ the rest of the content of Equation (10.42) being determined by the corresponding series of Equation (10.38) for the subgroup SO(3)). It is obvious that for p = P1P2 the Clebsch-
THREE-DIMENSIONAL ROTATION GROUPS 189 Gordan coefficients of 0(3) are given by ( ^mi1 PmJ22 | Pm 1 ) = mi^2’m'2 U1J2J,m), (10.43) where (Ji, mi, 7'2, m2 |ji, 7’2, J, m) are the Clebsch-Gordan coefficients of SO(3) described above. 6 Applications to atomic physics The ideas developed in the previous sections will now be applied to the “one- electron” theory of atomic structure, that is, to the theory in which each electron of the atom is assumed to move in the average field of all the other electrons, the resulting potential energy being assumed to be spherically sym- metrical. Although some important results will be derived in this section, the discussion is primarily intended to indicate the sort of conclusions that can be obtained. For more detailed accounts based on more elaborate models of atomic structure, the reader is referred to the monographs of Condon and Odabasi (1980), Wigner (1959) and Wybourne (1970). In this section the dynamical effects of the spin of the electron will be ne- glected. (For an introductory account in which they are included, see Chapter 12, Section 6, of Cornwell (1984)). With this assumption the wave function of the electron is a scalar function and the theory introduced in Chapter 1, Section 3, applies. In this situation the group of the Schrodinger equation is the group of all rotations in IR3. The theory of Chapter 1, Section 4 then shows that the eigenfunctions of the time-independent Schrodinger equation are basis functions of the irreducible representations of this group. These were deduced at the end of Section 4 above. Before proceeding further it is convenient to make a small change in no- tation to give that used in the physics literature. As noted in Equation (8.26), the operators P(ai), Р(аг) and Р(аз) are merely the orbital angu- lar momentum operators Lx, Ly and Lz multiplied by a factor (i/K). Con- sequently the basis functions ^m(r) °f Section 4 are, by Equations (10.31) and 10.32), eigenfunctions of the “total” angular momentum operator L2 (de- fined by L2 = L2 + L2 + L2) and Lz with eigenvalues h2 j(j + 1) and hm respectively. This first eigenvalue is normally denoted by h2l(l + 1) (with I = 0,1,2,...), so for the rest of the discussion of representations of SO(3) and 0(3), I will be used in place of j. The integer I is called the “azimuthal” or “orbital angular momentum quantum number”, and the integer m is known as the “magnetic quantum number” because of its role in the Zeeman ef- fect (which will be described shortly). With this change the eigenfunctions of the time-independent Schrodinger equation are labelled by I and m (with Z = 0,1,2,..., and m = Z, I — 1,..., —Z), together with another quantum num- ber, usually denoted by n and called the “total quantum number”, associated with the solution of the radial part of the time-independent Schrodinger equa- tion. Let ^nm(r) = Rni(r)Yim(9, ф) be such an eigenfunction. (It is common
190 GROUP THEORY IN PHYSICS to call states corresponding to Z = 0,1,2,3,..respectively s-states, p-states, d-states, f-states and so on.) The first application will be to the selection rules for optical transitions within the “dipole” approximation described in Chapter 6, Section 2. In the present case Qo (the invariance group of the unperturbed Hamiltonian) is the group of all rotations in IR3. Transitions from an initial state ф{,(т) to a final state ф/(Е) (both assumed to be basis functions of some irreducible representations of Qq) are forbidden if (</>/, С^фр) = 0, where Q = Ag.grad for absorption or induced emission in an electromagnetic field whose polarization is specified by the constant vector Aq and where Q = n.grad for spontaneous emission polarized in the direction of the unit vector n. Equations (5.29) show that the components of the operator grad trans- form as irreducible tensor operators of the representation of 0(3) defined by T(R) = R for all R e 0(3). This is equivalent to the irreducible representa- tion Г-1,1 of 0(3) in the notation introduced in Section 4. Indeed, Equations (8.23) show that [P(ai),£]=0, [P(ai),^] = -£, [P(ai),£] = ^, ' [Ла2),^] = £, [P(a2)4]=0, = > [Лаз),^] = -^, [Лаз)4] = £, [Р(аз),£] = 0. , (10.44) Taken with Equation (9.34) and the matrices for D1 given at the end of Section 3, Equations (10.44) imply that — / dx + id/dy), d/dz and (1/ д/2)(5/дх — id/dy) transform as irreducible tensor operators belonging to the m = 1, m = 0 and m = — 1 rows of the irreducible representation D1 of SO(3) and hence then transform according to the same rows of the representation Г-1,1 of 0(3). Now suppose that the initial state eigenfunction </>i(r) is the basis function ^nm(r) belonging to the row labelled by m of the irreducible representation rp,z of 0(3), where p = (—l)z. Suppose first that I > 1. Then, by Equation (10.42) (with pi = p(-l)z, Ji =l,P2 = “I, 72 = 1), rp’z 0 Г"1’1« r_p,z+1 e r_p’z e (10.45) so that the only possible final state eigenfunctions </>/(r) are basis functions of the irreducible representations on the right hand side of Equation (10.45). However, there are no basis functions transforming as Г-р,г (as with — p = — (—1/ the parity has the “wrong” value (see Section 4), so that </>/(r) can only be a basis function of r_p,z+1 or Г_р,/-1. That is, assuming I > 1, if <Mr) =<Z'm'(r), then I' = I + 1 or I - 1. (10.46) Similarly, if I = 0, Equation (10.42) gives ri,o0r-i,i (10.47)
THREE-DIMENSIONAL ROTATION GROUPS 191 so that </>/(r) can only transform as Г 1,1. That is, for I = О, V can only be given by 1/ = Z + l. (10.48) Further selection rules exist for polarized radiation. For Ao or n in the ^-direction, Q = d/dz, which transforms as the m = 0 row of Г-1,1. Then the Wigner-Eckart Theorem, taken with Equations (10.41) and (10.43), implies that m' = m. (10.49) Similarly, for Aq or n in the x- or ^/-direction, Q transforms as a combination of irreducible tensor operators belonging to the m = 1 and m = — 1 rows of Г-1’1, implying that mf = (m + 1) or (m — 1). (10.50) The selection rules of Equations (10.49) and (10.50) become observable if a small magnetic field H = (Яж, Hy, Hz) is applied to the system. The resulting theory provides an example of the general technique described in Chapter 6, Section 3. The perturbing term H' in the Hamiltonian H is Я' = -(е/2/хс){Яж£ж + HyLy + HZLZ}, (10.51) Lx, Ly and Lz being the orbital angular momentum operators and e and fi the charge and mass of the electron (Schiff 1968). Without loss of generality it may be assumed that the coordinate axes are chosen so that Оz is in the direction of H, so that Я' = -(e/2[ic)HzLz. (10.52) The analysis at the end of Section 4 shows that Я' is an irreducible tensor operator transforming as the m = 0 row of the irreducible representation Г1,1 of the invariance group Po = 0(3) of the unperturbed Hamiltonian ff0 = -(fl2/2M)V2 + V(r). Consequently the invariance group Q of the perturbed Hamiltonian Я(= Яо + Я') is the direct product of the group SO(2) of all proper rotations about Oz with the group {Я, I}. Consider the unperturbed energy level 6q corresponding to the eigenfunc- tion ^nm(r) that is a basis function of the irreducible representation Гр,г of 0(3) (withp = (—1/). This has degeneracy 2(2Z +1) (the factor 2 being due to the electron’s spin). However, the irreducible representations of SO(2) are all one-dimensional and are given by Г COSCJ — since 0 since 0 cosce 0 0 1 (10.53) for all integral values of m, both positive and negative (as SO(2) is isomor- phic to U(l)), and these are the irreducible representations of U(l). As the character xl of the irreducible representation Dz of SO(3) for this rotation is
192 GROUP THEORY IN PHYSICS Y^m=-i ехР(ршс^) (see Section 4), it follows immediately that the reduction of Dz of SO(3) into irreducible representations of SO(2) is simply dz « rz e r^1 e... e r~z+1 e гЧ (io.54) In fact, with Dz specified as in Sections 3 and 4, Dz is actually the direct sum of these irreducible representations of SO(2), no similarity transformation being needed to execute the reduction. (That is, the matrix S of Chapter 6, Section 3, is the identity matrix I2Z+1.) One therefore expects for I > 1 that the unperturbed energy level 60 will be split into (2Z + 1) different energy levels by the magnetic field (each having a two-fold degeneracy because of the electron’s spin). This prediction is easily confirmed. For I > 1 Equation (10.42) gives r1’10 rp’z « rp’z+1 e rp’z e r^"1, (10.55) the appearance of Гр,г on the right-hand side of Equation (10.55) indicating that the energy levels are perturbed to first order. The perturbed levels are the eigenvalues of the (21 + 1) x (21 + 1) matrix A whose elements are given (in the present notation) by Am'm = 4“ ('Фпт') 'Фпт) for т,т' = Z,I — 1,..., — I + 1, —I (cf. Equation (6.19)), 6q being the un- perturbed energy eigenvalue. However, by Equations (10.52) and (8.26), H' = — (eZzPz/2/zcz)P(a3), so that, by Equation (10.24) = -(е/гЯ2/2^с)т<то(г). Thus Am'm = (cfaHz/2/10)771}. Consequently the perturbed energy eigenvalues are 60 — (eTbHz/2/zc)m (10.56) for m = Z, Z — 1,..., —Z + 1, — I. (This analysis shows that because of the simple form of H' it is not necessary in this case to invoke the Wigner-Eckart Theorem to deduce the quantities (^m,,P'^m)).
Chapter 11 The Structure of Semi-simple Lie Algebras 1 An outline of the presentation The theory of semi-simple Lie algebras is worth studying in detail, not only because of its elegance and completeness but also because of its considerable physical applications, particularly in elementary particle theory. The present chapter is devoted to the study of the structure of semi-simple Lie algebras. Section 2 gives the definitions of simple and semi-simple Lie alge- bras and contains a very useful criterion of Cartan. The process of “complex- ification”, that is, of going from a real Lie algebra to a complex Lie algebra, is then investigated in Section 3, with particular emphasis on the semi-simple case. Most of the rest of the chapter is devoted to the structure of the semi- simple complex Lie algebras, for which the complete classification is presented. The semi-simple real Lie algebras are the subject of the last section of this chapter. Chapter 12 contains the basic ideas of representation theory of semi- simple Lie algebras and Lie groups together with examples. Appendix D contains some detailed information on the properties of simple Lie algebras. 2 The Killing form and Cartan’s criterion The developments of this section apply equally to real and complex Lie alge- bras (except where explicitly stated otherwise). The relationship between real and complex Lie algebras will be examined in the next section, particularly for the simple and semi-simple cases. Definition Simple Lie algebra A Lie algebra C is said to be “simple” if it is not Abelian and does not possess a proper invariant Lie subalgebra.
194 GROUP THEORY IN PHYSICS The definitions of the terms “proper” and “invariant” were given in Chap- ter 9, Section 2, the term “Abelian” having been introduced in Chapter 8, Section 4. Here (as throughout this book) the convention applies that every Lie algebra and subalgebra has dimension greater than zero. Definition Semi-simple Lie algebra A Lie algebra £ is said to be “semi-simple” if it does not possess an Abelian invariant subalgebra. The definitions imply that if £ is simple then £ is certainly semi-simple. However, the converse is not true, for if £ = su(2) ф su(3) then £ is semi- simple, but £ is not simple because it possesses invariant subalgebras isomor- phic to su(2) and su(3). If £ is Abelian then £ is neither simple nor semi-simple. (Such an algebra is barred explicitly from being simple, and is implicitly prevented from being semi-simple because it is an Abelian invariant subalgebra of itself.) As all one-dimensional Lie algebras are Abelian, simple and semi-simple Lie algebras must have dimension greater than one. Definition Simple linear Lie group A linear Lie group Q is said to be “simple” if and only if its real Lie algebra £ is simple. Definition Semi-simple linear Lie group A linear Lie group Q is said to be “semi-simple” if and only if its real Lie algebra £ is semi-simple Thus a simple linear Lie group is semi-simple, but the converse is not necessarily true. If Q is Abelian or possesses a proper Abelian invariant Lie subgroup, then Q is not semi-simple (see Theorem I of Chapter 9, Section 2). The treatment of examples will be deferred until the end of this section, when Cartan’s criterion will have been introduced. The “Killing form”, which will now be defined, provides not only a very convenient criterion for distinguishing semi-simple Lie algebras but also plays an important part in the analysis of the structure of such algebras. Definition Killing form The “Killing form” В (a, 6) corresponding to any two elements a and b of a Lie algebra £ is defined by B(a, 6) = tr {ad(a)ad(6)}, (11.1) where ad (a) denotes the matrix representing a E £ in the adjoint representa- tion of £ (as defined in Chapter 9, Section 5) and tr denotes the trace of the matrix product (see Appendix A, Section 1). If £ is a real Lie algebra, all the matrix elements of ad (a) are real for each
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 195 a e £,, so that in this case В (a, b) is real for all a, b E C. This will not be so when C is complex. Example I The Killing form of C = su(2) With the commutation relations (Equations (8.31)) of C = su(2), Equation (9.36) implies that 0 0 0 ’ ’00-1 ad(ai) = 0 0 1 , ad(a2) = 0 0 0 .° -i 0. 10 0 ’ 0 1 0 ‘ ad(a3) = -10 0 0 0 0 from which it follows by Equation (11.1) that B(ap,a9) — ^bpg for p, q = 1, 2, 3. Example II The Killing form of C = sl(2,IR) As noted in Table 8.1 of Chapter 8, Section 5, C = sl(2,IR) is the real Lie algebra of traceless real 2x2 matrices. A convenient choice of basis is bi4 0 -1 -1 0 ba=l 0 1 -1 0 b3 = 1 3 2 1 0 0 -1 (П-2) giving the basic commutation relations [bi,b2] =b3, [b2,b3] =bi, [b3,bi] = —b2. (It will be noted that the first two of these differ by a sign from the corre- sponding relations for su(2) (Equations (8.31)).) Thus, by Equation (9.36), " 0 0 0 " 0 0 1" ad(bi) = 0 0 1 , ad(b2) = 0 0 0 0 1 0 -1 0 0 _ '0-10 ad(b3) = -1 0 0 0 0 0 and consequently B(bx,bi) = 2, B(b2,b2) = —2, B(b3,b3) = 2, and, for p^q, (p,q = 1,2,3), B(bp,bQ) = 0.
196 GROUP THEORY IN PHYSICS The main properties of the Killing form are summarized in the following theorem: Theorem I The Killing form is a symmetric bilinear form. (See Appendix B, Section 5). That is (a) В (a, b) = В (b, a), for all a,6 G £; (b) В (aa. /3b) = a/3B(a, 6), for all a, b E a and (3 being any pair of real numbers if C is real or any pair of complex numbers if C is complex; (c) B(a,b + c) = B(a, 6) + B(a,c), for all a,b,c E C. Also (d) if is any automorphism of £, B(^(a), ^(6)) = В (a, b) for all a, b E (e) B([a, 6], c) = B(a, [6, c]), for all a,b,c E £; (f) if £! is an invariant subalgebra of £, and B/y denotes the Killing form of £! considered as a Lie algebra in its own right, then В (a, 6) = Bz/(a, b) for all a, b E £!. Proof See, for example, Appendix E, Section 6, of Cornwell (1984). The key to the whole theory of semi-simple Lie algebras is provided by “Cartan’s criterion for semi-simplicity”, which is as follows: Theorem II A Lie algebra £ is semi-simple if and only its Killing form is non-degenerate. That is, £ is semi-simple if and only if det В /= 0, where В is the n x n matrix whose elements are defined by Bpq = B(ap,ag) (for p, q = 1, 2,..., n), ai, a2,..., an being a basis for £. (The account of non-degenerate bilinear forms given in Appendix B, Sec- tion 5, shows the equivalence of the two conditions appearing in the statement of the theorem.) Proof See, for example, Appendix E, Section 6, of Cornwell (1984). The following theorem shows that the study of semi-simple algebras re- duces to the study of simple Lie algebras. Theorem III Every semi-simple Lie algebra is either simple or is the direct sum of a set of simple Lie algebras. That is, if £ is a semi-simple Lie algebra then there exists a set of invariant subalgebras £i, £2? • • •, &k (& > 1) which are simple, such that r = £1 er2e...ez:fc. (11.3) Moreover, this decomposition is unique.
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 197 Proof See, for example, Appendix E, Section 6, of Cornwell (1984). A further reason for studying a semi-simple Lie algebra by means of its adjoint representation is provided by the following theorem. Theorem IV If C is semi-simple then its adjoint representation ad is faith- ful Proof Suppose that ad is not faithful, so that there exist two elements a, b E C such that ad (a) = ad(6) but a b. Then ad(n — b) = 0. Hence, for any c 6 £, ad(a — 6)ad(c) = 0, so that В (a — 6, c) = 0. Thus the Killing form is degenerate, and hence C cannot be semi-simple. The adjoint representation has a further important property: Theorem V If C is simple then its adjoint representation is irreducible. Proof The carrier space V of the adjoint representation of C can be identified with C itself, with the operator ad(a) (a e £) acting in C defined for all b E C by ad(a) b = [a, b] (see Chapter 9, Section 5). If ad is reducible there must exist a non-trivial subspace V of V such that V V and ad(a)bf E V' for all a E C and b' E Vf. That is [a, b'] E Vf for all a E C and b' E Vf. Thus Vf is a proper invariant subspace of £, which is impossible if C is simple. Thus ad must be irreducible. Example III The groups SU(AT), N > 2 SU(V) is simple for all N > 2. For the case N = 2, this is a straightforward consequence of results already obtained. In fact Example I above shows that det В = (—2)3 = —8 (/= 0), so that Cartan’s criterion implies that su(2) is semi-simple. But su(2) must be simple, as otherwise su(2) has a decomposition of the form of Equation (11.3) with at least two members. This is not possible, as the dimension of su(2) is 3, while the dimension of every simple Lie algebra is greater than or equal to 2. For N 3 the “simplicity” of su(AT) can be demonstrated by identifying su(AT) with one of a set of simple Lie algebras. (See, for example, Appendix G, Section 1, of Cornwell (1984).) Example IV The groups SO(V), N > 2 (i) SO(2) is Abelian and therefore not simple. (ii) SO(3) is simple, as so(3) is isomorphic to su(2) (see Example III of Chapter 9, Section 3) which, as shown in Example III above, is simple. (iii) SO(4) is semi-simple but not simple, for it can be shown that SO(4) is homomorphic to SO(3) 0 SO(3). (iv) SO(V) is simple for N > 5.
198 GROUP THEORY IN PHYSICS Example V The groups U(AT), N > 1 (i) U(l) is Abelian and therefore not simple. (ii) For N > 2, as u(AQ = u(l) ф su(AT) (see Example I of Chapter 9, Section 6) and, as u(l) is Abelian, U(AT) is not semi-simple. Example VI The Euclidean group of JR3 Reference to Example II of Chapter 2, Section 7, shows that the subgroup of pure translations is an Abelian invariant Lie subgroup of this group, which cannot therefore be semi-simple. 3 Complexification The process of going from a real Lie algebra to a complex Lie algebra is known as “complexification”. The most straightforward situation is where the real Lie algebra consists of matrices or linear operators, and the basis elements are linearly independent over the field of complex numbers. (This was the situation encountered in Chapter 8, Section 4.) Suppose that £ is an n-dimensional real Lie algebra of matrices with basis ai, a2,..., an. Theorem I of Chapter 3, Section 1, shows that the only solution of 52p=1 Apap = 0 with Ai, A2,..., An all real is Ai = A2 = ... = An = 0. However, it is possible that Sp=i Apap = 0 with one or more of Ai, A2,..., An complex and non-zero, in which case ai, a2,..., an are not linearly independent over the field of complex numbers (Example II below provides an demonstration of this behaviour). Nevertheless, the simplest assumption to make is that ai,a2,... ,an are linearly independent over the field of complex numbers. (For a completely general treatment of complexification, see, for example, Chapter 13, Section 3, of Cornwell (1984).) With this assumption the set of matrices of the form 52p=1 Apap, where Ai, A2,..., An take arbitrary complex values constitute a complex Lie algebra £, the Lie product of which is given by n [a, b] = Ap/ZgCpgar, p,q,r=l where a = 52p=1 Apap and b = l^q^-qi anc^ where, in £, [ap,aQ] = 52r=1 cpgar? Cpq being the structure constants of £. £ is then the complexifi- cation of £. Clearly £ (considered as a complex vector space) and £ (consid- ered as a real vector space) have the same dimension n. Indeed ai, a2,..., an form a basis for both £ and £, and with this basis both Lie algebras have the same set of structure constants. Example I The complexification of £ = su(2) As the basis elements а1,а2,аз of £ = su(2) defined by Equations (8.30) are linearly independent over the field of complex numbers, by the above
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 199 construction the complexification £ of £ = su(2) may be taken to be the set of all 2 x 2 matrices of the form ]Cp=i \>ap, where Ai, A2, A3 are complex. (It should be noted that these are not necessarily Hermitian.) Example II Problems with the complexification of £ = sl(2, <D) As noted in Table 10.1. £ = sl(2, (C) is the set of all traceless 2x2 matrices. A convenient basis is 1 o’ ’ 0 1 ' ’ 0 0 ' ai = 0 -1 , a2 = ° ° , аз = 1 ° ' i O' ’ 0 г " ’ 0 0 ' a4 = 0 , a5 = 0 0 ? a6 = г 0 These are linearly independent over the real field, but as a4 = zai, as = Ш2 and as = газ, they are not linearly independent over the field of complex numbers. Definition Real form of a complex Lie algebra A “real form ~ of a complex Lie algebra £' is a real Lie algebra whose com- plexification £ is isomorphic to £'. The following example shows that a complex Lie algebra can have two (or more) real forms that are not isomorphic. Example III su(2) and sl(2, IR) as real forms of the same complex Lie al- gebra Let а1,а2,аз and Ь1,Ь2,Ьз be the bases of su(2) and sl(2,IR), defined by Equations (8.30) and (11.2) respectively. Then bi = zai, b2 = a2 and Ьз = газ, so the complexifications of su(2) and sl(2,IR) coincide. Thus su(2) and sl(2, IR) are both real forms of the same complex Lie algebra. This example indicates that the deduction of the real forms of a complex Lie algebra is not a trivial matter, even if the complex Lie algebra is simple. This problem will be examined in Section 10. Nevertheless, some straightfor- ward results do exist in this area, as the following very important theorem shows. Theorem I Let £ be the complexification of a real Lie algebra £. Then £ is semi-simple if and only if £ is semi-simple. Moreover, if £ is simple then £ is also simple. Proof See, for example, Chapter 13, Section 3, of Cornwell (1984). Although £ is necessarily simple if £ is simple, it should be noted that Hie converse is not true. However, it can be shown that if £ is simple and £ is not simple, then £ must be the direct sum of two simple complex Lie algebras that are isomorphic (Gantmacher 1939b).
200 GROUP THEORY IN PHYSICS It can be demonstrated quite easily that every d-dimensional representa- tion of C provides a d-dimensional representation of its complexification £, and vice versa. ~ Henceforth every complex semi-simple Lie algebra will be denoted by £, with C denoting a real Lie algebra. This notation is justified by the fact that every complex semi-simple Lie algebra is isomorphic to the complexification of some real Lie algebra. (Indeed every such complex Lie algebra is isomorphic to the complexification of at least two non-isomorphic real Lie algebras (see Section 10).) 4 The Cartan subalgebras and roots of semi- simple complex Lie algebras This section and all the remaining sections of this chapter up to the pen- ultimate section will be devoted to the study of the structure of semi-simple complex Lie algebras. The transition back to semi-simple real Lie algebras will be considered in the final section. The presentation will take the form of a series of theorems, which lead to the construction of the “canonical” form of Weyl (1925, 1926a,b). This facilitates the development (given here in outline only) of the complete clas- sification of all simple complex Lie algebras, which was originally given by Killing (1888, 1889a,b, 1890) and Cartan (1894). It will become clear in the next chapter on the representation theory of semi-simple Lie algebras that there are very considerable advantages in work- ing with the canonical form, so the construction of this form will be considered in detail for several physically important examples. Definition Cartan subalgebra TL ~ A “Cartan subalgebra” TL of a semi-simple complex Lie algebra £ is a subal- gebra of C with the following two properties: (i) TL is a maximal Abelian subalgebra of C (that is, TL is Abelian but every subalgebra of C containing TL as a proper subalgebra is not Abelian); (ii) ad(/i) is completely reducible for every h E TL. (Here ad denotes the adjoint representation of £.) (It is possible to give a definition of a Cartan subalgebra that applies to any Lie algebra, but it is necessarily rather more complicated than that just given above. Nevertheless, it can be shown that the general definition reduces to the above definition in the semi-simple case (see Goto and Grosshans 1978, Helgason 1962, 1978, or Samelson 1969). It requires a fairly^ lengthy proof to demonstrate that every semi-simple complex Lie algebra C does possess at least one Cartan subalgebra (see Hel- gason 1962, 1978). Also, although it is obvious that any automorphism of
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 201 C maps a Cartan subalgebra into another Cartan subalgebra, the proof that any Cartan subalgebra can be mapped into any other Cartan subalgebra by an automorphism of C is more difficult (see Helgason 1962). This latter result implies that all the Cartan subalgebras of a semi-simple Lie algebra have the same dimension, and so permits the following definition. Definition The rank of a semi-simple complex Lie algebra The “rank” I of a semi-simple complex Lie algebra C is defined to be the dimension of its Cartan subalgebras. Now let /ii, Л2, •.., Ы be~a basis of a Cartan subalgebra TL of a semi- simple complex Lie algebra C of rank I and dimension n. (For the present /11, hi,..., hi may be chosen quite arbitrarily, the only requirement being that they are each members of TL and are linearly independent.) Then, as TL is Abelian, the irreducible representations of TL are all one-dimensional. Conse- quently the matrices ad(/ij) for j = 1, 2,..., Z must not only be diagonalizable but must be simultaneously diagonalizable. As a similarity transformation ap- plied to ad corresponds to a change of basis of C (see Chapter 9, Section 5), there exists a basis /ii, /12, • • •, Ы, а'ъ a2,..., a'n_l of C such that [hj,a'k] = ак^)а'к for j = 1,2,..., Z, and к = 1, 2,..., n — Z, where ak(hfi) are a set of complex numbers. As TL is Abelian, [hj, hk\ — 0 for J, A; = 1, 2,..., Z. Moreover, as TL is a maximal Abelian subalgebra of £, for each к = 1,2,..., n — Z, there must exist at least one j(= 1, 2,..., Z) such that ak(hj) 0- Now let h = p^jhj be any element of TL and for each к = 1,2,..., n—Z, define a linear functional ak on TL by i ak(h) = ^Hjak{hj). (As always (see Appendix B, Section 6), a linear functional on a vector space is completely specified by its values on a basis of that space. Here /ii,/i2, •«• ,/iz are arbitrary complex numbers.) Then for all h E TL and for each к = 1,2,... ,n — Z, the linear functional ak is not identically zero (i.e. (*k(h) /= 0 for some h E TL) and [h, 4] = ak(h)ak. Each such linear functional is called a “non-zero root” of £. It is conceivable that two or more such roots may be identical, that is, possibly ak(h) = aw (Ji) for all h E TL and к /= k1. In fact the closer examina- tion that follows shows that this cannot happen, but the possibility will not be excluded for the present.
202 GROUP THEORY IN PHYSICS For any non-zero root a of £ the set of elements aa e C such that [ft, aa] = a(fi)aa (11.4) (for all h e TL) form a subspace of C which will be denoted by C(X and will be called the “root subspace” corresponding to^a. Then C is the vector space direct sum of TL and all the root subspaces Ca corresponding to non-zero roots. As [Л, hf] = 0 for all Л, h! e TL, it is sometimes convenient to regard TL as being the subspace of C corresponding to “zero root” and to write TL = £0. The set of distinct non-zero roots will be denoted henceforth by Д. Theorem I If aa e £a and E £,p, then G >Са+/3 if a + /? G Д, but [aa, ap] = 0 if a + (3 Д. Proof By Jacobi’s identity (Equation (8.14)), if h E TL, aa E £a and ap E Cp, [h, [aQ, a/?]] + [aa, [ap, Л]] + [a^, [h, aQ]] = 0, so that [h, [ac,,^]] = {a(/i) + (3(h)}[aa,a0], from which the stated result follows immediately. The conclusions of the following two theorems are rather technical, but are very useful for deducing the other theorems of the series. Theorem II If aa E £a and ap E Cp, and if a + (3 0, then -В(аа,а/з) = 0. Proof Suppose that n7 is any basis element of £7 for any root 7. Then (ad(aa)ad(ap))ay = [aa, [n^, a7]], which, by the preceding theorem, is a mem- ber of £q,+/3+7 if a+/?+7 is a root, but otherwise is equal to 0. In the first case £а+/з+7П£7 = 0 if a+(3 0. Thus, in both cases, (ad(nci)ad(a/3))a7 contains no part proportional to n7. Hence tr{ad(ncl)ad(a/3)} = 0 if a + (3 0. In particular, as TL = £0, it follows that B(h,aa) = 0 (11.5) for all h E TL and any aa E Ca (ae Д). Also if aa E Ca and a ^0, B(aQ:, aa) = 0. (11.6) Theorem III The Killing form of C provides a non-degenerate symmetric bilinear form on TL.
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 203 Proof All that has to be shown is that the Killing form В of £ is non- degenerate on TL, that is, if h! E TI and B(h\h) = 0 for all h E TI then hf = 0, for it is obvious that В is symmetric and bilinear on TL. Suppose therefore that h' E TI and B(h',h) = 0 for all h E TL. Then, by Equation (11.5), B(hf, a) = 0 for all a G £, and, as В is non-degenerate on C (as C is assumed to be semi-simple), it follows that hf = 0. It is now possible to associate with every linear functional a(fi) on Ti, and in particular with each root a E A, a unique element ha of TI by the definition B(ha,h) = a(fi) (11.7) for all h E TL (see Theorem I of Appendix B, Section 6). These elements ha play a very important role in the canonical basis of £. If a(ti) and /?(/i) are any two linear functionals on Ti, it follows from Equation (11.7) that ha+p = + hp. (11.8) Also, as В is symmetric, = /3(ha) = B(ha, hp}. (11.9) It is convenient to develop the notation a stage further and define {a, [3} by (a,/3} = B(ha,hp). (11.10) As В is symmetric, {a,f3} = {33, a}. (In fact (a, (3} can be regarded as a non-degenerate symmetric bilinear form on the dual space TL* of TL, that is, on the space of all linear functionals of TL. Angular brackets are used to emphasize that this is not an inner product.) Then Equation (11.9) can be rewritten as a(M = /?(M = (^,/?)- (11.П) Equation (11.4) then implies that (tV, fyaQi..) for all /?, a E A. ~ With the basis of C chosen so that each basis element is a member of some subspace £7, for any h E TL ad(h) is a diagonal matrix with zero diagonal elements corresponding to the basis elements of Cq = TL and with diagonal element corresponding to each basis element £7 (for 7 E A). Thus B(h,h'} = ^2(dimr7)7(h)7(h') (11.12) 7€A
204 GROUP THEORY IN PHYSICS for all h,h/ e TL. In particular, with h = ha and h' = hp. Equation (11.11) implies that {&,(?) = (dim£7)(a, 7) . (11.13) 7€A Theorem IV If a e A then —a e A. Proof Suppose a E A, aa E £a and —a qL A. Then, by Theorem II above, B(aQ:, a) = 0 for all a E £. As this is not possible because В is non-degenerate, —a must be a non-zero root. Before proceeding further with the general theory, it is useful to clarify these results by examining some physically important examples. Example I The Cartan subalgebra and roots of the complexification £ (= Ai J of £ = su(2) (and of £ = so(3)) As noted earlier, the real Lie algebras su(2) and so(3) are isomorphic (see Example III of Chapter 9, Section 3), so their complexifications are also iso- morphic. (They will be denoted by Ai in the general classification that will be given in Section 7.) For concreteness the argument will be given for £ = su(2). By Example I of Section 3, the complexification £ of su(2) can be taken to consist of all 2 x 2 matrices of the form Apap, where Ai, A2, A3 are arbitrary complex numbers and ai, a2, аз are defined by Equations (8.30). The subspace of matrices of the form Аза3 may be taken as a Cartan subalgebra TL. (This is certainly Abelian, and is maximal Abelian because if a = 52p=i pbpSLp is such that [a, A3 аз] = 0 then, by the commutation relations of Equations (8.31), /11 = /12 = 0. Moreover, from Example I of Section 2, ас!(Азаз) — A3 0 1 0 -10 0 0 0 0 which is diagonalizable and therefore completely reducible.) Thus the rank I is 1, and TL may be taken to have basis element hi = аз. From Equations (8.31), [hi, (ai+га2)] = z(ai+za2), 1 [hi, (ai - га2)] = -г(аг - га2), J so that there are two non-zero roots 07 and —ai, with ai(hi) = i. Thus £ai and £_ai are subspaces of matrices of the form A(ai + za2) and /z(ai — za2) respectively, A and /z being arbitrary complex numbers, so that both are one- dimensional. An explicit expression for hQ:i will now be found. From Equation (11.9), B(hai,hai) = ai(hQ1), so that with hQ:i = /йц, K2B(hi,hi) = KQi(hi).
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 205 But, from Example I of Section 2, B(hi, hi)(= B(a3, a3)) = —2, so к = — |г. Thus 1 4 1 0 0 -1 (11.14) It should be noted, as (ai,ai) = ai(hQ1) = KQi(hi), that (qi,Qi) = 1/2, which is real, positive and rational. Example II The Cartan subalgebra and roots of the complexification C (= A2) of£ = su(3) In his original paper on the SU(3) symmetry scheme for hadrons, Gell-Mann (1962) set up a basis for su(3) and its complexification, which has been widely used in the elementary particle literature ever since. It will be shown in the course of the following analysis how this basis is related to the canonical basis. Gell-Mann (1962) defined eight traceless Hermitian matrices Ai, A2,..., A8 by 0 1 0 0 Ai — 1 0 0 , a2 — i _ 0 0 0 _ _ 0 " 0 0 1 “ 0 A4 = 0 0 0 , A5 = 0 1 0 0 _ i ’ 0 0 0 ’ A7 = 0 0 —i -> a8 0 i 0 -i 0 0 0 0 0 _ 0 —i “ 0 0 0 0 - = (1M 3) A3 — A6 = ’ 1 0 0 1 0 0 1 0 - _ 0 " 0 ( 0 ( _ 0 : 0 " 0 -2 0 0 -1 0 0 0 ) 0 ’ ) 1 I 0 These satisfy the commutation relations 8 [Ap, Ag] = Ti,fpqr\ (11.15) where the fpqr are antisymmetric in all three indices, the non-zero values being listed in Table 11.1. A convenient basis for the real Lie algebra C = su(3) is then provided by the traceless anti-Hermitian matrices ai, a2,..., a8, defined by ap = гАр, p=l,2,...,8. As these are linearly independent over the field of complex numbers, then ai, a2,..., a8, or, alternatively, Ai, A2,. •, A8, may be taken as the basis of the complexification C (= Az). Direct calculation using Equation (11.15) shows that _B(ap,ag) — T26pq (p,q = 1,2,..., 8), the deeper significance of which will be explored in Exam- ple II of Section 10. Consequently, if В is the matrix introduced in Theorem II of Section 2, det В = (—12)8 0, so C and C are semi-simple.
206 GROUP THEORY IN PHYSICS pqr fpqr 123 147 156 246 257 345 367 458 678 1 1/2 -1/2 1/2 1/2 1/2 -1/2 ^3/2 ^3/2 Table 11.1: Non-zero values of the constants fpqr of su(3). The fpqr are antisymmetric under permutations of any two indices. A convenient choice of Cart an subalgebra is the subspace spanned by A3 and As, which implies that the rank I is 2. Then, with hi = A3 and 112 = As, [hi, (Ai + zA2)] = 2(Ai + zA2), [hi, (Аб + zAy)] = — (Аб + zAy), [hi, (A4 + гАб)] = (A4 + zAs), [hi, (Ai — zA2)] = — 2(Ai — zA2), [hi, (Ag — zAy)] = (Ag — zAy), [hi, (A4 — zA5)] = —(A4 — zA5), [112, (Ai + zA2)] = 0, [112, (Аб + zAy)] = \/3(Аб + zAy), [112, (A4 + ZA5)] = УЗ(А4 + zAs), [h2, (Ai — zA2)] = 0, [h2, (Аб — ZA7)] = — \/3(Аб — zAy), [h2, (A4 — ZA5)] = — УЗ(А4 — ZA5). Thus there are six non-zero roots: Qi, q2, <*3 and —Qi, — a2, —<^3, with ai(hi)=2, a2(hi) = -1, <*з(Ь1) = 1, aq(h2)=0, <^2(h2) = \/3, > a3(h2) = л/З. , Clearly £ai, Ca2, £аз, C_ai, C_a2, £-аз are all one-dimensional, with basis elements (Ai + zA2), (Аб + ZA7), (A4 + ZA5), (Ai — zA2), (Ae — zA?) and (A4 — ZA5) respectively. It should be noted that = Qi + a2- It remains to calculate explicit expressions for hQ1, htt2 and hQ3. Suppose hQ1 = ftihi + K2h2- Then, for j = 1,2, Equation (11.7) with h = h7 gives KiB(hi,hj) + к2В(Ь2,Ь7) = ai(hj), a pair of simultaneous linear equations for Ki and к2- As B(ap,ag) = — 126pq (p,q = 1,2, ...,8), it follows that B(hi, hi) = B(h2, h2) = 12 and B(hi, h2) = 0. Thus Kq = |, k2 = 0, and so , 1 1 hrv, = -hi = - ai 6 6 1 0 0 0-10 0 0 0 (11.16) Similarly, hcK2 11 1 “12hl+12^h2 = 6 0 0 0 0 1 0 0 0-1
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 207 ha3 - ^hx + ^V3h2 - i Finally, as hQ3 = hO1+a2 = hQ1 + hCV2, 1 0 0 0 0 0 0 0-1 It follows from Definition (11.10) that (^1,^1) 3? g> (a2,ai) = -|, (a2,a2) = (П-17) all of which are real and rational, and both (ai,ai) and are also positive. Indeed in this case (ai,ai) = {(*2, (*2)- Later developments (in Section 7) will also show the significance of the equality 2(ai,Q2)/(ai,ai) = -1, which is an immediate consequence of Equations (11.17). 5 Properties of roots of semi-simple complex Lie algebras The series of theorems on roots will now be continued. The notation is the same as in the preceding section. Theorem I If aa E and a_a E C_a then П—a] — a_a)ha. (11.18) Proof For any h E 'H, by part (e) of Theorem I of Section 2, B([aa, a_Q], h) = B(aa, [a_Q, h]). But, by Equations (11.4) and (11.7), [a_a, h] = — [/1, a_a\ = a(fi)a_a = B(ha, h)a_a, so that B([na, a_a], h) — B(aa, a_a)B(fia, h) = 0. That is, for any h E H, ^({[flco a] ce)^ce}? ^) = 0- As В is non-degenerate, it follows that {[aa,n_a] — B(aa, a_a)ha} = 0. Theorem II For each a E A and any aa E there exists an element of C-a such that B(na,a_a) 0.
208 GROUP THEORY IN PHYSICS Proof Suppose, to the contrary, that for some aa 0) of B(aa, a_a) = 0 for all a_a E C_a. Then by Theorem II of Section 4, B(ama) = 0 for all a E C. As the Killing form В is non-degenerate, this implies aa = 0, a contradiction. This theorem has the very important consequence that (as В is bilinear) for any aa E (a E A) and for any complex number Ba, there exists an element a_a E C_a such that B(aa,n_a) = Ba. Theorem III For every a, (3 E A, the quantities (a, /3) are real and rational. Moreover, for every a E A, (a, a) is positive. Proof See, for example, Appendix E, Section 7, of Cornwell (1984). Theorem IV H coincides with the subspace of C consisting of all elements of the form where the take all complex values. Proof Let Ti' be the latter subspace, which is clearly a subspace of H. Sup- pose that H' is a proper subspace of H. Then there exists a linear functional 7 (ft) of H such that у (Ji) is not identically zero on Ti but 7(/i) = 0 for all h E H'. Let 0) be defined (as in Equation (11.7)) by B(h^h) = y(h) for all h E H. Then, as 7(/ia) = 0 for all a E A, B(hyi ha) = 0 for all a E A, and hence, by Equation (11.9), a (Ay) = 0 for all a E A. Thus, by Equation (11.4), [Ay, a] = 0 for all a E so hy is the basis of a one-dimensional Abelian invariant subalgebra of £. As C is semi-simple, this is impossible, so Hf and H must coincide. This theorem implies that from the set of elements ha (a E A), a subset of I linearly independent elements may be selected and may be taken to form a basis for H. The elements of this set will be denoted by ..., (Z?i, , (3i £ A). (This set is not unique. For instance, for the complex- ification С (= A2) of su(3), Example II of Section 4 shows that (3^ = oq, /?2 = <^2 gives a basis for Ti, but one of several alternative bases is given by /31 = Qi, /З2 = Qi + Theorem V Every non-zero root a of A can be written in the form 1 where the coefficients ^2, • • •, k>i are all real and rational. Proof For any a E A, as heL , h@2 form a basis of Ti, it follows that ha = ч where Ki, ^2, • • •, Ki are a set of complex numbers. Then
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 209 a = Y^j=i so that, for к = 1,2,..., Z, i (@k, (fik > /3j}' j=l This is a set of I linear simultaneous algebraic equations for Ki, K2, • • •, k>i- As the quantities and (Д:, (3j) are all real and rational, then so must be Kl, K2, • • •, KT Let 'HfR denote the real vector space with basis , hg2,... The preceding theorem implies that, for any a E A, ha = J27-=1 Kjhe^ with Ki, K2, • • •, Ki real (and rational), so ha E Bjr for all a E A. Thus Tt^ is actually independent of the choice of the basis /77, /77,..., of TI. Theorem VI For any a E A, a(h) is real for all h E Tt^p>. Proof Suppose h E Tt^. Then there exists a set of real numbers /zi, //2, • • •, such that h = Y^3=i • Thus i i <*W = 7=1 i=l which is real by Theorem III of this section. Theorem VII The Killing form В of £ provides an inner product for the real vector space Tt^. Proof It has only to be shown that B(/z, h') is real for all Zz, h' E TIjr, and, for all h E that is non-negative and that B(fi,h) = 0 implies that h = 0 (see Appendix B, Sections 2 and 5). The first result follows immediately from Equation (11.12) and Theorem VI. Also from Equation (11.12), B(h, h) = ^27еД (dim £7) {7(71) }2, which is non-negative as each 7(/i) is real. Moreover, B(fi,h) = 0 only if 7(/i) = 0 for all 7 6 A, which is only possible if h = 0. Theorem VIII If a E A then dim£a = 1, and ka E A only if к = 1 or —1. Proof See, for example, Appendix E, Section 7, of Cornwell (1984). As a consequence of this theorem, Equation (11.12) simplifies to B(h,h') = ^y(hUh') (11.19) 7€A for all h,hf EH, and Equation (11.13) reduces to (<*,£) = 52 7) (11.20) 7<EA
210 GROUP THEORY IN PHYSICS for all a,/3 e Д. For each pair of roots a and — a of Д there exists a very useful three- dimensional simple subalgebra of C that can be constructed in the following way. Define Ha (e TC) by Ha = {2/(a,o)}ha, (11.21) and let Ea and E_a be elements of £a and £_a respectively such that B(Ea, E_a) = 2/(a, a). (11.22) Then, from Equations (11.4), (11.11), (11.18), (11.21) and (11.22), [Яа, Ea\ [Ea,E_a] 2Ea, -2E_a, > Ha. (11.23) These are precisely the commutation relations of the operators A3, A+ and A of Equations (10.12), (10.13) and (10.14), which played such an important role in angular momentum theory, the identification being 2A3 Ha. A+ Ea, A E_a. (11.24) Thus all the results on representations obtained in Chapter 10 apply immedi- ately to this subalgebra. This will prove very useful both in proving certain theorems and in constructing explicit representations of It should be noted that the analogue of Equation (11.8) is not true in general for the elements Ha. That is, in general, 7^ Ha 4“ Hp, because Equations (11.8) and (11.21) imply that Яа+/3 = {(a, a}/{a + /?, a + (3}}Ha + {(/?, (3}/{a + /?, a + fi)}Hp. Example I Application to C = the complexification o/su(2) As shown in Example I of Section 4, the complexification C = Ai of su(2) has only one pair of non-zero roots, namely and —Qi, and (ai,ai) = j. Thus, from Equations (11.14) and (11.21), HQ1 1 0 0 -1 Let Eai = K(ai + Ш2) and E_ai = /z(ai — Ш2). Example I of Section 2 then shows that B(Eai,E_ai) = + za2,ai - za2) = -4k/z,
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 211 so Equation (11.22) is satisfied with к/z = — 1. A convenient choice is к = /1 = — г, which gives (by Equations (8.30)) 0 1 0 0 0 0 1 0 E Ql This example shows that, for every semi-simple complex Lie algebra C and any non-zero root a of £, the elements Ha, Ea and E_a form the basis of a subalgebra that is isomorphic to the complexification Ai of su(2). Such a subalgebra will henceforth be called an “Ai subalgebra”, but in the math- ematical physics literature it is often referred to as an “su(2) subalgebra”. (Strictly speaking, this latter terminology is not accurate, as even if Equa- tions (11.23) are taken to be the commutation relations of a real Lie algebra, this real Lie algebra is not isomorphic to su(2) but is another real form of the complexification of su(2). Nevertheless, this slight misuse of language causes no confusion in practice.) Example II The Ai subalgebras (or “su(2) subalgebras”) of the complexifi- cation С (= A2) of su(3) As shown in Example II of Section 4, the complexification С (= A2) of su(3) has three pairs of non-zero roots, namely (a1? —ai), (q2? —(*2) and (аз, —a3). For the first pair, as (ai,ai) = Equations (11.16) and (11.21) imply that HQ1 1 0 0 0-10 0 0 0 With Eai = k;(Ai + гА2) and E_ai = /z(Ai — гА2), the third commutation relation of Equations (11.23), together with Equation (11.15), gives к/z = A convenient choice is к = /z = producing Eqj. — ’ 0 1 0 ’ 0 0 0 0 0 0 , E-Q1 — ’ 0 0 0 ’ 1 0 0 0 0 0 In the SU(3) symmetry scheme for hadrons this subalgebra is usually called the “I-spin” subalgebra, and a very common notation is defined by Ц = Eai, /_ = E_ai, h = (1/2)Яа1. Similarly, for the pair (a2, — ^2), ’ 0 0 0 ’ ' 0 0 0 ’ 0 0 0 0 1 0 , EQ2 — 0 0 1 , E-tt2 — 0 0 0 0 0-1 0 0 0 0 1 0 This subalgebra is called the “U-spin” or “L-spin” subalgebra (De Swart 1963), the identification of generators being U+ = L+ = EQ2, U_ = L_ = E_Q2, U3 = L3 = (1/2)Яа2.
212 GROUP THEORY IN PHYSICS Finally, for the pair (a3, —a3), a similar argument gives H«3 = '10 o’ 0 0 0 0 0-1 > — ’001’ 0 0 0 0 0 0 ? E-a3 — 0 0 0 ’ 0 0 0 1 0 0 the resulting subalgebra being called the “V-spin” or “К-spin” subalgebra (De Swart 1963), the identification of generators in this case being V±=K+ = Еаз, V-= K_ = E^ V3=K3 = -(1/2)Яаз. The next stage in the development involves the concept of a “string” of roots, the usefulness of which will become very clear in Section 7. Several other important results will appear as by-products of this notion. Definition The а-string of roots containing /? Suppose that a, E A. Then the “а-string of roots containing /?” is the set of all roots of the form /3 + ka, where k is an integer. Example III Strings of roots of the complexification C (= A%) o/su(3) As shown in Example II of Section 4, the non-zero roots of C (= Az) are ±a1? ±a2 and ±(ai + аг). Thus the ai-string containing аг consists of аг and (ai + аг), and the аг-string containing (ai +аг) consists of ai and (ai +аг). Theorem IX Let a, /3 E A. Then there exist two non-negative integers p and q (which depend on a and (3) such that (3-\-ka is in the а-string containing (3 for every integer к that satisfies the relation — p < к < q. Moreover, p and q are such that p — q = 2(/3, а)/(а, а). (11.25) Also 0-{2(0,a)/(a,a)}a (11.26) is a non-zero root. Proof See, for example, Appendix E, Section 7, of Cornwell (1984). Theorem X For all а,/? E Д, 2(/?, a)/(a, a) can take only the integral values 0, ±1, ±2 or ±3. (The quantities 2(/?, а)/(а, a) are called the “Cartan integers”.) Proof That 2(/?, a) / (a, a) must be an integer is an immediate consequence of Equation (11.25), as p and q are integers. Clearly, with (3 = ±a, 2(/?, a)/ (a, a) = ±2. Suppose then that /3 ±a, so that ha and h@ are linearly independent. As the Killing form В provides an inner product for applying the Schwarz inequality (see Appendix B, Section 2) gives < B(ha,ha) Bthfrhp).
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 213 Thus, by Equation (11.10), Ka,/?)|2 < {a,a){0,0), so that \2{a, 0)/{a,a}\ \2{a, 0)/{0,0}\ < 4, from which it follows that |2(a,/?)/(a, a)| can only take values 0, 1, 2 or 3. The final theorem shows that the Cartan subalgebra and root system to- gether completely specify a semi-simple complex Lie algebra (up to isomor- phism). As a preliminary, let TL and TC be two complex vector spaces and suppose that ф is a linear mapping of TL onto TC. Then а'(ф(1Т)) is a linear functional defined on TL if a' is a linear functional defined on TC. (This follows because, for any /ii, and any two complex numbers Kq and К2? a'((/>(/€i/ii + ^2^2)) = c/(fti<^(^i) + ^<^(^2)) (as ф is linear) = Kia7(</>(/&i)) + (</>(^2)) (as a' is linear).) Theorem XI Suppose that C and £,' are two semi-simple complex Lie al- gebras with Cartan subalgebras TL and TL1 respectively and non-zero root systems Д and Д' respectively. If ф is a linear mapping of TL onto TL1 such that a'(</>(/&)) e Д for every a1 E Д', then ф can be extended to become an isomorphic mapping of C onto . Proof See, for example, Helgason (1962, 1978) or Varadarajan (1974). 6 The remaining commutation relations It is now time to examine the commutation relations between elements of £a and where a,/? e Д and /3 —a. Suppose that a, /?, and a + /? E Д, and let ea, and еа+/з be basis elements of £,@ and £^+/3 respectively. Theorem I of Section 4 then implies that there exists a complex number Na^ such that [ea? — Na,pea+fl. (11.27) Theorem I If a, /? and a + (3 E Д, then Na^ 0. Proof See, for example, Chapter 13, Section 6, of Cornwell (1984). It will be obvious that the value of Na,p depends on the choice of the basis elements ea, ep and although it is always true that N/3,a = -Na,(3- (11.28)
214 GROUP THEORY IN PHYSICS However, Equation (11.6) shows that B(ea, ea) = 0, so that the Killing form В does not provide any natural normalization of ea. Several different sets of conventions for the choice of the ea will be found in the mathematics literature, and a detailed discussion of them is given in Chapter 13, Section 6, of Cornwell (1984). Here attention will be concentrated on just one choice of conventions, which will be adopted for the rest of this book, in which for all pairs a and —a of Д, B(ca,c_a) = -1, (11.29) and for all a, /3 E A, N-a,-p = Na,p. (11.30) With these conventions it can be shown that: (a) the Na,/3 are all real, (b) if a, /?, 7 e A and a + /3 + 7 = 0, then (11-31) (c) if a, /?, 7,6 E A are such that the sum of no two of them is zero, and if a + (3 + 7 + 6 = 0, then Na^N^s + + N^aN/3j = 0, (11.32) (d) for all a, /? e A, {Na,0}2 = {a, a}q(p + l)/2, (11.33) where p and q are such that the a-string containing (3 is (3 — pa, ..., /?, + qa. Moreover Equation (11.18) gives [еа,е_ q] = ha (11.34) for all a E A. Of course Equation (11.4) is valid with aa = ea, so that for all heH [h, ea] = a(Ji)ea, (11.35) and in particular [hp, ea] = a(hp)ea (= {a, (3)ea). (11.36) The basis elements Ea and E_a introduced in Section 5 may be defined in terms of the basis elements ea and by Ea = {2/(a, a) }1//2ea, 1 , E_a = f ( '' Then B(Ea, E_a) = 2/(a, a),
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 215 as in (11.22). In the case of a Lie algebra of matrices, the argument of Section 10 will show that, after an appropriate similarity transformation if necessary, the elements ha of Ti may be taken to be diagonal Hermitian matrices, while for each pair a and — a of Д the matrices ea and e_a may be chosen so that 6 — CK and correspondingly, by Equations (11.37), e_q = eJ. (11.38) (11.39) (Taking the adjoint of [h^,ea] = а(И/з)еа gives [h^ecJ] = — а(Ь/з)еа’*’, as = ha and a(h/3) (= (a,/?)) is real, implying that corresponds to root —a. (The argument of Section 10 confirms that, after an appropriate similar- ity transformation if necessary, can be taken to be a member of £.) Then taking the adjoint of [еа,е/з] = Na^Ca+p gives [еДе^] = so that Equation (11.38) is consistent with the convention of Equation (11.30).) With the choice of convention (11.38), Equation (11.29) becomes B(ea, eat) = 1, which determines ea up to a factor of modulus unity. Example I Basis vectors of the complexification C (= Az) o/su(3) As noted in Example II of Section 4, £ai, £Q2, £ai+cK2, £_ai, C-a2 and £_(ai+cK2) have basis elements (Ai + гА2), (Аб + гАу), (A4 + iA5), (Ai — гА2), (Аб — гАу) and (А4 — i\&) respectively. Thus eai — (Ai + гА2), ea2 = koc2(Aq + A7), ^(ai+ag) ^(^1+0:2) (^4 4“ ^5 ) 4 e_Qi tv—ai (Ai гА2), 6—a2 = a2(^6 ^7)? («1+0:2) (CK1+CK2) (^4 ^^5), where ка1,..., к_(а1+сК2) are a set of complex numbers. Then, as B(AP, Ag) = 126pq (p,q = 1,2,..., 8), the above conventions are satisfied with ( I/a/24, for a = ai,a2, (ai + a2), ( —1/л/24, for a = — ai, —a2, — (ai + a2). Consequently eQ1 = (1/V6) ea2 = (1/V6) ' 0 1 0 " 0 0 0 _ ° ° ° _ ' 0 0 0 ' 0 0 1 _ ° ° ° _ 0 0 1 ' 0 0 0 0 0 0 , e_Q1 = —(l/x/6) , e_Q2 = - (l/x/6) > e-(ai+a2) = — (1/v^) ' 0 0 0 " 1 0 0 _ ° ° ° _ ' 0 0 0 " 0 0 0 _ ° 1 ° _ ' 0 0 0 ' 0 0 0 1 0 0
216 GROUP THEORY IN PHYSICS With this choice, ^Vck!2 , —(cKl+<22) N_ (cKi+ск2),<^1 with the other non-zero structure constants being given by (11.30) and (11.28). The matrices hQ1 and htt2 of Example II of Section 4 complete the basis. In the theoretical physics literature (e.g. Behrends et al. 1962) a different choice of basis of the Cartan subalgebra TL is often encountered. Theorem VII of Section 5 shows that the Killing form В of £ provides an inner product for the real vector space TLjr. Consequently a basis may be set up in Bir that is ortho-normal with respect to the Killing form, and this basis is also a basis of TL. Thus there exist I elements Нъ H2,..., Hi of TLjr (and so of TL) such that В(Я,-,Я^) = ^, (11.40) for j, к = 1,2,..., I. Then, for any linear functional a(h) defined on TL the element ha (eTL) defined by Equation (11.7) is given by 1 ha (11.41) (As Я1, Я2,..., Hi form a basis for TL, ha = Y^j=i f°r some set of com- plex numbers • • • jA6?- But, by Equation (11.7), B(ha,Hk) = a(Hk), so Y^j=i Hk) = а(Яд.), whence Equation (11.40) implies that = a(Hk)). The advantage of this basis is that (a, /3) can always be expressed in a particularly simple form. Indeed for any two linear functionals a(h) and /?(/i) defined on Я, Equations (11.11) and (11.41) imply that 1 (а,(3)=^а(Н^Н^. (11.42) As noted earlier, every linear functional defined on TL is completely specified by its values on a basis of TL. In particular, a(h) is completely specified by the set of values а(Я1), а(Я2),..., а (Я/), which, by Theorem VI of Section 5, are all real when a is a root of C. This set can be written as an /-component vector a, that is a = а(Я2),..., а(Яг)). Consequently, if the scalar product ot./3 of two such /-component vectors ot and /3 is defined by 1 a.(3 = 3 = ^ that is, as the inner product of /-dimensional real Euclidean space, Equation (11.42) becomes simply ot.(3 = (a, /?). (11.43)
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 217 With this choice of basis, Equation (11.35) gives \Hj, Cftj Oi(^Hj)ea for each a E A, and Equations (11.34) and (11.41) give i [^cn a] ^(-Яу)-Яу f=i for each a E A, both relations involving only components of ol. These consid- erations explain why theoretical physicists tend to call a a “root”, whereas mathematicians prefer to attach this name to the corresponding linear func- tional It may be noted that as ad(Hj) is a diagonal matrix with zero diagonal elements corresponding to the basis elements of H and with diagonal element corresponding to the basis element B(Hj,Hk)= ^аШНк), so the condition in Equation (11.40) implies that ^(hfy)cv(hf/;;) = fijk, aGA for J, к = 1, 2,..., I. Example^II Ortho-normal basis of the Cartan subalgebra H of the complex- ification C (= Az) o/su(3) Applying the Schmidt orthogonalization process (see Appendix B, Section 2) to фд = haj (j = 1,2) with (</>j,</>fc) = B(haj,hak) = (а^ак), Equations (11.17) give Я1 = V3hai, 1 Hz = 2(hQ2 + |hQ1). J Thus, using Equations (11.17) again, «1 = |(2 A 0), 1 «2 = K-%/3,3), J and, using the explicit matrix representations of Example II of Section 4, (1/2УЗ)А3, > (1/2^3)Л8. The reader should be warned that there is a complete lack of uniformity of notation in the literature regarding the different choices of bases of Я, with the same symbols hj, haj, and Hj being sometimes defined quite differently by other authors.
218 GROUP THEORY IN PHYSICS 7 The simple roots Let , h/32,..., hfa be a set of I linearly independent elements of H, as defined in Section 5. Then, as shown in Theorem V of Section 5, every non-zero root a of Д can be written in the form i а = (11.44) j=i with the coefficients Ki, кг, • • •, *4 ah real and rational. Definition Positive root A non-zero root a of Д is said to be “positive” (with respect to the basis Z?i, Z?2, • • •, Z?z) if the first non-vanishing coefficient of the set Ki, K2, • • •, Ki ap- pearing in Equation (11.44) is positive. A similar definition of a “positive linear functional” can obviously be given for any linear functional a defined on ht for which Equation (11.44) is valid with the coefficients Ki, k2> • • •, all real. For brevity, the statement “a is positive” may be written as “a > 0”. Example I Positive roots of the complexification С (= A2) o/su(3) As shown in Example II of Section 4, the non-zero roots of C = A2 are ±ai, ±O!2 and ±(0;! + a2)- With respect to /?i = ai, /З2 = «2? the positive roots are ai, ci2 and a?[ + a?,- However, with respect to /31 = ai, /?2 = oq + <*2, the positive roots are ai, —a^ and + &2, (—0:2 being positive as —012 — ai — (oq + Q2) = /?i — /fe). This example shows that the question of whether or not a given root a e Д is positive depends entirely on the choice of the basis Normally once a choice of this basis is made it is adhered to, and with that understanding one can talk of a “positive root” without it being necessary to explicitly make reference to the basis Z?i, • • • ? Z?z• Clearly, if a E Д is positive then — a is not positive, so that exactly one half of the set of non-zero roots are positive roots. Also, if a > 0 and /3 > 0 then a + (3 > 0. The set of positive roots (defined relative to some fixed basis /?1, (З24 • • •, (3i) will be denoted by Д+. Definition Lexicographic ordering of roots Let a and /3 be any two roots of Д. Then if (a — /3) > 0 one says that a> (3. Clearly, if a then either a > (3 or (3 > a. If a = Y^j=i anc^ (3 = then a - (3 = ~ Thus a > /? if and only if the first non-vanishing coefficient rff — is positive. Put another way, a > /?, if, for some s with value 1,2,..., or Z, = к? for j = 1,2,..., s — 1 but kJ > к^. (The term “lexicographic” is used to describe the ordering
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 219 because it corresponds precisely to the conventional ordering of words in a dictionary, where for example “bat” appears before “cat” and “arm” before “art”.) Again the lexicographic ordering depends on the choice of the basis А, Аг,.. •, A- The lexicographic ordering definition can obviously be extended to any linear functionals defined on H for which Equation (11.44) is valid with the coefficients Ki, ft 2? • • •, all real. Definition Simple root of Д A non-zero root a of Д is said to be “simple” if a is positive but a cannot be expressed in the form a = (3 + 7, where /? and 7 are both positive roots of Д. Again all these statements are made relative to some chosen basis A, A, ..., A? and whether a given root a e Д is simple depends on this choice, as the following example shows. Example II Simple roots of the complexification £ (= A2) o/su(3) With A = ai, (З2 = it is obvious that ou and ot2 are the only simple roots. However, with A = ai, A = <^1 + 0:2, the simple roots are —ct2 and cei + a2- This follows because with this basis the set of positive roots is ai, —ot2 and cei + a2 (see Example I), but щ = (ai + Q2) + (—^2)? so 07 cannot be simple, whereas —a2 and щ + a2 are simple as they cannot be expressed as the sum of two other positive roots. For the rest of this section it will be assumed that for each C some choice of basis A? Аг, • • • , A has been made and is being strictly adhered to. Moreover, the simple roots that correspond to this basis will hereafter be denoted by ai, ci2,. •., 07- (This notation has already been anticipated in the discussions on C = Ai and A2 in the examples above and in the previous three sections. As Example II shows, for С = A2 and with the choice A = A = <*2, <*1 and c*2 are indeed simple. Similarly for £ = Ai, with A = <*1, the root Qi is positive and simple.) The first theorem that follows is of a rather technical nature, but the second is of crucial importance, for it shows that the set of simple roots form a basis of with very useful properties. Theorem I If a and (3 are two simple roots of Д, and a (3, then (a) a — (3 is not a root of Д; and (b) <«,/?>< 0. Proof See, for example, Chapter 13, Section 7, of Cornwell (1984). Theorem II If £ has rank I then £ possesses precisely I simple roots Qi, Q2, • • •, &i- They form a basis for the dual space (the space of all
220 GROUP THEORY IN PHYSICS linear functionals on H). Moreover, if a is any positive root of Д then where fci, fc2,..., fc/ is a set of non-negative integers. Proof See, for example, Appendix E, Section 9, of Cornwell (1984). With the properties of the simple roots established, the next stage is to introduce the “Cartan matrix”, which plays a crucial role in the developments that follow. Definition Cartan matrix~A The “Cartan matrix” A of £ is an I x I matrix whose elements Ajk are defined in terms of the simple roots ai, ce2, • •., сц of £ by = for J, к = 1, 2,..., I. Clearly Ajj = 2 for all J = Z, 2,..., Z, while Theorem X of Section 5 and part (b) of Theorem I above together imply that for j к the only possible values of Ajk are 0, —1, —2 or —3. Moreover, Ajk = 0 if and only if Akj = 0. (This follows because {o^ak} = ^Ajk^otj, aj) and (a/c,aJ)(= (aj,ak)) = ^Akj{oik,oik)- As (aj,aj) and (a^ak) are both non-zero, Ajk = 0 if and only if (aj,ak) = 0, that is, if and only if Akj = 0.) It can be shown that it is always true that det A 0. Example III The Cartan matrices of £ = A\ and £ = A2 (the complexifi- cations o/su(2) and su(3)) For £ = Ai, as I = 1, A is the lxl matrix A = [2]. For £ = A2, Equations (11.17) and (11.45) give As this example shows, it is elementary to construct the Cartan matrix A of £ from a knowledge of the root system of £. What is very remarkable is that the process can be reversed. In fact (i) it is possible to deduce a complete classification of all possible Cartan matrices merely from the theorems given above (without any a priori knowledge of the corresponding Lie algebras); and (ii) from the Cartan matrix A of £ it is possible to construct the complete system of roots Д of £, together with the whole set of quantities (aj, a&) for J, к = 1,2,..., I.
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 221 Indeed, as will be shown in Chapter 12, the Cartan matrix of £ also gives the irreducible representations of £. This process therefore provides a complete classification and specification of all the semi-simple complex Lie algebras. As Theorem III of Section 2 shows that every semi-simple complex Lie algebra is the direct sum of simple complex Lie algebras, attention can be concentrated on the Cartan matrices corresponding to the simple complex Lie algebras. No attempt will be made here to give the detailed analysis involved in this classification. (This may be found, for example, in the works of Jacobson (1962), Varadarajan (1974) and Goto and Grosshans (1978).) The argument is facilitated by the introduction for each £ of its “Dynkin diagram” (Dynkin 1947). In this diagram each simple root is assigned a point (or “vertex”) and AjkAkj lines are drawn between the vertices corresponding to aj and c^. Moreover, each vertex is assigned a “weight” cuj, defined by a>j = cu(aj,aj), where cj is a constant independent of j chosen so that the minimum value in the set cji,cj2, . • • ,cj/ is 1. Then, by Equation (11.45), cuj/cjfc = Akj/Ajk (provided Ajk 0) . As Ajk < 0 for j к, it follows that for j к Ajk = -{Л^ЛЛ172^/^}1/2, (11.46) so that it is obvious that the Dynkin diagram of C determines the Cartan matrix of £. The conclusion of the analysis is that there are four infinite sets of complex Lie algebras, (i) A, 1 = 1,2,3,..., (ii) Bh Z = 1,2,3,..., (iii) Chl = 1,2,3,..., (iv) = 3,4,5,..., that are known as the “classical simple Lie algebras”, and there are five others, denoted by Eq, E?, E%, and G2, which are called the “exceptional simple Lie algebras”. In each case the subscript is the rank of the algebra. The Dynkin diagrams are given in Figure 11.1. The diagrams for Ai, Bi and Ci are identical, from which it follows that the Lie algebras Al, Bi and Ci are isomorphic. The diagram for B2 differs from that for C2 only in the labelling of the simple roots, so B2 and C2 are also isomorphic. The same is also true of A3 and D3. Apart from these, all the Dynkin diagrams exhibited correspond to non-isomorphic simple Lie algebras. (It is permissible to consider the diagram given for Di when I — 2. It reduces to two vertices of weight 1 with no linking lines, from which it follows that D2 is not simple and is the direct sum of two A algebras.)
222 GROUP THEORY IN PHYSICS ( i ) A Hl- 1,2,3 .). 2 I (vni) E4 (ii) BL (/ =1,2,3 aL-\ al (iii) СД1 = 1,2,3 (iv) Z2z(Z = 3,4,5 (v) (vi) E7 (vii) 2 2 11 Z7| a2 ^3 <74 bx) G2 Figure 11.1: Dynkin diagrams of the simple complex Lie algebras. The weight ujj appears over the vertex labelled by a:r Example IV The Cartan matrices of £ = A± and deduced from their corresponding Dynkin diagrams For £ = Ai the Dynkin diagram is trivial, consisting only of one vertex with weight 1. The only inference is that I = 1, which, by Equation (11.45), implies that A = [2]. For £ = A2 the Dynkin diagram is given in Figure 11.2. Thus A12A21 = 1, cji = CJ2 = 1, so that, by Equation (13.62), 2 - 1 2 It is the equality of the Cartan matrices deduced here with those calcu- lated in Example III that implies that the complexifications £ of £ = su(2) and su(3) are Ai and A2 respectively in this scheme of classification. (This has been already anticipated in the notation used in some of the preceding discussions.) In Appendix D the Cartan matrices are displayed for all the classical simple complex Lie algebras, along with the complete and explicit specifications of all the roots and of the quantities (For similar information on the exceptional simple complex Lie algebras, see, for example, Appendix F of
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 223 *2 Figure 11.2: Dyukin diagram of A2. Cornwell (1984)). The method that enables these quantities to be deduced from the Cartan matrices will not be discussed her, but may be found, for example, in Chapter 13, Section 7, of Cornwell (1984). 8 The Weyl canonical form of £ It is convenient to now summarize the conclusions of the last four sections by gathering together various commutation relations that have been derived and the assumptions oi^which they have been based. Suppose that £ is a semi-simple complex Lie algebra of rank I. Let hai , ha2hai be the elements of the Cartan subalgebra TL corresponding (by Equation (11.7)) to the I simple roots ai, a2> • • •, 07 (as defined in Section 7). Let Д be the set of non-zero roots (as in Section 4), and let ea be the basis element of the subspace £a chosen so that B(ea,e_a) = -1 (11.47) (cf. Equation (11.23)). Then (i) for j, к = 1,2,..., Z, [ha^hak] = 0 (11.48) (as TL is Abelian); (ii) for any h E TL and any a E Д [/1, ea] = a(h)ea (11.49) (cf. Equation (11.35)) and in particular, for h = as a(/iaj) = (by Equation (11.10)), [haj, ea] = (a, aj)ea ; (11.50) (iii) for any a E Д, [ea,e_ q] = ha (11.51) (cf. Equation (11.34)), so that, with a = kfaj, 1 [eQ,e_a] = -^fc“^ ; (11.52) 3=1
224 GROUP THEORY IN PHYSICS (iv) for any a, /3 E A such that /3 —a, if a + /3 A, M/M (11.53) while if a + (3 E A, [^cn ^/?] -^Va:,/3^Ck:4-/3, (11.54) where Na,/3 is a non-zero real number that satisfies N_a,-p = Na,p (11.55) (cf. Equation (11.30)) and the identities in Equations (11.31) and (11.32), and whose square {Na,/3}2 is given by Equation (11.33). This canonical form with basis elements haj (j = 1,2,..., Z) and ea (a E A), and with these commutation relations, is known as the “Weyl canonical form” of £. 9 The Weyl group of £ For any linear functional (3 defined on TL and for any non-zero root a E A, define the linear functional Sa(3 on H by (Sa0)(h) = /3(/i) - {2{0,a)/{a,a}}a(h) (11.56) for all h E TL. This defines an operatorJSa that acts on linear functionals. In particular, if (3 is a non-zero root of £, then the last jpart of Theorem IX of Section 5 shows that Sa(3 is also a non-zero root of Put another way, each Sa maps the set of non-zero roots A of £ into itself. (It will be shown in Chapter 12, Section 2, that a similar result is true of the weights of each representation of £.) It is easy to verify that the following properties are valid for any a E A: (a) in the special case in which (3 = a Saa = —a ; (11.57) (b) Sa(Sa(3) = /? for any linear functional (3 on H, (c) for any linear functionals /3 and 7 on TL, (Sa(3,Sal) = (V,7) ; (11.58) (d) for any two linear functionals (3 and 7 on TL and any two complex num- bers ц and A, Sa(X(3 + M?) = ASa/3 +
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 225 Figure 11.3: The Weyl reflection Sa. The interpretation of these operators Sa becomes clear if the ortho-normal basis Bi, . •, Hi of TL (introduced in the last part of Section 6) is used. As noted there, each linear functional (3 on H has associated with it an l- component vector /3 defined by /3 = (/3(Я1),/3(Я2),...,/3(Яг)). Attention will be concentrated solely on those linear functionals /? for which the are all real. This includes all the roots (and all the weights) of Denoting the /-component vector corresponding to Safl by 5a/3, Equations (11.43) and (11.56) give Sa/3 = /3 - {2(/3.a)/(a.a)} a. Let 0 be the angle between the vectors ol and /3, so that /З.а = \/3\ |a| cos#, where |a| = (a.a)1/2 (see Figure 11.3). Then, if ol = a/|a| is the unit vector along a, {2(/3.a)/(a.a)}a = a{2|/3| cos#}, which is a vector in the direction of ol whose magnitude is twice the projection of (3 on ol. Consequently Sa{3 = /3 — a{2|/3| cos#} is the reflection of (3 in the plane perpendicular to ol (as is shown clearly in Figure 11.3). As the operators Sa were introduced by Weyl (1925, 1926a,b), they are known as the “Weyl reflections”. With the identity operator E defined by Ea = a (for any linear functional a on H) and with products such as SaS@ defined by SaS^ = S*a(S^7) (f°r any linear functional 7 on Tt), the set consisting of the identity operator, the Weyl reflections, and all products of Weyl reflections, forms a group, called the “Weyl group”, which will be denoted by W. A typical element of W wifl be denoted by S. As every element of W maps the set of non-zero roots of C into itself, W must be a subset of the set of all permutations of the non-zero
226 GROUP THEORY IN PHYSICS roots. Obviously this latter set is finite, so the Weyl group W of £ is a finite group. It can be shown (see Jacobson 1962) that every element of W can be expressed as a product of reflections S(yj associated with the simple roots. The construction of W is helped by the observations that if 5, T e W then S = T if and only if Saj = Taj for every simple root ay, and that (by Equations (11.45) and 11.56)) Safa:) A^ja^ for fik = 1,2,..., Z. (11.59) Example I The Weyl group W of £ = Az, the complexification of £ = su(3) Using the Cartan matrix of Example III of Section 7 and Equation 11.59), Sai cvi R&1 ^2 It then follows that S>CH2 S>OL 1^1 5.2^1^ = -ai = Qi + a2 —(ai + az) Ql Sa2ar = Sa2Q2 = S^ Sa2a\ Rdil R(y.2 ^2 ai + a2 —a2 = a2 = -(ai+a2) and further that Soc1 Sa2 Sai Qi Sa2 Sq^ Sa2 CV1 a2 5sv2 i ^2 SCY2 S^i Sa2 ^2 ^1 As every element of W is a string of products of Sai and Sa2, and {5ai }2 = {Sa2}2 = the fact that SaiSa2Sai = Sa2SaiSa2 implies that there are no further distinct elements in W. Thus W is the non-Abelian group of or- der 6 with elements {E, Sai, Sa2, Sai Sa2, Sa2 Sai, Sai Sa2 Sai}. (Incidentally, Equation (11.56) shows that Saij_a2 = SaiSa2Sai •) It is instructive to display these results using the ortho-normal basis Hi, Hz, Hi of H discussed above. As Z = 2 here, (3 is a two-component vector which can be taken to be the position vector of a point in a plane, as in Figure 11.4, the reflection planes of the general case becoming reflection lines in this plane. As noted in Example II of Section 6, «i = (l/\/3)(l, 0), a2 = (1/2х/3)(-1)Л/3), and so ai+a2 = (1/2^3) (1, ^3), so the reflection lines normal to <a2 and «i + a2 have directions (0,1), |(\/3,1) and |(\/3, ~1) respectively. If /3 does not lie on a reflection line, the six distinct elements S of W acting on /3 produce six distinct vectors 5/3. However, if /3 lies on a reflection line and /3^0, the set {5/3 | S E W} contains only three members, while if /3 = 0 then {5/3 | 5 E W} contains only the one member 0. The following two theorems give two other useful properties of Weyl re- flections. Theorem I If a is a positive root and a a3, then 5Qja > 0.
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 227 Figure 11.4: The actions of the elements S of the Weyl group W for C — A^ on a general 2-component vector corresponding to a linear functional (3 on H. Proof Suppose that a = 52i=i k^ai. Then it follows that = a - i = kiai + {kj-2{a,aj)/(aj,aj)}aj. If a /= aj and a is a positive root then, as naj is not a root (except when к = 1, —1 or 0), at least one of the coefficients ki must be positive for i j. Theorem II of Section 7 then implies that Saj oi must be positive. Theorem II If = Saj/3 for some j = 1,2,..., Z, then a = (3. Proof If = Saj/3, Equation (11.56) implies that a-/3= {2(a - /3, a j} / (a j, Thus a = /3 + кщ, where к is some complex number. Substituting back gives naj = {2(kqj, a.j}}oij = 2kqj, so к = 0, implying that a = (3.
228 GROUP THEORY IN PHYSICS These two theorems lead to a result that will prove useful in Chapter 12, Section 3: Theorem III If 8 = | (11.60) for all j = 1,2,..., I. Proof For any j = 1,2,...,/, Theorems I and II imply that 5Д maps the set of positive roots (excluding a3) into the set of positive roots (excluding оД so СКбA_|_ , = I 52 a~ (by Equation (n-57)) a^ctj = 6-aj. Then, by Equation (11.58), (Sa 5, оД = (5,аД so that (8 — = (5,аД and hence 2(5, — ay) = The order of W for each classical simple complex Lie algebra is listed in Appendix D. (The corresponding orders for the exceptional simple complex Lie algebras may be found, for example, in Appendix F of Cornwell (1984).) 10 Semi-simple real Lie algebras The process of complexification, that is,^of passing from a real Lie algebra C to its associated complex Lie algebra £, was described in Section 3. The opposite procedure will now be considered for the case in which C is semi- simple. In outline the procedure is quite straightforward^for if £ is a complex Lie algebra, with the basis elements «1,^2,... ,an of C chosen so that the structure constants crpq are all real, then real linear combinations of «1,^2,... and an form a real Lie algebra that is one of the “realjforms” of C. The problems essentially lie in constructing all the bases of C with the required property. There are essentially two stages in this construction. The first is to set up the “compact” real form of the complex Lie algebra £, and the second is to deduce the “non-compact” real forms from the compact real form. The construction and properties of the compact real form will be considered here in some detail, for these properties have wide implications. In particular, a remarkable result of Weyl shows that the compactness or otherwise of a semi-simple Lie group can be tested by algebraic properties of its associated real Lie algebra. Amongst other consequences, this result permits many of
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 229 the properties of unitary representations of compact semi-simple Lie groups to be carried over to non-unitary representations of non-compact semi-simple Lie groups, the discussion in Chapter 12, Section 1, of complete reducibility being a good example. The treatment of non-compact real forms here will be confined to an ex- plicit listing of all the simple non-compact real forms corresponding to classical semi-simple complex Lie algebras, but the method of construction, which is largely due to Cartan (1914, 1929) and Gantmacher (1939a,b), will be omit- ted. (A full account (using the involutive automorphisms of the compact real forms) appears, for example, in Chapter 14 of Cornwell (1984)). Definition Compact and non-compact semi-simple real Lie algebras A semi-simple real Lie algebra C is said to be “compact” if its Killing form is negative definite (that is, if В (a, a) < 0 for all a E C such that a 0), and is said to be “non-compact” otherwise. This terminology is used because compact semi-simple real Lie algebras correspond to compact Lie groups, as will be shown in Theorem III below. (The words “compact” and “non-compact” do not describe the topology of C itself which, being a vector space of dimension n, is homomorphic to the whole of IRn, which is unbounded and consequently always non-compact in the usual topology on IRn.) Example I C = su(2) as a compact semi-simple real Lie algebra Example I of Section 2, shows that for C = su(2), B(ap, ag) = — 28pq for p,q = 1,2,3. Thus, for a general element a of su(2), writing a = Pp&p with /zi, /12, Мз all real, it follows that В (a, a) = — /ip, so that B(a, a) < 0 for all a 0. Thus su(2) is a compact Lie algebra. As noted previously, su(2) is the real Lie algebra of the Lie group SU(2), which is a compact Lie group (see Example III of Chapter 3, Section 3). This provides a particular example of the general theorem stated below as Theorem III. If £ is a compact semi-simple real Lie algebra, then with (a, 6) defined for all a, b E C by (a, 6) = -B(a,6), (11.61) it follows that (a, 6) is an inner product and C is an inner product space (see Appendix B, Section 2). As usual, it is then possible to set up ortho-normal sets of basis elements in C. Then, if ai, «2,..., an is such an ortho-normal set, (ftp, Пд) = B(ap,aq) = 8pqi (11.62) for p, q = 1,2,..., n. The following two technical results will become useful in the subsequent development.
230 GROUP THEORY IN PHYSICS Theorem I If ai, a2, • • •, is an ortho-normal basis of a compact semi- simple real Lie algebra £, then the adjoint representation of C defined rela- tive to this basis consists of antisymmetric matrices. Put another way, with this basis the structure constants crpq are antisymmetric with respect to inter- changes of all pairs of indices (p, q), (q, r) and (r,p), i.e. cr = —cr = —cP = cP = —cP = cP pq QP Pr rP rQ Qr ’ (11.63) Proof By part (e) of Theorem I of Section 2, for any a e C and p,q = 1, 2,..., n, B([np, a], aq) = B(ap, [a, aq]). Thus, by Equations (9.14) and the linearity of B, n n ^{Hd(n)}rpB(nr, aQ) = ^{ad(n)}rgB(np, nr), r=l r=l from which it follows by Equation (11.62 that {ad(a)}gp = —{ad(a)}pg. The relation cpq = — cqp applies for any basis of any Lie algebra, while the other relations in Equations (11.63) are a consequence of the fact that {ad(ap)}r<? = crpq (Equation (9.37)). Theorem II If £' is a semi-simple real subalgebra of a semisimple real Lie algebra £, then £' is compact if C is compact. Proof See, for example, Appendix E, Section 10, of Cornwell (1984). The remarkable theorem of Weyl may now be stated: Theorem III A connected semi-simple linear Lie group Q is compact if and only if its corresponding real Lie algebra C is compact (that is, if and only if the Killing form of C is negative definite). Proof See, for example, Appendix E, Section 10, of Cornwell (1984). The most significant point is that this theorem provides a purely Lie al- gebraic criterion for the corresponding Lie groups to be compact. It should be noted that this is only valid in the semi-simple case. Indeed, it has al- ready been noted in Example II of Chapter 9, Section 3, that the Abelian non-compact Lie group IR+ and the Abelian compact Lie group SO(2) have isomorphic real Lie algebras. Consequently for Abelian groups there can be no Lie algebraic criterion for compactness. The following theorem shows that ever^semi-simple complex Lie algebra C has a compact real form C derivable from £ in a very straightforward manner. Theorem IV For each semi-simple complex Lie algebra £, if hag (j = 1,2,...,/) and eaie-a (a E Д+) are the basis elements of the Weyl canonical
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 231 form, then the set of elements iha:) (for j = 1,2,..., /), 1 ea + e_a (for each a e Д+), > i(ea — e_a) (for each a e Д+) J form the basis of a compact semi-simple real Lie algebra, which will be denoted by £c, and which will be called a “compact real form” of £. Proof See, for example, Appendix E, Section 10, of Cornwell (1984). Clearly £c consists of all elements of the form i a = i P'jh'a.j 4- (ea H- a) + iya c_q,)}, j = l ckGA_|_ where the /zj, xa and ya are all real. Writing za = xa 4- iya, so that za is a complex number, a typical element a E Cc can be expressed as i a = iEl^ha3+ 52 {^«e« + (n.64) y=l a e Ан- It can be shown that any two compact real forms of C are isomorphic, so Cc is unique up to isomorphism, and may be referred to as “the compact real form” of £. It is easily verified that the following elements provide a natural ortho- normal basis for £c with respect to the inner product of Equation (11.61): (for j = 1,2,...,/), (l/i/2)(ea + e-a) (for each a E Д+), > (l/i/2)^(ea — e_a) (for each a E Д+), Я1,Я2, • • • being the ortho-normal basis of TL (i.e. = 6jk for j, к = 1,2,...,/ (see Section 6)) and ea the basis element of £a in the Weyl canonical basis. In the case of Lie algebras of matrices, as every representation of a com- pact Lie group is equivalent to a unitary representation, the matrices of the corresponding compact real Lie algebra £c can (by part (f) of Theorem IV of Chapter 9, Section 4) be taken to be anti-Hermitian. The previous theorem then implies that hj = hQ and % for all a E Д (as mentioned earlier in Section 6).
232 GROUP THEORY IN PHYSICS Example II Ortho-normal basis for the compact real form Cc = su(3) of c = Az The canonical basis of the complex Lie algebra C = Az has been constructed in Example II of Section 4, in terms of the Gell-Mann matrices Ai, A2,..., As- Using this basis and the results of Examples I and II of Section 6, it follows that the ortho-normal basis of the compact real form £c of C = Az constructed by the above prescription is Ш1 (гУЗ/6)А3 ш2 (гд/3/6)Л8 (1/л/2)(еа1 + e_Q1) = (гд/3/6)А2 (l/\/^)^(eCk;i C—cq) (i^3/6)Ai (1/^) (ea2 4” e — CK2 ) = (i^3/6)A7 (1/V^)^(^a:2 ^—0:2) (г^З/6)А6 (l/V^)(eai+a2 4“ e—(CK1+CK2)) (гУЗ/6)А5 (l/V^)^(^ai+a:2 — (cq +<22)) (гу/3/6)А4 As the basis of C = su(3) chosen in Example II of Section 4 was ap = гАр, p = 1,2,..., 8, it follows that the compact real form Cc of C = Az constructed according to Theorem IV above is precisely the real Lie algebra C = su(3), and the ortho-normal basis of Cc is provided by (%/3/6)ap, p = 1,2,..., 8 (i.e. the ortho-normal basis differs from the original basis merely by a common factor \/3/6). Consequently _B(ap,ag) = -(6/^3)2<5M = -126pq for p, q = 1,2,..., 8, precisely as noted in Example II of Section 4. Moreover, Theorem I above shows that with this basis the structure constants cpq are antisymmetric in all pairs of indices, which is certainly so, as (with ap = zAp) cr = —2fpqr, fpqr being defined in Equation (11.15). For the classical simple complex Lie algebras Ац Вц Ci and Di the corre- sponding compact real forms are (see Table 8.1): (a) for C = Ai, Cc = su(Z + 1), I = 1,2,... (b) for C = Bi) £c = so(2Z + 1), Z = 1,2,... (c) for £ = Ci) Cc = sp(Z), Z = 1,2,... (d) for £ = Dt) £c = so(2Z), Z = 3,4,... (The last-mentioned result applies also to £ = Dz (which is semisimple but not simple), the corresponding compact real form being £c = so(4) (which is also semi-simple but not simple). See, for example, Appendix G, Section 2(b), of Cornwell (1984) for details.) The isomorphic mappings between Ai, B± and Ci (noted in Section 7) imply that their compact real forms su(2), so(3) and sp(l) are isomorphic.
STRUCTURE OF SEMI-SIMPLE LIE ALGEBRAS 233 Similarly, so(5) and sp(2) are isomorphic as a result of the isomorphism be- tween B2 and C2, and su(4) and so(6) are isomorphic because A3 and D3 are isomorphic. This in turn implies homomorphic mappings between the corre- sponding linear Lie groups. For example, there exist homomorphic mappings of SU(2) onto SO(3) and of SU(4) onto SO (6). For the classical simple complex Lie algebras Ai, Bi, Ci and Di the corre- sponding non-compact real forms are (see Table 8.1): (a) C = Ai (for I = 1,2,...) has non-compact real forms: (i) £ = su(Z + 1 -p,p) for p = 1,..., [|(Z + 1)], (ii) £ = sl(i + 1,IR), (iii) C = su*(Z + 1) for I odd only; (b) C = Bi (for I = 1,2,...) has non-compact real forms: so(2Z + 1 — 2p, 2p) for p = 1,..., Z; (c) C = Ci (for I = 1,2,...) has non-compact real forms: (i) C = sp(p, I -p) for p = 1,..., [|Z], (ii) C = sp(Z,IR); (d) C = Di (for I = 1,2,...) has non-compact real forms: (i) C = so(2Z — 2p, 2p) for p = 1,..., [|Z], (ii) C = so(2Z - 2p + 1,2p + 1) for p = 1,..., [|Z], (iii) C = so*(2Z). Here [a] denotes the largest integer not greater than a. Realizations in terms of matrices with complex entries have been given for all the compact and non-compact real forms of the exceptional simple com- plex Lie algebras Eq, E?, E3, and G2 by Cartan (1914) and Gantmacher (1939b). In addition to the above simple real Lie algebras that are generated from simple complex Lie algebras, there also exist the non-compact simple real Lie algebras that are^generated from поп-simple complex Lie algebras of the form £i ® £2, where £i and £2 are isomorphic simple complex Lie algebras. The resulting real forms for the classical algebras are: (a) for C = Ai e Ai, C = sl(Z + 1, (C), I = 1,2,... (b) for C = Bi e Bi, C = so(2Z + 1, (C), Z = 1,2,... (c) for £ = Q Ф G, £ = sp(Z, (C), Z = 1,2,... (d) for C = Di e Di, C = so(2Z, (C), I = 3,4,... Again, Table 8.1 contains the specifications of these Lie algebras. The following sets of non-compact real forms are isomorphic:
234 GROUP THEORY IN PHYSICS (i) su(l,l) « sl(2,IR) « so(l,2) « sp(l,IR), (ii) so(3,2) « sp(2,IR), (iii) so(l,4) « sp(l,l), (iv) su(2,2) « so(4,2), (v) su(3,l) « so* (6), (vi) so(6,2) « so*(8), (vii) sl(4,IR) « so(3,3), (viii) su*(4) « so(5,l), (ix) su(Z + 1 -p,p) « su(p, I + 1 -p) for p = 1,..., [|(Z + 1)], (x) so(2Z + 1 — 2p, 2p) « so(2p, 2Z + 1 — 2p) for p = 1,..., Z; (xi) so(2Z — 2p, 2p) « so(2p, 2Z — 2p) for p = 1,..., [|Z], (xii) so(2Z — 2p + 1,2p + 1) « so(2p + 1,21 — 2p + 1) for p = 1,..., [|Z], (xiii) sp(p, I - p) « sp(/ - p,p) for p = 1,..., [|i], (xiv) sl(2,(C) « so(3,(C) « sp(l,(C) « so(3,l), (xv) so(5,(C) « sp(2,(C), (xvi) sl(4,(C) « so(6,(C). (The origins of these isomorphisms are discussed, for example, in Chapter 14 of Cornwell (1984).)
Chapter 12 Representations of Semi-simple Lie Algebras 1 Some basic ideas The key idea in the representation theory of semi-simple complex Lie algebras is that of the “weights”. These are introduced in Section 2, where it will be shown that they have very similar properties to the “roots” of Chapter 11, and are related to them in several ways. Section 3 demonstrates how every irreducible representation can be determined from its “highest weight”, and how the highest weights themselves are to be found. A detailed study is made in Section 4 of the irreducible representations of (the complexification of su(3)), which not only demonstrate all the general features, but are of great importance in physical applications. Finally, in Section 5 the related idea of Casimir operators will be introduced. For a description of methods available for the explicit determination of all the matrices of a representation, the calculation of Clebsch-Gordan series and coefficients, and the specification of irreducible representations by Young tableaux, see, for example, Chapter 16 of Cornwell (1984). The computer programme “SimpLieTJVf” developed by Moody et al (1996) is very useful for carrying out such calculations. The close connection between the representations of a semi-simple complex Lie algebra, the representations of its various real forms, and the represen- tations of the Lie groups associated with these real Lie algebras means that certain statements proved in one situation can be readily transferred to the others. The following two theorems are interesting not merely for the re- sults that are stated, but also because they can be proved using this line of argument. Theorem I Every reducible representation of a semi-simple real or complex Lie algebra is completely reducible.
236 GROUP THEORY IN PHYSICS Proof See, for example, Chapter 15, Section 1, of Cornwell (1984). Theorem II Every reducible representation of a connected semi-simple Lie group is completely reducible. Proof This follows immediately from the previous theorem and part (d) of Theorem IV of Chapter 9, Section 4. 2 The weights of a representation Consider^a representation Г of dimension d of a semi-simple complex Lie algebra £. Because Г provides a representation of the compact form £c of C (see Chapter 11, Section 10), on Cc the representation Г is equivalent to a representation by anti-Hermitian matrices. Thus, for any h E TL^ as ih E Cci V(ih) is diagonalizable and consequently Г (ft) is diagonalizable for all h E Tdjp. Moreover, the matrices Г (ft) for each h E TL may be diagonalized by the same similarity transformation. Henceforth it will be assumed that any necessary similarity transformation has already been applied, so that Г (ft) is a diagonal matrix for each h E TL. Consider the diagonal elements for some fixed j (j = 1,2,... ,d). As Г (ah + bhf) = аГ(Л) + ЬГ(Л') for all h,hf eTL and any complex numbers a and 6, it follows that V(ah + bh')jj = aTfhfjj + bT(h')jj. Thus the diagonal elements are linear functionals defined on TL. These linear functionals are called the “weights” of the representation, so that a d-dimensional representation possesses d weights, some of which may be iden- tical. In terms of modules (see Chapter 9, Section 4), suppose that ^2, • • •? 'фй form a basis of the carrier space V of the representation Г and that Ф(а) is the operator defined for each a E C by d = y^r(a)fcjV>fc fc=i for J = 1,2,..., d. Then, for each h E TL, as Г (ft) is diagonal, = Г(Д)^ (12.1) for j = 1,2,..., d. Thus for each j = 1,2,..., d, and for all h E TL, T(Ji)jj is an eigenvalue of the operator Ф(/г), the corresponding eigenvector being ifj. Let define the weight Xj corresponding to the jth position in the representation. Denoting the corresponding eigenvector тДу by V'CVf), Equation (12.1) becomes (12.2)
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 237 for J = 1,2,..., d. When this position is unimportant the index j may be omitted, so that an arbitrary weight of the representation will be denoted by A. If the weight A appears m(A) times in the representation, m(A) is said to be the “multiplicity” of A. If m(A) = 1 then A is described as a “simple weight” of the representation. Later examples will show that weights need not be simple, even for irreducible representations. With this simplification of notation, Equation (12.2) can be written as Ф(/^(А) = А(/г)^(А) (12.3) for all h E TL, 'ф(Х) being any eigenvector of Ф(/г) with eigenvalue A(/i). The multiplicity m(A) is then the dimension of the subspace of V spanned by the eigenvectors 'ф(Х). Example I The adjoint representation of £, Let the basis elements ai, u2,..., an of C be chosen to be hai , ha2,..., hai , together with for each a E A. Then, by Equation (9.36) with a = h, it follows that ad(/i) is diagonal for each h E TI, the diagonal element corre- sponding to haj being 0 and the diagonal element corresponding to ea being a(h). Thus each non-zero root a of C is a weight of the adjoint representation of C and, as dim£a = 1, each such weight is simple. The only other weight A of the adjoint representation is such that A(/i) = 0 for all h E TL, and this has multiplicity I. Example II Weights of a three-dimensional representation of the complex- ification £,(= A2) ofsvdfS) Consider the explicit three-dimensional representation of C = A2 constructed in Example II of Chapter 11, Section 4. By inspection, Al (^CKl ) g 4 X2 (hai) = g, Xs(hai) = 0, Ai(hQ2) = 0, ^2(^2) — g? A3 (ha2) g, so the three weights Ai, A2 and A3 are all simple. It is interesting to note for future reference that, as aj(hak) = (aj,ak), Equations (11.17) imply that for h == hai and ha2, AiW = |cti(/i) + |q2(^), A2(h) = Ai(h) - ai(h), > A3(h) = Ai(h) - Qi(h) - a2(fi), (12-4) from which it follows that Equations (12.4) are true for all h ETC. It is now possible to derive some simple but important results. Theorem I If A is a weight of a representation, then A + a is also a weight of the same representation for each a E A such that Ф(еа)^(А) 0.
238 GROUP THEORY IN PHYSICS Proof As [Л, ea] = a(h)ea, then [Ф(Л), Ф(еа)] = а(й)Ф(еа). Consequently, as Ф(7г){Ф(еа)т/>(А)} = [Ф(/г),Ф(е«)]^(А) + Ф(еа){Ф(/г)^(А)}, Equation (12.3) implies that Ф(Л){Ф(еа)^(А)} = (а(Д) + А(Д)){Ф(еа)^(А)}, so Ф(еа)^(А) is an eigenvector of Ф(/г) with eigenvalue a(h) + A(/i), provided that Ф(еа)'ф(Х) 0. Theorem II For any weight A of any representation of C and for any root a of Д, 2(A,a)/(a,a) is an integer. Proof Let Г be any representation of C and let A be any weight of Г. Con- sider the three-dimensional Ai subalgebra with basis Ha, Ea and E_a defined in Equations (11.21) and (11.22), whose commutation relations are given by Equations (11.23). As noted in Chapter 10, Section 3, every irreducible rep- resentation of this subalgebra is equivalent to a representation in which the matrix representing Ha is diagonal and has only integral diagonal entries. But Г provides a representation of this subalgebra, which must be equivalent to a direct sum of such irreducible representations. Thus the diagonal elements of Г (Ha) must all be integers, so X(Ha) is an integer. However, by Equation (11.21), X(Ha) = {2/(a,a)}X(ha) = 2(A,a)/(a,a), from which the quoted result follows. Theorem III Every weight A can be written in terms of the simple roots Qi, Q2, • • • ? Oil'. I where the coefficients /ij are all real and rational. Consequently A(/i) is real for each h e Proof Theorem II of Chapter 11, Section 7, shows that ai, Q2, • • •? (*i form a basis for W*, so any linear functional A can certainly be written in the form of Equation (12.5) for some set of complex numbers /zi,/z2? • • • ,///• All that has to be shown is that these are real and rational if A is a weight. However, for each a e Д, by Equation (11.21), i i 2(A, q)/(q,q) X(Ha) jijaj (Ha) jij(cv?, q)/(q, q). The previous theorem shows that the left-hand side is an integer, while The- orem X of Chapter 11, Section 5, demonstrates that 2(aj, a)/(a,a) is also an integer for each j = 1, 2,..., I. As this is so for every a e Д, Mi, М2, • • •, El must be real and rational. Finally, for h E A(/i) = /zyctj (7ь), which
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 239 is real, as aj(h) is real for all J = 1,2,...,/, by Theorem VI of Chapter 11, Section 5. Theorem IV For each root a e A and each weight A of a representation Г of £, define the linear functional SaX on TL by (SaA)(/i) = X(fi) - {2(A,o)/(o,a)}a(/i) (12.6) for all h e TL. Then SaX is also a weight of the representation Г with the same multiplicity as A, i.e. m(SaX) = m(A). Proof See, for example, Chapter 14, Section 2, of Cornwell (1984). The weight SaX will be recognized as being a Weyl reflection of the weight A (see Chapter 11, Section 9), so the above theorem states that any Weyl reflection of any weight is a weight with the same multiplicity. Repeated application gives the result that 5A is also a weight with same multiplicity as A for every element S of the Weyl group W. In displaying the symmetries of weights in connection with the Weyl group, it is particularly convenient to use the ortho-normal basis , Hi of TL introduced at the end of Chapter 11, Section 6, and discussed further in Chapter 11, Section 9. Then to every weight A there corresponds an l- component vector A given by Л = (А(Я1),А(Я2),...,А(Яг)), the /-component vector corresponding to 5A being denoted by 5A. Of course, by Equation (12.3), W) = А(Я^(А) (12.7) for J = 1,2,...,/, so the components of A are the simultaneous eigenvalues of Ф(я1),Ф(я2),...,Ф(яг). The concept of a “string” can be extended from roots to weights and produces a generalization of Theorem IX of Chapter 11, Section 5. Definition The а-string of weights containing X Suppose that a is a root of C and A is a weight of some representation of C. Then the “а-string of weights containing A” is the set of all weights of that representation of the form A + ka, where к is an integer. Theorem V Let a be a non-zero root of C and A a weight of some repre- sentation of C. Then there exist two non-negative integers p and q (which depend on a and A) such that A + ka is in the а-string containing A for every integer к that satisfies the relations — p < к < q. Moreover, p and q are such that p — q = 2(A,o)/(o,a). (12.8)
240 GROUP THEORY IN PHYSICS Proof See, for example, Chapter 15, Section 2, of Cornwell (1984). It should be noted that the weights in the а-string containing A need not have the same multiplicity. As noted in Chapter 11, Section 10, if the basis ai, a2, • • •, of C is taken to be ihaj (j = 1,2,...,/) together with (ea + e_a) and i(ea — e_a) (for all a e Д+), the structure constants are all real, and this basis is also the basis of the compact real form £c of £. Thus, if Г is a d-dimensional rep- resentation of £, define the matrices Г*(а?) for each of these basis elements aj by Г*(а7-) = {r(aj)}* (j = 1,2, ...,n). Extending this to the whole of C by defining Г*(^2™=1 Hjaj) to be /Ъ‘Г*(<^) for any set of com- plex numbers /zi, ^2, • • •, Hb it is obvious that these matrices Г* form a rep- resentation of C of dimension d, and Г* is irreducible if and only if Г is irreducible. The notation here reflects the fact that the matrices Г* pro- vide a representation of the compact real Lie algebra Cc that is the com- plex conjugate of that provided by Г. However, it should be noted that {г(Е"=1Л;^)}* = = Е"=1Л|Г*(ау), which is not equal to r*(E7=i mafi) unless .. ,yi are all real. Consequently the description of Г* as being the “complex conjugate” of the representation Г has to be applied with caution, but is nevertheless very useful and widely adopted. Theorem VI A is a weight of the representation Г of £ if and only if —A is a weight of Г*. Moreover, the multiplicity of A in Г is the same as that of —A in Г*. (That is, the weights of Г* are the negatives of those of Г.) Proof By the above construction, for any simple root aj of £, T*(iha ) = {r(zAaJ}* = {zT(/ia )}* = —г{Г(/1а )}*. But Theorem III above implies that r(/ia ) is a real matrix, so Г*(гЛа ) = — iT(hag) = — T(ihaj), from which it follows that Г*(Л) = — Г(Л) for all h eH. Consequently the weights of Г* are the negatives of those of Г and the multiplicity of —A in Г* is equal to that of A in Г. One further simple property of representations and their weights is worth noting: Theorem VII For any representation Г of any semi-simple complex Lie algebra C tr Г (a) = 0 for all a e C. In particular, for a = h this implies that ^m(A)A(/i) = 0, A where the sum is over all the weights A of Г. That is, on taking the mul- tiplicities into account, the sum of the weights of any representation of C is zero.
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 241 Proof With the Weyl canonical basis of C given in Chapter 3, Section 8, Equations (11.49) and (11.51) show that for any basis element ap of C there exist two basis elements a'p and ap such that ap = /ap[ap1 ap\, where /ap is some constant. Thus Г(ар) = /хр[Г(ар),Г(ар)]. However, as 1г{Г(ар)Г(а")} = tr{r(u")r(«p)}, then tr r(up) = 0, and hence tr Г (a) = 0 for all a e C. With a = Zi, as the diagonal elements of T(/i) are the quantities A(/i), it follows immediately that m(A)A(/i) = 0. This result provides the final link in the proof of Theorem II of Chapter 4, Section 3: Theorem VIII If Q is a non-compact simple Lie group then Q possesses no finite-dimensional unitary representations apart from the trivial representa- tions in which Г(Т) = 1 for all T e Q. Proof Suppose that Q has a non-trivial d-dimensional unitary representation. Then the associated representation of £, its real Lie algebra, consists of d- dimensional anti-Hermitian matrices, which, by the previous theorem, must be traceless. This representation must be faithful as £ is simple, so £ must be isomorphic to a subalgebra of su(d). As su(l) is trivial, d must be greater than 1. As su(d) is compact and semi-simple for d > 2, Theorem II of Chapter 11, Section 10, shows that C must be compact. This contradicts the conclu- sion of Theorem III of Chapter 11, Section 10, if Q is non-compact, so the initial assumption that Q possesses a non-trivial finite-dimensional unitary representation must be false. 3 The highest weight of a representation In this section the very useful concept of the “highest” weight of a repre- sentation will be introduced. Each irreducible representation is uniquely and completely specified by its highest weight, all of its properties, such as its dimension and the other weights being easily deducible from it. Moreover, there is a straightforward procedure for constructing every possible highest weight. It is convenient to define for the weights of a representation of £ a lexico- graphic ordering (see Chapter 11, Section 7) relative to the basis consisting of the simple roots qi,q2, • • • ,07 of £• Then a weight A is said to be “pos- itive” if A = !Т)аэ the first non-vanishing component of the set {/zi,/z2> • • • positive, Theorem III of the previous section having shown that all the members of this set are necessarily real. Further, if A and A' are two weights of a representation of £, one says that A > A' if and only if A — A' > 0. (The second theorem of Chapter 11, Section 7, shows that a root a of £ is positive relative to /?i, /?2, • • •, A if and only if h is positive relative toai,a2,...,«i, so no confusion can arise from this new choice of a basis for lexicographic ordering.)
242 GROUP THEORY IN PHYSICS Definition Highest weight A. of a representation If Л is a weight of a representation of £, such that Л > A for every other weight A, then Л is said to be the “highest” weight of the representation. Theorem I If Л is the highest weight of an irreducible representation of a semi-simple complex Lie algebra £, then (а) Л is a simple weight (i.e. m(A) = 1); and (b) every other weight A of the representation has the form i (12.9) where Qi,.., qi are a set of non-negative integers. Proof See, for example, Appendix E, Section 12, of Cornwell (1984). Definition Fundamental weights of a semi-simple complex Lie algebra C The I “fundamental” weights Ai(/i),A2(/i), ..., A^(/i) of C are the I linear functionals on H defined by i Xj(h) = ^(A~1)kjak(h) (12.10) k=l for all h e Ti. (Here ai, a2> • • •, are the simple roots of C and A is the Cartan matrix of £.) It follows that = sjk, (12.11) because, by Equations (11.45) and (12.10), i 2(Aj,CVfc)/(afc,^fc) = £2(Л-‘) pj (ppi ^k) /(pk? Gtk) p=l I = У )pj^kp = djk- p=l Conversely, Equation (12.11) implies Equation (12.10), so Equation (12.11) could equally well be taken as the definition of the fundamental weights. It is worth noting that Equation (12.11) implies that = for j, к = 1,2,..., Z, where Hak is defined in Equation (11.21).
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 243 Theorem II For every irreducible representation of a semi-simple complex Lie algebra C the highest weight Л can be written as i 7=1 where {ni, П2,.. ., тц} is a set of non-negative integers and Ai, Л2,..., Aj are the fundamental weights of Moreover, to every set of non-negative integers {п1,П2,. • • ,n/} there exists an irreducible representation of C with highest weight A given by Equation (12.12), and this representation is unique up to equivalence. Proof See, for example, Chapter 15, Section 3, of Cornwell (1984) for a proof of the first part, for references to complete proofs, and a brief history. Example I The fundamental weights of the complexifications C = Ai and A2 o/su(2) and su(3) respectively From Example IV of Chapter 11, Section 7, for C = Ai, A-1 = [1/2], while for С = A2 A-i Г 2/3 1/3 ' [ 1/3 2/3 _ ' Thus the fundamental weight of C = Ai is Лг = (1/2)сц, and the fundamental weights of С = A2 are Ai = (2/3)ai + (1/3)а2? 1 A2 = (l/3)ai + (2/3)a2. J The irreducible representation of C with highest weight A specified by Equation (12.12) will be denoted by Г({п1,п2,... , n/}). Its dimension d is given by the following remarkably simple formula, known as “Weyl’s dimen- sionality formula”. Theorem III The dimension d of the irreducible representation of C with highest weight A is given by d= П {(Л + МЛМ}, (12-13) where <5 = | £а6д+ a- Proof Weyl (1925,1926a,b) deduced this result from his character formula. For details see, for example, Jacobson (1962), Samelson (1969), Humphreys (1972), or Varadarajan (1974).
244 GROUP THEORY IN PHYSICS Example ll Dimensions of the irreducible representations of the complexi- fications C = Ai and of su(2) and su(3) respectively For C = Ai, as Д+ consists only of Qi, Equation (12.13) involves only one product, and shows that the dimension d of r({ni}) is given by d = ni + 1. (Contact with the results of Chapter 10, Section 3, is obtained if ni is identified with 2/, for j = 0, 1, For С = A2, A+ = {aq,CE2,aq + a2}, so the dimension d of Г({п1,п2}) is given by d = {ui + l}{n2 + l}{(l/2)(ni + n2) + 1}. (12.14) In some very straightforward cases the other weights of an irreducible representation can be obtained from the highest weight by a simple application of Theorem IV of Section 2. For cases that are more complicated, but for which every weight is simple, Theorem V of Section 2 provides an elementary method for the deduction of the other weights. However, in general, some of the weights will not be simple, and a method is required for evaluating the multiplicities. “Freudenthal’s recursion formula”, which is stated in the next theorem, provides the necessary information. In this theorem the “level” q of a weight A = Л — qj(*j is defined to be q = Y^j=i Qj- I*1 particular, the highest weight Л is of level zero and is the only weight of this level. Theorem IV Consider an irreducible representation of C that has highest weight Л. Then the multiplicity m(A) of a possible weight A = Л — Qjaj (with Qi, q2, ..., q/ all non-negative integers) is given by {(A+<5,A+<5) — (A+<5,\-\-8)}m(X) = 2 m(A+to)(A+A;a, a), (12.15) aE Л-]- к where the second sum on the right-hand side is only over those values of к for which A + ka is a weight of the representation whose level is less than that of A, and where 8 = | a- In particular, if m(A) = 0 then A is not a weight of the representation. Proof See Humphreys (1972) or Jacobson (1962). As m(A) = 1, Freudenthal’s recursion formula allows the multiplicities of the weights to be obtained first for level 1, then for level 2, and so on. For example, every level-1 weight has the form A = A — aj, where aj is some simple root. The only non-zero term on the right-hand side of Equation (12.15) occurs with a = ce7, and к = 1, and is 2m(A)(A,aj) = 2(A,aj). Similarly, every level-2 weight multiplicity is given by Equation (12.15) in terms of the multiplicities of weights of level 1 and 0, and so on. It should be noted that if the linear functional A = A — ^=1 aj is not a weight, then Equation (12.15) gives m(A) = 0. Consequently these formulae
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 245 provide a self-contained and exhaustive procedure for finding all the weights and their multiplicities, by simply investigating every linear functional of the form Л — aj-> f°r everY set of non-negative integers {qi,q2, • • • ,Qi} for increasing values of q = Y^j=i stopping when the sum of the multiplicities reaches the value d. 4 The irreducible representations of £ = A2, the complexification of £ = su(3) As shown in Example I of Section 3, the fundamental weights of C = are Ai = (2/3)ai + (l/3)a2 Л2 = (l/3)«i + (2/3)a2- The corresponding two-component vectors are then Ai = (2/3)ai + (l/3)a2 Л2 = (l/3)«i + (2/3)a2, from which it follows from Example II of Chapter 11, Section 6, that Al = (1/6)(A 1), A2 = (l/6)(0,2). Weyl’s dimensionality formula implies that the dimension d of the irreducible representation Г({п1,п2}) is given by Equation (12.14), that is, d = {ui + l}{n2 + l}{(l/2)(ni + n2) + 1}. (12.16) If ni n2 a knowledge of the weights of Г({п1,п2}) yields those of Г({п2,П1}) immediately, for, if дцои + //2^2 is a weight of Г({п1,п2}), then /zia2 + //2^1 is a weight of r({n2,ni}) with the same multiplicity. Moreover the weights of T({n2,ni}) are the negatives of those of T({ni,n2}). (This will be clear from inspection of the examples given below. A general proof can be found, for example, in Chapter 15, Section 4, of Cornwell (1984).) In the elementary particle literature it is common to label each irreducible representation by its dimension d. When n2 Equation (12.16) shows that the irreducible representations Г({п1,п2}) and Г({п2,П1}) have the same dimension, one being the “complex conjugate” of the other in the sense of Section 2. In this case one representation is denoted by {d} and the other by {d*}, the usual convention being ^{^,712}) = I Ц (12.17) 1 J/ [ {d } 11 ni < n2. v 7 As shown in Example I of Chapter 11, Section 9, the Weyl group W of С = A2 consists of six elements S. As 5A is a weight with the same multiplicity as A for each S E W, the weights of an irreducible representation may be arranged in sets of six, three or one, the first occurring when A is not
246 GROUP THEORY IN PHYSICS A (H2) =-i3»l4“2 , A = g(/3,l) = tal*3°2 *~A (H|) ,,Se2SO|A = l(0,-2) =-i0|-|a2 Figure 12. 1: Weight diagram of the irreducible representation {3} (specified by «1 = 1, n2 = 0) of £ = A2. on a reflection line, the second when Л is on a reflection line and A 0, and the third when A = 0. The full sets of weights of the lower-dimensional irreducible representations of £ = A2 will now be examined. (а) Г({0,0}) = {1}: With m = n2 = 0, Equation (12.16) gives d = 1. Consequently this ir- reducible representation has only one weight, namely the highest weight A = 0. (b) Г({1,0}) = {3}: With щ = 1, n2 = 0, Equation (12.16) gives d = 3. The highest weight is Л = Ai, so A = Ai = |(%/3,1), which lies on a reflection line. Consequently this irreducible representation has two other simple weights 5aiA and 5a25Q1A, for which, by inspection of Figure 11.4, 5aiA = |(—\/3,1) and 5Q25aiA = |(0, —2). This implies that 5aiA = A — ai and Sa2SaiA = A — (ai +a2), so the weights of this irreducible representation are |aq + |a2, — |ai + |a2 and — |ai — |a2. The weight diagram is given in Figure 12.1. (с) Г({0,1}) = {3*}: The weights of {3*} are + |a2 (= Л = Л2), — |a2 and — |«i — |a2. The weight diagram is given in Figure 12.2. (d) Г({1,1}) = {8}: With m = n2 = 1, Equation (12.16) gives d = 8. The highest weight is A = Ai + Л2 = ai + a2, so A = ai+n2 = |(%/3,3). As this does not lie on a reflection line, there are five other simple weights obtained from it by Weyl reflections, which may be found by inspection using Figure 11.4. They are |(—\/3,3)(= a2), |(\/3, — 3)(= — a2),
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 247 л л (H2) " A=| (0,2) = 5»i* i°2 —(H I) S^A^US.-I) =-!»|-i“2 =5°|-3“2 Figure 12. 2: Weight diagram of the irreducible representation {3*} (specified by ni = О, П2 = 1) of £ = A2. |(-А-3)(= -ai - a2), |(2V3,0)(= aj and |(-2д/3,0)(= -ai). As only six of the eight weights are thereby accounted for, all that can remain is a weight 0 of multiplicity 2. Clearly this representation is the adjoint representation. The weight diagram is given in Figure 12.3. (e) r({2,0}) = {6}: With ni = 2, n2 = 0, Equation (12.16) gives d = 6. The highest weight is A = 2Ai, so A = 2Ai = |(2%/3,2), which lies on a reflection line. Exactly as for Г({1,0}) = {3} , Weyl reflections then produce two more simple weights |(—2%/3,2) and |(0, —4). This leaves three other weights to be determined. However, for the aq-string containing A, Equation (12.8) gives p — q = 2(A,ai)/(ai,ai) = 2{|(ai,ai) + |(a2, ai) }/(ai, ai) = |Ац + |Ai2 = 2. As A is the highest weight, it follows that q = 0, implying p = 2, and giving as the aq-string containing A the set {A, A — ai,A — 2aq}. But A — 2«i = |(—2%/3,2), which has been obtained already. However, A — aq = |(0,2), which is new. Weyl reflections applied to this weight then produce |(—\/3, — 1) and |(\/3, —1). Thus the weights are |ai + |a2, — joq + |a2, — jai — |a2, |aq + |a2, |ai — |a2 and — |aq — |a2, the weight diagram being given in Figure 12.4. (f) Г({0,2}) = {6*}: The weights of {6*} are the negatives of those of {6} and are obtained by interchanging the coefficients of aq and a2. The weight diagram is given in Figure 12.5. (g) Г({3,0}) = {10}: The argument is essentially the same as for the {6}, producing the ten simple weights shown in Figure 12.6.
248 GROUP THEORY IN PHYSICS A (H2) Sa, A =g (-/3,3) = a2 A = g (/3,3) = al + a2 S<z, S^S,, A 4 (-/3,-3) = -a, -a2 Sff2Sffl A = |(/3,-3) = -a2 Figure 12. 3: Weight diagram of the irreducible representation {8} (specified by ni = П2 = 1) of £ = A2. (Here о indicates a weight of multiplicity 2.) Л(Н2) S,,A=j(-2/3,2) =-|a|.|a2 S<7| S^2( A -Q|) = |(-/3,-l) _ 2 _ I _ --5a,-5a2 S^S^A--i(0,-4) 2 4 a. :-3al’3a2 Figure 12. 4: Weight diagram of the irreducible representation {6} (specified by ni = 2, n2 = 0) of £ = A2.
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 249 . A (H2) A = j(0,4) =!°i4a2 £(-/3,i). = -з°1*-э°2 £(-2/3,-2) «„ 2- заГза2 .£(/3,1) = зв|+за2 -----------------A (H ।) £(0,-2) £(2/3,-2) I 2 2 _ 2 _ = "заГза2 = 3al’3a2 Figure 12. 5: Weight diagram of the irreducible representation {6*} (specified by ni = О, П2 = 2) of £ = Л.2- > A (H2) £(-з/з,з) £(-/з,з) = ~ Q ।+ Q 2 = Q2 £(/з,з) а=£(з/з,з) = О|+а2 =2ара2 £(-2/3,0) = -«l (0,0) £(2/3,0) >Л(Н|) = 0 =в| £(-/3,-3) = -ага2 £(/з,-з) = -а2 , £(о,-б) = -О|-2а2 Figure 12. 6: Weight diagram of the irreducible representation {10} (specified by ni = 3, n2 = 0) of £ = A2.
250 GROUP THEORY IN PHYSICS M(H2) ’a4(o,6) = “lt2o2 |(-/3,3) = a2 g(/3,3) = al + a2 ±(-2/3,0) = -“l (0,0) ±(2/3,0) л(н1) - О =в| • • ±(-3/3,-3) ±(-/3,-3) = _ 2 Q| ~ (>2 =_Q|“Q2 ±(/3,-3) ±(3/3,-3) - “(>2 = ® I ~ Figure 12. 7: Weight diagram of the irreducible representation {10*} (specified by ni = 0, n2 = 3) of £ = A2. (h) Г({0,3}) = {10*}: The weights of {10*} are the negatives of those of {10} and are obtained by interchanging the coefficients of ai and a2. The weight diagram is given in Figure 12.7. As will become clear in Chapter 13, Section 3, the only higher-dimensional irreducible representations that are of interest in the su(3) symmetry scheme for hadrons are those for which (ni — n2) is divisible by three. Only one example will be considered in detail. Its mathematical interest lies in the fact that it provides the first case of a non-zero weight that is not simple. (i) Г({2,2}) = {27}: The highest weight is A = 2Ai + 2A2 = 2ai + 2q2. Application of Freudenthal’s recursion formula (Equation (12.15)) shows that the weights of level 1 are A — ai(= Qi + 2q2) and A — a2(= 2cei + a2), and that both are simple. Now consider the only possible weight of level 2, namely A = ai + a2. As 6 = |{ai + a2 + (ai + a2)} = ai + a2, Freudenthal’s recursion formula (Equation (12.15)) gives {(3(oq + a2), 3(oq + a2)) — (2(oq + a2), 2(oq + a2))}m(ai + a2) = 2{m(2«i + 2a2)(2(aq + a2), + a2) +m(ai +2a2)(ai + 2q2,q2) +m(2«i + q2)(2qi +a2,ai)},
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 251 £(-2/3,6) - 2 a g , Я(н2) 5 (0,6) * A =£(2/3,6) =Q| + 2a2 =2q| + 2q2 • ® £(-з/з,з) £(-/з,з) £(/з,з) £(з/з,з) - “Q| +Q2 =Q 2 = Q|+Q2 =2Q|+Q2 * 0 Si 42) * * । 11 । \ £(-4/3,0) £(-2/3,0) & . . Л (H|) (0,0) £(2/3,0) £(4/3,0) =-2a, =-O| = 0 - a। - 2 a । • ® £(-з/з,-3) £(-/з,-з) £(/з,-з) £(з/з,-3) = -2 ap a2 ra2 - -Q2 =a|-a2 £(-2*/3,-6) ’£(0,-6) £’(2/3,-6) z-2a|~2a2 --Q|-2a2 ~ ~ 2(>2 Figure 12. 8: Weight diagram of the irreducible representation {27} (specified by ni = 2, П2 = 2) of C = Az- (Here о and @ indicate weights of multiplicity 2 and 3 respectively.) and, as (ai, ai) = |, (ai, az) = — | and m(2a:i +2^2) = m(ai + 2az) = m(2ai + az) = 1, this gives m(ai + a2) = 2. Repetition of this type of argument gives the weight diagram of Figure 12.8. Some useful Clebsch-Gordan series for Az are: {3} 0 {3*} « {8}e{i}> {3}0{3}0{3} « {io}e2{8}e{i}, . > {8} 0 {8} « {27} e {10} e {10*} e 2{8} e {1} (12.18) (In the latter two Clebsch-Gordan series the expressions “2{8}” indicate that in each case the irreducible representation {8} occurs with multiplicity 2. The Clebsch-Gordan coefficients for Az have been discussed in detail by de Swart (1963). For an introduction to the derivation of the Clebsch-Gordan series and coefficients for Az see, for example, Chapter 16, Sections 5 and 6, of Cornwell (1984).) 5 Casimir operators In the analysis given in Chapter 10 of the representation theory of the su(2) (and so(3)) Lie algebras, a very important part was played by an operator A2
252 GROUP THEORY IN PHYSICS defined in Equation (10.16). As noted in Equations (10.22), when written in the language of angular momentum theory, this is (l/Й2) times the operator J2. Casimir (1931) showed how a similar operator can be defined for any semi- simple Lie algebra. His prescription produced an operator of second order in the basis elements. This is appropriately called the “second-order Casimir operator” and is denoted by С2. For I > 1 semi-simple Lie algebras possess other similar operators constructed using higher-order products. These will be called the “higher-order Casimir operators” The basic properties of the second-order Casimir operator are summarized in the following theorem. Theorem I Let eq, <22, • • • , be a basis of a semi-simple Lie algebra C (either real or complex) and V be the carrier space of some representation Г of C whose linear operators are Ф(а) (aE£). Then: (a) The second-order Casimir operator C2 specified by n C2=^ (12.19) p,Q=l is well defined and is independent of the choice of basis eq, a2> • • •, (Here В is the n x n matrix with elements Bpq = B(ap, aq).) In partic- ular, for a basis of a complex (or compact real) semi-simple Lie algebra C such that B(ap, aq) = —Spq n С2 = -^Ф(ар)2. (12.20) P=1 (b) (?2 commutes with Ф(а) for all a e £. (c) If Г is an irreducible representation of £, then C2 is a constant times the identity operator. If Г has highest weight A, this constant will be written as 62(A). Then, for any E V, C2^ = С2(Л)^, so that 62(A)) may be described as the “eigenvalue of C2 in the irre- ducible representation with highest weight A”. (d) This eigenvalue is given by the expression C2(A) = (A, A + 25), (12.21) where <5 = 1 Ya- (12-22) ckGA+
REPRESENTATIONS OF SEMI-SIMPLE LIE ALGEBRAS 253 (e) For the adjoint representation ad, С2(Л) = 1. Proof See, for example, Chapter 16, Section 2, of Cornwell (1984). Example I Second-order Casimir operator of £, = su(2) (= so(3)) With the operators Ai, A2 and A3 introduced in Chapter 10, Section 3, in terms of the basis eq, n2, n3 of su(2) given in Equations (8.31), Ap = — гФ(ар) (p = 1,2,3). However, with this basis B{apiaq) = —28pq (see Example I of Chapter 11, Section 2), so (B-1)pg = — ^8pq. Thus, from Equation (12.19), C2 = (A2 + A\ + A2)/2 = A2/2. (12.23) In terms of the angular momentum operator J2, Equations (10.22) and (12.23) give C2 = (1/2/z2)J2. (12.24) For the irreducible representation with highest weight Л = niAi = jniai, as 8 = |ai for C = Ai, Equation (12.21) gives С2(Л) = ((l/2)n1a1, (l/2)n1a1 + oq) = (l/4)m(ni + 2)(aq, oq) = (l/8)ni(ni+2), as (oq, ai) = | (see Example I of Chapter 11, Section 4). Then, with щ = 2j C'2(A)=j(j + 1)/2, so that the eigenvalues of A2 are j(j + 1) (these being, of course, exactly the values found in Chapter 10, Section 3). Example II Second-order Casimir operator of C = su(3) (and of C = A2) For C = su(3) (and C = A2) 8 = qi+q2 and Ai = |ai + |a2, Л2 = |ai + |a2. Thus, for the irreducible representation Г({п1,п2}) with highest weight A = niAi +n2A2, A + 28 = {(2/3)ni + (l/3)n2 + 2}aq + {(l/3)ni + (2/3)n2 + 2}q2. Hence, by Equations (11.17) and (12.21), С2(Л) = (n2 + П2 + П1П2 + 3ni + 3n2)/9. (As expected, <72(A) = 1 for the adjoint representation Г{1,1}).) As an irreducible representation of a complex semi-simple Lie algebra C is determined by its highest weight A, which itself depends on I (integer) parameters щ, n2,..., n/, one would not expect that specification of <72(A) would be sufficient to fix the irreducible representation. This expectation was
254 GROUP THEORY IN PHYSICS confirmed by Racah (1950, 1951), who showed that, if C has rank I that is greater than 1, then C possesses a set of higher-order Casimir operators whose eigenvalues do completely specify irreducible representations. See also Gruber and O’Raifeartaigh (1964), Okubo (1977), and Englefield and King (1980) for further work in this area.
Chapter 13 Symmetry schemes for the elementary particles 1 Leptons and hadrons The starting point of all symmetry schemes for the elementary particles is the observation that there appear to be four fundamental interactions between these particles. These are, in decreasing order of strength: (i) the strong interaction, first discussed in the context of the binding of the nucleons in the nucleus; (ii) the electromagnetic interaction; (iii) the weak interaction (which, for example, is responsible for beta decay); (iv) the gravitational interaction. (Recent developments suggest that these interactions may not be distinct, but may be manifestations of a single fundamental interaction.) In terms of these four interactions it is possible to divide the observed particles into two major categories, the “leptons” (and “antileptons”) which never experience strong interactions, and the “hadrons” (and “antihadrons”) which, at least in some circumstances, interact through the strong interaction. In addition there are the “intermediate” particles that are the carriers of the interactions (of which the photon, W± and Z° have actually been observed at the time of writing). The category of hadrons can be further divided into two classes, those whose intrinsic spin j is an integer (= 0,1, 2,...) being called “mesons” and the others (for which j = |, |,...) being referred to as “baryons”. The “lepton number” and “baryon number” may then be defined for all the presently observed particles by 1, -1, 0, L = if the particle is a lepton, if the particle is an antilepton, for any other type of particle,
256 GROUP THEORY IN PHYSICS and B=l -1, I °’ if the particle is a baryon, if the particle is an antibaryon, for any other type of particle. 2 The global internal symmetry group SU(2) and isotopic spin The object of this section is to introduce the concept of isotopic spin and present the basic ideas in such a way that they are easily generalizable to other internal symmetries. Consider first the case of the proton (p) and the neutron (n). Their rest masses mp and mn are almost identical (mpc2 = 938.3 MeV, mnc2 = 939.6 MeV), and their interactions with each other (that is p-p, p-n and n-n) are independent of how they are paired (provided that they are always coupled into the same state of total spin and parity). It is as though there is only one particle, the “nucleon” (N), which might exist in either of two states, one corresponding to the proton and the other to the neutron, these two states being distinguished only by an electromagnetic field. This is a similar situation to that of an atom in a state with orbital angular momentum I subjected to a small magnetic field H. As noted in Chapter 10, Section 6, if all the effects of the electrons’ spins are neglected (including degeneracies caused by them) then the energy eigenvalue of a state with angular momentum I is (2/ + l)-fold degenerate in the absence of the field, but splits into (21 + 1) different values when the field is applied. Naturally one does not regard these as being (2Z +1) different atoms, but rather they are thought of as (21 + 1) different states of the same atom. The correspondence between these two situations depends on the connec- tion between energy and mass in the special theory of relativity. It leads to the proposal that the nucleon N should be assigned an “isotopic spin” I with value | (this value being chosen so that 21 + 1 = 2, so that it can exist in 21+1 (=2) different states, one corresponding to the proton and one to the neutron. Further, it is suggested that in the absence of electromagnetic inter- actions (that is, in a universe with no electromagnetic interactions) the proton and the neutron would be identical, and each of their interactions, which are all “strong”, would also be identical. Developing this further, one can introduce three self-adjoint linear opera- tors Zi, Z2 and Z3 that satisfy the commutation relations р2,2з] Рз,^] г?з, Hi, il2- (13.1) That is, more briefly, Epqr^r, (13.2)
ELEMENTARY PARTICLE SYMMETRY SCHEMES 257 for p, q = 1, 2, 3. These are identical to the commutation relations in Equation (10.9). Indeed one can write, by analogy with Equation (10.7), — ^Ф(^р)? (13.3) (for p = 1,2,3), where eq, and are basis elements of the real Lie algebra su(2). The analogy may be extended so that Zi, and Z3 may be regarded as being operators corresponding to the measurement of the “components” of isotopic spin in three mutually perpendicular directions in an “isotopic spin space”. Introducing the linear operator Z2 by T2 = (zj2 + (Z2)2 + (Z3)2 (13.4) (by analogy with Equation (10.16)), it is clear that all the properties of the operators Ai, A2, A3 and A2 considered in Chapter 10, Section 3, apply equally to the operators Zi, Z2, Z3 and Z2. In particular, the operator Z2 has eigenvalues of the form 7(7+1), where 7 takes one of the values 0, |, 1, |,.... This quantity I is then regarded as the “isotopic spin”, and the possible values of its “component in the third direction in isotopic spin space” associated with the operator Z3 are given by the eigenvalue 7з of Z3, which assume any of the (27 + 1) values 7,7 — 1,..., —7. The simultaneous eigenvector of Z2 and Z3 with eigenvalues 7(7 + 1) and 7з may be denoted (by analogy with Equations (10.23) and (10.24)) as ?/J3, so that ^ + 1)<, W3. (13.5) Indeed, for any element a of the su(2) Lie algebra spanned by the basis ele- ments ni, a2 and a3 of Equation (13.3), Ф(а)^/3 = 52 D^a>)^L П=-1 (13.6) where DJ is the irreducible representation of su(2) introduced in Chapter 10, Section 3. It may also be assumed that all these isotopic spin operators commute with all the operators corresponding to space-time transformations, so that the state vector of each hadron is the direct product of a function of space- time and one of the vectors . Each value of 7з corresponds to a particle, the set of (27 +1) particles associated with a particular value 7 being said to form an “isotopic multiplet”. It is implied from Equation (13.6) that the vectors form the basis of the (27 + l)-dimensional irreducible representation DJ of su(2). In the case of the nucleons, the proton is assigned the value 7з = | and the neutron the value 7з = — |. These considerations imply that all the particles in an isotopic multiplet must have the same intrinsic spin and parity, as well as the same baryon number (and other quantum numbers, such as strangeness and charm).
258 GROUP THEORY IN PHYSICS isotopic multiplet В Y I I3 Q particle -1 -1 7Г-, p~ 7Г, p 0 0 1 0 0 Tp,p° к 1 1 7T+,P+ к, K* 0 1 1 2 1 2 1 2 0 1 №,K*° K+,K*+ 77, ф, Ш 0 0 0 0 0 N 1 1 1 2 1 2 1 2 0 1 n p _3 2 -1 Д- Д 1 1 3 2 < 1 2 1 2 0 1 A° Д+ к 3 2 2 Д++ Л 1 0 0 0 0 Л0 -1 -1 £- S 1 0 1 0 0 S° к 1 1 s+ 2 1 -1 1 2 1 2 1 2 -1 0 s° n 1 —2 0 0 -1 Table 13.1: Isotopic spin, hypercharge and baryon number assignments of some of the most important hadrons. It is assumed that all hadrons can be classified within this scheme. His- torically, the earliest particles to be incorporated in this scheme after the nucleons were the three pions 7г+, 7г° and 7г“, which were assigned by to an isotopic multiplet with I = 1, the values of /3 being 1, 0 and —1 respectively. For both the nucleons and the pions the electric charge Qe of the particle is given by Q = IS + ±B, (13.7) where B, the baryon number introduced in the previous section, has value 1 for the nucleons and 0 for the pions. In fact Equation (13.7) holds only for all non-strange and un-charmed hadrons, the generalization for strange hadrons being given later in Equation (13.9). A list of isotopic spin assignments for some of the most important hadrons is contained in Table 13.1.
ELEMENTARY PARTICLE SYMMETRY SCHEMES 259 The essential assumption underlying the above analysis is that the SU(2) group corresponding to the Lie algebra su(2) is the invariance group of the strong interaction Hamiltonian. This implies that this Hamiltonian and the corresponding T-matrix are irreducible tensor operators transforming as the one-dimensional identity irreducible representation. This enables predictions to be made of ratios of cross-sections and similar dynamical quantities using the Wigner-Eckart Theorem and the Clebsch-Gordan coefficients for su(2). (See, for example, Chapter 18, Section 2, of Cornwell (1984) for an introduc- tory detailed analysis). 3 The global internal symmetry group SU(3) and strangeness The present account of the su(3) symmetry scheme for hadrons is intended to introduce its most significant features and to emphasize the role of the group- theoretical and Lie-algebraic arguments developed in earlier chapters. There have been many long and detailed reviews of the su(3) scheme, and to these the reader is referred for more specific information on certain topics. The fol- lowing list gives a selection of these: Behrends et al. (1962), Behrends (1968), Berestetskii (1965), Carruthers (1966), Charap et al. (1967), de Franceschi and Maiani (1965), de Swart (1963, 1965), Dyson (1966), Emmerson (1972), London (1964), Gatto (1964), Gell-Mann and Ne’eman (1964), Gourdin (1967), Kokkedee (1969), Lichtenberg (1978), Mathews (1967), Ne’eman (1965), O’Raifeartaigh (1968) and Smorodinsky (1965). The concept of the strangeness quantum number was developed out of the “associated production” hypothesis of Pais (1952) to explain the observation that certain hadrons are created by strong interactions, but decay through the weak interaction (Gell-Mann 1953, Nakano and Nishijima 1953, Nishi- jima 1954, Gell-Mann and Pais 1955). The proposal was that every hadron possesses a “strangeness quantum number” 5, which is assumed to be an inte- ger, and that production or decay takes place through the strong interaction if and only if the quantity AS, defined by bAS = {sum of initial values of S} — {sum of final values of S}, (13.8) is zero, that is, if and only if strangeness is additively conserved. The gener- alization of Equation (13.7) is given by the “Gell-Mann-Nishijima formula” Q = 13 + (1/2)B + (1/2)S, (13.9) (which is consistent with Equation (13.7), as nucleons and pions are assigned the value S = 0). This formula indicates that it is more convenient to work with the “hypercharge” Y defined by Y = В + S, (13.10)
260 GROUP THEORY IN PHYSICS in terms of which Equation (13.9) becomes Q = h + (1/2)K (13.11) Assuming that В is conserved, the selection rule for strong interactions is that they act if and only if NY = 0. (13.12) Table 13.1 gives the assignment of hypercharge for some of the most important hadrons. It is natural to assume that the possible values of Y are eigenvalues of a self-adjoint linear operator y. As all the particles in an isotopic multi- plet are assumed to have the same value of У, and as Y is assumed to be simultaneously measurable with /3, it is necessary that [J,2p]=0 (13.13) for p = 1, 2,3, implying that [J,Z2]=0 (13.14) as well. Moreover, Y is assumed to be unchanged by space-time transforma- tions. As Y is an integer for all observed particles, it is reasonable to assume that iy is the basis element of a real Lie algebra that is isomorphic to a u(l) real Lie algebra (the corresponding basis element of u(l) being [г]).) (As the unitary irreducible representations of the corresponding Lie group U(l) are all one-dimensional and are given by Ги(1)([е“]) = where p = 0, ±1, ±2,..., and where x is real, it follows that the corresponding irreducible representations of u(l) are such that rU(i)(W) = M Then the eigenvalues of У take the values p = 0, ±1, ±2,....) Consequently the set consisting of £Ti, H2 and H3 forms the basis of a u(l) ф su(2) real Lie algebra (the commutation relations being Equations (13.1) and (13.13)). However, this alone does not imply any correlation between the eigenvalues of У and Z3. To obtain this it is necessary to make the further assumption that this u(l) ф su(2) Lie algebra is the proper subalgebra of a larger real Lie algebra. The natural candidates to consider are the rank-2 compact semi-simple real Lie algebras, because all their relevant properties are known. Being com- pact, all the finite-dimensional representations of their associated Lie groups are equivalent to unitary representations, which the isotopic spin arguments of the previous section suggest to be a desirable feature. A rank-2 algebra is appropriate because it can accommodate two mutually commuting operators
ELEMENTARY PARTICLE SYMMETRY SCHEMES 261 (corresponding to У and Z3) in its Cartan subalgebra. The non-simple can- didate su(2) Ф su(2) can be eliminated because it would leave the values of У and Z3 unrelated, so the choice is narrowed to the rank-2 compact simple real Lie algebras. The analysis of Chapter 11 shows that there are only three non-isomorphic algebras with the required properties, namely su(3) (the com- pact real form of A2), so(5) (which is the compact real form of B2 and C2, as these are isomorphic), and the compact real form of G2. It is now clear that the scheme based on su(3) agrees well with experimental observation, and that this is not the case for the schemes based on the other algebras. Consequently the present account will be confined solely to the su(3) scheme. Even with su(3) selected as being the appropriate algebra, there still remains the question of the precise relationship of У and Z3 to the basis elements of the Cartan subalgebra of A2. This is equivalent to the problem of assigning particles to multiplets, which was resolved by Gell-Mann (1961, 1962) and Ne’eman (1961), and which will be discussed shortly. The basic philosophy of the su(3) scheme is that У and Z3 are members of the Cartan subalgebra of A2, and their eigenvalues Y and I3 are determined by the weights of the irreducible representations of A2. The set of hadrons corresponding to a particular irreducible representation is said to form a “uni- tary multiplet” and the hadrons involved are assumed to be identical apart from their values of У, I3 and I, so that they all have the same spin, par- ity and baryon number. Moreover, it is assumed that in an ideal universe there is only one type of interaction, the strong interaction, and that all the particles in a unitary multiplet have exactly the same mass. At this point there is a problem, because it will become apparent that in the real world the particles in a unitary multiplet have masses that are only very roughly equal. The situation is quantitatively quite different from that in the iso- topic spin scheme, where the masses within an isotopic multiplet differ by at most a few per cent, and where the difference can be attributed to the weaker electromagnetic interaction. It is clear that the considerable mass-splittings between isotopic multiplets in a unitary multiplet cannot be attributed to the electromagnetic interaction, so that it is necessary to make the assumption that there are two types of strong interaction. The weaker version, which will be called the “medium-strong interaction”, is assumed to be responsible for these mass-splittings. The stronger version will still be referred to as “the” strong interaction The first priority is to establish the relationship of J2, Zi, Z2 and Z3 to the basis elements /io,2, e_ cq, and of the Weyl canonical basis of A2. The requirements are that: (i) Zi, Z2, Z3 satisfy the commutation relations in Equations (13.1); (ii) У satisfies the commutation relations in Equation (13.13); and (iii) if any particle in a unitary multiplet has integral electric charge (that is, if Q is an integer), then all the particles in the multiplet must have integral electric charge.
262 GROUP THEORY IN PHYSICS 2 3 Figure 13. 1: Values of /3 and Y for the irreducible representation^}. These requirements lead to the assignments: У = |Ф(Яа1) + |Ф(Яа2) = 2Ф(Ла1) + 4Ф(Ла2) = 2Ф(Я2), Т, = |Ф(ад + |Ф(Я_а1) = ^{Ф^) - Ф(е_а1)}, 12 = -^Ф(ЯО1) + И(Я_а1) = -г072{Ф(еа1) + Ф(е_а1)}, Т3 = ±Ф(Яа1) = ЗФ(Ла1) = ^Ф(Я1), (13.15) (where Н± and Н2 are the ortho-normal basis elements of the Cartan sub- algebra of A2 of Example II of Chapter 11, Section 6). (The detailed argument that leads to Equations (13.15) may be found, for example, in Chapter 18, Section 3, of Cornwell (1984)). The irreducible representations of A2 were investigated in detail in Chapter 12, Section 4. For a weight A = /ZiQi + /12^2, (13.16) the associated eigenvalues /3 and Y of the operators Z3 and У are given by I3 = 1И- 5M2, Y = /Z2. The argument is simply that, by Equations (13.15) above, /3 = A(3/iai) = 3{/zi(ai,ai) + /z2(^i,^2)} = Vi ~ (l/2)/z2, and Y = A(2/iai + 4/ia2) — ai) + 4(ai, (*2)} + M2{2(ai, (*2) + 4(ci2, 012)} = /12- (13.17)
ELEMENTARY PARTICLE SYMMETRY SCHEMES 263 "I 2 0 2 I -------------------0-----------------.------- Z3 Figure 13. 2: Values of I3 and Y for the irreducible representation^}. The resulting pairs of eigenvalues /3 and Y for the irreducible representations {3}, {8}, {6} and {10} can be read off Figures 12.1, 12.3, 12.4 and 12.6, and are displayed in Figures 13.1, 13.2, 13.3 and 13.4. For the representations {3*}, {6*} and {10*} the values of I3 and Y are the negatives of those of {3}, {6} and {10} respectively. By Equation (13.11) the corresponding values of the electric charge Qe are given by Q = /zi. (13.18) The weight of multiplicity 2 of the irreducible representation {8} may be thought of as being associated with two eigenvectors, one corresponding to the eigenvalues I = 0, /3 = 0, Y = 0, and the other to I = 1, /3 = 0, Y = 0. The best-established non-trivial unitary multiplets are indicated in Figures 13.5, 13.6, 13.7 and 13.8. In each case the figure on the right hand side is the quantity me2, quoted in MeV, where m is the average rest mass of the corresponding isotopic multiplet. (The members of an isotopic multiplet necessarily lie in the same horizontal line in each of these figures.) To each of the baryon multiplets {8} and {10} there correspond antibaryons transforming as {8} and {10*} respectively ({8} being identical to its complex conjugate). At the time that this scheme was proposed all the particles of the baryon decuplet had already been observed, except for the Q-. The subsequent discovery of this particle with precisely the predicted quantum numbers (and a rest mass as predicted by the Gell-Mann-Okubo mass formula) was a triumph for the theory. In addition to the hadrons listed in the figures, there are a
264 GROUP THEORY IN PHYSICS Figure 13. 3: Values of Z3 and Y for the irreducible represent at ion{6}. Y I 3 1 л 1 . 3 "2 -I "2 0 2 I 2 -I f-2 Figure 13. 4: Values of Z3 and Y for the irreducible representation 10}.
ELEMENTARY PARTICLE SYMMETRY SCHEMES 265 i .p (udd) (uud) 939 S’ Z°(uds) ( - I “I 0 2 —-----------------e-------- (dds) Л0 (uds) I 1193 (uus) 1115 a (dss) H° (uss) 1318 Figure 13.5: The baryon octet {8} with j = | and parity +. (The quark contents are in parentheses. The figures on the right hand side give me2 (in Mev), where m is the average mass of the isotopic multiplet to its left.) (u’JT1232 ,2:+ з 1385 I г (uds) (uus) ^3 a (dss) (uss) 1530 П" •-2 (sss) 1672 Figure 13.6: The baryon decuplet {10} with j = j and parity +. (The quark contents are in parentheses. The figures on the right hand side give me2 (in Mev), where m is the average mass of the isotopic multiplet to its left.)
266 GROUP THEORY IN PHYSICS (uu,dd) -|T~ 4 OzJ° g I ”•* 137 (d*0) (uu,dd,ss) (*d) K73 □49 K" (su) -I K° (sd) 496 Figure 13.7: The meson octet {8} with j = 0 and parity —. (The quark contents are in parentheses. The figures on the right hand side give me2 (in Mev), where m is the average mass of the isotopic multiplet to its left.) number of singlets (belonging to the irreducible representation {1}). One point that is immediately apparent from Figure 13.1 is that for the irreducible representation {3} the values of Q are |, — | and — | i.e. they are not integers. This is actually a special case of the general result that the eigenvalues Q for the unitary multiplet belonging to the irreducible rep- resentation r({ni, П2}) are integers if and only if (щ — П2)/3 is an integer. (The argument is that, by Equations (12.9) and (12.10), every weight A in Г({п1,П2}) is of the form Л = П1Л1 + П2Л2 - Qiai - q2a2 = Efe=i{Sj=i nj(A-1)fcj “ Qk}otk , so that, from Equations (13.16) and (13.18), 2 Q = = (2/3)ni + (l/3)n2 —Qi = -(l/3)(ni-n2) +«i ~qi- J = 1 As ni, П2, Qi, and q2 are all integers, this expression is an integer if and only if (ni — П2)/3 is an integer.) The most fruitful proposal for dealing with this observation was made by Gell-Mann (1964) and Zweig (1964), and is that the particles corresponding to the irreducible representations {3} and {3*} do exist, and are the basic constituents of all the observed hadrons. Gell-Mann (1964) called the particles of the {3} “quarks”, so that those of the {3*} become “antiquarks”. The assumption is that the quarks have baryon number В = | while the antiquarks
ELEMENTARY PARTICLE SYMMETRY SCHEMES 267 (du) k*° K*+ • I • (ds) (us) (uu,dd) i 2 (uu,dd,ss) 892 । p* 770 (ud) 783 73 K*- (su) K*° (sd) 892 Figure 13.8: The meson octet {8} with j = 1 and parity —. (The quark contents are in parentheses. The figures on the right hand side give me2 (in Mev), where m is the average mass of the isotopic multiplet to its left.) correspond to В = — |. The three quarks are now usually called the tq d and s quarks (u corresponding to isotopic spin “up”, d to isotopic spin “down”, and s to non-zero strangeness), and the associated antiquarks are denoted by tq d and s. The properties of the quarks are summarized in Table 13.2. In the simplest model the mesons are made of qq pairs (i.e. quark and antiquark pairs). As (10.38) shows that D1/20D1/2 « D^-^D0, two particles with intrinsic spin | combine to produce composites with spin 1 and spin 0. Moreover, as noted in (12.18), for A2 {3} 0 {3*} « {8} ф {1}, so that the qq pairs transform as the {8} and the {1}. This explains very neatly the observation that there exist su(3) meson octets and singlets with both spin 1 and spin 0. For baryons the simplest assumption is that each baryon consists of three quarks (and so each antibaryon consists of three antiquarks). As three parti- cles with intrinsic spin | couple to produce a composite with intrinsic spin | quark в I h Y S Q и 1/3 1/2 1/2 1/3 0 2/3 d 1/3 1/2 1/3 0 -1/3 s 1/3 0 0 —2/3 -1 -1/3 Table 13.2: Quantum numbers of the quarks tq d and s.
268 GROUP THEORY IN PHYSICS or | (because, by Equation (10.38)), D1/2 0 D1/2 0 D1/2 « (D1 © D°) 0 D1/2 « (D1 0 D1/2) © (D° 0 D1/2) ~ (D3/2 © D1/2) © D1/2, and, as was noted in (12.18), for A2 {3}0{3}0{3} « {10} ©2{8} ©{1}, this provides a simple explanation of the existence of baryon octets of spin | and baryon decuplets of spin j. The quark contents suggested by the considerations are indicated in Fig- ures 13.5, 13.6, 13.7 and 13.8. When the unitary spin parts of the state vectors for the baryons are in- vestigated along the lines indicated above for mesons, one very significant feature emerges. It can shown that the triple products of {3} basis vectors that form basis vectors for the {10} are symmetric with respect to the inter- change of indices. Also, as D3/2 corresponds to the highest weight appearing in D1/2 0 D1/2 0 D1/2, the intrinsic spin part of the state vectors for the | spin composites are symmetric products of the spin parts of the constituents. As the generalized Pauli Exclusion Principle states that fermion state vectors must be antisymmetric with respect to interchanges such as these, it follows that, if the only distinguishing labels for the quarks are those already intro- duced, then the orbital part of the three-quark wave functions for the spin-j decuplet baryons must be antisymmetric. While this is not impossible, it is contrary to experience with ground state configurations in other systems. The dilemma can be avoided by making the further assumption that each of the three quarks u, d and s comes in three varieties that are distinguished by a further feature, which is called “colour”. (It will be appreciated that this is purely a matter of terminology, and that it has nothing to do with “colour” in the normal sense of the word.) If each of the three quarks of a spin-| decuplet has a different colour, then the internal symmetry part of the state vector is no longer symmetric, and so the problem with the orbital part does not arise. This idea forms the basis of the “SU(3) colour symmetry scheme” and thence of “quantum chromodynamics”. In this scheme the strong interaction takes place through the exchange of 8 “gluons”, which belong to the irreducible representation {8} of the SU(3) colour group. This introduction will be concluded by noting that it has proved very fruitful to extend the above considerations in various directions. The most straightforward generalization, from su(3) to su(4), produces a scheme with the additional quantum number “charm”. The more sophisticated suggestion that symmetry breaking is “spontaneous” in origin gives rise to problems within “global” schemes (Goldstone 1961, Goldstone et al. 1962). However, as was shown by Higgs (1964a,b, 1966), when incorporated in a gauge theory (Yang and Mills 1954, and Shaw 1955) these difficulties not only disappear but permit mass generation of the intermediate particles, thereby allowing the construction of a unified theory of weak and electromagnetic interactions (c.f. Salam 1980, Weinberg 1980, and Glashow 1980), based on a u(l) © su(2) algebra.
APPENDICES

Appendix A Matrices The object of this appendix is to give the definitions, notations and terminol- ogy for matrices that are used in this book, together with a brief but coherent account of their relevant properties. 1 Definitions An m x n “matrix” A is defined as a rectangular array of mn elements Ajk (l<j<m,l<fc<n), each of which is a real or complex number, arranged in m rows and n columns. That is >111 A12 Aln >121 A22 • • A2n A = Ami Am2 A • • ^mn A matrix whose elements are all zero is called a “null matrix” or “zero matrix” and is denoted by 0. When m = n, as is the case for most matrices encountered in this book, the matrix is said to be “square”. In this case the elements Ajk with j = к are called the “diagonal” elements, while those with j < к are referred to as being in the “upper off-diagonal” positions. If Ajk = 0 for j к then A is said to be a “diagonal matrix”, the most important example is the “unit matrix” 1, which is defined by m _ л _ J 1, if J = ^, - °jk - o, if J / k, 8jk being the Kronecker delta symbol. (The dimension of 1 is usually clear from its context, but when this is not so the m x m unit matrix will be denoted by lm.) The “sum” of two m x n matrices A and В is defined to be another m x n matrix A + В such that (A + B)^ = Ajk + Bjk (1 < j <m,l<k<n).
272 GROUP THEORY IN PHYSICS Similarly, the “scalar product” of an m x n matrix A with a real or complex number A is an m x n matrix A A defined such that (AA)^ = XAjk (1 < j < m, 1 < к < n). The “matrix product” of an m x n matrix A and an n x p matrix В is defined as an m x p matrix AB whose elements are given by n = (A.l) k=l When m = n = p, both AB and BA exist and have the same dimensions, but even so, in general, AB BA. However, if A and В are both diagonal matrices, then necessarily AB = BA. Of course, for any m x n matrix A, Al = 1A = A. The “transpose” A of an m x n matrix A is defined to be the n x m matrix whose elements are given by (A)j& = A/cj, so that if C = AB then C = BA. The “complex conjugate” A* of A is defined as the m x n matrix such that (A*)jfc = (Ajk)*, the * here denoting complex conjugation. Combining these concepts gives the “Hermitian adjoint” A’*’ of A, which is the n x m matrix defined by A’*’ = (A)*. (In the mathematical literature this A’*’ is often referred to as the “associate” of A, the term “adjoint” being reserved for another matrix.) The “determinant” det A of an m x m matrix A is the real or complex number defined by det A = 52(-l)pAiC1 A2c2 •.. АШСгп, where (ci, c2,..., cm) is a permutation of (1,2,..., m), p being the number of transpositions required to bring (ci, c2,..., cm) to the “natural” order (1,2,..., m), and the sum is over all such permutations. Then det 1 = 1, and det A = det A, (A.2) det A* = (det A)*, (A.3) det(AB) = (det A)(det B), (A.4) Moreover, if В is a matrix obtained from A by interchanging all the elements of a pair of rows (or a pair of columns), then det В = — det A. The “inverse” A-1 of an m x m matrix A is defined as the m x m matrix such that A-1 A = AA-1 = 1. A-1 exists if and only if det A 0, in which case A is described as being “non-singular”. If A and В are two non-singular m x m matrices, then (AB)”1 =B-1A-1. Table A.l gives the definitions of a number of important special types of matrix. It should be noted that a matrix that is both real and symmetric is necessarily Hermitian, and a matrix that is both real and orthogonal is necessarily unitary. For an orthogonal matrix A, as A A = 1, Equations (A.2) and (A.4) imply that (det A)2 = 1, so that det A = +1 or — 1. (A.5) Similarly, for a unitary matrix A, as A^A = 1, Equations (A.2), (A.3) and (A.4) imply that | det A|2 = 1, so that det A = ехр(ш), (A.6)
APPENDIX A 273 Description of matrix symmetric antisymmetric (or skew-symmetric) real orthogonal Hermitian (or self-adjoint) anti-Hermitian unitary anti-unitary Defining property A = A A = —A A* = A A = A-1 Af = A Af = -A Af = A-1 Af = -A'1 Table A.l: Definitions of special types of matrix. where a is some real number. The “trace” tr A of an m x m matrix A is defined by tr A = Ajj-> that is, it is the sum of the diagonal elements of A. If A and В are any two m x m matrices, it follows immediately from Equation (A.l) that tr(AB) = tr(BA). Also tr(A + B) = tr A + tr B, and, for any complex number a, tr(aA) = a(tr A). Moreover, if A,B, C are any three m x m matrices, Equation (A.l) gives tr(ABC) = tr(BCA) = tr(CAB). If tr A = 0, then A is said to be “traceless”. If A and A' are m x m matrices related by a so-called “similarity transformation” A' = S-1AS, where S is any m x m non-singular matrix, then tr A' = tr A. A matrix may be “partitioned” into submatrices by inserting dividing lines between arbitrarily chosen adjacent pairs of rows and columns. For example, A = All А21 >112 >113 >122 >123 >114 >124 A31 >132 >133 A34 ’ А11 А12 А13 ' А21 А22 А23 is a partitioning of a 3 x 4 matrix A into six submatrices Aj/c, defined by А11 = >111 >12! , а12 = >112 >122 >113 >123 , А13 = >114 А24 А21 = [А31], А22: = Нз2 Азз], А23 = [А34]. In terms of submatrices, the matrix product C = AB of an m x n matrix A with an n x p matrix В has a remarkably simple property, provided that the “column” partitioning of A is chosen to be the same as the “row” partitioning of B. Explicitly, if A, В and C are partitioned into st submatrices AjA:, tu submatrices BA:Z and su submatrices respectively, where 1 < j < s < m, 1 < k < t < n, 1 < I < и < p, and where AjA:, BH and have dimensions rrij хпк, rik xpi and rrij xpi respectively (where J2j=i mj = 521=i nk = and Pi = Pi then t Cjl = Y,AjkBkl. (A.7) k=l
274 GROUP THEORY IN PHYSICS There is a striking similarity of form with Equation (A.l). It is as though the submatrices Aj/c, Bkl and GyZ can be regarded as being matrix elements, but with the product of AjA: and Bkl being given by Equation (A.l). For clarity, the various submatrices have been distinguished here by superscripts. However, it is sometimes convenient to use subscripts instead, as, for example, in Chapter 4, Section 4. Moreover, the dividing lines will be omitted when there is no possibility of confusion, so that for the above example - A11 A12 A13 - A21 A22 A23 The “direct product” (or “Kronecker product”) of an m x m matrix A and an n x n matrix В is defined to be an mn x mn matrix A 0B, whose rows and columns are each labelled by a pair of indices in such a way that (A 0 = AjkBst (for 1 < j, к < m and 1 < s, t < n). (A.8) In order to express such matrices in the usual form in which rows and columns are each labelled by a single index it is necessary to put the set of mn pairs (J, s) (1 < J < m; 1 < s < n) into one-to-one correspondence with a set of mn integers p (1 < p < mn), with an identical correspondence between the pairs (k,t) and a set of integers q. The most convenient choice is p = n(j — 1) + s, q = n(k — 1) +1. With this prescription, the matrix A 0 В for m = 2 and n = 2 would be displayed as (A 0 B)ii5n (A 0 B)i2,n (A 0 B)2i,n (A 0 B)22,ll (A 0 В)цд2 (A 0 В)12Д2 (A 0 В)21д2 (A 0 B)22,12 (A 0 B)nj2i (A 0 B)i2j2i (A 0 B)2i,2i (A 0 B)22,2i (A 0 B)nj22 (A 0 B)i2522 (A 0 B)21522 (A 0 B)22 22 Clearly, the diagonal elements of A 0 В in the pair-labelling scheme are those for which j = к and s = t. If A and В are both diagonal, then A 0B is also diagonal (for if Ajk = aj^jk and Bst = bs6st, then (A 0 В)7-5^ = ft j bs6jk&st)’ If A and A' are both m x m matrices and В and B' are both n x n matrices, then (A 0 B)(A' 0 B') = (AA') 0 (BB'), (A.9) where all products other than those indicated by the symbol 0 are ordinary matrix products (as defined in Equation (A.l)). The proof of Equation (A.9) is straightforward, for the (Js, kt) element of the right-hand side is m n (AA')jfc(BB')st = ££ AjtA'lkBsuB'ut 1=1 W=1 while the (Js, kt) element of the left-hand side is m n m n ^0 0 B)jsju(A 0 В )iu,kt = ^2 AjiBsuAikBut. 1=1 гг=1 1=1 гг=1
APPENDIX A 275 Finally, if A and В are both unitary, then A 0 В is also unitary. (It follows directly from Equation (A.8) that (A 0 B)^ = A’*’ 0 Bt, and, as AtA = AA^ = lm, B^B = BB^ = ln and 1TO 0 1n = lm+n, the unitary property of A 0 В is an immediate consequence of Equation (A.9).) 2 Eigenvalues and eigenvectors If A is an m x m matrix and A is a real or complex number which, together with an m x 1 “column” matrix c (c/0), satisfies the equation Ac = Ac, (A.10) then A is said to be an “eigenvalue” of A and c is said to be an “eigenvector” corresponding to A. Equation (A. 10) has a non-trivial solution if and only if det(A - Al) = 0, (A.ll) which is often referred to in the mathematical physics literature as a “secular equation”. The left-hand side of Equation (A.ll) is a polynomial P(A) of degree m, known as the “characteristic polynomial”, whose coefficients are determined by explicit evaluation of the determinant. The eigenvalues are given therefore by the “characteristic equation” P(A) = 0, (A.12) that is, they are the roots of P(A). Suppose that P(A) has R distinct roots Ai, A2,..., Ад, and that Xj has multiplicity rj, j = 1, 2,..., P, so that P(A) = (A - Х±Г (A - A2)"2 ... (A - Ад)"-. (A.13) Then the Cayley-Hamilton Theorem states that the matrix A also satisfies the characteristic equation (Equation (A.12)), that is, P(A) = (A - Ai)"1 (A - A2)"2 ... (A - Ад)"- = 0. (A.14) If A' is related to A by a similarity transformation A' = S-1AS, where S is any non-singular m x m matrix, then Equation (A. 10) can be written as A'(S-1c) = A(S-1c). (A.15) Thus A and A' have the same set of eigenvalues, with identical multiplicities, and, if c is an eigenvector of A corresponding to A then c' = S-1c is an eigenvector of A' corresponding to A and vice versa. The question now arises as to whether A' can be made diagonal by an appropriate choice of S. If this is so then A is said to be “diagonalizable”. This is certainly true if A is Hermitian or unitary, and in both of these cases S can be chosen to be unitary (Gantmacher 1959). As the operators corresponding to physical observables in quantum mechanics are self-adjoint,
276 GROUP THEORY IN PHYSICS whenever the corresponding operator eigenvalue equation is cast in matrix form (as in Appendix B, Section 4) the matrix involved is Hermitian. Thus all the matrices occurring in such a context are automatically diagonalizable. However, non-diagonalizable matrices do occur in various contexts, so it is worthwhile analysing the question of diagonalizability in more detail. The most important result is embodied in the following theorem. Theorem I An m x m matrix A is diagonalizable if and only if A possesses m linearly independent eigenvectors. Proof Suppose first that A is diagonalizable and that A' = S-1AS is the diagonal form. Then each diagonal element of A' is an eigenvalue of A' (and of A) and the eigenvector of A' corresponding to the eigenvalue A'- can be taken to be c'-, where / / \ f 1, if 1’ = k, / л -1 ^kl ~ { 0, if j/ k. (A-16) Thus A' possesses m linearly independent eigenvectors c'-, j = 1,2,..., m, and hence A possesses m linearly independent eigenvectors Sc'-, j = 1,2,..., m. Conversely, suppose that A possesses m linearly independent eigenvectors cj, j = 1, 2,..., m. Define the m x m matrix S by S — [ Ci | C2 | • • • | ], so that det S 0. By virtue of Equation (A.7), if Cj is defined by Equations (A.16) then Sc'- = cy, so c'- = S-1cy for j = 1,2, But the set c'- (J = 1,2,... ,m) can be eigenvectors of A' (as required by Equation (A. 15)) only if A' is diagonal, so A must be diagonalizable. This theorem implies that if A is diagonalizable there are rj linearly in- dependent eigenvectors corresponding to each eigenvalue of multiplicity rj, whereas if A is non-diagonalizable at least one eigenvalue has less linearly independent eigenvectors than its multiplicity. that The following examples illustrate the two possible situations. Suppose first 0 1 0 1 0 0 0 0 1 A = (A.17) which is Hermitian and hence diagonalizable. The characteristic polynomial is P(A) = (A — 1)2(A+ 1). For the eigenvalue A = 1, the (1,1) and (2,1) com- ponents of Equation (A.10) both give C21 = сц, while the (3,1) component is trivially satisfied. Thus 1 1 0 and 0 0 1 are two linearly independent eigenvectors corresponding to eigenvalue A = 1. Similarly, for the eigenvalue A = —1, the (1,1) and (2,1) components of
APPENDIX A 277 Equation (A.10) give C21 = -сц, while the (3,1) component gives C31 = 0. Thus 1 ' -1 0 is the only linearly independent eigenvalue corresponding to eigenvalue A = -1. As an example of a non-diagonalizable matrix, consider 1 0 0 0 A = (A.18) for which P(A) = A2. For the eigenvalue A = 0, the (1,1) component of Equation (A. 10) gives C21 = 0, while the (2,1) component is trivially satisfied. Thus although A = 0 is an eigenvalue of multiplicity 2, ’ 1 ’ 0 is the only linearly independent eigenvector. There exists a useful criterion for diagonalizability involving the “minimal polynomial” M(A) of A, which is defined as the polynomial of lowest degree in A such that M(A) = 0. In some cases the minimal polynomial is identical to the characteristic polynomial, but otherwise its degree is less than m. It can be shown (Gantmacher 1959) that M(A) is unique and has the form M(A) = (A - Ai)S1(A - A2)S2 ... (A - XRy\ where 1 < Sj < rj, j = 1, 2,..., P, and where Xj and rj are as defined in Equation (A. 13). (In particular this implies that every distinct eigenvalue of A appears in a factor of M(A).) Moreover, A is diagonalizable if and only if Sj = 1 for every j = 1, 2,..., P, that is, if and only if M(A) consists only of linear factors. This has the corollary that A is necessarily diagonalizable if every eigenvalue of A has multiplicity 1. The examples that were considered above demonstrate this criterion very neatly. For the matrix A of Equation (A. 17), M(A) = A2 - 1 = (A - 1)(A + 1), which has only linear factors, so that A must be diagonalizable. By con- trast, for the matrix A of Equation (A.18), M(A) = A2, so this A is non- diagonalizable. Even when A is not diagonalizable, it can be transformed by an appro- priate similarity transformation into a standard form, the “Jordan canonical form”. More precisely, for any m x m matrix A there exists an m x m matrix S such that all the elements of A' = S-1AS are zero except possibly the A'rr
278 GROUP THEORY IN PHYSICS elements (for r = 1,2,..., m) and the AJ. +1 elements (for r = 1,2,..., m—1). Moreover, A^r+1 = 0 or 1 for all r = 1,2,..., m — 1, and {0 or 1, if Afrr = A^+l r+1, 0, if Arr 7^ . For example, A' = ” 2 1 0 0 0 " 0 2 0 0 0 0 0 5 1 0 0 0 0 5 1 0 0 0 0 5 is in Jordan canonical form. So too is the matrix A of Equation (A. 18). Obviously the eigenvalues of A' are just the set of diagonal elements A'rr, r = 1,2,..., m. (Clearly the case in which A' is diagonal is merely a special case in which A'r r+1 = 0 for all r = 1,2,..., m — 1.)
Appendix В Vector Spaces This appendix is intended both to provide an introduction to vector spaces and to give the various notations and conventions that are used throughout this book. 1 The concept of a vector space A general vector space is obtained by selectively abstracting certain properties of vectors of the three-dimensional Euclidean space IR3. A vector Ф of IR3 may be specified by a triple of real numbers x±, ж2 and x%, that is, Ф — (^1,^2, ^з)- (In elementary treatments of IR3 it is conventional to indicate a vector by using bold type, but for treatments of higher-dimensional spaces this convention is discontinued. To help avoid confusion, as far as pos- sible vectors in this appendix will be denoted by the Greek letters y,... and scalars (that is, real or complex numbers) by a, 6, c,....) The product of a vector of IR3 with a real number a is defined to be another vector а/ф such that а/ф = (n^i, ax2^ ахз), (B.l) from which it follows that if b is any other real number Ъфмф') = (Ъа\ф. (B.2) The sum of two vectors ф> = (#i, ^3) and ф = (г/i, у 2, Уз) of Ш3 is defined by Ф + ф = (a?i + ?/i,a;2 + У2,х3 + 2/3), (В.З) so that ф + ф = ф + ф. (В.4) Similarly, if x — (^i, 22, ^3) is any other vector of IR3, Ф + (Ф + x) = (V> + Ф) + X- (B.5)
280 GROUP THEORY IN PHYSICS With these definitions it is easily verified that a(V> + ф) = аф + аф (В. 6) for any real number a and any two vectors ф and ф. Similarly, (a + Ь)ф = аф + Ьф (В. 7) for any real numbers a and b and any ф. Finally, there exists a vector 0 = (0,0,0) such that ф + 0 = ф (B.8) for every ф of IR3. In a general vector space V it is assumed that multiplication of vectors by scalar and vector addition can always be defined (though not necessarily by Equations (B.l) and (B.3)) in such a way that the properties in Equations (B.2), (B.4), (B.5), (B.6), (B.7) and (B.8) are retained. The precise definition is as follows: Definition Vector space A “vector space” V is a collection of elements ф, ф, у, • • • (called vectors) for which “scalar multiplication” аф is defined for any “scalar” a from a certain set and for any vector ф, and for which “vector addition” ф + ф is defined for all vectors ф and ф, such that а(Ьф) = (аЬ)ф, ф + ф = ф + ф, Ф + (</> + %) = ФФ + Ф) + х, > а(ф + ф) = аф + аф, (а + Ъ)ф = аф + Ьф. Moreover, there must exist in V a “zero vector” 0 such that ф + 0 = ф for all ф E V. If the set of scalars consists of all real numbers then V is said to be a “real vector space”. Similarly, if the set of scalars consists of all complex numbers, V is called a “complex vector space”. The set of scalars is often referred to as the “field”. A set of vectors ф±, ф2,..., фа of V is described as being linearly dependent if there exists a set of non-zero scalars (of the appropriate set) «1,^2,... , см such that ^1Ф1 + «2^2 + • • • + аафа = 0. (B.9) Thus the set ф\, ф2 , • . ., фа is linearly independent if the only solution of Equa- tion (B.9) (in the appropriate set of scalars) is a± = a2 = ... = aa = 0. If V contains a set of d linearly independent vectors, but every set of (d+1) vectors is linearly dependent, then V is known as a “d-dimensional space”. (For ex- ample, the vectors of IR3 form a three-dimensional real vector space.) If there
APPENDIX В 281 is no limit on the number of linearly independent vectors then V is said to be an “infinite-dimensional space”. Finite-dimensional spaces are much easier to deal with and fortunately most of the vector spaces encountered in this book will be of this type. Let V be a vector space of finite dimension d and let ^1, ^2, • • •, ^d be any set of linearly independent vectors of V. Then any ф eV can be uniquely expressed in terms of ^1, ^2, • • •, ^d by Ф> = «1^1 + «2^2 + • • • + (B.10) where ai, a2, • • •, «d are a set of scalars that depend on *ф. (This follows from the definition just given, for ^1, ^2, • • • , ^d must be linearly dependent, so there exists a set of scalars b, £>i, £>2, • • •, &d such that bip + 61^1 + 62^2 + ... + bd^d = 0, and not all members of the set 6, 61,b2, • • • ,bd are zero. As b must be non-zero, dividing by b and putting aj = —bj/Ъ gives Equation (B.10) immediately. The decomposition (Equation (B.10)) is unique because if = Uj=i aj^j and = SjLi aj^j then _ aj№j = 0, f°r which the only solution is aj = a'j for j = 1,2,..., d, as the vectors ^1, ^2, • • •, ^d are assumed to be linearly independent.) The set ^1, ^2, • • •, ^d is therefore said to provide a “basis” for V. In IR3 a very convenient basis is provided by 'фг = (1,0,0), ^2 = (0,1,0) and ^3 = (0,0,1) for, if <ф = (ж1,ж2,^з), then 'Ф = Хуф1 + Х2ф2 + ^3^3- The set of d vectors , V4, • • •, V>d defined in terms of the basis ^1, , • • -, ^d by d 'Фп = $тпфт m=l for n = 1,2,... ,d form a linearly independent set if and only if S is non- singular. Thus when S is non-singular ^1, ^2, • • •, ^d provides an alternative basis for V. It is occasionally convenient to regard a d-dimensional complex vector space as a 2d-dimensional real vector space. If ^1, ^2, • • •, ^d is a basis for the complex space, then Vh, ^2, • • •, ^d, together with г^1, г^2, • • •, ^d, form a basis for the real space. (It should be noted that 'фд and ify are linearly independent elements of the real space (although they are linearly dependent in the complex space) as aipj + biipj = 0 has no solution for real numbers a and 6, other than a = b = 0.) The following examples show some of the widely differing forms of vector space that are encompassed by the definition. Example I The three-dimensional complex vector space (C3 (C3 consists of the set of triples <ф = (^1,^2,^з), where ^1,^2 and x% are complex numbers. Scalar multiplication by an arbitrary complex number a is defined by а*ф = (a#i, ax2, ax%) and vector addition is defined by Equa- tion (B.4). The set of vectors (1,0,0), (0,1,0) and (0,0,1) again provides a convenient basis.
282 GROUP THEORY IN PHYSICS Example II The set of all N xN traceless anti-Hermitian matrices (AT > 1) Let A and В be any two N x N traceless anti-Hermitian matrices (see Table A.l). Then the scalar product a A and vector sum A + В may be taken to be the scalar product and matrix sum defined in Appendix A, Section 1. Then a A and A + В are N x N traceless anti-Hermitian matrices, provided that a is real. Thus the set of such matrices form a real vector space, the N x N matrix 0 (all of whose elements are zero) providing the zero element. It should be noted that even though this vector space is real the elements of matrices involved may be complex! As will be seen in Example II of Chapter 8, Section 5, this vector space has an additional structure and forms the Lie algebra su(AT) of the linear Lie group SU(AT). It is shown there that the dimension of this space is (A2 — 1). Example III Set of all functions defined in IR3 Let ^(r) and ф(т) be any two complex-valued functions defined for all r E IR3. Then ф + ф is defined in the natural way by (ф + </>) (r) = ^(r) + ф(т) for all r E IR3 and, for any complex number а, аф is defined by (a^)(r) = а(ф(г)) for all r E IR3. The set of all such functions then forms an infinite-dimensional complex vector space, the zero vector being defined to be the function that is zero for all r E IR3. A “subspace” of a vector space V is a subset of V that is itself a vector space. The subspace is said to be “proper” if its dimension is less than that of V. V is said to be the “direct sum” of two subspaces Vi and V2 if every ф E V can be written uniquely in the form ф = Ф1 + фэ, where ф-i E Vi and Ф2 E V2. This implies that Vi and V2 have only the zero element of V in common. If фъфъ, • • • ^фа is a basis for V and 1 < d' < d, then ^1,'fe • • •, V’d' and V’d'+i, • • • > фа are bases for two subspaces of V of dimensions d' and (d — d') respectively. Moreover, V is the direct sum of these two subspaces, because if ф = aj^:h then ф = ^1 + ^2, where Ф1 = and Ф2 = <иФр this decomposition being unique because the set ai, a2,..., «d depends uniquely on ф. The concept of a direct sum can be generalized to more than two subspaces in the obvious way. 2 Inner product spaces Many vector spaces have the additional attribute of being endowed with an “inner product”. Consider first the example of vectors of IR3, in which the inner product is the familiar scalar product. Thus, if ф = (ж1,Ж2,^з) and Ф = (x/i7 Z/2, Z/з) are any two vectors of IR3, their inner product (ф,ф) is the real number defined by (V1, </>) = + Х2У2 + X3y3. The “length” of0 = (жх,Ж2,жз) is given by {(жх)2 + (Ж2)2 + (Ж3)2}1/2, which is real and non-negative. Indeed it is only zero when ф = 0, the zero vector.
APPENDIX В 283 It may be denoted by ||^|| and will be called the “norm”” of ?/>. Clearly IWI = Ш^)}1/2- In (C3 (see Example I of the previous section) it is natural to again require that the norm ||^|| be always real and non-negative and also that ||^|| =0 only when?/; = 0. This is achieved by the definition \\ф\\ = {|a?i|2 + |ж2|2 + |^зI2}1/2. The identity \\ф\\ = {(ф^ф)}1^2 can be retained if the inner product of any two vectors ф = (ж1,Ж2,^з) and ф = (x/i, X/25 Z/з) (where the components are now complex numbers) is defined by фф, Ф) = Х1У1 + Х2У2 + Х3У3. With this definition = (Ф,ФУ\ and for any two complex numbers a and b (аф, Ьф) = а*Ьфф, ф). Also, if x = (21, 23), then (V> + Ф,х) = (Vsx) + (Ф,х)- A general “inner product space” is a vector space possessing an inner product that has the properties exhibited by these examples (even though the definition of this inner product may be quite different). The precise re- quirements are as follows. Definition Inner product space A complex vector space V is said to be an “inner product” space if to every pair of vectors ф and ф of V there corresponds a complex number (^, ф) (called the inner product of ф with ф) such that: (a) (V>, 0) = (b) (аф, Ъф) = а*Ь(ф, ф) for any two complex numbers a and 6; (c) (V> + ф,х) = (Ф, X) + (Ф, x) for any X £ V; (d) (ф,ф) > 0 for all ф, and (e) (ф,ф) = 0 if and only if ф = 0, the zero vector. If V is a real vector space the inner product is required to be a real number, and in (b) a and b are restricted to being real numbers, but otherwise the requirements (a) to (e) are the same as for a complex space. An “abstract” inner product space is a space that satisfies all the axioms without possessing a “concrete” realization for the inner product. It should be noted that (a) and (c) imply that (х.-ф + ф) = (х,^) + (х,0)>
284 GROUP THEORY IN PHYSICS so that, by (b), (%, a/ф + Ьф) = a(x, V’) + b(x, </>), (оф + Ъф,x) = а*(ф, x) + &*(</>, x)- Also, (a) implies that (ф,ф) is necessarily real (which is implicit in the re- quirement (d)). For any inner product space the “norm” \\ф\\ may be defined by 1М1 = Ш0}1/2- It follows from (b) that 1Ы1 = Ы11< (B.ll) Two other properties that are easily proved (Akhiezer and Glazman 1961) are the “Schwarz inequality” (with strict inequality applying if ф and ф are linearly independent), \(Ф,Ф)\ <MM, and the “triangle inequality” \\Ф + Ф\\ <M + Ml, both valid for any ф and ф of an inner product space. By analogy with the situation in IR3, the “distance” (1{ф,ф) between two vectors ф and ф in a general inner product space may be defined by <1(ф,ф) = \\ф-ф\\. (В.12) Then it follows immediately that (i) (1(ф,ф) = (1{ф.ф\ (ii) d(^, Ф) = 0; (iii) d(^, ф) > 0 if ф ф\ and (iv) d(^, ф) < Нфф, x) + d(x, Ф) for any ф,ф,х £ V> all of which are essential for the interpretation of d(^, ф) as a distance. The distance function йфф^ф) is often called the “metric”. Example I The d-dimensional complex vector space Cd Cd is the set of d-component quantities ф = (rri,^2, • • •, Xd), where ^1,^2, • • •, Xd are complex numbers. It is a complex vector space of dimension d. The inner product of (Cd may be defined by d ('ФтФ) = Yx*jyi’ (B-13) J=1
APPENDIX В 285 where ф = (rri,^2, • • •, хф) and ф = (yi^yz-) • • •, уф)-, which satisfies all the requirements for Cd to form an inner product space. From Equations (B.12) and (B.13) it follows that = {E^=i \xj - %l2}1/2 • Example II The set of all m x m matrices The set of all m x m matrices with complex elements forms a complex vector space of dimension m2, provided that the scalar product and vector sum are taken to be the scalar product and matrix sum defined in Appendix A, Section 1. The inner product of two such matrices A and В may be defined by m m (A,B)=££A^BJfc, (B.14) j=l k=l which again satisfies all the requirements for the vector space to form an inner product space. Moreover, Equations (B.12) and (B.14) imply that d(A,В) = {E7=1 ЕГ=11Afc - BJfc|2}V2 . This explains the origin of the metric of Equation (3.1). Comparison of Equa- tions (B.12) and (B.13) shows that this inner product space is essentially just c™2. Two elements ф and ф of an inner product space are said to be “orthog- onal” if (^, ф) = 0. (In IR3 this coincides with the usual geometric notion of orthogonality.) A vector ф is described as being “normalized” if \\ф\\ = 1. An “ortho-normal” set is then a set of vectors ф^ ф2,. •. such that (ф^фк) = Sjk for фк = 1,2,.... From any set of linearly independent vectors </>i, </>2,... an ortho-normal set ф\, ф2,... can be constructed by taking appropriate linear combinations. The procedure, often called the “Schmidt orthogonalization process”, is as follows. First let = Ф1\ $2 = Ф2 — {(0i,<fe)/(0i,0i)}0i; Оз = Фз~ {(01,<^з)/(01,01)}01 _ {(#2,фз)/(#2, ^2)}^; and so on. The vectors are then mutually orthogonal. Finally let фj = {HfyИ-1}#? for j = 1,2,..., so that, by Equation (B.ll), ||^|| = \\0j II-1 II0j || = 1. Then ф1, ф2,... form an ortho-normal set. Ortho-normal sets are particularly useful as bases. If V is an inner product space of dimension d and the basis ^1,^2, • • • ,фа of Equation (B.10) is an ortho-normal set, then forming the inner product of both sides of Equation (B.10) with фj gives d d = '^2ак1Фз,'Фк) = XiakS3k = ai- k=l k=l
286 GROUP THEORY IN PHYSICS Thus Equation (B.10) can be rewritten as d .7 = 1 (B.15) If 1, ^2 > • • • > Фа another basis for V and the d x d matrix S is defined by d Фп = ^тпФт m=l for n = 1,2,..., d, then ^1? • • • > a^so form an ortho-normal set if and only if S is a unitary matrix. This follows from the fact that d d M = YYs^snk^m,M = &s)jk. m=l n=l 3 Hilbert spaces For an infinite-dimensional inner product space it is natural to enquire whether the expansion in Equation (B.15) is valid with the finite sum re- placed by an infinite sum. This immediately poses questions of convergence for such infinite series. With the metric introduced in Section 2 one may say that the infinite sequence </>i, </>2, • • • of vectors in an inner product space V tends to a limit ф of V (i.e. фп -+ ф as n сю) if and only if d(^n, ф) —> 0 as n -> 00. Then for an infinite series one may say that JSjli converges to ф if the sequence of partial sums defined for n = 1,2,... by фп = Y^j=i Фз converges to ф, so that all such questions are reduced to questions about sequences. A sequence </>i, (/>2, • • • for which lim d(^n,^m) = 0 m,n—>00 (where m and n tend to infinity independently) is called a “Cauchy se- quence”. It follows immediately from property (iv) of the metric d(^, ф) that, if <^>i, • • • tends to some limit ф, then </>i, (/>2, • • • must be a Cauchy se- quence. Unfortunately, examples can be constructed which demonstrate that, in general, the converse is not true. This makes the general investigation of convergence very difficult, for while it is easy to test whether a sequence is a Cauchy sequence or not, direct examination of the definition of convergence requires some presupposition about the possible limit ф. This problem can be completely avoided by confining attention to those spaces for which every Cauchy sequence converges, that is, to “Hilbert spaces”. The definition will be given for complex inner product spaces, as the only infinite-dimensional spaces that will be met in this book are of this type.
APPENDIX В 287 Definition Hilbert space A “Hilbert space” is a complex inner product space in which every Cauchy sequence converges to an element of the space. The following further restriction is required in order that Equation (B.15) may be generalized to the desired form. Definition Separable Hilbert space A Hilbert space V is said to be “separable” if there exists a countable set of elements 5 contained in V such that every vector ф e V has some element ф e S arbitrarily close to it. That is, for any ф eV and any 6 > 0 there must exist а ф e S such that d(^, ф) < e. The set 5 is then said to be “dense” in V. It is easily shown that every finite-dimensional complex inner product space is a separable Hilbert space. Definition Complete ortho-normal system An ortho-normal set of vectors Vh, , • • • of a Hilbert space is said to be “complete” if there is no non-zero vector that is orthogonal to every j = 1,2,.... Obviously in an infinite-dimensional Hilbert space a complete set of vec- tors necessarily contains an infinite number of elements. The following two theorems then provide the required extension of Equation (B.15). Theorem I If an infinite-dimensional Hilbert space is separable, then the space contains a complete ortho-normal system, and every complete ortho- normal system in the space consists of a countable number of vectors. Theorem II If the vectors ф±, ф%,... form a complete ortho-normal system of an infinite-dimensional Hilbert space, then any vector ф of the space can be written as oo = (B.16) J=1 Moreover, oo 11< = £М,<, (B.17) Equation (B.17) is often called “Parseval’s Relation”. Proofs of both the- orems may be found in the book of Akhiezer and Glazman (1961). Example I The separable Hilbert space L2 L2 is defined to be the set of all complex-valued functions ф(г) (defined for
288 GROUP THEORY IN PHYSICS all r E IR3) such that exists and is finite, the integral here being the Lebesgue integral (see below). The inner product of L2 may be defined by (B.18) where the integral is again the Lebesgue integral. With addition and scalar multiplication defined as in Example III of Section 1, it can be shown that L2 is an infinite-dimensional separable Hilbert space (cf. Akhiezer and Glazman 1961). Equation (B.18) implies that For a proper development of the concept of the Lebesgue integral, the reader is referred to specialized texts such as that of Riesz and Sz.-Nagy (1956). However, for the understanding of the present book no detailed knowledge is required. It is sufficient to be aware that the definition of the Lebesgue integral is more general than that of the more familiar Rie- mann integral, so that functions that are not Riemann-integrable may still be Lebesgue-integrable. Nevertheless, the generalization is such that every Riemann-integrable function is Lebesgue-integrable and the values of the two integrals coincide. Also, if /(r) = 0 except on a “set of measure zero” then = 0. (It is difficult to give a concise characterization of sets of measure zero, but two important facts are easily stated. Firstly, the set of points r in any sphere |r — ro |2 < S of IR3 has non-zero measure provided 6 > 0. Secondly, a set consisting of a finite or a countable number of points has measure zero.) Two functions /(r) and <?(r) that are equal except on a set of measure zero are said to be equal “almost everywhere”. For such functions Consequently two functions ^(r) and ф(т) that are equal almost everywhere are to be regarded as being identical members of L2. 4 Linear operators Let D be a subset of a separable Hilbert space V. If for every ф e D there exists a unique element ф E V, one can write ф = Аф, thereby defining the
APPENDIX В 289 “operator” A. D is called the “domain” of A, and the set Д consisting of all ф = Аф, where ф runs through all of D. is known as the “range” of A. Two operators A and В are then said to be “equal” if they have the same domain D, and if Аф = Вф for all ф E D. If the mapping ф = Аф is one-to-one, the inverse operator A-1 may be defined by А~гф = ф if and only if ф = Аф. Clearly the domain and range of A-1 are Д and D respectively. Definition Linear operator An operator A is said to be “linear” if its domain D is a linear manifold (a set D such that if ф,ф E D then (аф + Ьф) E D for all complex numbers a and 6) and if А(аф + Ьф) = аАф + ЬАф for all ф,ф E D and any two complex numbers a and b. There is no requirement in general that D be the whole Hilbert space, so the definition accommodates such operators as d/dx acting in V = L2, for which D is the set of functions of L2 that are differentiable with respect to x. Definition Bounded linear operator A linear operator A is said to be “bounded” if there exists a positive constant К such that ||Аф\\ < А'Ц'^Ц for all ф E D. Theorem I If A is a linear operator acting in a finite-dimensional inner product space V and D = V, then A is necessarily bounded. Definition Unitary operator An operator U is said to be “unitary” if D = Д = V and (ифХФ) = (ф,ф) for all ф,ф EV. It is easily shown that every unitary operator is a bounded linear op- erator. It is obvious that if V’b'fe-- - form a complete ortho-normal set then ф'^ = иф^ j = 1,2,... also form a complete ortho-normal set. Con- versely, if ^i, Ф2, • •.. and фф ф'2,... are two complete ortho-normal sets in a Hilbert space V, then there exists a unitary operator U such that ф^ = иф^ 7 = 1,2,.... For a general treatment of linear operators the reader is referred to the books of Akhiezer and Glazman (1961), Simmons (1963) and Riesz and Sz. Nagy (1956). However, as all the operators associated with finite-dimensional representations of groups and Lie algebras are either unitary or act on finite- dimensional spaces, attention here will henceforth be concentrated exclusively on bounded linear operators whose domain is the whole Hilbert space V.
290 GROUP THEORY IN PHYSICS If A is such an operator there exists an “adjoint” operator whose do- main is also V such that (A^, ф) = (</>, Аф) for all ф.ф e V. It is easily shown that (AB)’*’ = B’*’A’*’, (A’*’)’*’ = A, and iff = B-1 for a unitary operator U. Definition Self-adjoint operator A bounded linear operator A whose domain is the whole Hilbert space V is said to be “self-adjoint” if A = A\ that is, if (Аф,ф) = (ф,Аф) (B.19) for all ф,ф e V. If for a bounded linear operator A there exists a non-zero vector ф and a complex number A such that Аф = Аф, (B.20) then ф is said to be an “eigenvector” of A and A is referred to as the cor- responding “eigenvalue”. If there exist d linearly independent eigenvectors ф1,ф2,... ,фа of A with the same eigenvalue A, then A is said to have “mul- tiplicity d” or to be “d-fold degenerate”. In that case any linear combination (bi^i + 62^2 + ... + bdffd) is also an eigenvector with the same eigenvalue A. For the special case of self-adjoint operators there are three important theorems: Theorem II The eigenvalues of a self-adjoint operator are all real. Proof Suppose that Аф = Аф, where ф 0. Then, if A is self-adjoint, A(^, ф) — (ф, Аф) = (Аф, ф) = А*(ф, ф), so А = А*. Theorem III Eigenvectors of a self-adjoint operator belonging to different eigenvalues are orthogonal. Proof Suppose that Аф± = Ai^i and Аф% = А^ф^ where Ai 7^ A2. Then, if A is self-adjoint, Ai(^2,^i) = (^2, A^i) = (A^2,^i) = ^(^2,^1) = M'feV'i), so that (Ai — A2)(V>2, Ф1) = 0- As Ai — A2 / 0, it follows that (^2, Ф1) = 0. Theorem IV If A is a self-adjoint operator and В is a unitary operator, then A! = B-1 AU is also self-adjoint and possesses exactly the same eigenvalues as A. Proof A! is self-adjoint because (A')t = (B“1AB)t = BtAt(B“1)t = B-1AB = A'.
APPENDIX В 291 Now suppose that is an eigenvector of A! with eigenvalue A', so that А!ф’ = Х'ф'. Then U~xAU^r = A'?//, so that А(ифг) = A'(C7?//), showing that иф' is an eigenvector of A with the same eigenvalue A'. Every bounded operator has a matrix representation. Indeed, the operator eigenvalue equation (Equation (B.20)) can be re-cast in the form of the matrix eigenvalue equation (Equation (A. 10)). For convenience, the argument will be presented for a finite-dimensional inner product space V of dimension d, but the results generalize in the obvious way to bounded operators acting on a separable infinite-dimensional Hilbert space, although in that case all the matrices involved are infinite-dimensional. Let — be an ortho-normal set of V. Taking the inner product of both sides of Equation (B.20) with any and invoking Equation (B.15) gives d d (= ACV’fc.V’)) (B-21) .7=1 J=1 for к = 1,2,..., d. Let A be the d x d matrix defined by А^ = (фк,Аф^ (B.22) for J, к — 1,2,..., d, and let c be the d x 1 column matrix whose elements are specified by c?i = (тДу, ^), j = 1, 2,..., d. Then Equation (B.21) can be rewritten as Ac = Ac, that is, as Equation (A. 10). It should be noted that if Афп is expanded in terms of the ortho-normal set, then, by Equation (B.15), d d Афп = (V^m; Агап'фга. (В.23) т=1 т=1 It will be observed that the ordering of indices is exactly as in Equations (4.1) and (4.4). If A is a self-adjoint operator then its corresponding matrix A is Hermi- tian, as, by Equations (B.19) and (B.22), Akj = (фк,Аф^ = (Афк,ф^ = (ф^Аф^* = A*k. Similarly, if U is a unitary operator and U is its corresponding matrix, then U is a unitary matrix. (This follows as (^, ифф = (иф^ фк)* = (тДу, so that Ukj = ((U-1)jfc)*.) Finally, if A, В and C are three bounded operators such that C = AB, and if A, В and C are their corresponding matrices, then C = AB. (As ABi/jj = Cifj for each ф^, j = 1,2, ...,d, then Ckj = {фк,Сф^ = (фк,АВф^. But, from Equation (B.23), Вфз = Y^m=i^rn, Вф^фт, so Ckj = Т^тп=^кЛФтп){Фтп,Вф^ = Й=1Лтп^.) This is the origin of the duality between operators and matrices that is used repeatedly, partic- ularly in Chapter 1, Section 4, and Chapter 4, Section 1.
292 GROUP THEORY IN PHYSICS 5 Bilinear forms Even when a vector space does not possess an inner product it may possess a symmetric non-degenerate bilinear form which gives rise to rather similar properties. In particular this is true of semi-simple Lie algebras (see Chapter 11). Definition Symmetric bilinear form A complex vector space V possesses a symmetric bilinear form В if to every pair of vectors ф and ф of V there corresponds a complex number В(ф,ф) such that (а) В(ф,ф) = В{ф,ф), (b) В(аф, Ьф) = аЬВ(ф) </>), for any two complex numbers a and 6, (с) В(ф + ф) x) = В(ф, x) + В(ф, x), for any х^У- If V is a real vector space the bilinear form В(ф, ф) is required to be real for all 'ф) ф E Vand in (b) a and b are restricted to being real numbers, but otherwise the conditions (a) to (c) are the same as for a complex space. It should be noted that (a), (b) and (c) imply that В(х, аф + Ьф) = aB(x, ф) + ЬВ(х, Ф) and В(аф + Ьф) х) = аВ(ф) х) + ЬВ(ф, х)- There is no requirement that B(^,^>) be real (unless V is a real vector space), and even then В(ф,ф) could be negative, or could be zero with ф 7^ 0. Thus a symmetric bilinear form does not in general have the properties of an inner product (as defined in Section 2). Conversely, if V is a complex inner product space then the inner product is not a symmetric bilinear form (because the right-hand sides of parts (a) and (b) of the definition in Section 2 of an inner product involve complex conjugation, whereas the corresponding parts (a) and (b) of the definition of a symmetric bilinear form do not do so). However, if V is a real inner product space these particular distinctions disappear, so in this case an inner product is also a symmetric bilinear form. Let ф±) ф2, • • •, Фа be a basis for V, and let В be the dxd matrix defined by Bpq = В(фр,фд), p,q = 1,2,... ,d. (B.24) Then, if ф = “jV’J and Ф = ЪкФк, d d В(ф,ф) = ££ Bjka:)bk. (B.25) j=l k=l
APPENDIX В 293 Suppose that ^1? ^2> • • • > another basis for V, with d Фп = ^гпп'фгп-) m=l (n = 1,2,... , d), so that (as noted in Section 1) S is a d x d non-singular matrix. Let B' be the corresponding matrix for the bilinear form defined for this basis, that is, let B'pq = B^'p,^q), p,q — 1,2,... ,d. Then a very straightforward argument shows that B' = SBS. This implies that detB' = (det S)2 det B. Consequently detB' = 0 if and only if det В = 0. If V is a real vector space it can be shown (Gantmacher 1959) that S may be chosen so that B' is diagonal with diagonal elements 1,-1, or 0 only. Then, if the basis ^2> • • • > ordered so that the first d+(> 0) members correspond to 1, the next d_(> 0) to —1, and the remaining do(> 0) to 0, and if and Ф = Цф'р then d+ d_|-+d_ m</o = Ea^ - E a'N (b.26) .7 = 1 j=d+ + l Matrices S with this property can be chosen in an infinite number of ways, but all choices give the same values of the dimensions d+, d_ and do (Gantmacher 1959). The invariant quantity a = d+ — d_ is called the “signature” of the bilinear form. Definition Degenerate and non-degenerate symmetric bilinear forms A symmetric bilinear form В is said to be “degenerate” if there exists in V some 7^ 0 such that В(ф,ф) = 0 for all ф eV. Conversely, a symmetric bilinear form is “non-degenerate” if, for each ф e V, the condition В(ф, ф) = 0 for all </> e V implies that ф = 0. Theorem I The symmetric bilinear form В is non-degenerate if and only if detB 7^ 0, where В is the d x d matrix defined in Equation (B.24). / Proof It should be noted that as detB' = 0 if and only if detB = 0, this condition for non-degeneracy is actually independent of the choice of basis, as is to be expected.
294 GROUP THEORY IN PHYSICS Suppose that there exists a ф E V such that B(V>, Ф) — 0 for all ф E V. By Equation (B.25) this is so if and only if Bj^ajbk = 0 for all sets 61,62, • • • > bd, that is, if and only if ^jkaj = 0 for each к = 1,2,..., d. As this set of d simultaneous linear equations for ai,a2,...,«d has a non- trivial solution (i.e. a solution other than ф = 0) if and only if det В = 0, the quoted result follows. 6 Linear functionals The theory of linear functionals will be considered here only for finite- dimensional vector spaces and inner product spaces. The results will be needed in the discussions of Lie algebras in Chapter 13. The generalization to infinite-dimensional Hilbert spaces may be found in the books of Akhiezer and Glazman (1961) and Riesz and Sz. Nagy (1956). Definition Linear functional If to every member ф of a complex finite-dimensional vector space V a complex number Ф(ф) is assigned in such a way that Ф(аф + Ьф) = аФ(ф) + ЬФ(ф) (В.27) for every ф,ф Е V and any two complex numbers a and 6, then Ф is said to be a “linear functional” on V. Likewise, a linear functional on a real finite- dimensional vector space V is an assignment of a real number Ф(ф) to every ф E V such that Equation (B.27) holds for every ф,ф E V and any two real numbers a and b. If Ф and Ф are any two linear functionals defined on a finite-dimensional vector space V, then (Ф + Ф) may be defined by (Ф + Ф)(ф) = Ф(^) + for all ф E V. Similarly, пФ may be defined by (аФ)(ф) = а(Ф(ф)) for all ф E V, a being any real or complex number as appropriate. Then the set of linear functionals on V themselves form a vector space V*, called the “dual” of V. (The zero of V* is the functional whose value is 0 for all ф E V.) V* is real when V is real and is complex when V is complex. Suppose that V has dimension d and ф2,..., фа is a basis for V. Then each linear functional Ф on V is completely specified by the d numbers Ф(^Д j = 1,2,..., d. (Any ф E V can be written in the form of Equation (B.10) as ф = ^Jj=i aj^j^ so> by Equation (B.27), Ф(ф) = ^2j=i аз®(Фз)-) Let Ф&, к = 1, 2,..., d, be a set of linear functionals defined by ЫФФ=$зк (B.28) for all j.k = 1,2,..., d. The functionals of this set are obviously linearly independent. Moreover, if Ф is any linear functional on V, then Equation (B.28) implies that Ф = Ф(фФ)Фк, that is, Ф depends linearly on Ф1, Ф2,..., Фй. Thus the dual space V* has the same dimension d as V, and Ф1, Ф2,.. •, Фа provide a basis for V*.
APPENDIX В 295 If V is equipped with a symmetric non-degenerate bilinear form, or is an inner product space, the following theorems show that every linear functional is given by a remarkably simple expression. Theorem I Each linear functional Ф on a finite-dimensional vector space equipped with a symmetric non-degenerate bilinear form can be expressed in the form Ф(^) = В^'ф) (B.29) for all ф E V, where В(?/>ф, ф) is the bilinear form, and фф is an element of V which is uniquely determined by the functional Ф. Proof Suppose that фф = has the required property, ^2, • • being a basis for V. Then Equation (B.29) can be written in the form Ф(^) = ^(V^> гФ)аэ-> so that for each к = 1,2,..., d, d Thus if Ф and a are the dx 1 matrices with elements Ф(^) and respectively (fc = l,2,...,d), and В is defined by Equation (B.24), as В is symmetric these equations can be written as Ф = Ba. The linear functional Ф fixes Ф. This equation has a unique solution a when det В 0, namely a = В-1Ф, which then determines фф uniquely. Theorem II Each linear functional Ф on a Hilbert space V can be expressed in the form Ф(^) = (V^VO for all ф E V, where (^ф,^) is the inner product of V, and фф is an element of V which is uniquely determined by the functional Ф. Proof If V is finite-dimensional, a proof can be given along the lines of that of the previous theorem. For the infinite-dimensional case see Akhiezer and Glazman (1961) or Riesz and Sz. Nagy (1956). 'This latter theorem is often called the “Riesz Representation Theorem”. It is easily verified that if фф is as specified in this theorem and Vh, V>2, • • • is an ortho-normal basis for V, then = 52 j 7 Direct product spaces Let Vi and be two complex inner product spaces of dimensions d± and d^. Let i/jj (j = 1,2,..., di) and ф8 (s = 1,2,..., d2) be ortho-normal bases for
296 GROUP THEORY IN PHYSICS Vi and V2 respectively. Then the “direct product” or “tensor product” space Vi 0 V2 may be defined as the complex vector space having the set of did2 “products” ipj 0 фз as its basis, so that Vi 0 V2 is the set of all quantities в of the form 0 = ajs^j 0 Фв, (В.30) j = 1 S=1 where the djs are a set of complex numbers. The direct product of any two elements ф = ^2jLi and Ф = ^2^1сзФз °f ^1 and V2 is defined to be ф 0 ф = У2 У? bjCs^>j 0 ф8, (В.31) j = 1 s=l so that the set of such products is a subset of Vi 0 V2. It is easily verified that all the requirements for Vi 0 V2 to be a vector space are satisfied. (The zero vector of Vi 0 V2 corresponds to ajs = 0 for all j = 1,2,... ,di, and s = 1,2,..., d2.) The products ф^фз are assumed to be linearly independent, so that Vi 0 V2 has dimension did2. (This is assumed to be the case even when Vi and V2 are identical, when one could take фj = for j = 1,2,..., di (= d2), implying that the products фj 0 ф3 and ф3 0 фj are linearly independent.) An inner product can be defined on Vi 0 V2 by assuming that the basis elements фj 0 ф8 are ortho-normal, i.e. that (jpj 0 Фв, Фк 0 Ф^ — djk&st' (В.32) Then if 0 is defined as in Equation (B.30), and di dz x=‘Wj 0 (B.33) j = 1 S=1 it follows that d± dg = (B.34) j=l S=1 This inner product has all the required properties of an inner product space. The definition of Vi 0 V2 and its inner product of Equation (B.34) is actually independent of the choice of the ortho-normal bases of Vi and V2. To see this let ф'к (k = 1,2,..., di) and ф^ (t = 1,2,..., d2) be another pair of ortho-normal bases for Vi and V2 respectively. Then (see Section 2) there exists a di x di unitary matrix F and a d2 x d2 unitary matrix G such that dr k=l and dz t=l
APPENDIX В 297 Then, for any 0 of Vi ® V2, defined as in Equation (B.30), d± d2 d = ^^а'ыф'к®ф'ъ k=l t=l where d± d^ a kt EE ^kl^lu^tu 1 1 = 1 71=1 hereby demonstrating that the set 'ф'к0ф'1. forms an alternative basis for V1012- Moreover, as the vector x of Equation (B.33) can similarly be rewritten as di d2 X= '/’t, k=l t=l with di d2 d*kt EE FkldkuGtu-) 1=1 14=1 and as F and G are unitary, it follows that di d2 M = £^ktd’kt, k=l t=l showing that the inner product is independent of the choice of basis (see Equation (B.34)). In the physics literature the 0 sign is often omitted in products such as ® Фвч but it will be retained throughout this book as a warning that the product is not ordinary multiplication. For abstract inner product spaces Vi and V2, the product 0 in фу 0 ф3 neither requires nor is amenable to any further specification. However, in concrete examples this product can be defined quite naturally. As a first example, suppose Vi = (C3 and V2 = C2, with = (1,0,0), ф2 = (0,1,0), фз = (0,0,1) and ф1 = (1,0), ф2 = (0,1). Then the products фj 0 фв can be defined as the six-component quantities: Ф1&Ф1 = (1,0, 0,0,0,0), фг ®ф2 = (0,0,0,1,0,0), ' ф20ф1 = (0,1, 0, 0,0,0), ф2 ®ф2 = (0,0, 0,0,1,0), > фз^>Ф1 = (0,0,1,0,0,0), фз 0 ф2 = (0,0,0,0,0,1), so that Equation (В.32) is satisfied. Clearly (C3 0 (C2 can be identified with (C6. As a second example, let Vi and ~V2 be subspaces of L2, the elements of Vi being functions ф(гф) of iq and the elements of V2 being functions ф(г2) of r2. Then the elements of Vi 0 V2 are linear combinations of products ф(гф)ф(г2ф the inner product for which may be defined by = f f J f f J dxi (fy], dx2 dyz dz2 V,/*(rl)^>'*(r2)V’(rl)</,(r2)-
298 GROUP THEORY IN PHYSICS As a final example, let Vi be a subspace of L2, consisting of functions V>(r), and let V2 be the two-dimensional space of 2 x 1 matrices with constant entries. Let ^(r) (J = 2,..., di) be an orthonormal basis of Vi and let Ф1 = 1 0 Ф2 = 0 1 be an ortho-normal basis of V2 (it being assumed that the inner product of two elements a b and a' b' of V2 is (a*a' + 6*6'))- Then Vi 0 V2 is the set of two-component quantities V'(r) ^'(r) (^(r),^'(r) e Vi) with inner product defined by = / / /{^*(rMr) + ф'*(r)ri'(r)} dx dy dz for all ^(r),^'(r),7/(r),7/'(r) £ Ti- lt is shown in Chapter 5, Section 4, that if Vi and V2 are both subspaces of L2 consisting of functions of the same variable r, then it may not be possible to identify Vi 0 V2 with the set of linear combinations of ordinary products of members of Vi and V2 and at the same time retain the inner product of L2 as the inner product of Vi 0 V2-
Appendix С Character Tables for the Crystallographic Point Groups The 32 crystallographic point groups of three-dimensional Euclidean space IR3 will be listed in roughly decreasing order of complexity. For each group the following details are given: (a) The group elements. The notation for rotations is as in Chapter 1, 2(a), Cnj denoting a proper rotation through 2тгIn in the right-hand screw sense about the axis Oj and I denoting the spatial inversion operator. All the axes involved are indicated in Figures C.l and C.2. The matrices R(T) for every relevant proper rotation are specified in Table C.l. The rotations are listed in classes. (b) The character table. Several alternative systems of labelling are given, the first column merely giving an arbitrary listing. In the labelling of the second column one-dimensional representations are denoted by 4 or B, two-dimensional irreducible representations by E and three-dimensional irreducible representations by T, in all cases with subscripts and/or su- perscripts attached. (The subscripts g and и (standing for gerade and ungerade) indicate representations that are even and odd under I re- spectively.) However, each member of a pair of one-dimensional complex conjugate representations is given the same label, as they correspond to degenerate eigenvalues. (See, for example, Chapter 6, Section 5(a), and Chapter 7, Section 3(f), of Cornwell (1984).) For a point group that is isomorphic to a group (7o(k) (see Chapter 7, Section 7(a)) for and C^, the third column gives the labelling convention of Bouckaert et al. (1936). In such cases the corresponding к-vector is as defined in Tables 7.2, 7.3, and 7.4. As described in Chapter 7, Section 7(b), it is possible for two or more k-vectors in different stars
300 GROUP THEORY IN PHYSICS Figure C.l: The axes Oa, Ob, Oc, Od, Oe, О f, Oa, 0(3,0-y and 06. Figure C.2: The axes О A, OB, OC, OD. (All these axes lie in the plane Oxy.)
APPENDIX С 301 to have point groups (7o(k) that are isomorphic. The group elements for each such (7o(k) belonging to Of^ and are specified when this occurs. (c) Matrices for the irreducible representations of dimension greater than one. (Of course these are only unique up to a similarity transformation). For one-dimensional representations the characters themselves are the matrix elements. The notation employed for the point groups is that of Schonfliess (1923). More information on these groups may be found in the book of Koster et al. (1964), which is wholly devoted to this subject, and the articles of Altmann (1962, 1963). (1) Oh: (a) Classes [for <70(k) °f Г, Я, and R\: G = E- C2 = C3a, Сзв, C3y, C3S, C3-1, C3-/, C3-1, Сз,1; C3 = C2x, C2y, C2z; C4 = C4x, C4y, C4z, Cib) Cic-) Cie-) Cq = I] C7 = IC3a,IC3e,IC3y, IC3S,IC^,IC~^ 1С3у\1С-/-, Cs = IC2x,IC2y, ic2z-, C9 = IC4x,IC4y, IC4z, IC4x\lC4v\lC^-, C10 = IC2a, IC2b, IC2c, IC2d, IC2e,IC2f. (b) The character table is given in Table C.2. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of Oh'. r3(Gv) = r8(CnJ) = r"(Cnj- г3(КЛч) = -rs(icnj) = r"(CnJ r4(C„y) = r9(Cnj) = R(GJ ПЛ?»,) = -r9mj = Wni) r5(cnj) = r10(Cnj) = r'(Cnj.) г5(ад = -r10(/cnj) = Г'(СпУ) where the matrices r'((7nj), P"(Cnj) and R(<7nj) are given in Tables С.3, C.4, and C.l. (2) DQh-. (a) Classes [for <7o(k) °f Г, Я, and R\: = E-C2 = Cez, C^1; Сз = C3z, C3-1; C4 = C2Z1 C5 = Cix, C2A1 С2В1 Св = C2y, C2C1 C2D1 C7 = I-,Cs = IC6z,IC&-, C9 = IC3z,IC^-, C10 = IC2z, Cn = IC2x,IC2A,IC2B-, C12 = IC2y,IC2c,IC2D. (b) The character table is given in Table C.5.
302 GROUP THEORY IN PHYSICS 1 0 0 " 1 0 0 R(£) = 0 1 0 R(C2x) = 0 -1 0 0 0 1 0 0 - 1 -1 0 0 -1 0 0 R(C2v) = 0 1 0 R(c2z) = 0 -1 0 0 0 -1 0 0 1 1 0 0 ' ‘10 0 R(C4x) = 0 0 1 R(^) = 0 0-1 > 0 -1 0 0 1 0 0 0 - -1 0 0 1 R(C4a) = 0 1 0 5 R(c4”1) = 0 1 0 > 1 0 0 -10 0 0 1 0 ' 0 -1 0 R(C4z) = -1 0 0 = 1 0 0 0 0 1 0 0 1 0 1 0 " 0 -1 ( ) R(C2a) = 1 0 0 R(C2b) = -1 0 0 5 0 0- -1 0 0 -1 0 0 1 0 0 1 R(C2c) = 0 -1 0 R(CM) = 0 -1 ( 1 5 1 0 0 -1 0 ( 1 -1 0 0 ’ ' -1 0 1 Э R(C2e) = 0 0 1 R(C2/) = 0 0 1 5 0 1 0 0 -1 1 0 — 0 1 0 " 0 -1 0 R(C3a) = 0 0 -1 R(C3/3) = 0 0- 1 > -1 0 0 1 0 0 •4 0 -1 0 0 1 0 R(C37) = 0 0 1 R(C3«) = 0 0 1 -1 0 0 1 0 0 0 0 -1 0 0 1 R(^) = 1 0 c ) R(^) = -1 0 0 > 0 -1 c ) 0 -1 0 0 0 -1 ‘ 0 0 1 R^’1) = -1 0 c ) R(C3-X) = 1 0 0 0 1 c ) 0 1 0 _ 1 2 1 / 2 V 3 0 “ 1 1 2 2 3 0 ' R(c3z) = 1 2 0 R(C3-1) = 1 2 0 5 0 0 1 0 0 1 = 1 2 1 / 2 V 3 0 " —If 1 tsO II-* 3 0 " R(C6z) = 1 2 0 R^e"1) = |V3 1 2 0 5 0 0 1 0 0 1 “2 2^ 0 _1 _ 2 1 2 'з 0 " R(C2A) = iV3 1 2 0 , R(C2b) = -|V3 2 0 0 0 -1 0 0 1 = 1 2 1 / 2 V 3 0 ' : 11/ 2 2V "3 0 ' R(C2c) = — • 1 2 0 , R(C2d) = - 1 2 0 0 0 - -1 0 0 -1 Table C.l: The matrices R(T) for the proper rotations T appearing in various crystallographic point groups.
APPENDIX С 303 Ci C2 C3 C4 C5 Cq C7 Cs Cq Сю Г1 ^1.9 Г1 Hi Ri 1 1 1 1 1 1 1 1 1 1 Г2 Г2 H2 R2 1 1 1 -1 -1 1 1 1 -1 -1 Г3 Eg Г12 H12 R12 2 -1 2 0 0 2 -1 2 0 0 Г4 Tig Г15' H15f R15f 3 0 -1 1 -1 3 0 -1 1 -1 Г5 Ъд Г25' ^25' ^25' 3 0 -1 -1 1 3 0 -1 -1 1 Г6 Гр Яр Яр 1 1 1 1 1 -1 -1 -1 -1 -1 Г7 Аъи Г2' Ну R2' 1 1 1 -1 -1 -1 -1 -1 1 1 Г8 Eu Г12' #12' R12' 2 -1 2 0 0 -2 1 -2 0 0 Г9 Tlu Г15 Я15 Я15 3 0 -1 1 -1 -3 0 1 -1 1 plO T2u Г25 Н25 Я25 3 0 -1 -1 1 -3 0 1 1 -1 Table С.2: Character table for (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of Deh‘. T\Cnj) r5(ICnj) Г6(С^) r6(/Cnj) Г11^) = V'(Cnjy, -rn(ICnj) = D\Cnjy r12(CnJ) = D"((7nJ); -r12(lCnJ) = D"(Cnj); where the matrices D'(<7nj) and D"((7nj) are given in Tables C.6 and C.7 respectively. (3) Td: (a) Classes [for Po(k) of P]: G = E-, C2 = C3a, C30, C37, C3S, C3-1, C3-;, C3-1, Сз,1; c3 = c2x,c2y,c2z-, C4 = IC4x,IC4y,IC4z,IC^,IC^ICZ1-, c5 = IC2a,IC2b, IC2c, IC2d, IC2e, IC2f. (b) The character table is given in Table C.8. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of 7^: r3(CnJ) = r"(CnJ); r4(OJ = R(CnJ); Г5(СП,) = E'(Cnjy and for any improper rotation ICnj of T^. P\icnj) = -r"(CnJ); r4(dC'nj = -R(C'nJ); Г5(/Сп>) = -p\cnjy where the matrices T^Cnj), r"(Cnj) and R(Cnj) are given in Tables С.3, C.4, and C.l.
304 GROUP THEORY IN PHYSICS 1 0 0 " 1 0 0 Г(Е) 0 1 0 Г'(С2ж) = 0 1 0 _ 0 0 1 _ 0 0- -1 -1 0 0 " 1 0 0 Г'(С2у) = 0 -1 0 , Г'(С2г) = 0 -1 0 0 0 1 _ 0 0 - -1 0 0 1 ’ 0 0 - -1 r'(<x) = 0 -1 0 , mV) = 0 -1 0 -1 0 0 _ 1 0 0 " 0 -1 0 " - 0 1 0 r'(cu = 1 0 0 , mV) = 1 0 0 _ 0 0 - -1 _ 0 0- -1 -1 0 0 " - 1 0 0 Г'(С4г) = 0 0 - -1 , mV) = 0 0 1 0 1 0 _ 0 -1 0 1 0 0 ’ 1 0 0 ' Г'(С2а) = 0 0 - -1 , Г'(С2Ь) = 0 0 1 _ 0 -1 0 _ _ 0 1 0 . 0 -1 0 " 0 1 0 ’ Г'(С2с) = -1 0 0 , r'(C2d) = 1 0 0 0 0 1 _ 0 0 1. 0 0 - -1 0 0 1 ' Г'(С2е) = 0 1 0 , Г'(С2/) = 0 1 0 -1 0 0 _ _ 1 0 0. 0 -1 0 " 0 1 0 Г'(Сза) = 0 0 1 , Г'(Сз^ = 0 0- -1 _ -1 0 0 _ — 1 0 0 0 -1 0 0 1 0 Г'(Сз7) = 0 0 - -1 , Г'(СЗЙ) = 0 0 1 1 0 0 _ 1 0 0. 0 0 - -1 ’ 0 0 - -1 т-1) = -1 0 0 > mV) = 1 0 0 0 1 0 _ _ 0 -1 0 0 0 1 " 0 0 1 ' mV) = -1 0 0 > mV) = 1 0 0 0 -1 0 0 1 0 Table C.3: The matrices Г' for the proper rotations of Oh, Ta, O, Th, and T.
APPENDIX С 305 Г"(Е) = Г"(С2х) = Г"(С2г/) =Г"(С2г) 1 о О 1 = Г"(Сз7) = Г"(Сзб) Г"(Сза) = Г"(С3/3) VW») =Г"(С37) г"(<М =Г"(С4-1) г"(см =г"(с4-1) Г"(С42) =Г"(С4'1) = г'ЧСз-1) = r"(c37: = r"(C2e) = Г"(С2/) = r"(C2c) = r"(C2d) = r"(C2a) = Г"(С2Ь) Table С.4: The matrices Г" for the proper rotations of Oh, Tfj, O, Th, and T. С1 с2 Сз с4 с5 Сб Ст Cs Сэ Сю Сц Ci2 г1 Alg, 1 1 1 1 1 1 1 1 1 1 1 1 г2 1 1 1 1 -1 -1 1 1 1 1 -1 -1 г3 Bi^ 1 -1 1 -1 1 -1 1 -1 1 -1 1 -1 г4 #2<7 1 -1 1 -1 -1 1 1 -1 1 -1 -1 1 г5 Е29 2 -1 -1 2 0 0 2 -1 -1 2 0 0 г6 Е1д 2 1 -1 -2 0 0 2 1 -1 -2 0 0 г7 А1и 1 1 1 1 1 1 -1 -1 -1 -1 -1 -1 г8 Л2и 1 1 1 1 -1 -1 -1 -1 -1 -1 1 1 г9 В1и 1 -1 1 -1 1 -1 -1 1 -1 1 -1 1 р,10 Въи 1 -1 1 -1 -1 1 -1 1 -1 1 1 -1 г11 Еъи 2 -1 -1 2 0 0 -2 1 1 -2 0 0 j--, 12 Е±и 2 1 -1 -2 0 0 -2 -1 1 2 0 0 Table C.5: Character table for D^h-
306 GROUP THEORY IN PHYSICS D'(E) = D'(C22) D'(C6z) = D7C3-1) Ж,1) = D'(C3z) = D'(C2j,) D'(C2a) =D'(C2O) D'(C2B) = D'(C2B) 1 0 0 1 Table C.6: The matrices D' for the proper rotations of -Рвл.> D3h, C3v, and D6. D"(C2B)= 1 1 1 Ю|м 1 Ю|н-1 со] -1^3 1 2 D"(C2C) = 1 2 —-л/З L 2 _ 1 2 D"(C2a) = D"(<72j,) = D"(C6z) = D"(C3z) = D"(C2z) = -i 2 2 175 1 2 V° 2 -1 0 ’ 0 1 Table C.7: The matrices D" for the proper rotations of Cqv1 Dq and D3.
APPENDIX С 307 Table С.8: Character table for Tj and O. (4) O: (a) Classes: C1 = e-,c2 = c3a, c3/3, c3y,c3S, Сз = Czy, C2z, ^4 = , (>4^, C4z, C±x , C4y , C4z C5 = Clb, Cici Cid, CieiCif. (b) The character table is given in Table C.8. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of O: r3(CnJ) = Г"(СПД r4(CnJ) = Г'(СП>); T5(Cnj) = R(Cnj); where the matrices r'(Cnj), r"(Cnj) and R(C'nj) are given in Tables С.3, C.4, and C.l. (5) Th: C1=E-C2 = C3a, C3p, C3y,C3S-, C3 = С^,С^,С^,С^; c4 = C2x, Cly, Ciz', С3 = Cq = IC^IC^ IC^IC^ C7 = Cs = IC2x,IC2y,IC2z. (b) The character table is given in Table C.9. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of 7\: r4(Cnj) = r8(CnJ) = r'(Cnj); r4(/cnJ) = -r8(/cnj = r'(cnjy, where the matrices r'((7nj) are given in Table C.3.
308 GROUP THEORY IN PHYSICS Ci C2 C3 C4 C5 Се C7 Cs Г1 Г2 Г3 Г4 Г5 Г6 Г7 Г8 Ag E3 e9 T3 Au Eu Eu Tu 11111111 1 ф Ф2 1 1 ф Ф2 1 1 ф2 Ф 1 1 ф2 Ф 1 3 0 0 -1 3 0 0 -1 111 1 -1 -1 -1 -1 1 ф Ф2 1 -1 -ф -Ф2 -1 1 ф2 Ф 1-1 -ф2 -Ф -1 3 0 0 -1 -3 0 0 1 Table C.9: Character table for Th (ф = ехр(|тгг)). £i C2 C3 C4 C5 Св С? С s Cq Сю Г1 Xi, Mi 1 1 1 1 1 1 1 1 1 1 Г2 Big %2, M2 1 1 1 -1 -1 1 1 1 -1 -1 r3 B2g X3,M3 1 -1 1 -1 1 1 -1 1 -1 1 r4 A2g X4,M4 1 -1 1 1 -1 1 -1 1 1 -1 r5 Eg X5, M5 2 0 —2 0 0 2 0 —2 0 0 Г6 Xu, Myi 1 1 1 1 1 -1 -1 -1 -1 -1 Г7 Blu X^f 1 XI<2 > 1 1 1 -1 -1 -1 -1 -1 1 1 Г8 B2u JC3/, М31 1 -1 1 -1 1 -1 1 -1 1 -1 Г9 a2u X41, 1 -1 1 1 -1 -1 1 -1 -1 1 plO Eu X5f,M5f 2 0 —2 0 0 —2 0 2 0 0 Table С. 10: Character table for D^h. (6) Dih-. (a) (i) Classes [for f7o(k) of X]: Ci = E-, C2 = C2x, C2y; C3 = C2z', C4 = Ciz,C^-, C5 = C2a, C2b; Ce = I-,C7 = IC2x,IC2y-, Cs = IC2z-, C9 = IC4z,IC4z\ C10 = IC2a,IC2b. (ii) Classes [for <70(k) of M]: Ci = E; C2 = C2yi C2Z1 C3 = C2x; C4 = C4X) C5 = C2ei C2f\ Cq = I\ C7 = IC^y, IC2Z1 Cs = Ю2Х1 C9 = IC4x,IC4I\ C10 = IC2e,IC2f. (b) The character table is given in Table C.10. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of D^: r5(Cnj) = r10 (ii)(Cn>) = D(Cnj); r5(/Cn>) = - r10(/Cnj) = D(C'nj); where for <70(k) of M the matrices D((7nj) are given in Table C.ll, while for Po(k) of X the matrices D((7nj) are given in Table C.12.
APPENDIX С 309 D(E) = 1 0 ’ 0 1 D(C2x) = ’ -1 0 0 -1 , D(C2v) = ’ 1 0 0 -1 D(C2z) = ’ -1 0 0 1 , D(C-ta) = 0 1 ’ -1 0 , d(c4-;) = ’ 0 -1 1 0 D(C2e) = ’ 0 1 1 0 D(C2/) = 0 -1 -1 0 Table C.ll: The matrices D for the proper rotations of D^h for Po(k) of M. D(S) = 1 0 ’ 0 1 D(C2J = ’ 1 0 ’ 0 -1 D(C2V) = -1 0 ’ 0 1 D(C2z) = ' -1 0 0 ’ -1 , D(C4z) = 0 1 ’ -1 0 D^1) = ’ 0 -1 ’ 1 0 D(C2a) = ’ 0 1 1 0 D(C2b) = 0 -1 -1 0 Table C.12: The matrices D for the proper rotations of for Po(k) of (7) D3h' (a) Classes: Cl = E; C2 = Сз^Сз”1; Сз = С’2ж, С'2Л, C'2B; C4 = IC2z-, C5 = IC6z,IC^-, C6 = IC2y,IC2C,IC2D. (b) The character table is given in Table C.13. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of D^- T3(Cnj) = r6(Cnj) = D'(Cnjy, and for any improper rotation ICnj of D^: V3(ICnj) = -V\icnj) = D'(Cnj); where the matrices D'((7nj) are given in Table C.6. (8) D3d\ (a) Classes [for Po(k) of L\: Ci = E, C2 = Сзб,(73/; Сз = Сгь,C2d,С2/; С4 = /; С5 = 1Сы,1С-^, Сб = IC2b,IC2d,IC2f. (b) The character table is given in Table C.13.
310 GROUP THEORY IN PHYSICS D^h Dzd Cqv D6 Cl c2 ^3 Ci c5 c6 Г1 A'i Alg Ai Ai Z1 1 1 1 1 1 1 Г2 A2 A2g A2 A2 l2 1 1 -1 1 1 -1 Г3 E' e9 e2 e2 L3 2 -1 0 2 -1 0 Г4 A'/ -din Bi Ei' 1 1 1 -1 -1 -1 Г5 A2u B2 B2 L2! 1 1 -1 -1 -1 1 Г6 E" Eu Ei Er L31 2 -1 0 —2 1 0 Table C.13: Character table for Cqv and Dq. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of D^d- r3(Cnj) = r6(CnJ) = D"(Cnj); r3(/Gy) = -r6(/Cnj) = D"(Cnj); where the matrices D"(Cnj) are given in Table C.4. (9) C6v: (a) Classes: Ci = E- C2 = C3z,C^-, C3 = IC2x,IC2A,IC2B-, c4 = c2z-, c5 = c6z,c^-, C6 = IC2y,ic2C,IC2D. (b) The character table is given in Table C.13. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of Cqv: V\cnj) = V"(Cnj)-, r3(Cnj) = D'(Cnj); and for any improper rotation ICnj of Cqv: r6(ICnj) = D"(Cnj); r3(/CnJ) = D'(Cnjy, where the matrices D'((7nj) and D"(Cnj) are given in Tables C.6 and C.7 respectively. (10) C6h: (a) Classes: Ci = E- C2 = C3z, c3 = c3z-, Ci = C2z; C5 = C3-1; C6 = C^1; c7 = i-,c8 = ic6z-, c9 = ic3z; Clo = IC2z, Cn = ic£-, C12 = IC£. (b) The character table is given in Table C.14.
APPENDIX С 311 Ci C2 Сз C4 C5 Св С7 Cs Сд Сю Сц С12 Г1 ^9 1 1 1 1 1 1 1 1 1 1 1 1 Г2 B9 1 -1 1 -1 1 -1 1 -1 1 -1 1 -1 Г3 Eg 1 DJ cj2 -1 -CJ -CD2 1 DJ Ш2 -1 -CJ -cj2 Г4 E'g 1 -cj2 -CJ -1 cj2 CD 1 -cj2 -UJ -1 cj2 DJ Г5 E'' 1 cj2 -CJ 1 cj2 -UJ 1 w2 -UJ 1 cj2 -DJ Г6 E'' 1 -CJ cj2 1 -CJ cj2 1 -UJ cj2 1 -CJ cj2 Г7 Au 1 1 1 1 1 1 -1 -1 -1 -1 -1 -1 Г8 Bu 1 -1 1 -1 1 -1 -1 1 -1 1 -1 1 Г9 E’u 1 DJ CD2 -1 -CJ -UJ2 -1 -CJ 1 DJ Ш2 p,10 E'u 1 -CD2 -UD -1 CD2 UJ -1 w2 DJ 1 -Ш2 -DJ Г11 E" 1 cj2 -UJ 1 cj2 -UJ -1 -cj2 UJ -1 -cj2 UJ p,12 E” 1 -CJ 1 -CJ UJ2 -1 DJ -cj2 -1 DJ -CD2 Table C.14: Character table for C^h (w = exp(| 7TZ)). (И) В6: (a) Classes: Ci = Е\ С2 = Сз2, С3/; Сз = Сг®, Сгл, Сгв; С4 = (?22; С5 = Cqz1 Cqz‘, Cq = С2у> C2C1 C2D- (b) The character table is given in Table C.13. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of Cqv: P\Cnj) = D"(Cnj); r3(Cni) = D'(C^); where the matrices D'(C'nj) and D"((7nj) are given in Tables C.6 and C.7 respectively. (12) T: (a) Classes: Cl = E; C2 = C3a, C3f3, C3y, C3S-, C3 = C3-1, C3-1, C37; C4 = С2Х4 C2l/, Cj2z' (b) The character table is given in Table C.15. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of T: r4(Cnj) = r'(Cnj); where the matrices r'((7nj) are given in Tables C.3.
312 GROUP THEORY IN PHYSICS Ci C2 C3 C4 Г1 Г2 Г3 Г4 A E E T 1111 1 ф ф2 1 1 ф2 ф 1 3 0 0 -1 Table C.15: Character table for T (ф = ехр(|тгг)). Ci C2 Сз c4 C5 Ce C7 C8 г1 Alg, Ni 1 1 1 1 1 1 1 1 г2 Big n2 1 -1 1 -1 1 -1 1 -1 г3 В2д N3 1 -1 -1 1 1 -1 -1 1 г4 Взд n4 1 1 -1 -1 1 1 -1 -1 г5 A in n2, 1 1 1 1 -1 -1 -1 -1 г6 Blu Nv 1 -1 1 -1 -1 1 -1 1 г7 В2и 1 -1 -1 1 -1 1 1 -1 г8 Взи N3, 1 1 -1 -1 -1 -1 1 1 Table C.16: Character table for Z>2/i- (13) D2h or Vh: (a) Classes [for Po(k) °f ^]: Ci = E; C2 = C2x-, C% = <72e; C4 = C2/; C5 = I; Cq = IC2x\ C7 = IC2e- Cs — IC2f. (b) The character table is given in Table C.16. (14) C4v: (a) Classes [for (7o(k) of Д]: Ci = E- c2 = c2z, c3 = ciz, c£-, C4 = IC2x, ic2y; C5 = IC2a, ic2b. (b) Classes [for <70(k) of T]: c4 = e-c2 = c2x, c3 = cix,c£-, C4 = IC2y,IC2z, C5 = IC2e,IC2f. (c) The character table is given in Table C.17. (d) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of C^v: r5(CnJ) = О(С^); and for any improper rotation ICnj of C^v: г5(/Ои) = -d(c^); where for (7o(k) of Д the matrices D((7nj) are given in Table C.12 and for Po(k) of T they are given in Table C.ll.
APPENDIX С 313 Cl c2 c3 c4 cs Г1 Ai Д1Д1 VZ1 1 1 1 1 1 Г2 Bl A.2.T2 W2, 1 1 -1 1 -1 Г3 a2 Ai',Ti' W2 1 1 1 -1 -1 Г4 B2 Д2' ,T2, 1 1 -1 -1 1 Г5 E Д5Д5 w3 2 —2 0 0 0 Table С. 17: Character table for Сфу, and D2d. (15) D4: (a) Classes: Ci = E; C2 = C2y; C% = C±y, C4 = C2x,C2z, C5 = C2c,C2d. (b) The character table is given in Table C.17. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of Z)4: r5(Cn>) = S(Cnjy, where the matrices S((7nj) are given in Tables C.18. (16) D2d or Vd: (a) Classes [for Po(k) of W]: G = E; c2 = c2y-, C3 = IC4y,IC^-, c4 = ic2x,ic2z-, C5 = C2c,C2d. (b) The character table is given in Table C.17. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of D2d\ r5(Cnj) = S(CnJ); and for any improper rotation ICnj of D2d: r5(ICnj) = -S(Cnjy, where the matrices S(Cnjj are given in Tables C.18. (17) C4h: (a) Classes: G = E; C2 = С4г; C3 = C2z; C4 = C^1; C5 = I; C6 = IC4z\ C7 = IC2z-, C8 = IC£. (b) The character table is given in Table C.19.
314 GROUP THEORY IN PHYSICS S(£?) = S(C2.) = S(C2c) = S(C2d) = 0 -1 ’ -1 0 , S(C4J,) = ' 0 -1 ’ 1 0 , S(C4-X) = S(C2.) = ' 1 o’ 0 -1 ’ S(C22) = ' -1 0 ’ 0 1 0 1 1 0 0 1 -1 0 1 0 0 1 Table C.18: The matrices S for the proper rotations of D4 and Did- Cl c2 Сз c4 c5 c6 C7 Cs Г1 Ag 1 1 1 1 1 1 1 1 Г2 Bg 1 -1 1 -1 1 -1 1 -1 Г3 Eg 1 i -1 —i 1 i -1 —i Г4 e9 1 —i -1 i 1 —i -1 i Г® -du 1 1 1 1 -1 -1 -1 -1 Г6 Bu 1 -1 1 -1 -1 1 -1 1 Г7 Eu 1 i -1 —i -1 —i 1 i Г8 Eu 1 —i -1 i -1 i 1 —i Table C.19: Character table for D^h- (18) C3h: (a) Classes: Cl = E; C2 = IC3z; C3 = C3z, C4 = IC2z; C5 = Cg-1; C6 = IC&. (b) The character table is given in Table C.20. (19) CV (a) Classes [for (7o(k) of A]: Ci = E; C2 = C3S,C^-, C3 = IC2b,IC2d,IC2f. (b) Classes [for (7o(k) of F]: Ci = E; C2 = C3a, C3-1; Сз = IC2b,IC2c,IC2e. c3h c3l c6 Ci C2 Сз C4 C5 Cq Г1 Г2 Г3 Г4 Г5 Г6 Af Ag A A" Au В E" Eu E' E" Eu Ef E’ Eg E" Ef Eg E" 1 1111 1 1-1 1-1 1-1 1 CJ cj2 — 1 —CJ —CJ2 1 —CJ2 —CJ —1 CJ2 CJ 1 CJ2 —CJ 1 CJ2 —CJ 1 —CJ CJ2 1 —CJ CJ2 Table C.20: Character table for C3h, C3t and Cq (<v = ехр(|тгг)).
APPENDIX С 315 Г1- г2 г3 Ai Аъ FiAi F2A2 F3A3 С1 С2 Сз 1 1 1 -1 -1 О Table С.21: Character table for C3v and D3. (c) The character table is given in Table C.21. (d) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of C^v\ T\Cnj) = Y'\Cnjy, and for any improper rotation ICnj of C^v: Ts(ICnj) = Г"(СПД where the matrices r"(Cnj) are given in Tables C.4. (20) D3: (a) Classes: Ci = E\ C2 = Сзг1С3^\ C3 = C2X1C2A1 C2B- (b) The character table is given in Table C.21. (c) Matrices for irreducible representations of dimension greater than one, for any proper rotation Cnj of D3: r\Cnj) = D'(Cnjy, where the matrices D'((7nj) are given in Tables C.6. (21) С3г or 56: (a) Classes: Cl = E-, C2 = /С3-1; Сз = C3z, C4 = /, C5 = C3-1, C6 = IC3z. (b) The character table is given in Table C.20.
316 GROUP THEORY IN PHYSICS C2v C2h D2 Cl Ai A9 Ai Si Si Zi Gi Bi A2 Au Bi s2 s2 Z2 G2 B2 Bi Bu B2 S3 S3 z3 G3 B3 B2 Bg B3 s4 S4 Z4, Gt B4 C2 Сз C4 1 1 1 1 -1 -1 -1 -1 1 -1 1 -1 Table C.22: Character table for C^, C2h and D2. (22) C6: (a) Classes: Ci = E- c2 = cez-, c3 = c3z, c4 = c2z, c5 = C3-1, c6 = c£. (b) The character table is given in Table C.20. (23) C2v-. (a) Classes [for Po(k) of S]: Ci = e-c2 = c2e-, C3 = IC2x, C4 = IC2f. Classes [for (7o(k) of D] : Ci = E; C2 = C2x; C3 = IC2e, C4 = IC2f. Classes [for Po(k) of S]: Ci = E; C2 = C2a; C3 = IC2z, C4 = IC2b. Classes [for Po(k) °f Ci = E, C2 = C2y, C3 = IC2z, C4 = IC2x. Classes [for <70(k) of G]: Ci = E; C2 = C2f, C3 = IC2x, C4 = IC2e. (b) The character table is given in Table C.22. (24) C2h: (a) Classes: Ci = E, C2 = C2z, C3 = I, C4 = IC2z. (b) The character table is given in Table C.22. (25) D2 or V: (a) Classes: Ci = E, C2 = Сз = C2y, C4 = C2z. (b) The character table is given in Table C.22. (26) C4: (a) Classes: Ci = E\ C2 = (742; C3 = C2zi C4 = С4/.
APPENDIX С 317 Ci C2 C3 Г2 Г3 Г4 В 1 -1 1 El i -1 E 1 -i -1 C4 1 -1 —i i Table C.23: Character table for C4 and 54. Cl T^ r2 r3 A E 1 1 1 C2 C3 1 1 Ф Ф2 Ф2 Ф Table C.24: Character table for C3 (ф = exp(|тгг)). (b) The character table is given in Table C.23. (27) S4: (a) Classes: Cl = E- C2 = ICiy, Сз = C2y, C4 = IC^. (b) The character table is given in Table C.23. (28) C3: (a) Classes: Ci = E; C2 = C3z; C3 = . (b) The character table is given in Table C.24. (29) Cs or Cih'- (a) Classes: Ci = E; C2 = IC2z- (b) The character table is given in Table C.25. Г2 cs C2 A' A A" В Сг Ад Аи Qi Q2 Ci 1 1 C2 1 -1 Table C.25: Character table for (7S, C2 and Ci.
318 GROUP THEORY IN PHYSICS (30) C2: (a) Classes [for <7o(k) of Q]: Ci = E; C2 = C2d. (b) The character table is given in Table C.25. (31) Ci or S2: (a) Classes: G = E- C2 = I. (b) The character table is given in Table C.25. (32) Ci: (a) This group consists of E alone. (b) X(E) = 1.
Appendix D Properties of the Classical Simple Complex Lie Algebras 1 The simple complex Lie algebra А/, I > 1 (a) The Dynkin diagram is given in Figur e D.l. (b) The Cartan matrix of Ai is "2-1 0 .. 0 0 0 " -1 2 -1 .. 0 0 0 0-1 2 .. 0 0 0 A = 0 0 0.. 2-1 0 0 0 0.. . -1 2-1 0 0 0.. 0-1 2 (c) (i) Ai has |Z(Z + 1) positive roots, namely Ylp=j ap f°r = 1, 2,..., I with j < k. (ii) It is sometimes convenient to introduce (Z + 1) auxiliary linear functionals 61,61,..., 6/+i on H. In terms of these the positive l a a2 Figure D.l: The Dynkin diagram of Ai (I = 1, 2,3,...).
320 GROUP THEORY IN PHYSICS roots are (cp — (p, q = 1, 2,..., I +1; p < q), with aj = 6j — 6j+i (for j = 1,2,...,/). (d) The dimension n of Ai is given byn = (Z + l)2 — 1. (e) The quantities (а7-,ад.) defined in Equation (11.10) are given for Ai by ( 1/(Z + 1), for j = = 1,2,...,/); (aj,ak) = < — 1/{2(Z + 1)}, for j = к ± 1, (J, к = 1, 2,..., /); [ 0, for other values of j, A;, (j, к = 1,..., Z). (f) The order of the Weyl group W for Ai is (Z + 1)!. (g) The fundamental weights of Ai are {1/(г + 1)}£р=1(/ + 1-р)аР’ forj = l; - {i/G + +1 - p)aP + Ep=? N +1 _ „ for j = 2,3,..., I, as A”1 = {l/(Z + l)}x I (Z - 1) (Z - 2) (Z - 3) ... 3 2 1 (Z - 1) 2(Z - 1) 2(Z - 2) 2(Z - 3) ... 6 4 2 (Z - 2) 2(Z - 2) 3(Z - 2) 3(Z - 3) ... 9 6 3 (Z - 3) 2(Z - 3) 3(Z - 3) 4(Z - 3) ... 12 8 4 3 6 9 12 ... 3(Z - 2) 2(Z - 2) (Z - 2) 2 4 6 8 ... 2(Z - 2) 2(Z - 1) (Z - 1) 1 2 3 4 ... (Z - 2) (Z - 1) I (h) The compact real form of Ai is su(Z + 1). (i) The adjoint representation is Г({2}) for Ai and Г({1,0,..., 0,1}) for A, Z > 1. 2 The simple complex Lie algebra Bi, I > 1 (a) The Dynkin diagram is given in Figure D.2. (b) The Cartan matrix of Bi is 2-1 0 -1 2 -1 0-1 2 0 0 0 0 0 0 0 0 0 A = 0 0 0 0 0 0 0 0 0 2-1 0 -1 2 -1 0-2 2
APPENDIX D 321 In particular, for Z = 2 and I = 3 respectively A = 2 —2 -1 2 2-1 0 -1 2 -1 0-2 2 and A = (c) (i) Bi has Z2 positive roots, namely Ep=jaP’ for j = 1,2,..., I; Ep=}ap + 2Yfp=kapy for = i,2,...,Z; j <k; * forj,k = 1,2,...,I; j < k. , (ii) It is sometimes convenient to introduce I auxiliary linear function- als 6i, 6i,..., 6i on H, with aj = 6j — 6j+i (for j = 1,2,..., Z — 1) and ai = 6i. The pattern of positive roots then appears more regular, as it consists of tj O' = 1,2,..., Z) and (e,- ± ek) (j, к = 1,2,... ,1; j < k). (d) The dimension n of Bi is given by n = 1(21 + 1). (e) The quantities (aj,ak) (as defined in Equation (11.10)) are given for Bi by (aj, afc) — < 1/(2Z-1), 1/{2(2Z-1)}, -1/{2(2Z - 1)}, o, for J = A;, (> = 1, 2,..., Z — 1); for j = к = I; for j = к ± 1, (j, к = 1, 2,..., Z); for all other J, A;, (j, к = 1,..., Z). (f) The order of the Weyl group W for Bi is 2ll\. (g) The fundamental weights of Bi are 52p=i ap-> Ep=ipaP + Ep=p'aP> 1 2Ep=lP«P> for j = 1; for j = 2,3,..., Z - 1; for j = I; Figure D.2: The Dynkin diagram of Bi (Z = 1, 2,3,...).
322 GROUP THEORY IN PHYSICS as 1111... 1 1 I 1 2 2 2... 2 2 1 1 2 3 3... 3 3 j 1 2 3 4... 4 4 2 1 2 3 4 ... (Z-2) (Z-2) |(Z —2) 1 2 3 4 ... (Z-2) (Z-l) |(Z-1) 1 2 3 4 ... (Z-2) (Z-l) |Z (h) The compact real form of Bi is so(2Z + 1). (i) The adjoint representation is Г({2}) for B1 and Г({0,1,0,..., 0}) for Bi for Z > 1. 3 The simple complex Lie algebra I > 1 (a) The Dynkin diagram is given in Figure D.3. (b) The Cartan matrix of Ci is 2-1 0 -1 2 -1 0-1 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 2-1 0 -1 2 —2 0-1 2 In particular, for Z = 2 and I = 3 respectively 2 —2 and A = 1 2 2-1 0 -1 2 —2 0-1 2 a I a2 Figure D.3: The Dynkin diagram of Ci (Z = 1, 2,3,...).
APPENDIX D 323 (c) (i) Ci has Z2 positive roots, namely EpZjap, for j,k = 1,2, j < к-, ap + 2 SpLfe ap + ab for j, к = 1,2,..., I - 1; j < к-, Y!p=jaP + ab for j = 1,2,... ,1 - 1; ► 2 EpLj otp + at, for j = 1,2,..., I - 1; ai. (ii) It is sometimes convenient to introduce I auxiliary linear function- als 6i, 6i,..., €i on H, with aj = ej — 6j+i (for j = 1, 2,..., Z — 1) and ai = 2б/. The pattern of positive roots then appears more regular, as it consists of (cj е/г), (j\ к 1, 2,..., Z, j <Z &), 1 (^ + 6fc), (7, k = 1, 2,..., Z; j < к). J (d) The dimension n of Ci is given by n = Z(2Z + 1). (e) The quantities {aj^ak) (as defined in Equation (11.10)) are given for Ci by {aj,ak} = < WG + 1)}, 1/G + l), -1/{4(Z + 1)}, for j = k,(j = 1,2,..., Z — 1); for j = к = I; for j = к ± 1, (7, к = 1, 2,..., I - 1); for j = I — 1, к = Z, & j = Z, к = I — 1; for all other J, A;, (7, к = 1,..., Z). 0, (f) The order of the Weyl group W for Ci is 2ll\. (g) The fundamental weights of Ci are < y-u—1 2_-/p=i ap + for j = 1; Aj — < as yU —1 < 2-/p=i " 1 1 1 1 PaP + Hlp~=j iaP + pap + ^lai, 1 1 1 ... 1 2 2 2 ... 2 2 3 3 ... 3 2 3 4 ... 4 for j = 2,3,..., I - for j = Z; 1 1 2 2 3 3 4 4 A”1 1 1 1 _ 2 2 3 4 ... (Z-2) 2 3 4 ... (Z-2) 1 | 2 ... |(Z —2) (Z - 2) (Z - 2) (Z - 1) (Z - 1) 2 G — f) 2^
324 GROUP THEORY IN PHYSICS Figure D.4: The Dynkin diagram of Di (Z = 3,4,5,...). Figure D.5: The Dynkin diagram of D2. (h) The compact real form of Ci is sp(Z). (i) The adjoint representation is Г({2,0,...,0}). 4 The simple complex Lie algebra I > 3 (and the semi-simple complex Lie algebra D2) (a) The Dynkin diagram for Di for I > 3 is given in Figure D.4 and the corresponding diagram for D2 is given in Figure D.5. (b) The Cartan matrix of Dt is 2 -1 0 .. 0 0 0 0 -1 2 -1 .. 0 0 0 0 0 -1 2 .. 0 0 0 0 A = 0 0 0 .. 2 -1 0 0 0 0 0 .. . -1 2 -1 -1 0 0 0 .. 0 -1 2 0 0 0 0 .. 0 -1 0 2
APPENDIX D 325 In particular, for Z = 3 and I = 4 respectively A = For Z = 2 2 -1 -1 -12 0 -10 2 and A = 2-100 -1 2 -1 -1 0-120 0-102 (c) (i) Di has Z(Z — 1) positive roots, namely (1) for J, k = 1,2,..., I - 2; j <k,(l> 3): ^2p=j aP +2 Y^lp=k ap + \~^k — l Z_^p=j Otpi (2) for j = 1,2,..., Z — 2; j < fc; (Z > 3): El — 2 . . > p=j ap + (ty-i + a/, 2^P=j aP + v—U—2 . ^p “h ^l ? 22р=7аР> J (3) a/_i and ai. (ii) It is sometimes convenient to introduce I auxiliary linear function- als 6i, 6i,..., 6i on H, with aj = 6j — 6j+i (for j = 1, 2,..., Z — 1) and ai = 6/-i +б/. The pattern of positive roots then appears more regular, as it consists of (cj ± 6&) for J, к = 1,2,..., Z with j < k). (d) The dimension n of Di is given by n = 1(21 — 1). (e) The quantities (aj, ak) (as defined in Equation (11.10)) are given for Di by i/{2(z -1)}, -1/{4(Z-1)}, (cVj, ak) for J = k,(j = 1,2, ...,Z); for j = к ± 1, (J, к = 1,2,..., I - 3); and for j = I — 2 with к = I — 1, Z; and for к = I — 2 with j = I — 1, Z; (all for I > 3) for all other J, A;, (j, к = 1,..., I). (f) The order of the Weyl group W for Di is 2l 1Z!.
326 GROUP THEORY IN PHYSICS (g) The fundamental weights of Di are Aj = < f 2 I 1 I 1 Z-^p=l aP “1“ + 2°^’ EpEi PaP + Ep=j jap + Уа1-1 + for j = 2,3,..., I - 2; |{Ep=iPaP + + (|г - l)ai}, k 2<Ep=iP«p "I" (2^ — 1)аг-1 + for j = 1; for j = I — 1; for j = I; as 1 1 1 1 ... 1 1 2 2 2 ... 2 1 2 3 3 ... 3 1 2 3 4 ... 4 A 1 = 1 1 2 2 1 1 3 3 2 2 2 2 1 2 3 4 ... (Z-2) J(Z —2) |(Z —2) 11!!... 1(1-2) 11 |(l-2) J 1 J 2 ... 1(1-2) |(Z —2) 11 (h) The compact real form of Di is so(2Z). (i) The adjoint representation for Di for I > 4 is Г({0,1,0,..., 0}) and for D3 isr({0,l,l}).
References Adams, J. F. (1969). “Lectures on Lie Groups”. W. A. Benjamin, New York. Ado, I. D. (1947). Uspeki Mat. Nauk (N.S.) 2, No. 6(22), 159-173. [Trans- lation (1962): Amer. Math. Soc. Trans. Series 1, 9, 308-327.] Agrawala, V. K. (1980). J. Math. Phys. 21, 1562-1565. Akhiezer, N. I. and Glazman, I. M. (1961). “Theory of Linear Operators in Hilbert Space”, Vol. I. Frederick Ungar, New York. Altmann, S. L. (1957). Proc. Camb. Phil. Soc. 63, 343-367. Altmann, S. L. (1962). Phil. Trans. A255, 216-240. Altmann, S. L. (1963). Rev. Mod. Phys. 35, 641-645. Altmann, S. L. and Cracknell, A. P. (1965). Rev. Mod. Phys. 37, 19-32. Baker, H. F. (1905). Proc. London Math. Soc. (2) 3, 24-47. Barut, A. O. and Raczka, R. (1977). “Theory of Group Representations and Applications”. Polish Scientific Publishers, Warsaw. Bateman, H. (1932). “Partial Differential Equations of Mathematical Phy- sics”. Cambridge U.P., Cambridge, England. Behrends, R. E. (1968). In “Group Theory and its Applications” (Ed. E. M. Loebl), Vol. I, 541-627. Academic Press, Orlando, New York and London. Behrends, R. E., Dreitlein, J., Fronsdal, C. and Lee, W. (1962). Rev. Mod. Phys. 34, 1-40. Bell, D. G. (1954). Rev. Mod. Phys. 26, 311-320. Berestetskii, V. B. (1965). Soviet Phys. Uspekhi 8, 147-176. Bertrand, J. (1966). Ann. Inst. Henri Poincare, 5, 235-256. Biedenharn, L. C. and Louck, J. D. (1979a). “Angular Momentum in Quan- tum Mechanics”. Addison-Wesley, Reading, Massachusetts. Biedenharn, L. C. and Louck, J. D. (1979b). “The Racah-Wigner Algebra in Quantum Theory”. Addison-Wesley, Reading, Massachusetts. Bloch, F. (1928). Z. Phys. 52, 555-600. Born, M. and von Karman, T. (1912). Phys. Z. 13, 297-309. Bouckaert, L. P., Smoluchowski, R. and Wigner, E. P. (1936). Phys. Rev. 50, 58-67. Brennich, R. H. (1970). Ann. Inst. Henri Poincare 13, 137-161.
328 GROUP THEORY IN PHYSICS Burns, G. and Glazer, A. M. (1978). “Introduction to Space Groups for Solid State Scientists”. Academic Press, Orlando, New York and London. Butler, P. H. (1981). “Point Group Symmetry Applications, Methods and Tables”. Plenum, New York. Callaway, J. (1958). “Electron Energy Bands in Solids”. In “Solid State Physics, Advances in Research and Applications”, Vol. 7 (Eds. F. Seitz and D. Turnbull), 100-212. Academic Press, Orlando, New York and London. Callaway, J. (1964). “Energy Band Theory”. Academic Press, Orlando, New York and London. Campbell, J. E. (1897a). Proc. London Math. Soc. (1) 28, 381-390. Campbell, J. E. (1897b). Proc. London Math. Soc. (1) 29, 14-32. Carruthers, P. (1966). “Unitary Symmetry of Strong Interactions”. Wiley Interscience, New York. Cartan, E. (1894). “Sur la Structure des Groupes de Transformations Finis et Continus”. Thesis, Paris. [Reprint (1952): In “Oeuvres Completes”, Partie I, Vol. I, 137-287. Vuibert, Paris.] Cartan, E. (1914). Ann. I’Ecole Norm. Sup. 3-eme serie 31, 263-355. Cartan, E. (1929). J. Math. Pure Appl. 8, 1-33. Casimir, H. (1931). Proc. R. Acad. Amstd. 34, 844-846. Charap, J. M., Jones, R. B. and Williams, P. G. (1967). Repts. Progr. Phys. 30, 227-283. Chelikowsky, J. R. and Cohen, M. L. (1976). Phys. Rev. B14, 556-582. Chevalley, C. (1946). “Theory of Lie Groups”, Vol. 1. Princeton U.P., Princeton. Cohen, E. R. (1974). “Tables of the Clebsch-Gordan Coefficients”. North American Rockwell Science Center, Thousand Oaks, California. Coleman, A. J. (1968). “Induced and Subduced Representations”. In “Group Theory and its Applications” (Ed. E. M. Loebl), 57-118. Aca- demic Press, Orlando, New York and London. Condon, E. U. and Odabasi, H. (1980). “Atomic Structure”. Cambridge U.P. Cambridge, England. Condon, E. U. and Shortley, G H. (1935). “The Theory of Atomic Spectra”. Cambridge U.P., Cambridge, England. Cornwell, J. F. (1969). “Group Theory and Electronic Energy Bands in Solids”. North Holland, Amsterdam. Cornwell, J. F. (1971). Phys. Status Solidi 43(b), 763-767. Cornwell, J. F. (1972). Phys. Status Solidi 52(b), 275-283. Cornwell, J. F. (1984). “Group Theory in Physics”. Vols. I and II. Academic Press, Orlando, New York and London. de Franceschi, G. and Maiani, L. (1965). Fortschr. Phys. 13, 279-384. de Swart, J. J. (1963). Rev. Mod. Phys. 35, 916-939. de Swart, J. J. (1965). Rev. Mod. Phys. 37, 326.
REFERENCES 329 Donnay, J. D. H. and Nowacki, W. (1954). “Crystal Data”. Geological Society of America, Memoir 60, New York. Dynkin, E. B. (1947). Uspeki. Mat. Nauk. (N.S.) 2, No. 4(20), 59-127 [Translation (1950): Amer. Math. Soc. Trans, series 1, 9, 328-469.] Dynkin, E. B. and Oniscik, A. L. (1955). Uspeki. Mat. Nauk. (N.S.) 10, No. 4(66), 3-74. [Translation (1962): Amer. Math. Soc. Trans. Series 2, 21, 119-192.] Dyson, F. (1966). “Symmetry Groups in Nuclear and Particle Physics”. W. A. Benjamin, New York. Edmonds, A. R. (1957). “Angular Momentum in Quantum Mechanics”. Princeton U.P., Princeton. Emmerson, J. McL. (1972). “Symmetries in Particle Physics”. Oxford U.P., Oxford. Englefield, M. J. and King, R. C. (1980). J. Phys. A13, 2297-2317. Fleming, W. (1977). “Functions of Several Variables”, 2nd edition. Springer- Verlag, New York, Heidelberg and Berlin. Fletcher, G. C. (1971). “The Electron Band Theory of Solids”. North Hol- land, Amsterdam. Freudenthal, H. and de Vries, H. (1969). “Linear Lie Groups”. Academic Press, Orlando, New York and London. Gantmacher, F. R. (1939a). Rec. Math. (Mat. Sbornik) N.S. 5(47), 101- 144. Gantmacher, F. R. (1939b). Rec. Math. (Mat. Sbornik) N.S. 5(47), 217- 250. Gantmacher, F. R. (1959). “The Theory of Matrices”, Vol. 1. Chelsea Publishing, New York. Gatto, R. (1964). Nuovo Cimento Suppl. 4, 414-464. Gel’fand, I. M. Minlos, R. A. and Shapiro, Z. Ya. (1963). “Representations of the Rotation and Lorentz Groups and their Applications”. Pergamon Press, London. Gell-Mann, M. (1953). Phys. Rev. 92, 833-834. Gell-Mann, M. (1961). California Institute of Technology Report CTSL-20, unpublished. Gell-Mann, M. (1962). Phys. Rev. 125, 1067-1084. Gell-Mann, M. (1964). Phys. Letts. 8, 214-215. Gell-Mann, M. and Ne’eman, Y. (1964). “The Eightfold Way”. W. A. Benjamin, New York. Gell-Mann, M. and Pais, A. (1955). In “Proceedings of the 1954 Glasgow Conference on Nuclear and Meson Physics” (Eds. E. H. Bellamy and R. G. Moorhouse), 342-352. Pergamon Press, London and New York. Glashow, S. L. (1980). Rev. Mod. Phys. 52, 539-543. Goldstone, J. (1961). Nuovo Cimento 19, 154-164. Goldstone, J., Salam, A. and Weinberg, S. (1962). Phys. Rev. 127, 965-970.
330 GROUP THEORY IN PHYSICS Goto, M. and Grosshans, F. D. (1978). “Semi-simple Lie Algebras”. Marcel Dekker, New York and Basel. Gourdin, M. (1967). “Unitary Symmetries”. North Holland, Amsterdam. Gruber, B. and O’Raifeartaigh, L. (1964). J. Math. Phys. 5, 1796-1804. Haar, A. (1933). Ann. Math. 34, 147-169. Halmos, P. R. (1950). “Measure Theory”. D. Van Nostrand, New York. Halpern, F. R. (1968). “Special Relativity and Quantum Mechanics”. Pren- tice-Hall, Englewood Cliffs, New Jersey. Hausdorff, F. (1906). Leipz. В er. 58, 19-48. Helgason, S. (1962). “Differential Geometry and Symmetric Spaces”. Aca- demic Press, Orlando, New York and London. Helgason, S. (1978). “Differential Geometry, Lie Groups, and Symmetric Spaces”. Academic Press, Orlando, New York and London. Henry, N. F. M. and Lonsdale, K. (1965). International Tables for X-ray Crystallography, Vol. I. International Union of Crystallography, Kynoch Press, Birmingham, England. Herring, C. (1937). Phys. Rev. 52, 365-373. Hewitt, E. and Ross, K. A. (1963). “Abstract Harmonic Analysis”, Vol. I. Springer-Verlag, Berlin. Higgs, P. W. (1964a). Phys. Letts. 12, 132-133. Higgs, P. W. (1964b). Phys. Rev. Letts. 13, 508-509. Higgs, P. W. (1966). Phys. Rev. 145, 1156-1163. Howarth, D. J. and Jones, H. (1952). Proc. Phys. Soc. A65, 355-368. Humphreys, J. E. (1972). “Introduction to Lie Algebras and Representation Theory”. Springer-Verlag, New York, Heidelberg and Berlin. Indnu, E. and Wigner, E. P. (1952). Nuovo Cimento. 9, 705-718. Jacobson, N. (1962). “Lie Algebras”. Interscience, New York. Killing, W. (1888). Math. Ann. 31, 252-290. Killing, W. (1889a). Math. Ann. 33, 1-48. Killing, W. (1889b). Math. Ann. 34, 57-122. Killing, W. (1890). Math. Ann. 36, 161-189. Klimyk, A. U. (1971). Teor. and Mat. Fiz. (USSR) 8, 55-60. Klimyk, A. U. (1972). Teor. and Mat. Fiz. (USSR) 13, 327-342. [Transla- tion (1972): Theor. and Math. Phys. 13, 1171-1182.] Klimyk, A. U. (1975). Repts. Math. Phys. 7, 153-166. Kokkedee, J. J. J. (1969). “The Quark Model”. W. A. Benjamin, New York. Koster, G. F., Dimmock, J. O., Wheeler, R. E. and Statz, H. (1964). “Prop- erties of the Thirty-two Point Groups”. M.I.T. Press, Cambridge, Mas- sachusetts. Koster, G. F. (1957). In “Solid State Physics, Advances in Research and Ap- plications” (Eds. F. Seitz and D. Turnbull), Vol. 5,173-256. Academic Press, Orlando, New York and London. Kudryavtseva, N. V. (1967). Fiz. Tverd. Tela 9, 2364-2368. [Translation (1968): Soviet Physics, Solid State 9, 1850-1853.]
REFERENCES 331 Lichtenberg, D. B. (1978). “Unitary Symmetry and Elementary Particles”. Academic Press, Orlando, New York and London. Lipsman, R. L. (1974). “Group Representations”, Lecture Notes in Mathe- matics, Vol. 388. Springer-Verlag, Berlin Lomont, J. S. (1959). “Applications of Finite Groups”. Academic Press, Orlando, New York and London. London, G. W. (1964). Fortschr. Phys. 12, 643-666. Loomis, L. H. (1953). “An Introduction to Abstract Harmonic Analysis”. D. Van Nostrand, Princeton. Luehrmann, A. W. (1968). Adv. Phys. 17,1-77. Mackey, G. W. (1963). Bull. Amer. Math. Soc. 69, 628-686. Mackey, G. W. (1968). “Induced Representation of Groups and Quantum Mechanics”. W. A. Benjamin, New York. Mackey, G. W. (1976). “The Theory of Unitary Group Representations”. University of Chicago Press, Chicago and London. Mathews, P. T. (1967). In ’’High Energy Physics” (Ed. E. H. Burhop), Academic Press, Orlando, New York and London. Miller, W. (1964). Commun. Pure Appl. Math. 17, 527-540. Moody, R.V., Patera, J. and Rand, D. (1996) “SimpLie™” software. Centre de Recherches Mathematiques, Montreal. Naimark, M. A. (1957). Amer. Math. Soc. Trans. Series 2, 6, 379-458. [Original (1954): Uspeki Mat. Nauk (N.S.) 9, No. 4(62), 19-93.] Naimark, M. A. (1964). “Linear Representations of the Lorentz Group”. Pergamon Press, London. Nakano, T. and Nishijima, K. (1953). Progr. Theor. Phys. 10, 581-582. Ne’eman, Y. (1961). Nucl. Phys. 26, 222-269. Ne’eman, Y. (1965). In “Progress in Elementary Particle and Cosmic Ray Physics” (Eds. J. G. Wilson and S. A. Wouthysen), 69-118. North > Holland, Amsterdam. Niederer, U. H. and O’Raifeartaigh, L. (1974a). Forts, der Physik 22, 111- 129. Niederer, U. H. and O’Raifeartaigh, L. (1974b). Forts, der Physik 22, 131- 157. Nishijima, K. (1954). Progr. Theor. Phys. 12, 107-108. Okubo, S. (1977). J. Math. Phys. 18, 2382-2394. O’Raifeartaigh, L. (1968). In “Group Theory and its Applications” (Ed. E. M. Loebl), Vol. I, 469-540. Academic Press, Orlando, New York and London. Pais, V. (1952). Phys. Rev. 86, 663-672. Peter, F. and Weyl, H. (1927). Math. Ann. 97, 737-755. Phillips, J. C. (1956). Phys. Rev. 104, 1263-1277. Pincherle, L. (1960). Rep. Progr. Phys. 23, 355-394. Pincherle, L. (1971). “Electronic Energy Bands in Solids”. McDonald, Lon- don.
332 GROUP THEORY IN PHYSICS Pontrjagin, L.S. (1946). “Topological Groups”. Princeton U.P., Princeton. Pontrjagin, L.S. (1986). “Topological Groups”. In “L.S.Pontrjagin, Selected Works” Vol. 2. Gordon and Breach, New York. Price, J. F. (1977). “Lie Groups and Compact Groups”. Cambridge U.P., Cambridge, England. Racah, G. (1942). Phys. Rev. 62, 438-462. Racah, G. (1950). Rend. LinceiS, 108-112. Racah, G. (1951). Ergebnisse der Exacten Naturwissenschaften, 37, 28-84. Rashba, E. I. (1959). Fiz. Tvend. Tela 1, 407-421. [Translation (1959): Soviet Phys. Solid State 1, 368-380.] Reitz, J. R. (1955). In “Solid State Physics, Advances in Research and Applications”, Vol. 1 (Eds. F. Seitz and D. Turnbull), 1-95. Academic Press, Orlando, New York and London. Riesz, F. and Sz.-Nagy, B. (1956). “Functional Analysis”. Blackie and Son, London and Glasgow. Rose, M. E. (1957). “Elementary Theory of Angular Momentum”. J. Wiley and Sons, New York. Rotman, J. J. (1965). “The Theory of Groups”. Allyn and Bacon, Boston. Sagle, A. A. and Walde, R. E. (1973). “Introduction to Lie Groups and Lie Algebras”. Academic Press, Orlando, New York and London. Salam, A. (1980). Rev. Mod. Phys. 52, 525-538. Samelson, H. (1969). “Notes on Lie Algebras”. Van Nostrand Reinhold, New York. Schiff, L. I. (1968). “Quantum Mechanics”. McGraw-Hill, New York and London. Schlosser, H. (1962). J. Phys. Chem. Solids 23, 963-969. Schonfliess, A. (1923). “Theorie der Kristallstruktur”. Bomtraeger, Berlin. Schwinger, J. (1952). “On Angular Momentum”. U.S. Atomic Energy Com- mission, NYO-3071. Shaw, R. (1955). “The Problem of Particle Types and Other Contributions to the Theory of Elementary Particles”. Ph.D. Thesis, Cambridge Uni- versity, unpublished. Sheka, V. I. (1960). Fiz. Tvend. Tela, 2, 1211-1219. [Translation (1960): Soviet Phys. Solid State 2, 1096-1104.] Shephard, G. C. (1966). “Vector Spaces of Finite Dimension”. Oliver and Boyd, Edinburgh and London. Shubnikov, A. V. and Koptsik, V. A. (1974) “Symmetry in Science and Art”. Plenum, New York. Simmons, G F (1963). “Topology and Modern Analysis”. McGraw-Hill, New York and London. Smorodinsky, Ya. A. (1965). Soviet Phys. Uspekhi 7, 637-655. Stein, E. M. (1965). In “High Energy Physics and Elementary Particles” (Ed. A. Salam), 563-584. International Atomic Energy, Vienna.
REFERENCES 333 van den Broek, P. M. and Cornwell, J. F. (1978). Phys. Stat. Sol. 90(b), 211-224. van Hove, L. (1953). Phys. Rev. 89, 1189-1193. Varadarajan, V. S. (1974). “Lie Groups, Lie Algebras, and their Represen- tations”. Prentice-Hall, Englewood Cliffs, New Jersey. Voisin, J. (1965a). J. Math. Phys. 6, 1519-1529. Voisin, J. (1965b). J. Math. Phys. 6, 1822-1832. Voisin, J. (1966). J. Math. Phys. 7, 2235-2237. von der Lage, F. C. and Bethe, H. A. (1947). Phys. Rev. 71, 612-622. Warner, F. W. (1971). “Foundations of Differentiable Manifolds and Lie Groups”. Scott, Foreman and Co. Glenview, Illinois and London. Weinberg, S. (1980). Rev. Mod. Phys. 52, 515-523. Weyl, H. (1925). Math. Zeits. 23, 271-309. Weyl, H. (1926a). Math. Zeits. 24, 328-376. Weyl, H. (1926b). Math. Zeits. 24, 377-395. Wigner, E. P. (1939). Ann. Math. 40, 149-204. Wigner, E. P. (1959). “Group Theory and its Application to the Quantum Mechanics of Atomic Spectra”. Academic Press, Orlando, New York and London. Wondratschek, H. and Neubuser, J. (1967). Acta. Cry st. 23, 349-352. Wood, J. H. (1962). Phys. Rev. 126, 517-527. Wybourne, B. G. (1970). “Symmetry Principles and Atomic Spectroscopy”. J. Wiley and Sons, New York. Wyckoff, R. W. G. (1963). “Crystal Structures 1”. Interscience, New York. Wyckoff, R. W. G. (1964). “Crystal Structures 2”. Interscience, New York. Wyckoff, R. W. G. (1965). “Crystal Structures 3”. Interscience, New York. Yang, C. N. and Mills, R. L. (1954). Phys. Rev. 96, 191-195. Zweig, G. (1964). CERN Repts. TH 401 and TH 402.

Index Ai, Cartan matrix, 220 from Dynkin diagram, 222 Casimir operator, 253 Clebsch-Gordan coefficients, 75, 175, 186-188 Clebsch-Gordan series, 175, 186- 188 complexification of su(2), as, 177- 178, 198-199, 204-205, 222 explicit forms of basis elements, roots, structure constants, and Cartan subalgebra, 204- 205, 210-211 isomorphism with Bi and Ci, 221, 232, 234 positive simple root, 219 representations, adjoint, 320 irreducible, 177-183, 244 Weyl’s dimensionality formula, 244 weights, fundamental, 243 Aa, Ai subalgebras, 211-212 Cartan matrix, 220 Casimir operators, 220 Clebsch-Gordan coefficients, 251 Clebsch-Gordan series, 251 complexification of su(3), as, 205, 222 Dynkin diagram, 223 explicit form of basis elements, roots, structure constants, and Cartan subalgebra, 205- 207, 211-212, 215-216, 217 ortho-normal basis of Cartan sub- algebra, 217 positive roots, 218 real forms, 232, 233 representations,205-207, 211-212, 235 adjoint, 247 irreducible, 245-251, 262-265 Weyl’s dimensionality formula, 244-245 roots, simple, 219 strings of, 212 weights, 237, 245-251, 262-265 fundamental, 243, 245 Weyl group, 226-227, 245-246 Аз, isomorphism with Вз, 221, 233 real forms, 233 Л (/>1), Cartan matrix, 319 dimension, 320 Dynkin diagram, 222, 319 properties, summary of, 319-320 real forms, 232, 233, 320 representations, adjoint, 320 explicit, 320 roots, 319-320 su(7 + 1) as compact real form, 232, 320 weights, fundamental 320 Weyl group, 320 Abelian group, classes, 22 definition, 3-4 finite, as direct product of cyclic groups, 85-86 irreducible representations, 57, 85-86, 109 irreducible representations, 57
336 GROUP THEORY IN PHYSICS Lie, as being neither simple nor semi- simple, 194 see for particular cases: addi- tive group of real numbers, multiplicative group of real numbers, S0(2) and U(l) Abelian Lie algebra, as being neither simple nor semi- simple, 194 definition, 145 irreducible representations, 162 Accidental degeneracies - see Energy eigenvalues, degeneracies Additive group of real numbers, 2 Adjoint, of a matrix, 272 of an operator, 290 Ado’s theorem, 142 Allowed k-vectors, 110 Analytic curve, 146 Analytic function, 37 Angular momentum, operators as irreducible tensor op- erators of 0(3), 186 representation theory of SU(2) and SO(3), relationship to, 143- 144, 175, 178-179, 252-253 Anti-Hermitian matrices, definition, 273 forming a real vector space, 282 Antiparticles, 255, 256, 263, 266, 267. Associated production, 259 Associative law, 1 Atomic physics, including electron’s spin, 189 neglecting electron’s spin, 189-192 Automorphic mapping, of a group, 30 of a Lie algebra, 155, 170-171, 196, 200-201 Azimuthal quantum number, 189 Bi, isomorphism with Ai and Ci, 221, 232, 234 real forms, 232, 234 B2, Cartan matrix, 321 isomorphism with C2, 221, 232, 234 real forms, 232, 234 Вз, Cartan Matrix, 321 Cartan matrix, 320 dimension, 321 Dynkin matrix, 222, 320-321 properties, summary of, 320-322 real forms, 221, 232, 234, 322 representations, adjoint, 322 roots, 321 so(2Z + 1) as compact real form, 322 weights, fundamental, 321 Weyl group, 321 Baryon, 255-256 Baryon number, 256, 258-260, 267 Basis functions of a representation - see Representations of a group, basis functions Basis of an inner product space, 285- 286 Basis of a vector space, 281 Bilinear forms, 292-294 Bloch functions, 110 Bloch’s theorem, 107, 109-111 Born cyclic boundary conditions, 103, 107-109, 120 Bravais lattices, 103-107 Brillouin zones, definition and general properties, 111-115 general points, 122 symmetry axes, 122 symmetry planes, 122 symmetry points, 122 Ci (complex Lie algebra), isomorphism with Ai and Bi, 221, 232, 234 real forms, 232, 234 Ci (crystallographic point group), 318 Cifi — see Cs C2 (complex Lie algebra), Cartan matrix, 322 isomorphism with B2, 221, 233
INDEX 337 real forms, 232, 234 C2 (crystallographic point group), 318 C2h, 316 C2v, 316 C3 (complex Lie algebra), Cartan matrix, 322 C3 (crystallographic point group), 317 C3 (complex 3-dimensional space), 281 C3h, 314 C3i, 315 C3v, 314-315 C4 (crystallographic point group), 316- 317 C4h, 313-314 C4v, 312-313 Cq (crystallographic point group), 316 C6h, 310-311 C6v, 310 C3 (complex d-dimensional space), direct product space, as, 297 inner product space, as, 284-285 Ci, 318 С (z>i), Cartan matrix, 322 dimension, 323 Dynkin diagram, 222, 322 properties, summary of, 322-324 real forms, 221, 232, 234, 324 representations, adjoint, 324 roots, 323 sp(Z) as compact real form, 221, 324 weights, fundamental, 323 Weyl group, 323 Cs, 317 Campbell-Baker-Hausdorff formula, 137- 138 Cartan matrix, 220-223 Cartan subalgebra, 200-207 Cartan’s criterion, 196 Casimir operators, 251-254 Cauchy sequence, 286-287 Cayley-Hamilton theorem, 275 Character of a group element in a rep- resentation, 59-64 Character projection operators, 69-70 Characteristic equation of a matrix, 275 Characteristic polynomial of a matrix, 275 Charm, 268 Class of a group, 21-23, 59, 61, 62 Clebsch-Gordan coefficients, definition, 74, 80-81, 167-168 for a particular group of Lie al- gebra see appropriate group or Lie algebra Clebsch-Gordan series, determining selection rules, 99 general definition, 72 for a particular group of Lie al- gebra see appropriate group or Lie algebra Colour, 268 Commutation of group elements, 3-4 Commutator of matrices, 136 Commutator of a Lie algebra, 141- 142, 144 Compact Lie group - see Lie group, compact Compact set, 42-43 Compatibility relations, 132-134 Complete ortho-normal system, 287 Complex Lie algebra, 135, 144-145 definition, 144 dimension, 144, 154, 194 real forms, 199-200, 228-234, 320, 322, 324, 326 semi-simple, Ai (or su(2)) subalgebras, 211- 212 Cartan matrix, 220-223, 242, 320-326 classification, 220-223 Clebsch-Gordan coefficients, 235 Clebsch-Gordan series, 235 definition, 193-194 dimension, 320, 321, 323, 325 Killing form, 194-196, 202-204, 207-210, 214-217 notation, 200, 217 rank, 201 real forms, 228-234, 320, 322, 324, 326 representations, 193, 197, 200, 224, 235-254; adjoint, 197, 237, 253, 320, 322, 324, 326; complete reducibility, 235-
338 GROUP THEORY IN PHYSICS 236; complex conjugate, 240; irreducible, 241-245, 251-254; Weyl’s dimensionality formula, 243-245 roots: definition, 201-202, 216- 217; positive, 218, 319, 321, 323, 325; properties, 202-228, 237-239, 241-242; simple, 218- 223, 241-243; string, 212 root subspaces, 202 structure, 200-234 weights, 235-251; Freudenthal’s recursion formula, 244-245, 250; fundamental, 242-243, 245,320-323, 326; highest, 242- 254; multiplicity, 237, 239, 240, 242, 244-245; positive, 241; simple, 237; string, 239- 240 Weyl canonical form, 223-224 Weyl group, 224-228, 239, 245- 251, 320, 321, 323, 325 simple, classical, 221, 319-326 classification, 220-223 definition, 193-194 Dynkin diagram, 220-223, 319, 320, 322, 324 exceptional, 221-222 isomorphisms, 221, 232-234 representations: irreducible, 197, 235-236, 243-254 structure, 136, 193-194, weights, 235 structure constants, 144 see also Lie algebra Conjugacy class - see Class of a group Connected component - see Lie group, linear, connected components Coset of a group, 24-28, 41 Coset representative, 26 Critical points of electronic energy bands, 133-134 Crystal class, 118-119 Crystal lattices - see Bravais lattices Crystalline solids, translational sym- metry of, 103-117 Crystallographic point groups, character tables, 299-318 irreducible representations, 299- 318 specification, 118, 299-318 Crystallographic space groups, 118-121 invariance groups, as, 105, 125- 126 symmorphic, definition, 119 irreducible representations, 87, 91,121-134 semi-direct product groups, as, 34, 121 see for particular space groups O\, Oh and under O\, Oh and O9h Cubic lattice, body-centred, Brillouin zone, 113-114 lattice vectors, 104-106 reciprocal lattice vectors, 113 space group - see face-centred, Brillouin zone, 114-115 lattice vectors, 104, 106-107 reciprocal lattice vectors, 115 space group - see simple, Brillouin zone, 113-114 lattice vectors, 104, 106 reciprocal lattice vectors, 113 space group - see OAh Cyclic group, definition, 85 irreducible representations, 85 £>2 (complex Lie algebra), Al e Al, as, 221 Cartan matrix, 325 Dynkin diagram, 324 £>2 (crystallographic point group), 86, 316 D2d, 313 D2h, 312 D3 (complex Lie algebra), Cartan matrix, 325 isomorphism with A3, 221, 233 real forms, 233 D3 (crystallographic point group), 315 D3d, 309-310
INDEX 339 D^h, 309 Z>4 (complex Lie algebra), Cartan matrix, 325 LU (crystallographic point group), basis functions for representations, 17, 67-69 classes, 21-22 Clebsch-Gordan coefficients, 75- 76, 99-100 Clebsch-Gordan series, 72-73, 99- 100 cosets, 24, 28 definition, 7-9, 313 factor groups, 28 homomorphic mappings, 28-29 invariant subgroups, 24, 26 irreducible tensor operators, 76- 77, 99-100 optical selection rules, 99-100 representations, 16, 48, 61, 62- 63, 70, 72, 313 subgroups, 20 308 Dq. 311 Dqk. 301, 303 Di (Z >2), Cartan matrix, 324 dimension, 325 Dynkin diagram, 222, 324 properties, summary of, 324-326 real forms, 232-234, 326 roots, 325 so(2Z) as compact real form, 232, 326 weights, fundamental, 326 Weyl group, 325 Degeneracy of an eigenvalue, energy eigenvalue - see Energy eigenvalues, degeneracy general definition, 290 Degenerate symmetric bilinear form, 293-294 Diagonalizability of a matrix, 275-278 Dipole approximation, 98-100, 190-192 Direct product group, definition, 31-33 representations, 83-86 structure when constituents are finite or linear Lie groups, 84 Direct product, of matrices, 70, 274-275 of vector spaces, 79-80, 295-298 Direct sum, of vector spaces, 282 of Lie algebras, 171-173 Distance between vectors in an inner product space, 284 Domain of an operator, 288-289 Dual of a vector space, 294 Dynkin diagram, 221-223 Eq. as exceptional simple Lie alge- bra, 221 Dynkin diagram, 222 real forms, 233 ^7, as exceptional simple Lie alge- bra, 221 Dynkin diagram, 222 real forms, 233 Es, as exceptional simple Lie alge- bra, 221 Dynkin diagram, 222 real forms, 233 Eigenvalues, of Hamiltonian operator - see En- ergy eigenvalues of matrices, 275-278, 291 of operators, 290-291 Eigenvectors, of matrices, 275-278, 291 of operators, 290-291 Electric dipole transitions, 98-100, 190- 192 Electron spin, 11, 189 Electronic energy bands, 94, 107, 115- 118, 126-134 Elementary particles, baryons, 255-256 hadrons, 255-258 intermediate particles, 255, 268 internal symmetries, gauge theories, 268; spontaneous symmetry breaking, 268; uni- fied theory of weak and elec- tromagnetic interactions, 268
340 GROUP THEORY IN PHYSICS global theories, 255-268; SU(2) scheme, 255-268; SU(3) scheme, 259-268; symmetry breaking: intrinsic, 261, spontaneous, 268 leptons, 255 mesons, 255 Equivalent k-vectors, 112-113 Energy eigenvalues, calculation, 93-97, 100-102 definition, 10-11 degeneracies, 17-18, 96-97, 100- 102 Euclidean group of IR3, definition, 33-34 non-semi-simple group, as, 198 representations, 86, 87 semi-direct product group, as, 33- 34 as exceptional simple Lie alge- bra, 221 Dynkin diagram, 222 real forms, 233 Factor group, 26-28 Fermi energy, 117 Fermi surface, 117-118, 130 First homomorphism theorem, 29-31 Forbidden transition, 91 Freudenthal’s recursion formula, 244- 245, 250 gl(N, C), (N > 1), definition, 150 gl(N, IR), (N > 1), definition, 150 G2, as exceptional simple Lie alge- bra, 221 Dynkin diagram, 222 real forms, 233 GL(N,(D), (N > 1), definition, 150 GL(N,IR), (N > 1), definition, 150 Gauge theories - see Elementary par- ticles, internal symmetries, gauge theories Gell-Mann-Nishijima formula, 259-260 Gluons, 268 Group, Abelian - see Abelian group automorphic mapping of - see Au- tomorphic of a group axioms, 1-2 class of - see Class of a group coordinate transformations, of, in IR3, 4-10 coset of - see Coset of a group cyclic - see Cyclic group definition, 1-2 direct product - see Direct prod- uct group factor - see Factor group finite, 4 isomorphism - see Isomorphic map- ping of groups Lie - see Lie group multiplication table, 4 order of, 4 proper rotations in IR3, of all - see SO(3) Rearrangement theorem, 20-21 representations of - see Repre- sentations of a group rotations in IR3, of all - see 0(3) Schrodinger equation, of the, basis functions of representa- tions, relationship of to en- ergy eigenfunctions, 17-18, 51, 96-97 definition and introduction, 10- 11 matrix elements of Hamiltonian operator for basis functions, 79 perfect crystal, for a, 105 semi-direct product - see Semi- direct product group for a particular group or type of group see appropriate group or type of group Hadrons, definition, 255 tabulation, 258 Hamiltonian operator, eigenvalues - see Energy eigen- values invariance group, of - see Group, Schrodinger equation, of the irreducible tensor operators, as, 76, 79, 94-96 matrix elements, 79, 94-102
INDEX 341 quantum mechanics, role in, 10- 11 Heine-Borel theorem, 42 Hermitian adjoint, 272 Hermitian matrix, 272-273, 291 Hidden symmetries, 97 Hilbert space, 286-288, 295 L2, 14, 287-288, 298 Homomorphic mapping, of groups, general definition and proper- ties, 28-31 kernel, 30 Lie groups, case of, 155-160 of Lie algebras, 154-160 Hydrogen atom, degeneracies of energy eigenval- ues, 97 hidden symmetries, 97 Hypercharge, 259-267 I-spin, 211 Ideal - see Subalgebra of a Lie algebra, invariant Idempotent method, 67 Idempotent operator, 67 Inner product spaces, 282-286 Interactions, fundamental, 255 Internal symmetries for elementary par- ticles - see Elementary par- ticles, internal symmetries Invariance group, Hamiltonian opera- tor, of, for an electronic sys- tem - see Group, Schrodinger equation, of the Invariant integration, 44-46 Inversion, spatial, 6 Iron, electronic energy band structure, 116, 118, 128 Irreducible tensor operators, 76-79, 81- 83, 98-100, 168 Isomorphic mapping, of groups, general definition and proper- ties, 7, 29-31 Lie groups, case of, 156-160 of Lie algebras, 155-160 Isotopic multiplet, 257 Isotopic spin, 180, 256-267 Jacobi’s identity, 141 Jordan canonical form of a matrix, 227-228 к-space, 111 Kernel of a homomorphic mapping - see Homomorphic mapping of groups, kernel Killing form, 194-196 Kronecker product of matrices - see Direct product of matrices Kubic harmonics, 126 L2, 14, 287-288, 298 L.C.A.O. method, 94 Lattice points, 103 Lattice vectors, 103 Lebesgue integration, 288 Legendre functions, 185 Lepton number, 255 Leptons, 255 Lie algebra, Abelian - see Abelian Lie algebra abstract, 142 automorphic mapping - see Au- tomorphic mapping of Lie algebras automorphism groups, 155, 171 commutative - see Abelian Lie algebra complex - see Complex Lie alge- bra direct sums, representations, 172-173 structure, 171-173 homomorphic mapping - see Ho- momorphic mapping of Lie algebras isomorphic mapping - see Isomor- phic mapping of Lie alge- bras linear operators, of, 142-145 real - see Real Lie algebra representations - see Representa- tions of a Lie algebra Lie group, 4, 35-46 Abelian - see Abelian group, Lie compact, definition, 42-44 elements, expressibility of in terms of exponentiation of
342 GROUP THEORY IN PHYSICS Lie algebra elements, 148- 149, 151 invariant integration, 45-46 representations, 52, 56, 57-61, 63, 66-67, 72-75, 78, 165 semi-simple, 228-231 Wigner-Eckart theorem, 78, 82 linear, analytic homomorphism, 155- 160 canonical coordinates, 148 compact - see Lie group, com- pact connected, 41 connected components, 40-42 continuous homomorphism, 156 definition, 35-40 direct product group, 171-173 discrete subgroup, 157 invariant integration - see In- variant integration one-parameter subgroup - see Subgroup, one-parameter real Lie algebra, relationship to, 135-136, 145-151, 157, 171-173 representations: adjoint, 168- 170; analytic, 48, 162-165; continuous, 48, 162-165; re- lationship to representations of the corresponding real Lie algebra, 162-168 non-compact, expressibility of in terms of exponentiation of Lie algebra elements, 148- 149, 151 semi-simple, algebraic criterion for compact- ness, 43, 228-230 compact, 228-233 definition, 194 invariant integration, 46 representations: irreducible, 86, 193, 228-229; complete re- ducibility of, 235-236 Wigner-Eckart theorem, 78, 82 simple, definition, 194 irreducible representations, uni- tary in non-compact case, 52, 241 see also under Lie group, semi- simple unimodular, 46 universal covering group, 158 Lie product, 141, 144 Linear functionals, 294-295 Linear independence, 280 Linear Lie group - see Lie group, lin- ear Linear operator, 288-291 Little group, 88 Magnetic quantum number, 189 Matrices, definitions and properties, 271-278 Matrix exponential function, 136-139 Matrix representations, of a group - see Representations of a group of a Lie algebra - see Represen- tations of a Lie algebra of operators, 291 Maximal point group of a crystal lat- tice, 103-104, 118 Medium-strong interaction, 261 Meson, 255 Metric of an inner product space, 284 Minimal polynomial of a matrix, 277 Module - see Representations of a group, module, and Representations of a Lie algebra, module Multiplicative group of positive real numbers, connected components, 40-41 definition, 2 homomorphism with SO(2), 158 linear Lie group, as, 38 non-compactness, 43 representations, irreducible, 52- 53 Multiplicity, eigenvalue, of, 290 Neutron, 256, 258, 265 Norm of a vector, 282-283 Normal subgroup - see Subgroup, in- variant Normalised vectors, 285 Nucleon, 256, 258 (9, 307
INDEX 343 0(2), compactness, 43-44 connected components, 41 definition, 38-39 linear Lie group, as, 38-39 0(3), basis functions, 186 classes, 22-23 Clebsch-Gordan coefficients, 175, 188-189 Clebsch-Gordan series, 175, 188, 190-192 compactness, 43-44 direct product group, as, 33, 186 irreducible representations, 186, 190 irreducible tensor operators, 186, 190 properties, summary of, 175 rotations in IR3, relationship to group of all, 7, 175 Schrodinger equation, as group of the, for spherically sym- metric system, 12, 190 O(N), (N > 2), compactness, 43-44 definition, 3, 150 linear Lie group, as, 40 O(p, q) (p > 1,<7 > 1), 150 O(7V,(D), (TV > 2), 150 Oh, 310, 303-305 ok, irreducible representations, 126- 129, 133-134 structure, 119 Ol, irreducible representations, 126- 129, 133-134 label changing of irreducible rep- resentations due to change of origin, 134 structure, 119 symmetry points, axes, and planes, 122 irreducible representations, 126- 129, 133-134 structure, 119 symmetry points, axes, and planes, 122 symmetry properties of electronic energy bands, 130, 132 O.P.W. method, 94 Orbit, 88 Orbital angular momentum quantum number, 189 Orientation dependence of the sym- metry labelling of electronic states, 134 Origin dependence of the symmetry labelling of electronic states, 134 Orthogonal groups - see 0(2), 0(3), O(TV), SO(2), SO(3), SO(4), SO(6), SO(TV), O(TV, 0) and SO(TV, 0) Orthogonal matrix, 272-273 Orthogonality of vectors, 285 Ortho-normal set of vectors, 285-287 Parity, 186, 257, 261 Parceval’s relation, 287 Partitioning of matrices, 273-274 Pauli exclusion principle, 117, 268 Pauli spin matrices, 30-31, 159, 176 Perturbation theory, time-dependent, 97-100, 190-191 time-independent, 100-102, 191- 192 Photon, 255 Pions, 258, 266 Poincare groups, 87 Point group, allowed k-vector, of, 121 crystallographic - see Crystallo- graphic point group space group, of the, 118 Primitive translations - see Transla- tions in IR3, primitive Projection operators, 65-70, 95 Proton, 256-258, 265 Pseudo-orthogonal groups - see O(p, q), SO(p,g) Pseudo-unitary groups - see Ufj^q), Quantum chromodynamics, 268 Quarks, 265-268
344 GROUP THEORY IN PHYSICS Quasicrystal, 8, 118 Range of an operator, 289 Real Lie algebra, 4, 36, 38, 42, 135- 136, 140-151 complexification, 135-136, 144-145, 198-200, 228 definition, 141 dimension, 141, 154, 194 generators, 147 labelling convention, 147 Lie groups, relationship to, 135- 136, 140-151, 156-160, 171- 173 semi-simple, compact, 228-233 definition, 194 Killing form, 194-196, 229-230 non-compact, 229, 230, 233- 234 representations, 193, 197, 235- 236; adjoint, 197, 230; com- plete reducibility, 235-236 structure, 136, 193-200, 228- 234 universal linear group, 158 simple, compact, 228-233 definition, 193 isomorphisms, 232-234 non-compact, 228-230, 233-234 representations, 197, 235-236 structure, 136, 193-194, 199 universal linear group, 158 structure constants, 142 see also Lie algebra Reciprocal lattice vectors, 111, 120- 121 basic, 111 Reduced matrix elements, 78-79, 82 Representations of a group, 47-91 analytic, of a Lie group, 48, 162- 165 basis functions, definition, 16 energy eigenfunctions for the group of the Schrodinger equa- tion, relationship to, 17-18, 51, 94-97 expansion of arbitrary function, 65-67 ortho-normality, 53-54 basis vectors, 48 carrier space, definition, 48 invariant subspace, 55 characters, 59-64 character table, 62-63 completely reducible, 55-56 decomposable, 56 definition, 16, 29, 47 direct product representations, 70- 73 direct sum, 56 equivalent representations, 49-51 faithful, 29, 47 identity, 47-48 induced, 86-91, 121-129 infinite-dimensional, 49, 86 irreducible, 49, 55-58, 60-63 Kronecker product, 71 module, 48-49, 165-166 orthogonality theorems for char- acters, 60-62 orthogonality theorems for ma- trices, 57-58 reducible, 54-56 Schur’s lemmas, 57 tensor product, 71 unitary, 52-54 for a particular group or type of group see appropriate group or type of group Representations of a Lie algebra, adjoint, 168-170, 194-195, 197, 200, 230, 237 carrier space, 161 completely reducible, 162, 235 complexification, effect of, 200, 235 definition, 160-161 direct product representations, 166- 168 irreducible, 161-162 Kronecker product, 167 module, 161, 165-168, 236-237 properties, 160-171 reducible, 161-162 Schur’s lemmas, 162
INDEX 345 for a particular algebra or type of algebra see appropriate al- gebra or type of algebra Riemann integral, 288 Riesz representation theorem, 295 Rotations in IR3, 5-9 group of all - see 0(3) proper, 6 group of all - see S0(3) pure, 10 sl(2,C), complexification, 199 isomorphism with so(3,l), 234 sl(7V,(D), (TV > 2), 150, 233 sl(2,IR), Ai, as a real form of, 199 adjoint representation, 195 isomorphism with su(l,l), so(2,l), and sp(l,lR), 234 Killing form, 195 sl(4, IR), isomorphism with so(3,3), 234 sl(7V,IR), (TV > 2), Ajv-i, as a real form of, 233 definition, 150 so(2), isomorphism mapping onto Lie algebra of the multiplicative group of positive real num- bers, 158 isomorphism with u(l), 157 representations, 164 so(3), angular momentum, connection with quantum theory of, 142- 144, 177-183, 186-188, 251- 253 basis elements, 141-143 Casimir operator, 253 Clebsch-Gordan coefficients, 186- 189 Clebsch-Gordan series, 187 commutation relations, 142-143 complexification, 204-205 deduction from the group SO(3), 140-141 isomorphism with su(2) and sp(l), 159, 164-165, 175, 232 representations, 164-165, 177-183 structure constants, 143 so(5), isomorphism with sp(2), 233 so(6), isomorphism with su(4), 233 so(7V), (N > 3), compact real form of B(_/v-i)/2 or Dn/2') as, 233, 322, 326 definition, 147, 150 so* (6), isomorphism with su(3,l), 234 so* (8), isomorphism with so(6,2), 234 so* (IV), (N even), 150, 133 so(2,l), isomorphism with sl(4, IR), su(l,l) and sp(l,lR), 234 so(3,l), isomorphism with sl(2,C), 234 so(3,2), isomorphism with sp(2,lR), 234 so(3,3), isomorphism with sl(4,lR), 234 so(4,l), isomorphism with sp(l,l), 234 so(4,2), isomorphism with su(2,2), 234 so(5,l), isomorphism with su*(4), 234 so(6,2), isomorphism with so* (8), 234 so(p, g), (p > 1, q > 1), 150, 233 so(7V, (D), (N > 2), 150, 233 sp(l), isomorphism with so(3) and su(2), 232 sp(2), isomorphism with so(5), 233 sp(7V/2), (TV even), 150, 232, 324 sp(TV/2, C), (N even), 150, 233 sp(l,lR), isomorphism with so(2,l), su(l,l) and sl(2,lR), 234 sp(2,lR), isomorphism with so(3,2), 234 sp(TV/2,IR), (N even), 150, 233 sp(l,l), isomorphism with so(4,l), 234 sp(r, s), (r > 1, s > 1), 150, 233 su(2), Ai, as real form of, 204, 222 adjoint representation, 195 angular momentum, connection with quantum theory of, 142- 144, 177-183, 186-188, 251- 253 basis elements, 147 Casimir operator, 253 Clebsch-Gordan coefficients, 186- 189 Clebsch-Gordan series, 187 commutation relations, 147 compact real Lie algebra, as, 229 complexification, 198-199, 204-205 definition, 147-149 generators, 147
346 GROUP THEORY IN PHYSICS irreducible representations, 177- 183, 244 isomorphism with so(3) and sp(l), 159, 164-165, 175, 232 isotopic spin, relationship to, 256- 259 Killing form, 195 simple and semi-simple, as be- ing, 197 su(3), Aa, as real form of, 205, 222 Casimir operator, 253 Clebsch-Gordan coefficients, 251 Clebsch-Gordan series, 251 complexification, 205-207, 222, 232 Gell-Mann basis, 205-207, 232 ortho-normal basis, 232 irreducible representations, 244, 245-251, 253, 262-264 role in strong interaction physics, 262-264 semi-simple Lie algebra, as, 205 su(2) subalgebras, 211-212 su(4), isomorphism with so(6), 233 su(7V), (N > 2), Ajv-i, as compact form of, 232, 320 definition, 147, 149, 150 simple, as being, 197 structure, 282 su*(4), isomorphism with so(5,l), 234 su*(7V), (N even), 150, 233 su(l,l), isomorphism with sl(2,lR), so(2,l) and sp(l,lR), 234 su(2,2), isomorphism with so(4,2), 234 su(3,l), isomorphism with so* (6), 234 su(p,Q), (p > 1, q > 1), 150, 233 S2 - see Ci S-4, 317 Sq — see C^i SL(7V,C), (N > 2), 150, 233 SL(7V,IR), (N > 2), 150, 233 SO(2), analytic isomorphic mapping onto U(l), 157 compactness, 43 connected component, 41 definition, 38-39 homomorphic image of multiplica- tive group of positive real numbers, as, 158 irreducible representations, 191- 192 linear Lie group, as, 38-39 one-parameter subgroup of SO(3), as, 139 representations obtained by ex- ponentiation of those of so(2), 164 SO(3), angular momentum, connection with quantum theory of, 142- 144, 177-183, 186-188, 251- 253 basis functions of irreducible rep- resentations, 144, 184-185 characters, 177 classes, 176-177 Clebsch-Gordan coefficients, 75, 175, 186-189 Clebsch-Gordan series, 175, 186- 189 derivation of real Lie algebra so(3), 140-145, 177 elements expressed as matrix ex- ponential functions, 136-137, 139, 149 homomorphic image of SU(2), as, 30-31, 159, 233 irreducible representations, 144, 175, 183-185, 189 one-parameter subgroups, 139, 177 parametrizations, 176 proper rotations in IR3, relation- ship to group of all, 7, 140, 175 properties, summary of, 175 representations obtained by ex- ponentiation of those of so(3), 164-165, 175 simple Lie group, as, 197 SO(4), homomorphism with SO(3)®SO(3), 197 semi-simple but not simple Lie group, as a, 197 SO(6), as homomorphic image of SU(4), 233 SO(7V), (N > 2), compactness, 43-44
INDEX 347 connected linear Lie group, as, 42 definition, 3, 150 linear Lie group, as, 40 simple, (for N = 3 and N > 5), 197 SO*(IV), (N even), 150 SO(p,g), (p > 1, q > 1), 150 SO(W,(D), (N>2), 150 Sp(AT/2), (TV even), 150 Sp(TV/2,C), (TV even), 150 Sp(TV/2,lR), (TV even), 150 Sp(r, s), (r > 1, s > 1), 150 SU(2), angular momentum, connection with quantum theory of, 142- 144, 177-183, 186-188, 251- 253 basis functions of irreducible rep- resentations, 144, 184-185 characters, 177, 183-184 classes, 176-177 Clebsch-Gordan coefficients, 75, 175, 186-189 Clebsch-Gordan series, 175, 186- 189 compactness 44, 229 definition, 39-40 dimension, 149 derivation of real Lie algebra so(3), 140-145, 177 homomorphic mapping onto SO(3), 30-31, 159, 233 irreducible representations, 144, 175, 183-185, 189 Lie subgroup of SU(3), as, 160 linear Lie group, as, 39-40 parametrizations of whole group, 41-42, 176 symmetry scheme for hadrons, 256- 259 SU(3), Clebsch-Gordan coefficients, 251 Clebsch-Gordan series, 251 dimension, 149 irreducible representations, 244, 245-251, 253, 262-264 symmetry scheme for hadrons, “colour” model, 268 “flavour” model, 259-268 SU(4), homomorphic mapping onto SO (6), 233 symmetry scheme for hadrons, 268 SU(7V), (N > 2), compactness, 44 connected linear Lie group, as, 42 definition, 3, 150 dimension, 40, 149 linear Lie group, as, 40 simple Lie group, as, 197 SU*(7V), (N even), 150 SU(p,g), (p > 1, q > 1), 150 Scalar field, 12 Scalar transformation operator P(T), 12-15 Schmidt orthogonalization process, 285- 286 Schrodinger equation, group of - see Group, Schrodinger equation of the solution using group theoretical methods, 93-97 Schur’s lemmas, for groups, 57 for Lie algebras, 162 Schwarz inequality, 284 Secular equation, 96, 275 Selection rules, 97-100 for optical transitions in atoms, 190-191 Self-adjoint operator, 290-291 Semi-direct product group, definition, 33-34 representations, 87-91 Semi-simple Lie group - see Lie group, semi-simple Separable Hilbert space, 287-288 Set of measure zero, 288 Signature of a bilinear form, 293 Silicon, electronic energy band struc- ture, 117-118 Similarity transformation, 50, 161, 275 Simple Lie group - see Lie group, sim- ple Single-particle approximation, 10-11, 117 Special orthogonal groups - see SO(2), SO(3), SO(4), SO(6) and SO(7V) Special pseudo-orthogonal groups - see SO(3,1) and SO(p,Q)
348 GROUP THEORY IN PHYSICS Special unitary groups - see SU(2), SU(3), SU(4), SU(5) and SU(7V) Spherical harmonic, 185 Spontaneous symmetry breaking, 268 Star of k, 122, 130-131 Strangeness, 259 Strong interaction, 255, 259, 261 Subalgebra of a Lie algebra, Cartan - see Cartan subalgebra definition, 153-154 dimension, 154 invariant, 154 proper, 154 Subgroup, connected, 41 criterion for a subset of a group to be a subgroup, 19-20 definition, 19 invariant, definition and properties, 23- 24, 26-27, 41 relationship to invariant Lie sub- algebra, 154 Lie, compactness, 43 definition, 40 relationship to Lie subalgebra, 154 normal - see invariant one-parameter, 135, 139-140 proper, 19 Subspace of a vector space, 282 Symmetric bilinear form, 292-294 Symmetry points of Brillouin zone, ИЗ- 115, 122 Symmetry system of crystal lattices, 104 Symplectic groups, complex - see Sp(7V/2, C) pseudo-unitary - see Sp(r, s) real - see Sp(7V/2,lR) unitary - see Sp(7V/2) T, 311-312 Td, 303 Th, 307 Tensor operators, irreducible - see Ir- reducible tensor operators Tensor product of vector spaces - see Direct product of vector sp- aces Total quantum number, 189 Trace of a matrix, 273 Transformation operators, scalar - see Scalar transformation oper- ators Transition probabilities, general pre- diction, 97-100 Translation groups of a crystal lattice, 103, 107 irreducible representations, 109- 111 Translational symmetry of crystalline solids, 107-115 Translations in IR3, 9-10 primitive, 103 pure, 10 Triangle inequality, 284 u(l), irreducible representations, 260 isomorphism with so(2), 157 u(2), as direct sum of u(l) and su(2), 172 (N > 1), being isomorphic to u(l)®su(JV) (for N > 2), 172 definition, 147, 150 u(p,g), (p > 1, q > 1). 150 U(l), analytic isomorphic mapping onto SO(2), 157 parametrization, 40 U(2), shown not to be a direct prod- uct group, 172 UW, (N > 1), compactness, 44 connected linear Lie group, as, 42 definition, 3, 150 linear Lie group, as, 40 non-semi-simple group, as, 198 not isomorphic to U(1)®SU(7V), as being, 172 u(p,g), (p > 1, q > 1), 150 U-spin, 211-212 Unified gauge theories, weak and elec- tromagnetic interactions, 268 Unitary multiplet, 261 Unitary groups - see U(l), U(2), and U(7V)
INDEX 349 Unitary matrix, 272-273, 291 Unitary operator, definition, 289 properties, 289-291 Unitary symplectic groups - see Sp(7V/2) Universal covering group, 158 Universal linear group, 158 V - see D2 Vd - see D2d Vh - see D2h Vector spaces, 279-298 Weak interaction, 255, 268 Weak intermediate vector bosons, 255 Weight functions, 44-46 Weyl canonical form, 223-224 Weyl group, 224-228, 239, 245-251, 320, 321, 323, 325 Weyl reflection, 225 Weyl’s dimensionality formula, 243- 245 Wigner-Eckart theorem, 71, 73-83, 97- 102 Zeeman effect, 189
No longer do physicists regard group theory merely as providing a valuable tool for the elucidation of the symmetry aspects of physical problems. Recent developments, particularly in high-energy physics, have transformed its role so that it now occupies a crucial and central position Group Theory in Physics - An Introduction is an abridgement and revision of Volumes I and II of the author’s previous three volume work Group Theory in Physics. It has been designed to provide a succinct introduction to the subject for advanced undergraduate and postgraduate students, and for others approaching the subject for the first time. It aims to present all the relevant important mathematical developments in a form that is easy for physicists to understand and appreciate. The treatment starts with the basic concepts and is carried through to some of the most significant developments in atomic physics, electronic energy bands in solids and the theory of elementary particles. No prior knowledge of group theory is assumed, and for convenience, various relevant algebraic concepts are summarised in appendices. The intention has been to concentrate on introducing and describing in detail the most important basic ideas and the role that they play in physical problems. Nevertheless, the mathematical coverage goes outside the strict confines of group theory itself, and includes a study of Lie algebras, which, though related to Lie groups, are often developed by mathematicians as a separate subject. ACADEMIC PRESS HARCOURT BRACE & COMPANY, PUBLISHERS SAN DIEGO • LONDON BOSTON • NEW YORK SYDNEY • TOKYO • TORONTO PRINTED IN GREAT BRITAIN ISEN 0-12-189800-8 9 780121 898007