322

Pierre Ramond Group Theory a Physicists Survey 2010

Embed Size (px)

Citation preview

Page 2: Pierre Ramond Group Theory a Physicists Survey 2010

This page intentionally left blank

Page 3: Pierre Ramond Group Theory a Physicists Survey 2010

GROUP THEORY

A Physicist’s Survey

Group theory has long been an important computational tool for physicists, but,with the advent of the Standard Model, it has become a powerful conceptual toolas well. This book introduces physicists to many of the fascinating mathematicalaspects of group theory, and mathematicians to its physics applications.

Designed for advanced undergraduate and graduate students, this book gives acomprehensive overview of the main aspects of both finite and continuous grouptheory, with an emphasis on applications to fundamental physics. Finite groupsare extensively discussed, highlighting their irreducible representations and invari-ants. Lie algebras, and to a lesser extent Kac–Moody algebras, are treated in detail,including Dynkin diagrams. Special emphasis is given to their representationsand embeddings. The group theory underlying the Standard Model is discussed,along with its importance in model building. Applications of group theory to theclassification of elementary particles are treated in detail.

P I E R R E RA M O N D is Distinguished Professor of Physics at the University ofFlorida and Director of the Institute for Fundamental Theory. He has held posi-tions at FermiLab, Yale University and Caltech. He has made seminal contributionsto supersymmetry, superstring theory and the theory of neutrino masses, and is aMember of the American Academy of Arts & Sciences.

Page 4: Pierre Ramond Group Theory a Physicists Survey 2010
Page 5: Pierre Ramond Group Theory a Physicists Survey 2010

Group Theory

A Physicist’s Survey

PIERRE RAMONDInstitute for Fundamental Theory, Physics Department

University of Florida

Page 6: Pierre Ramond Group Theory a Physicists Survey 2010

CAMBRIDGE UNIVERSITY PRESSCambridge, New York, Melbourne, Madrid, Cape Town, Singapore,São Paulo, Delhi, Dubai, Tokyo

Cambridge University PressThe Edinburgh Building, Cambridge CB2 8RU, UK

First published in print format

ISBN-13 978-0-521-89603-0

ISBN-13 978-0-511-77474-4

© P. Ramond 2010

2010

Information on this title: www.cambridge.org/9780521896030

This publication is in copyright. Subject to statutory exception and to the provision of relevant collective licensing agreements, no reproduction of any partmay take place without the written permission of Cambridge University Press.

Cambridge University Press has no responsibility for the persistence or accuracy of urls for external or third-party internet websites referred to in this publication, and does not guarantee that any content on such websites is, or will remain, accurate or appropriate.

Published in the United States of America by Cambridge University Press, New York

www.cambridge.org

eBook (EBL)

Hardback

Page 7: Pierre Ramond Group Theory a Physicists Survey 2010

To Richard SlanskyWho would have written this book

Page 8: Pierre Ramond Group Theory a Physicists Survey 2010
Page 9: Pierre Ramond Group Theory a Physicists Survey 2010

Contents

1 Preface: the pursuit of symmetries page 12 Finite groups: an introduction 4

2.1 Group axioms 52.2 Finite groups of low order 62.3 Permutations 192.4 Basic concepts 22

2.4.1 Conjugation 222.4.2 Simple groups 252.4.3 Sylow’s criteria 272.4.4 Semi-direct product 282.4.5 Young Tableaux 31

3 Finite groups: representations 333.1 Introduction 333.2 Schur’s lemmas 353.3 The A4 character table 413.4 Kronecker products 443.5 Real and complex representations 463.6 Embeddings 483.7 Zn character table 523.8 Dn character table 533.9 Q2n character table 563.10 Some semi-direct products 583.11 Induced representations 613.12 Invariants 643.13 Coverings 67

4 Hilbert spaces 694.1 Finite Hilbert spaces 69

vii

Page 10: Pierre Ramond Group Theory a Physicists Survey 2010

viii Contents

4.2 Fermi oscillators 704.3 Infinite Hilbert spaces 72

5 SU (2) 785.1 Introduction 785.2 Some representations 825.3 From Lie algebras to Lie groups 865.4 SU (2) → SU (1, 1) 895.5 Selected SU (2) applications 93

5.5.1 The isotropic harmonic oscillator 935.5.2 The Bohr atom 955.5.3 Isotopic spin 99

6 SU (3) 1026.1 SU (3) algebra 1026.2 α-Basis 1066.3 ω-Basis 1076.4 α′-Basis 1086.5 The triplet representation 1106.6 The Chevalley basis 1126.7 SU (3) in physics 114

6.7.1 The isotropic harmonic oscillator redux 1146.7.2 The Elliott model 1156.7.3 The Sakata model 1176.7.4 The Eightfold Way 118

7 Classification of compact simple Lie algebras 1237.1 Classification 1247.2 Simple roots 1297.3 Rank-two algebras 1317.4 Dynkin diagrams 1347.5 Orthonormal bases 140

8 Lie algebras: representation theory 1438.1 Representation basics 1438.2 A3 fundamentals 1448.3 The Weyl group 1498.4 Orthogonal Lie algebras 1518.5 Spinor representations 153

8.5.1 SO(2n) spinors 1548.5.2 SO(2n + 1) spinors 1568.5.3 Clifford algebra construction 159

8.6 Casimir invariants and Dynkin indices 1648.7 Embeddings 168

Page 11: Pierre Ramond Group Theory a Physicists Survey 2010

Contents ix

8.8 Oscillator representations 1788.9 Verma modules 180

8.9.1 Weyl dimension formula 1878.9.2 Verma basis 188

9 Finite groups: the road to simplicity 1909.1 Matrices over Galois fields 192

9.1.1 PSL2(7) 1979.1.2 A doubly transitive group 198

9.2 Chevalley groups 2019.3 A fleeting glimpse at the sporadic groups 205

10 Beyond Lie algebras 20810.1 Serre presentation 20810.2 Affine Kac–Moody algebras 21010.3 Super algebras 216

11 The groups of the Standard Model 22111.1 Space-time symmetries 222

11.1.1 The Lorentz and Poincaré groups 22311.1.2 The conformal group 231

11.2 Beyond space-time symmetries 23511.2.1 Color and the quark model 239

11.3 Invariant Lagrangians 24011.4 Non-Abelian gauge theories 24311.5 The Standard Model 24411.6 Grand Unification 24611.7 Possible family symmetries 249

11.7.1 Finite SU (2) and SO(3) subgroups 24911.7.2 Finite SU (3) subgroups 252

12 Exceptional structures 25412.1 Hurwitz algebras 25412.2 Matrices over Hurwitz algebras 25712.3 The Magic Square 259Appendix 1 Properties of some finite groups 265Appendix 2 Properties of selected Lie algebras 277References 307Index 308

Page 12: Pierre Ramond Group Theory a Physicists Survey 2010
Page 13: Pierre Ramond Group Theory a Physicists Survey 2010

1

Preface: the pursuit of symmetries

Symmetric objects are so singular in the natural world that our ancestors must havenoticed them very early. Indeed, symmetrical structures were given special magicalstatus. The Greeks’ obsession with geometrical shapes led them to the enumerationof platonic solids, and to adorn their edifices with various symmetrical patterns.In the ancient world, symmetry was synonymous with perfection. What could bebetter than a circle or a sphere? The Sun and the planets were supposed to circlethe Earth. It took a long time to get to the apparently less than perfect ellipses!

Of course most shapes in the natural world display little or no symmetry, butmany are almost symmetric. An orange is close to a perfect sphere; humans arealmost symmetric about their vertical axis, but not quite, and ancient man musthave been aware of this. Could this lack of exact symmetry have been viewed as asign of imperfection, imperfection that humans need to atone for?

It must have been clear that highly symmetric objects were special, but it is acurious fact that the mathematical structures which generate symmetrical patternswere not systematically studied until the nineteenth century. That is not to say thatsymmetry patterns were unknown or neglected, witness the Moors in Spain whodisplayed the seventeen different ways to tile a plane on the walls of their palaces!

Évariste Galois in his study of the roots of polynomials of degree larger than four,equated the problem to that of a set of substitutions which form that mathemati-cal structure we call a group. In physics, the study of crystals elicited wonderfullyregular patterns which were described in terms of their symmetries. In the twenti-eth century, with the advent of Quantum Mechanics, symmetries have assumed acentral role in the study of Nature.

The importance of symmetries is reinforced by the Standard Model of elemen-tary particle physics, which indicates that Nature displays more symmetries inthe small than in the large. In cosmological terms, this means that our Universeemerged from the Big Bang as a highly symmetrical structure, although most ofits symmetries are no longer evident today. Like an ancient piece of pottery, some

1

Page 14: Pierre Ramond Group Theory a Physicists Survey 2010

2 Preface: the pursuit of symmetries

of its parts may not have survived the eons, leaving us today with its shards. Thisis a very pleasing concept that resonates with the old Greek ideal of perfection.Did our Universe emerge at the Big Bang with perfect symmetry that was pro-gressively shattered by cosmological evolution, or was it born with internal defectsthat generated the breaking of its symmetries? It is a profound question which somephysicists try to answer today by using conceptual models of a perfectly symmetricuniverse, e.g. superstrings.

Some symmetries of the natural world are so commonplace, that they are diffi-cult to identify. The outcome of an experiment performed by undergraduates shouldnot depend on the time and location of the bench on which it was performed. Theirresults should be impervious to shifts in time and space, as consequences of timeand space translation invariances, respectively. But there are more subtle mani-festations of symmetries. The great Galileo Galilei made something of a “trivial”observation: when your ship glides on a smooth sea with a steady wind, you canclose your eyes and not “feel” that you are moving. Better yet, you can performexperiments whose outcomes are the same as if you were standing still! Today, youcan leave your glass of wine while on an airplane at cruising altitude without fearof spilling. The great genius that he was elevated this to his principle of relativity:the laws of physics do not depend on whether you are at rest or move with constantvelocity! However, if the velocity changes, you can feel it (a little turbulence willspill your wine). Our experience of the everyday world appears complicated by thefact that it is dominated by frictional forces; in a situation where their effect canbe neglected, simplicity and symmetries (in some sense analogous concepts) arerevealed.

According to Quantum Mechanics, physics takes place in Hilbert spaces. Bizarreas this notion might be, we have learned to live with it as it continues to beverified whenever experimentally tested. Surely, this abstract identification of aphysical system with a state vector in Hilbert space will eventually be found to beincomplete, but in a presently unimaginable way, which will involve some otherweird mathematical structure. That Nature uses the same mathematical structuresinvented by mathematicians is a profound mystery hinting at the way our brains arewired. Whatever the root cause, mathematical structures which find natural repre-sentations in Hilbert spaces have assumed enormous physical interest. Prominentamong them are groups which, subject to specific axioms, describe transformationsin these spaces.

Since physicists are mainly interested in how groups operate in Hilbert spaces,we will focus mostly on the study of their representations. Mathematical conceptswill be introduced as we go along in the form of scholia sprinkled throughout thetext. Our approach will be short on proofs, which can be found in many excellenttextbooks. From representations, we will focus on their products and show how

Page 15: Pierre Ramond Group Theory a Physicists Survey 2010

Preface: the pursuit of symmetries 3

to build group invariants for possible physical applications. We will also discussthe embeddings of the representations of a subgroup inside those of the group.Numerous tables will be included.

This book begins with the study of finite groups, which as the name indicates,have a finite number of symmetry operations. The smallest finite group has onlytwo elements, but there is no limit as to their number of elements: the permuta-tions on n letters form a finite group with n! elements. Finite groups have foundnumerous applications in physics, mostly in crystallography and in the behavior ofnew materials. In elementary particle physics, only small finite groups have foundapplications, but in a world with extra dimensions, and three mysterious familiesof elementary particles, this situation is bound to change. Notably, the sporadicgroups, an exceptional set of twenty-six finite groups, stand mostly as mathematicalcuriosities waiting for an application.

We then consider continuous symmetry transformations, such as rotations byarbitrary angles, or open-ended time translation, to name a few. Continuous trans-formations can be thought of as repeated applications of infinitesimal steps,stemming from generators. Typically these generators form algebraic structurescalled Lie algebras. Our approach will be to present the simplest continuous groupsand their associated Lie algebras, and build from them to the more complicatedcases. Lie algebras will be treated à la Dynkin, using both Dynkin notation anddiagrams. Special attention will be devoted to exceptional groups and their repre-sentations. In particular, the Magic Square will be discussed. We will link back tofinite groups, as most can be understood as subgroups of continuous groups.

Some non-compact symmetries are discussed, especially the representationsof space-time symmetries, such as the Poincaré and conformal groups. Group-theoretic aspects of the Standard Model and Grand Unification are presented aswell. The algebraic construction of the five exceptional Lie algebras is treated indetail. Two generalizations of Lie algebras are also discussed, super-Lie algebrasand their classification, and infinite-dimensional affine Kac–Moody algebras.

I would like to express my gratitude to the Institute for Advanced Study andthe Aspen Center for Physics for their hospitality, where a good part of this bookwas written. I would like to thank Professors L. Brink and J. Patera, as well asDrs. D. Belyaev, Sung-Soo Kim, and C. Luhn, for their critical reading of themanuscript, and many useful suggestions. Finally, I owe much to my wife Lillian,whose patience, encouragements, and understanding made this book possible.

Page 16: Pierre Ramond Group Theory a Physicists Survey 2010

2

Finite groups: an introduction

Symmetry operations can be discrete or continuous. Easiest to describe are the dis-crete symmetry transformations. In this chapter we lay out notation and introducebasic concepts in the context of finite groups of low order. The simplest symmetryoperation is reflection (parity):

P : x → x ′ = − x .

Doing it twice produces the identity transformation

I : x → x ′ = x,

symbolically

P P = I.

There are many manifestations of this symmetry operation. Consider the isoscelestriangle. It is left the same by a reflection about its vertical axis. In geometry thisoperation is often denoted as σ . It takes place in the x–y plane, and you will see itwritten as σz which denotes a reflection in the x–y plane.

The second simplest symmetry operation is rotation. In two dimensions, it isperformed about a point, but in three dimensions it is performed about an axis.A square is clearly left invariant by an anti-clockwise rotation about its center by90◦. The inverse operation is a clockwise rotation. Four such rotations are akin tothe identity operation. Generally, anti-clockwise rotations by (2π/n) generate thecyclic group Zn when n is an integer. Repeated application of n such rotations is theidentity operation. It is only in two dimensions that a reflection is a 180◦ rotation.

A third symmetry operation is inversion about a point, denoted by i . Geomet-rically, it involves a reflection about a plane together with a 180◦ rotation in theplane. Symbolically

i = σh Rot(180◦).

4

Page 17: Pierre Ramond Group Theory a Physicists Survey 2010

2.1 Group axioms 5

Reflections and inversions are sometimes referred to as improper rotations. Theyare rotations since they require a center of symmetry.

The symmetry operations which leave a given physical system or shape invariantmust satisfy a number of properties: (1) a symmetry operation followed by anothermust be itself a symmetry operation; (2) although the order in which the symmetryoperations are performed is important, the symmetry operations must associate;(3) there must be an identity transformation, which does nothing; and (4) whateveroperation transforms a shape into itself, must have an inverse operation. Theseintuitive considerations lead to the group axioms.

2.1 Group axioms

A group G is a collection of operators,

G : { a1, a2, . . . , ak, . . . }with a “�” operation with the following properties.

Closure. For every ordered pair of elements, ai and a j , there exists a uniqueelement

ai � a j = ak, (2.1)

for any three i, j, k.

Associativity. The � operation is associative

( ai � a j ) � ak = ai � ( a j � ak ). (2.2)

Unit element. The set G contains a unique element e such that

e � ai = ai � e = ai , (2.3)

for all i . In particular, this means that

e � e = e.

Inverse element. Corresponding to every element ai , there exists a unique elementof G, the inverse (ai )

−1 such that

ai � (ai )−1 = (ai )

−1 � ai = e. (2.4)

When G contains a finite number of elements

(G : a1, a2, . . . , ak, . . . , an )

it is called a finite group, and n is called the order of the group.

Page 18: Pierre Ramond Group Theory a Physicists Survey 2010

6 Finite groups: an introduction

In the following we will discuss groups with a finite number of elements, butwe should note that there are many examples of groups with an infinite number ofelements; we now name a few.

• The real numbers, including zero, constitute an infinite group under addition (� → +).Its elements are the zero, the positive and the negative real numbers. Closure is satisfied:if x and y are real numbers, so is their sum x + y. Each x has an inverse −x , such that

x + (−x) = 0, x + 0 = 0 + x,

and we see that the zero plays the role of the unit element.• The real numbers also form a group under multiplication (� → ×). Indeed xy is a real

number. The inverse of x is 1/x , and the unit element is 1. In this case, zero is excluded.• The rational numbers of the form n

m , where m and n are non-zero integers also form agroup under multiplication, as can easily be checked.

2.2 Finite groups of low order

We begin by discussing the finite groups of order less than thirteen (see Ledermann[14]). In the process we will introduce much notation, acquaint ourselves withmany useful mathematical concepts, and be introduced to several ubiquitous groupsand to the different aspects of their realizations.

Group of order 2

We have already encountered the unique group, called Z2, with two elements. Oneelement is the identity operation, e, and the second element a must be its owninverse, leading to the following multiplication table.

Z2 e a

e e aa a e

It can serve many functions, depending on the physical situation: a can be theparity operation, the reflection about an axis, a 180◦ rotation, etc.

Group of order 3

There is only one group of order 3. This is easy to see: one element is the identity e.Let a1 be its second element. The group element a1 � a1 must be the third elementof the group. Otherwise a1 � a1 = a1 leads to a1 = e, or the second possibilitya1�a1 = e closes the group to only two elements. Hence it must be that a1�a1 = a2,the third element, so that a1 � a2 = a2 � a1. It is Z3, the cyclic group of order three,defined by its multiplication table.

Page 19: Pierre Ramond Group Theory a Physicists Survey 2010

2.2 Finite groups of low order 7

Z3 e a1 a2

e e a1 a2a1 a1 a2 ea2 a2 e a1

It follows that

a1 � a1 � a1 = e,

which means that a1 is an element of order three, and represents a 120◦ rotation.It should be obvious that these generalize to arbitrary n, Zn , the cyclic group of

order n. It is generated by the repeated action of one element a of order n

Zn : { e , a , a � a , a � a � a , . . . , (a � a � · · · a � a)n−1 },with

(a � a � · · · a � a)k ≡ ak, an = e.

If we write its different elements as a j = a j−1, we deduce that ai � a j = a j � ai .A group for which any two of its elements commute with one another is calledAbelian. Groups which do not have this property are called non-Abelian.

Groups of order 4

It is equally easy to construct all possible groups with four elements{ e, a1, a2, a3 }. There are only two possibilities.

The first is our friend Z4, the cyclic group of order four, generated by 90◦ rota-tions. We note that the generator of this cyclic group has the amusing realization

a : z → z′ = i z, (2.5)

where z is a complex number. It is easy to see that a is an element of order fourthat generates Z4.

The second group of order four, is the dihedral group D2, with the followingmultiplication table.

D2 e a1 a2 a3

e e a1 a2 a3a1 a1 e a3 a2a2 a2 a3 e a1a3 a3 a2 a1 e

It is Abelian, since the multiplication table is symmetrical about its diagonal. Itis sometimes called V (Vierergruppe) or Klein’s four-group.

Page 20: Pierre Ramond Group Theory a Physicists Survey 2010

8 Finite groups: an introduction

It is the first of an infinite family of groups called the dihedral groups of order2n, Dn . They have the simple geometrical interpretation of mapping a plane poly-gon with n vertices into itself. The case n = 2 corresponds to the invariance groupof a line: a line is left invariant by two 180◦ rotations about its midpoint, one aboutthe axis perpendicular to the plane of the line, the other about the axis in the planeof the line. It is trivially invariant under a rotation about the line itself. The mul-tiplication table shows three elements of order two, corresponding to these three180◦ rotations about any three orthogonal axes in three dimensions. It has manyother realizations, for instance in terms of (2 × 2) matrices(

1 00 1

),

(1 00 −1

),

(−1 00 1

),

(−1 00 −1

). (2.6)

An interesting realization involves functional dependence. Consider the fourmappings (functions)

f1(x) = x, f2(x) = −x, f3(x) = 1

x, f4(x) = −1

x. (2.7)

It is easily verified that they close on D2. These two groups are distinct since Z4

has an element of order four and D2 does not.Groups of order four share one feature that we have not yet encountered: a subset

of their elements forms a group, in this case Z2. The subgroup within Z4 is Z2,generated by ( e, a2 ), expressed as

Z4 ⊃ Z 2.

As for D2, it contains three Z2 subgroups, generated by ( e, a1 ), ( e, a2 ), and( e, a3 ), respectively. Since a2 � a1 = a1 � a2, the elements of the first two Z2

commute with one another, and we can express D2 as the direct product of the firsttwo commuting subgroups

D2 = Z2 × Z2. (2.8)

This is the first and simplest example of a general mathematical construction.

Scholium. Direct product

Let G and K be two groups with elements {ga}, a = 1, . . . , ng and {ki },i = 1, . . . , nk , respectively. We assemble new elements ( ga, ki ), with multiplica-tion rule

( ga, ki ) ( gb, k j ) = ( ga � gb, ki � k j ). (2.9)

They clearly satisfy the group axioms, forming a group of order ngnk called thedirect (Kronecker) product group G×K. Since G and K operate in different spaces,

Page 21: Pierre Ramond Group Theory a Physicists Survey 2010

2.2 Finite groups of low order 9

they can always be taken to be commuting subgroups, and the elements of the directproduct can be written simply as gaki . This construction provides a simple way togenerate new groups of higher order.

Scholium. Lagrange’s theorem

This theorem addresses the conditions for the existence of subgroups. Consider agroup G with elements {ga}, a = 1, 2, . . . , N , that contains a subgroup H with nelements ( h1, h2, . . . , hn ) ≡ {hi }, i = 1, 2, . . . , n < N ,

G ⊃ H.

Pick an element g1 of G that is not in the subgroup H. The n elements of the formg1 � hi are of course elements of G, but not of the subgroup H. If any of them werein H, we would have for some i and j

g1 � hi = h j ,

but this would imply that

g1 = h j � (hi )−1

is an element of H, contradicting our hypothesis: the two sets {hi } and {g1 � hi }have no element in common. Now we repeat the procedure with another elementof G, g2, not in H, nor in g1H. The new set {g2 � hi } is distinct from both {hi } and{g1 � hi }, for if they overlap, we would have for some i and j

g1 � hi = g2 � h j .

This would in turn imply that g2 = g1 � hk, contradicting our hypothesis. Weproceed in this way until we run out of group elements after forming the last set{gk � hi }. Hence we can write the full G as a (right) coset decomposition

G = {hi } + {g1 � hi } + · · · + {gk � hi }≡ H + g1 �H + · · · + gk �H, (2.10)

in which none of the sets overlap: the order of G must therefore be a multiple ofthe order of its subgroup. Hence we have Lagrange’s theorem.

If a group G of order N has a subgroup H of order n, then N is necessarily an integermultiple of n

The integer ratio k = N/n is called the index of H in G. This theorem is about tosave us a lot of work.

Page 22: Pierre Ramond Group Theory a Physicists Survey 2010

10 Finite groups: an introduction

Let a be an element of a finite group G, and form the sequence

a, a2, a3, . . . , ak, . . . ,

all of which are elements of G. Since it is finite not all can be different, and wemust have for some k > l

ak = al → ak−l = e,

that is some power of any element of a finite group is equal to the identity element.When an = e, we say that a is an element of order n. Let k be the order of anyelement b of G; it generates the cyclic subgroup Zk of G. By Lagrange’s theorem,k must be a multiple of n, the order of the group G.

As a second application, let G be a group of prime order p. Since a prime hasno divisor, the order of any of its elements must be either one or p itself. Hence itmust be that G = Zp: we do not need to construct groups of order 5, 7, 11, . . .,they are all cyclic, and Lagrange’s theorem tells us that the cyclic groups of primeorder have no subgroup.

Groups of order 6

From what we have just learned, we know that there are at least two groups of ordersix: the cyclic group Z6, generated by 60◦ rotations, and the direct product groupZ2 × Z3, but they are the same. To see this, let a and b be the generators of Z3

and Z2, respectively, that is a3 = b2 = e, and ab = ba. Consider the element abof Z2 × Z3. Clearly, (ab)3 = b, so that (ab)6 = (b2) = (e), and ab is of order six.Hence both have an element of order six, and necessarily the two groups must beisomorphic to one another

Z6 = Z2 × Z3. (2.11)

This is true only because the two factors are relatively primes.Any other group of order six must contain an order-three element a (a3 = e). If

b is a different element, we find six elements ( e, a, a2, b, ab, a2 b ). It is easilyseen that all must be distinct: they must form a group of order six.

In particular the element b2 must be e, a or a2. The latter two choices implythat b is of order three, and lead to a contradiction. Hence b must be an element oforder two: b2 = e. Now we look at the element ba. It can be either ab or a2b. Ifba = ab, we find that ab must be of order six, a contradiction. Hence this group isnon-Abelian. By default it must be that ba = a2b; the multiplication table is nowfixed to yield the following dihedral group.

Page 23: Pierre Ramond Group Theory a Physicists Survey 2010

2.2 Finite groups of low order 11

D3 e a1 a2 a3 a4 a5

e e a1 a2 a3 a4 a5a1 a1 a2 e a4 a5 a3a2 a2 e a1 a5 a3 a4a3 a3 a5 a4 e a2 a1a4 a4 a3 a5 a1 e a2a5 a5 a4 a3 a2 a1 e

This is the last multiplication table you will see! There is a much better way topresent the information it contains. We note that all group elements can be obtainedfrom one order-three element a1 ≡ a, and one order-two element a3 ≡ b; a and bare the generators of the group.

Scholium. Presentation

A presentation of a finite group is a list of its generators, their order, and whateverfurther relations are needed to specify the group. The same group may have severalpresentations. For example, D3 has the simple presentation

< a, b | a3 = b2 = e ; bab−1 = a−1 > . (2.12)

With two generators, D3 is said to have rank two. D3 is also the symmetry groupof the equilateral triangle.

������

� �

�O

��

��

��

A B

C

Its element of order three describes the 60◦ rotation about O , the center of thetriangle. Its three elements of order two are the reflections about the three axesO A, O B, and OC , from each vertex to its center.

There is yet a better way to represent this group, since mapping the triangle intoitself amounts to permuting its three vertices: the group operations are permuta-tions on the three letters A, B, C which label its vertices. The reflection aboutthe vertical OC axis which interchanges the vertices A and B is the permuta-tion on two letters, called a transposition, and denoted by the symbol (A B). Theother two reflections are simply (B C) and (A C). The 120◦ rotations about O are

Page 24: Pierre Ramond Group Theory a Physicists Survey 2010

12 Finite groups: an introduction

permutations on three letters, denoted by the symbol (A B C), meaning the trans-formations A → B → C → A. We call transpositions two-cycles, and we callpermutations like (A B C) three-cycles. The six elements of D3 break up into theunit element, three transpositions, and two three-cycles. This cycle notation willprove ubiquitously useful. We note that these operations generate all 3! permu-tations on three objects. There are in general n! permutations on n objects; theyobviously form a group called the permutation group Sn . In this case we see thatD3 = S3.

We can also think of the three vertices as components of a vector on which thecycles act. The group operations are then represented by (3×3) matrices acting ona column vector with entries labeled A, B, and C . We make the identification

b = (A B) →⎛⎝0 1 0

1 0 00 0 1

⎞⎠ , a = (A B C) →⎛⎝0 1 0

0 0 11 0 0

⎞⎠ . (2.13)

Group action is simply matrix multiplication. For instance

ab = (A B C)(A B) = (B C) →⎛⎝0 1 0

0 0 11 0 0

⎞⎠⎛⎝0 1 01 0 00 0 1

⎞⎠ =⎛⎝1 0 0

0 0 10 1 0

⎞⎠ ,

(2.14)

and so on.This geometrical interpretation suggests a generalization to the dihedral group

Dn with 2n elements, as the symmetry group of n-sided plane polygons.

- The n-polygon is clearly left invariant by the n (2π/n) rotations about its center.- It is also left invariant by n mirror reflections. If n is odd as it is for the triangle, the

reflections are about the polygon’s n symmetry axes drawn from its center to each vertex.If n is even (as in the square shown below), the reflections are about the (n/2) symmetryaxes drawn from the center to the vertices, and (n/2) axes drawn from its center to themiddle of the opposite faces. Correspondingly, its presentation is

Dn : < a, b | an = b2 = e ; bab−1 = a−1 > . (2.15)

Groups of order 8

By now it should be obvious how to construct many groups with eight elements.First of all we have the cyclic group Z8, and the dihedral group D4, the symmetrygroup of the following rectangle.

Page 25: Pierre Ramond Group Theory a Physicists Survey 2010

2.2 Finite groups of low order 13

� �

O

� �

A B

CD

We can also form the direct products Z2 × Z2 × Z2 = D2 × Z2, and Z4 × Z2.Are all these groups different? Clearly, the direct product Z4×Z2 which contains

an element of order four, is different from Z2 × Z2 × Z2 which has only elementsof order two. Furthermore it does not contain any order eight element and cannotbe Z8.

Any other group of order eight must have at least one element of order four. Callit a, with (a)4 = e. Let b be a different element, which is necessarily of order twoor four. In this way we generate eight distinct elements

{ e, a, a2, a3, b, ab, a2b, a3b }

which must form a group. We consider two possibilities.The first is where b2 = e. There are three possibilities here: ba = a2b leads to a

contradiction; ba = ab yields the Abelian Z4 × Z2; and finally ba = a3b, and theelement ab is of order two, generating the dihedral group D4.

The second is where b4 = e. In that case we must have b2 = a2 which leadsalso to three possibilities: again, ba = a2b leads to a contradiction, while ba = abyields the Abelian group Z4 × Z2. The third possibility ba = a3b generates anew group called the quaternion group, Q. Rather than showing its cumbersomemultiplication table, it suffices to give its presentation in terms of two order-fourgenerators:

Q : < a, b | a4 = e, a2 = b2, bab−1 = a−1 > . (2.16)

Q is a rank two group.Why is it called the quaternion group? Quaternions are a generalization of real

and complex numbers which have the property that the norm of a product of tworeal or complex numbers is equal to the product of their norm, the Hurwitz property.The norm of a complex number z = x + i y is defined as

N (z) = √z z, (2.17)

Page 26: Pierre Ramond Group Theory a Physicists Survey 2010

14 Finite groups: an introduction

where z = x − i y. For any two complex numbers w and z, we have

N (z w) = N (z)N (w). (2.18)

This property is trivial for the real numbers. Quaternions are generalizations of realand complex numbers which satisfy the Hurwitz property. A quaternion is like acomplex number with three imaginary units

q = x0 + e1 x1 + e2 x2 + e3 x3, q = x0 − e1 x1 − e2 x2 − e3 x3. (2.19)

Defining

N (q) = √q q, (2.20)

it is easy to verify that, for any two quaternions,

N (q q ′) = N (q)N (q ′), (2.21)

as long as

(e1)2 = (e2)

2 = (e3)2 = −1, e1 e2 = −e2 e1 = e3, (2.22)

plus cyclic combinations, producing Hamilton’s famous quaternion algebra. It canbe realized in terms of the (2 × 2) Pauli (Felix Klein) matrices, as

e j = −i σ j , σ j σk = δ jk + i ε jklσl, (2.23)

with

σ1 =(

0 11 0

), σ2 =

(0 −ii 0

), σ3 =

(1 00 −1

). (2.24)

The generators of the quaternion group are the multiplication of quaternions by theimaginary units

a : q → q ′ = e1 q, b : q → q ′ = e2 q. (2.25)

The order-eight quaternion group is, as we shall soon see, the first of the infinitefamily of the finite dicyclic (binary dihedral) groups.

Groups of order 9

Rather than dragging the reader through a laborious algebraic procedure, we relyon a theorem that a group (to be proved later) whose order is the square of a primemust be Abelian. There are two different groups of order nine

Z9 ; Z3 × Z3. (2.26)

Page 27: Pierre Ramond Group Theory a Physicists Survey 2010

2.2 Finite groups of low order 15

Groups of order 10

There are no new types of groups at this order. One finds only one cyclic group

Z10 = Z5 × Z2,

since two and five are relative primes, and the dihedral group D5 which leaves thepentagon invariant.

Groups of order 12

First of all, we have the obvious possibilities

Z12; D6; Z6 × Z2; Z4 × Z3; D3 × Z2; D2 × Z3,

but not all are different: since four and three are relative primes, the isomorphism

Z12 = Z4 × Z3

obtains, and also

Z2 × Z6 = Z2 × Z2 × Z3 = D2 × Z3,

resulting in only two Abelian groups of order twelve, Z4 × Z3, and Z6 × Z2, andthe non-Abelian

D6 = D3 × Z2. (2.27)

There are two new non-Abelian groups at this order. The first is the tetrahedralgroup, T , the symmetry group of the regular tetrahedron.

���������

���

��

��

��

� ���

��

���

����������

O

A C

D

B

Each of its four faces is an equilateral triangle left invariant by two rotations aroundthe four axes O A, O B, OC , and O D, for a total of eight symmetry operations oforder three, as well as three order-two rotations about an axis linking oppositeedges, e.g. AD and BC . With the identity, this makes for a size-twelve group withelements of order two and three.

Page 28: Pierre Ramond Group Theory a Physicists Survey 2010

16 Finite groups: an introduction

An elegant way to describe this (and any other) finite group is to think of itsoperations as permutations of the four letters A, B, C, D which label the verticesof the tetrahedron. We have already seen this type of construction applied to theequilateral triangle.

The twelve elements T neatly break down as the product of two transpositions

(A B)(C D), (A C)(B D), (A D)(B C), (2.28)

and rotations about the axes from the center to each vertex, written as

(A B C), (A C B); (A B D), (A D B); (2.29)

(A C D), (A D C), (B C D), (B D C). (2.30)

Together with the unit element, these form the tetrahedral group, T . We easilyverify that it is generated by the two (4 × 4) matrices

a = (1 2)(3 4) →

⎛⎜⎜⎝0 1 0 01 0 0 00 0 0 10 0 1 0

⎞⎟⎟⎠ ; b = (1 2 3) →

⎛⎜⎜⎝0 1 0 00 0 1 01 0 0 00 0 0 1

⎞⎟⎟⎠ , (2.31)

which generate the four-dimensional matrix representation of T . We will seeshortly that this procedure generalizes since any group operation can be viewedas a permutation. For completeness, we can write its presentation as

T : < a, b | a2 = b3 = (ab)3 = e > . (2.32)

The tetrahedral group turns out to be isomorphic to a member of an infinite familyof groups, called the alternating groups An (T = A4), generated by the evenpermutations of n letters.

The second non-Abelian group of order twelve, , is generated by two elementsa and b with presentation

: < a, b | a6 = e; b2 = a3, bab−1 = a−1 > . (2.33)

The generators can be written as Pauli spin matrices, with

a = 1

2(σ0 + i

√3σ3 ) = 1

2

(1 + i

√3 0

0 1 − i√

3

); b = iσ1 =

(0 ii 0

). (2.34)

This group is part of an infinite family of dicyclic groups Q2n , with presentation

Q2n : < a, b | a2n = e, b2 = an, bab−1 = a−1 > . (2.35)

Our limited exploration of groups of low order has produced the low-lying mem-bers of five infinite families of finite groups: the cyclic family, Zn generated by one

Page 29: Pierre Ramond Group Theory a Physicists Survey 2010

2.2 Finite groups of low order 17

Table 2.1. Finite groups of lowest order

Order Group Isomorphism Type

2 Z2 Cyclic3 Z3 Cyclic4 Z4 Cyclic4 D2 Z2 × Z2 Dihedral, Klein’s four-group5 Z5 Cyclic6 Z6 Z3 × Z2 Cyclic6∗ D3 S3 Dihedral, permutation7 Z7 Cyclic8 Z8 Cyclic8 Z4 × Z28 Z2 × Z2 × Z2 D2 × Z28∗ D4 Dihedral8∗ Q Q4 Quaternion, dicyclic9 Z9 Cyclic9 Z3 × Z3

10 Z10 Z5 × Z2 Cyclic10∗ D5 Dihedral11 Z11 Cyclic12 Z12 Z4 × Z3 Cyclic12 Z6 × Z2 D2 × Z312∗ D6 D3 × Z2 Dihedral12∗ Q6 Dicyclic, binary dihedral12∗ T A4 Tetrahedral, alternating

element of order n, the dihedral groups Dn of order 2n, generated by one elementof order n and one of order 2. Dn is non-Abelian for n > 2. At order six, we foundthe permutation group on three objects, S3, a member of the infinite family Sn ofpermutations on n objects. At order eight we found the quaternion group Q4, ofthe dicyclic group family Q2n . At order twelve, we found another dicyclic groupQ6, as well as the tetrahedral group T . Later we will identify it as A4, a member ofthe infinite family An , the group of even permutations on n objects. At each orderwe encountered in addition groups that were direct products of groups of lesserorder.

We summarize our results in two tables. In Table 2.1, non-Abelian groups aremarked with a star. We show the names of the group as well as isomorphisms.Table 2.2 gives the total number of finite groups up to order 200, with the numberof Abelian groups indicated in brackets, following Lomont [15].

Table 2.2 is pretty amazing. We note, as expected, that there is only one groupof prime order. Another regularity is the existence of only two Abelian groups ofsquared prime order. What about the order equal to twice a prime? Then there aresome stupendous numbers, where there are large numbers on non-Abelian groups.

Page 30: Pierre Ramond Group Theory a Physicists Survey 2010

18 Finite groups: an introduction

Table 2.2. Number of groups of order up to 200

Order N [NAbel ] Order N [NAbel ] Order N [NAbel ] Order N [NAbel ] Order N [NAbel ]1 1 [1] 41 1 [1] 81 15 [5] 121 2 [2] 161 1 [1]2 1 [1] 42 6 [1] 82 2 [1] 122 2 [1] 162 55 [5]3 1 [1] 43 1 [1] 83 1 [1] 123 1 [1] 163 1 [1]4 2 [2] 44 4 [2] 84 15 [2] 124 4 [2] 164 5 [2]5 1 [1] 45 2 [2] 85 1 [1] 125 5 [3] 165 2 [1]6 2 [1] 46 2 [1] 86 2 [1] 126 16 [2] 166 2 [1]7 1 [1] 47 1 [1] 87 1 [1] 127 1 [1] 167 1 [1]8 5 [3] 48 52 [5] 88 12 [3] 128 2328 [15] 168 57 [3]9 2 [2] 49 2 [2] 89 1 [1] 129 2 [1] 169 2 [2]

10 2 [1] 50 2 [2] 90 10 [2] 130 4 [1] 170 4 [1]11 1 [1] 51 1 [1] 91 1 [1] 131 1 [1] 171 5 [2]12 5 [2] 52 5 [2] 92 4 [2] 132 10 [2] 172 4 [2]13 1 [1] 53 1 [1] 93 2 [1] 133 1 [1] 173 1 [1]14 2 [1] 54 15 [3] 94 2 [1] 134 2 [1] 174 4 [1]15 1 [1] 55 2 [1] 95 1 [1] 135 5 [3] 175 2 [2]16 14 [5] 56 13 [3] 96 230 [7] 136 15 [3] 176 42 [5]17 1 [1] 57 2 [1] 97 1 [1] 137 1 [1] 177 1 [1]18 5 [2] 58 2 [1] 98 5 [2] 138 4 [1] 178 2 [1]19 1 [1] 59 1 [1] 99 2 [2] 139 1 [1] 179 1 [1]20 5 [2] 60 13 [2] 100 16 [4] 140 11 [2] 180 37 [4]21 2 [1] 61 1 [1] 101 1 [1] 141 1 [1] 181 1 [1]22 2 [1] 62 2 [1] 102 4 [1] 142 2 [1] 182 4 [1]23 1 [1] 63 4 [2] 103 1 [1] 143 1 [1] 183 2 [1]24 15 [3] 64 267 [11] 104 14 [3] 144 197 [10] 184 12 [3]25 2 [2] 65 1 [1] 105 2 [1] 145 1 [1] 185 1 [1]26 2 [1] 66 4 [1] 106 2 [1] 146 2 [1] 186 6 [1]27 5 [3] 67 1 [1] 107 1 [1] 147 6 [2] 187 1 [1]28 4 [2] 68 5 [2] 108 45 [6] 148 5 [2] 188 4 [2]29 1 [1] 69 1 [1] 109 1 [1] 149 1 [1] 189 13 [3]30 4 [1] 70 4 [1] 110 6 [1] 150 13 [2] 190 4 [1]31 1 [1] 71 1 [1] 111 2 [1] 151 1 [1] 191 1 [1]32 51 [7] 72 50 [6] 112 43 [5] 152 12 [3] 192 1543 [11]33 1 [1] 73 1 [1] 113 1 [1] 153 2 [2] 193 1 [1]34 2 [1] 74 2 [1] 114 6 [1] 154 4 [1] 194 2 [1]35 1 [1] 75 3 [2] 115 1 [1] 155 2 [1] 195 2 [1]36 14 [4] 76 4 [2] 116 5 [2] 156 18 [2] 196 17 [4]37 1 [1] 77 1 [1] 117 4 [2] 157 1 [1] 197 1 [1]38 2 [1] 78 6 [1] 118 2 [1] 158 2 [1] 198 10 [2]39 2 [1] 79 1 [1] 119 1 [1] 159 1 [1] 199 1 [1]40 14 [3] 80 52 [5] 120 47 [3] 160 238 [7] 200 52 [6]

It gets worse; for example, there are 20 154 non-Abelian groups of order 384, butonly 11 of order 390.

To go further, more systematic methods need to be developed. Can groups bederived from basic building blocks? Are some groups more equal than others?Believe it or not, none of the groups we have constructed so far are viewed as

Page 31: Pierre Ramond Group Theory a Physicists Survey 2010

2.3 Permutations 19

“fundamental” by mathematicians: the smallest such fundamental non-Abeliangroup happens to have sixty elements, corresponding to the even permutations onfive objects!

2.3 Permutations

The systematic study of groups was started by Évariste Galois as permutations ofroots of polynomial equations (more on this later). Permutations do play a centralrole in the study of finite groups because any finite group of order n can be viewedas permutations over n letters. We have seen several examples of this. In this sectionwe develop the idea systematically.

A permutation on n identified objects, 1, 2, 3, . . . , k, . . . , n, shuffles them intoa different order, say, a1, a2, a3, . . . , ak, . . . , an . This transformation k → ak , k =1, 2, . . . , n, is represented by the symbol(

1 2 · · · na1 a2 · · · an

). (2.36)

It clearly satisfies all the group axioms, for instance(1 2 · · · na1 a2 · · · an

)(a1 a2 · · · an

b1 b2 · · · bn

)=

(1 2 · · · nb1 b2 · · · bn

), (2.37)

and so on. Every permutation has a unique inverse, and the identity permutationleaves the array invariant. Clearly, the n! permutations on n letters form a group,the symmetric group, denoted by Sn .

We use the more concise cycle notation for permutations, which we have alreadyencountered. A permutation that shuffles k < n objects into themselves, leaving therest untouched, is called a k-cycle. Using permutations on four objects as examples,we write (

1 2 3 42 3 4 1

)∼ (1 2 3 4). (2.38)

Reading from left to right it reads 1 → 2, 2 → 3, 3 → 4, 4 → 1. Anotherfour-cycle element is (

1 2 3 43 4 2 1

)∼ (1 3 2 4). (2.39)

On the other hand, we have(1 2 3 42 1 3 4

)∼ (1 2)(3)(4), (2.40)(

1 2 3 42 1 4 3

)∼ (1 2)(3 4), (2.41)

Page 32: Pierre Ramond Group Theory a Physicists Survey 2010

20 Finite groups: an introduction

as well as (1 2 3 43 1 2 4

)∼ (1 3 2)(4). (2.42)

One-cycles, included here for clarity, will hereon be omitted.The degree of a permutation is the number of objects on which it acts. Per-

mutations can be organized into several types called classes (see later for aformal definition). On these four objects the classes are the transpositions on twoobjects, (xx); the product of two transpositions, (xx)(xx), three-cycles, (xxx); andfour-cycles, (xxxx).

Any permutation can be decomposed in this manner. Consider an arbitrary per-mutation on n objects. Pick one object, say k, and follow what happens to thesequence k → a[1]

k → a[2]k → · · · → k as we go through the permutation. The

sequence terminates and k goes back into itself since the number of objects isfinite. One of two things can happen. Either we need to go over all n objects, andthe permutation is written as an n-cycle

(k a[1]k a[2]

k · · · a[n−1]k ), (2.43)

or it terminates earlier, after only r iterations. In that case, let m be one of (n − r)objects unaccounted for, and follows its fate through the permutation m → a[1]

m →a[2]

m → · · · → a[s−1]m → m, such that it returns to itself after s iterations. Again

one of two things can happen. If s = (n − r), all objects are accounted for and thepermutation is written as

(k a[1]k a[2]

k · · · a[r−1]k )(m a[1]

m a[2]m · · · a[n−r−1]

m ). (2.44)

If s < (n−r), we need to account for the remaining (n−r−s) objects. By repeatingthis procedure until we run out of objects, we arrive at the cycle decomposition

(k a[1]k a[2]

k · · · a[r−1]k )(m a[1]

m a[2]m · · · a[s−1]

m ) · · · (t a[1]t a[2]

t · · · a[u−1]t ), (2.45)

with r + s +· · ·+ u = n. This decomposition is unique because it does not dependon the element we started from, nor on the choices we made at each stage. We justproved the following.

Every permutation can be uniquely resolved into cycles which operate on mutuallyexclusive sets

Any permutation can also be decomposed as a (non-commutative) product oftranspositions

(a1 a2 · · · an) = (a1 a2)(a1 a3) · · · (a1 an). (2.46)

Page 33: Pierre Ramond Group Theory a Physicists Survey 2010

2.3 Permutations 21

If the number of transpositions in this decomposition is even (odd), the permutationis said to be even (odd). The product of two even permutations is an even permuta-tion, and even permutations form a subgroup of order n!/2, called the alternatinggroup An .

In our example, T = A4, the even permutations are the product of two transpo-sitions like (1 2)(3 4), and three-cycles like (1 2 3) = (1 2)(1 3), also the product oftwo transpositions. The elements of S4 not in the subgroup A4 are the odd permu-tations, that is those containing one transposition (xx), and one four-cycle (xxxx).These do not form a group since the product of two odd permutations is an evenpermutation.

Consider a permutation on n objects described by α1 one-cycles, α2 two-cycles, . . . , αk k-cycles, so that

n =k∑

j=1

jα j . (2.47)

We want to find out the number of permutations with this cycle structure. Thisis like finding the number of ways of putting n distinct objects into α1 one-dimensional boxes, α2 two-dimensional boxes, . . .. There will be two types ofredundancies: with α j boxes of the same type, we expect α j ! redundancies fromshuffling the boxes. In addition, there are j ways to put the objects in eachj-dimensional box, and since there are α j such boxes this will produce a fur-ther redundancy of jα j . It follows that the number of permutations with this cyclestructure is

n!∏kj jα j α j !

. (2.48)

For example, the twenty-four elements of S4 break up into the identity, six two-cycles, three double transpositions, eight three-cycles, and six four-cycles.

Let G be a group with elements ga , a = 1, 2, . . . , n. Pick an element gb, andform the n elements g1 gb, g2 gb, . . . , gn gb, and associate to gb the permutation

Pb =(

g1 g2 g3 · · · gn

g1gb g2gb g3gb · · · gngb

). (2.49)

This can be done for every element of G, to produce n distinct permutations. Inaddition, this construction leaves the multiplication table invariant,

gagb = gc → Pa Pb = Pc, (2.50)

and the {Pa} form a group representation of G. Since the Pa satisfy the groupaxioms, they form a subgroup of the permutations on n objects. Thus we haveCayley’s theorem.

Page 34: Pierre Ramond Group Theory a Physicists Survey 2010

22 Finite groups: an introduction

Every group of finite order n is isomorphic to a subgroup of the permutation group Sn .

The permutations Pa of degree n form a representation called the regular rep-resentation, where each Pa is a (n × n) matrix acting on the n objects arrangedas a column matrix. These are very large matrices. The regular representation ofthe order-eight quaternion group is formed by (8 × 8) matrices, yet we have con-structed a representation of the same group in terms of (2×2) matrices. Clearly theregular representation is overkill. Such a representation is said to be reducible. Wewill soon develop systematics to find the irreducible representations of the finitegroups, but first we need to introduce more concepts that will prove useful in thecharacterization of finite groups.

2.4 Basic concepts

In this section, we introduce many of the concepts and tools which have provencrucial for a systematic study of finite groups (see Carmichael’s [2], and Hall’s [9]books). In addition we provide scholia (marginal explanatory notes), of the typefound in ancient scientific books and texts.

2.4.1 Conjugation

There is a very important operation which has the property of preserving any mul-tiplication table. Let G be a group with elements ga . We define the conjugate ofany element ga with respect to any other group element g as

ga ≡ g ga g−1. (2.51)

Clearly, such a transformation maps the multiplication table into itself

ga gb = gc → ga gb = gc, (2.52)

since

g ga g−1 g gb g−1 = g(ga gb)g−1 = g gc g−1. (2.53)

This is an example of a mapping which leaves the multiplication table invariant,called a homomorphism. In addition, since it is one-to-one, it is a special homo-morphism called an automorphism. This particular mapping is called inner since itis generated by one of its elements. If two elements ga and gb commute, then ga isself-conjugate with respect to gb and vice versa.

Conjugacy has the important property that it does not change the cycle structureof a group element. To see this, consider a group element, written as a permutation

Page 35: Pierre Ramond Group Theory a Physicists Survey 2010

2.4 Basic concepts 23

P = (k l m p q) that cycles through five elements of a group of order n. To verifythat

g P g−1 = (g(k) g(l) g(m) g(p) g(q)), (2.54)

where g(k) is the action of group operation g on group element k, etc., it sufficesto note that since

g P g−1g(k) = g P(k) = g(l), (2.55)

conjugacy by g replaces k → l by g(k) → g(l). We conclude that conjugacy doesnot alter the cycle decomposition of the permutation: conjugate elements have thesame cycle structure.

Scholium. Classes

We can use conjugacy to organize the elements of any group into non-overlappingsets. Pick any element gb, and conjugate it with respect to every element of G

Cb : gb = gagbg−1a , ∀ ga ∈ G, (2.56)

thus generating a set of elements Cb, called a class. Now pick a second element gc

of G that is not in Cb, and form a new class

Cc : gc = gagcg−1a , ∀ ga ∈ G. (2.57)

The classes Cb and Cc have no elements in common, for if they did, it would implythat gc is in Cb, which contradicts our hypothesis. The procedure continues untilwe have exhausted all the group elements. We conclude that any finite group oforder n can be decomposed into k < n classes.

Since this construction does not depend on our choice of elements gb, gc, . . ., itis unique, and k, the number of classes is a characteristic of the group. Any Abeliangroup of order n has n classes, each containing one element. The identity elementis always in a class by itself, traditionally denoted as C1.

Elements of a given class have the same cycle structure, but permutations withthe same cycle structure may belong to different classes.

Scholium. Normal subgroup

Let H with elements hi be a subgroup of G

G ⊃ H.

Some subgroups have a very special property. If the conjugate of hi with respectto every element g of G remains in H, that is

g hi g−1 = h j ∈ H, ∀ g ∈ G, (2.58)

Page 36: Pierre Ramond Group Theory a Physicists Survey 2010

24 Finite groups: an introduction

the subgroup H is said to be a normal subgroup. In that case, mathematicians usethe symbol

H � G, (G � H). (2.59)

Most groups have normal subgroups; those without normal subgroups, calledsimple groups are very special, and form the building blocks for all the others.

Scholium. Quotient group

To understand what is so special about a normal subgroup, consider the homomor-phic mapping of the group G into a smaller group K,

G → K, ga → ka,

which preserves the multiplication table: gagb = gc implies that kakb = kc. Let Hbe the subset of elements of G that are mapped into the identity (the kernel of themapping)

H → 1.

The set hi ∈ H clearly is a subgroup of G, and furthermore, since gahi g−1a →

ka k−1a = 1, it is a normal subgroup.

Now take the argument the other way. Start from G and assume it has a normalsubgroup H. Consider any elements gahi , and gbh j of two different cosets gaHand gbH, and form their product

(gahi )(gbh j ) = gagbg−1b hi gbh j = gagb(g

−1b hi gb)h j = gagbhi h j , (2.60)

where the conjugate hi is an element of the normal subgroup H. Hence the productof cosets

gaH ⊗ gbH = gcH, (2.61)

satisfies the group property. To complete the group structure, the unit element isthe subgroup H itself, and if gagb = hi , the cosets gaH and gbH are inverses ofone another. The group axioms are satisfied (associativity follows easily), and wehave constructed a group with ng/nh elements. This group is called the quotient(factor) group,

G/H.

It is a homomorphic image of G, but not a subgroup of G.Suppose H is not the largest subgroup of G. We split its generators in the form

G : { ga , gα , hi }, (2.62)

Page 37: Pierre Ramond Group Theory a Physicists Survey 2010

2.4 Basic concepts 25

with

G ⊃ H′ � H, H′ : { gα , hi }, (2.63)

since H is a normal subgroup of H′. We can form two factor groups

G/H : { ga H , gα H , H(= 1) }, K ≡ H′/H : { gαH , H(= 1) }. (2.64)

Clearly K is a subgroup of G/H

K ⊂ G/H. (2.65)

Since H is a normal subgroup, conjugation satisfies

gα H = gα H, gα = gagαg−1a . (2.66)

Thus if gα is an element of H′ (∈ {gα}), K is a normal subgroup of the quotientgroup; otherwise it is not. But then, H′ must itself be a normal subgroup of G: G/Hhas a normal subgroup only if H is not maximal.

The quotient group G/H has no normal subgroup as long as H is the maximalnormal subgroup of G. This fact opens the way for a systematic decomposition ofall groups in terms of simple groups.

2.4.2 Simple groups

Finite groups with normal subgroups are dime-a-dozen, and not viewed as funda-mental by mathematicians. Galois, the founder of group theory, split groups intotwo types: simple groups which have no normal subgroup, and the rest.

This was a pretty important distinction to Galois, since he related the simplicityof A5, the alternating group on five objects, to the impossibility of finding a formulathat solves the quintic equation by radicals (using square roots, cube roots, . . .)!

So far we have encountered only one family of simple groups, the cyclic groupsof prime order Zp, which have no subgroups. One of the great triumphs of modernmathematics has been the comparatively recent classification of all simple groups.Typically simple groups are very large: for example, A5 contains 60 elements, andthe next largest (not counting the cyclic groups) has 168 elements! We shall returnto their classification once we have covered the theory of Lie algebras. For themoment we just list the types of simple groups.

• Cyclic groups of prime order, Zp.• Alternating groups on five or more objects, An , n ≥ 5.• Infinite families of groups of Lie type (to be defined later).• Twenty-six sporadic groups.

Page 38: Pierre Ramond Group Theory a Physicists Survey 2010

26 Finite groups: an introduction

All finite groups can be systematically constructed from simple groups, by meansof a procedure we now proceed to describe.

Scholium. Composition series

We can decompose any group in terms of its normal subgroups by seeking itslargest normal subgroup H1. If it has a normal subgroup, the group elements splitinto H1 and the quotient group G/H1. We repeat the procedure applied to H1: weseek its maximal normal subgroup H2, and splits its elements into H2, and H1/H2,and so on. This procedure yields the composition series of the group. It alwaysterminates on the identity element

G � H1 � H2 � · · · � Hk (simple subgroup) � e, (2.67)

generating the quotient subgroups

G/H1,H1/H2, . . . ,Hk . (2.68)

In general quotient groups need not be simple, but those in the composition seriesare because they are constructed at each step out of the maximal normal subgroup.This procedure dissects a group down in term of its simple constituents. That is thereason mathematicians consider the simple groups to be the fundamental entitiesout of which all other finite groups can be built. In the special case where all thequotient groups are the cyclic groups of prime order, the group is said to be solubleor solvable (Galois’s terminology).

The wonderful thing is that, although a group can have more than one composi-tion series, i.e. several normal subgroups to choose from, composition series havesome invariant features: the number of steps to the identity is always the same, andthe order of the quotient groups, called the composition indices are also the same!They are therefore characteristics of the group.

The quaternion group has several subgroups, and they are all normal andAbelian; the largest is Z4, and three different Z2. Its three-step compositionseries is

Q � Z4 � Z2 ⊃ e, (2.69)

with quotient groups

Q/Z4 = Z2; Z4/Z2 = Z2; Z2. (2.70)

Its composition indices are 2, 2, 2, and the length is three.We encourage the reader to determine the composition series and normal

subgroups of S4 and A4.

Page 39: Pierre Ramond Group Theory a Physicists Survey 2010

2.4 Basic concepts 27

Scholium. Derived (commutator) subgroup

Let a and b be any two elements of a group G. We define their commutator as

[a, b] ≡ a−1b−1ab, (2.71)

which has the virtue to be equal to the unit element if they commute. Furthermore,since [b, a] = ([a, b])−1 the product of all possible commutators form a group,G ′ called the derived or commutator subgroup. It is easy to see that the derivedsubgroup is a normal subgroup. We have

[a, b] = g(a−1b−1ab

)g−1 = [a, b], (2.72)

where

a = g ag−1, b = g bg−1, . . . , (2.73)

so that conjugation maps the derived subgroup into itself. As an example, to com-pute D′

n , the derived subgroup of the dihedral group with elements of the formam, amb, m = 1, 2, . . . , n with b2 = e and aba = b. Explicitly, we find

[am, akb] = a−2m, [amb, akb] = a2(m−k), (2.74)

all other commutators being either the unit element or the inverse of the above.Hence the commutator subgroup of Dn is the cyclic group Zn/2 for n even, and Zn

if n is odd.There are three possibilities: G ′ is the same as G, then G is a perfect group;

G ′ = e, G is Abelian; otherwise G ′ �G, and G is not a simple group. With a distinctcommutator subgroup, Dn is not a simple group. Hence a group can be perfect (thefact that it is its own commutator subgroup does not mean it does not have normalsubgroups), while a non-Abelian simple group must be perfect!

The derived subgroup of Sn is An . That of A4 is the Klein four-group, and thederived subgroup of the quaternion group Q is Z2. None of these groups is simple.We leave it to the reader to find the derived subgroup of , the other group of ordertwelve.

2.4.3 Sylow’s criteria

Let p be a prime number. A p-group is a group for which the order of every elementis a power of p. They can be Abelian, such as the cyclic groups, or non-Abelian:D4 whose elements are of order two and four, is a Sylow two-group.

Let G be a group of order n. Decompose its order in powers of primes, say,n = pmr , where p is a prime, and r is integer not multiple of p. Then, there is aseries of theorems, due to Ludwig Sylow that state the following.

Page 40: Pierre Ramond Group Theory a Physicists Survey 2010

28 Finite groups: an introduction

• G contains n p (Sylow) p-subgroups of order pm , Gip, i = 1, 2 . . . , n p.

• All Gip are isomorphic to one another, related by G j

p = gGkpg−1, g ∈ G.

• n p is a divisor of r .• n p = 1 mod(p).

These theorems put very strong restrictions of the possible groups of a givenorder.

Consider groups of order n = pq where both p and q are prime. Sylow tells usthese groups have n p subgroups Gp of order p, and nq subgroups Gq of order q.These are necessarily cyclic groups. Sylow requires that n p divide q, that is n p = 1or q. In addition, n p = 1, 1+ p, . . . , so that if p > q, the unique solution is n p = 1.On the other hand, nq = 1, 1 + q, 1 + 2q, . . . are possible, as long as nq dividesp. There is always one solution with nq = 1, corresponding to

Zpq = Zp × Zq .

We have already seen this case: if a, b are elements of Zp and Zq respectively, theelement (a, b) is clearly of order pq.

The second possibility is when nq = p, but this is possible only if p = 1mod(q). This proves that there are at most two groups of order pq, if p = 1mod(q), and only one otherwise. If there is only one Sylow subgroup (n p = 1), itis a normal subgroup.

Let us consider a few applications. There is only one group of order 15 = 3 · 5,since 5 �= 1 mod(3), Z15 = Z3 × Z5.

On the other hand, there are two groups of order 21 = 7 · 3, since 7 = 1mod(3). The second group is called the Frobenius group. It is a subgroup of thepermutation group on seven objects, with one Sylow seven-subgroup generated by(1 2 3 4 5 6 7), and seven Sylow three-subgroups generated by (2 3 5), and (4 7 6).

Sylow analysis was very useful in the hunt for simple groups. The reason is thata group with only one Sylow subgroup cannot be simple, since conjugation mapsit into itself. For instance, any group of order 84 = 22 · 3 · 7 contains n7 Sylowseven-subgroups, where n7 is required to divide 12 and be 1, 8, 15, . . .. The onlysolution is n7 = 1, and there are no simple groups of order 84.

2.4.4 Semi-direct product

There is another way to construct a group out of two groups. It is more subtle thanthe direct product construction. It requires two groups G and K, and the action ofone on the other.

Page 41: Pierre Ramond Group Theory a Physicists Survey 2010

2.4 Basic concepts 29

Assume that G acts on K in the sense that it sends any of its elements k ∈ K intoanother one

G : k → k ′ ≡ k g, (2.75)

where g ∈ G. For any two group elements

(k g1) g2 = k g1g2 = kg′. (2.76)

For any two elements ka, kb ∈ K,

(kakb)g = k g

a k gb , (2.77)

so that the K-multiplication table is preserved (homomorphism)

ka kb = kc → (ka)g(kb)

g = k gc . (2.78)

Then we can verify that the elements (ka, gi ) form a group with composition law

( ka , gi ) · (kb , g j ) ≡ ( (ka)g j kb , gi g j ). (2.79)

This new group of order (nk ng) is called the semi-direct product of G with K, anddenoted by

K � G. (2.80)

The funny symbol is to distinguish it from the direct product construction.For finite groups, one can think of several possibilities: a natural choice is to take

G to be the automorphism group of K, sometimes called Aut K, in which case it iscalled the holomorph of the group

Hol(K) ≡ K � Aut K. (2.81)

A second possibility is to take K to be a normal subgroup of G. Then G acts on Kby conjugation.

Examples can illustrate the construction. Consider D2, with Aut D2 = S3. Thisallows us to form the semi-direct product of D2 with S3 or any of its subgroups. Ingeneral, we have

Dn = Zn � Z2. (2.82)

The presentation

Dn : ( an = e; b2 = e; bab−1 = a−1 )

shows that Z2 generated by b acts on Zn by conjugation.

Page 42: Pierre Ramond Group Theory a Physicists Survey 2010

30 Finite groups: an introduction

Scholium. Transitivity

The notion of transitivity plays an important role in the classification offinite groups. Permutations are described by cycles acting on sets of objects[ a1, a2, . . . , an−1, an ]. The n! permutations in Sn generate n! distinct arrangements[ b1, b2, . . . , bn−1, bn ]. Label the objects by number, and start from one arrange-ment, [ 1, 2, . . . , (n − 1), n ]. Clearly any other arrangement can be attained by apermutation in Sn: Sn is said to be (singly) transitive. Groups without this prop-erty are called intransitive. For example, the permutation group on four objects,generated by the cycles (1 2), (3 4) is intransitive because it does not contain thepermutation that maps [ 1, 2, 3, 4 ] to [ 3, 2, 1, 4 ]. Since Sn maps any set of nobjects into another, it is called n-ply transitive. Put in another way, if we definethe orbit of an element g as the set of all elements generated from g by group oper-ation, a transitive group has only one orbit. Obviously, any permutation in Sn is afortiori k-ply transitive with k < n. The correspondence

even permutations → 1, odd permutations → −1, (2.83)

maps Sn into Z2, which is a subgroup of Sn . As we have seen, An the group of evenpermutations form the group the kernel of this mapping. It is a normal subgroup,since conjugation by an odd permutation leaves the parity unaltered. Hence Sn

is not a simple group. Since An is the largest subgroup of Sn , we verify that itsquotient group is indeed simple, as

Sn/An = Z2, (2.84)

Consider the array of (n −1) objects [ a2, a3, . . . , an−1, an ] which does not containa1. The only way to generate the (n −1) array [ a1, a3, . . . , an−1, an ], where a2 hasbeen replaced by a1 is by the transposition (a1 a2):

[ a2, a3, . . . , an−1, an ] −→ [ a1, a3, . . . , an−1, an ]. (2.85)

(a1 a2)

This is an odd permutation, and we conclude that An is not (n − 1)-ply transitive.Now look at the set of (n−2) objects [ a3, a4, . . . , an−1, an ] which contains neithera1 nor a2. The array [ a1, a4, . . . , an−1, an ] where a3 has been replaced by a1 canalso be reached by a transposition (a1 a3) as before, but since we have one “idle”object a2, we can also reach it by means of the even permutation (a3 a1 a2)

[ a3, a4, . . . , an−1, an ] −→ [ a1, a4, . . . , an−1, an ]. (2.86)

(a3 a1 a2)

The set where both a3 and a4 are replaced by a1 and a2 is clearly obtained by the(even) product of two transpositions. Hence An is (n − 2)-ply transitive.

Page 43: Pierre Ramond Group Theory a Physicists Survey 2010

2.4 Basic concepts 31

Other than Sn and An , there are relatively few multiply transitive groups: severalinfinite families of doubly transitive groups, and one infinite family of triply tran-sitive groups. All have interesting interpretations in terms of discrete geometries.In addition, there are several isolated doubly transitive groups, one triply transitive,two quadruply transitive, and two quintuply transitive groups. All but one of theseare sporadic groups.

2.4.5 Young Tableaux

Finally, using S5, the first meaty permutation group on five objects as exemplar, wepresent a general method to organize its 5! = 120 permutations. We begin with thefollowing cycle decomposition.

( x x x x x ) 5 24 +( x x x x )( x ) 4+1 30 −( x x x )( x x ) 3+2 20 −( x x x )( x )( x ) 3+1+1 20 +( x x )( x x )( x ) 2+2+1 15 +( x x )( x )( x )( x ) 2+1+1+1 10 −( x )( x )( x )( x )( x ) 1+1+1+1+1 1 +

Here, ( x x x x x ) denotes a cycle with arbitrary entries. The number of differentcycles is in one-to-one correspondence with the partitions of 5, as indicated in theabove table, where we also list the number of permutations in each partition, as wellas their parity. To each partition of n corresponds an irreducible representation ofSn . It is common practice to express the partitions in terms of an ordered set of nintegers {λ1 ≥ λ2 ≥ · · · ≥ λn}, defined in terms of αk , the number of cycles oforder k as

λ1 =n∑

i=1

αi , λ2 =n∑

i=2

αi , . . . , λn = αn. (2.87)

For instance, ( x x x x )( x ) with α1 = 1, α4 = 1, yields the five integers{2, 1, 1, 1, 0}, in short {2, 13}.

Partitions can also be represented graphically, by means of Young Tableaux. TheYoung Tableau associated with the partition {λ1, λ2, . . . , λn} is obtained by draw-ing horizontally λ1 boxes side to side, then just below λ2 boxes, down to the lowestvalue of λk . This rule associates a unique Young Tableau to each partition and classof S5, as follows.

Page 44: Pierre Ramond Group Theory a Physicists Survey 2010

32 Finite groups: an introduction

(x)(x)(x)(x)(x) (x x)(x)(x)(x) (x x)(x x)(x) (x x x)(x x)

(x x x)(x)(x) (x x x x)(x) (x x x x x)

This representation of the partitions in terms of Young Tableaux turns out to bevery useful in other contexts, especially in determining tensor symmetries, andtransformation properties of products of representations.

Page 45: Pierre Ramond Group Theory a Physicists Survey 2010

3

Finite groups: representations

The representation theory of finite groups addresses the number of ways the oper-ations of a certain group can be written in terms of matrices acting on some vectorspace. It answers questions such as how many distinct representations does a grouphave, and how to distinguish between them? We have already seen that the groupactions of any finite group G of order n can be represented in terms of (n × n)matrices, forming the regular representation. In the following, we show how tobreak up the regular representation in terms of irreducible blocks of irreduciblerepresentations (irreps). Their number and type characterize the group, and are thestarting point for physical applications. For more details, the reader is directed toclassic books by Carmichael [2], Hall [9], and Hammermesh [10].

3.1 Introduction

Consider a finite vector space V, spanned by an orthonormal set of N bras and kets,

〈 i | j〉 = δi j ,

N∑i=1

| i 〉〈 i | = 1. (3.1)

We wish to study the action of an order-n group G on this vector space. We assumethat the action of g ∈ G in V is represented by (N × N ) non-singular matricesacting as

| i 〉 → | i(g) 〉 = Mi j (g)| j 〉, (3.2)

with

Mi j (g−1) = M−1

i j (g), (3.3)

forming an N -dimensional representation R of G. The regular representationwould correspond to n = N . Under what conditions can this representation be bro-ken down into smaller irreducible representations? We already know there is at

33

Page 46: Pierre Ramond Group Theory a Physicists Survey 2010

34 Finite groups: representations

least one, the trivial one-dimensional representation where every group action ismultiplication by one.

Assume that R is reducible, with a d1-dimensional matrix representation R1 of Gin V, acting on a smaller set | a 〉, a = 1, 2, . . . , d1 < N , spanning the subspace V1.The remaining | m 〉, m = 1, 2, . . . , (N − d1) span its orthogonal complement V⊥

1 .Assemble the kets in an N -dimensional column, with the group matrices acting as(

M[1](g) 0N (g) M[⊥](g)

). (3.4)

These matrices keep their form under multiplication, with

M[1](gg′) = M[1](g)M[1](g′), M[⊥](gg′) = M[⊥](g)M[⊥](g′), (3.5)

and

N (gg′) = N (g)M[1](g′) + M[⊥](g)N (g′). (3.6)

A clever change of basis, (1 0V 1

), (3.7)

where

V = 1

n

∑g

M[⊥](g−1)N (g), (3.8)

brings the representation matrices to block-diagonal form(M[1] 0

0 M[⊥]

). (3.9)

In this basis, the original representation R is said to be fully reducible, that is

R = R1 ⊕ R⊥1 .

Clearly, the process can be continued if R⊥1 is itself reducible, until one runs out

of subspaces. In general, a representation R on V will be expressed as a sum ofirreducible representations

R = R1 ⊕ R2 ⊕ · · · ⊕ Rk, (3.10)

corresponding to

V = V1 ⊕ V2 ⊕ · · · ⊕ Vk . (3.11)

The same irreducible representation may appear several times in this decomposi-tion. Application of this decomposition to the regular representation is the key ofrepresentation theory.

Page 47: Pierre Ramond Group Theory a Physicists Survey 2010

3.2 Schur’s lemmas 35

3.2 Schur’s lemmas

Express R1 of G in matrix form

| a 〉 → | a(g) 〉 = M[1]ab (g)| b 〉. (3.12)

Since both | a 〉 and | i 〉 live in V, they are linearly related

| a 〉 = Sai | i 〉, (3.13)

where S is a (d1 × N ) array. This enables us to view the group action in two ways:

| a 〉 → | a(g) 〉 = M[1]ab (g)| b 〉 = M[1]

ab (g)Sbi | i 〉, (3.14)

or

| a 〉 = Sai | i 〉 → Sai | i(g) 〉 = SaiMi j (g)| j 〉. (3.15)

Comparing the two and using the completeness of the {| i 〉}, we deduce that thereducibility of R implies

SaiMi j (g) = M[1]ab (g)Sbj . (3.16)

Conversely, starting from this equation, we see that the {| a 〉} form a smaller rep-resentation of G: R is reducible. Thus if R is irreducible, there is no such S . Wecan state the result as Schur’s first lemma.

If the matrices of two irreducible representations of different dimensions are relatedas SaiMi j (g) = M[1]

ab (g)Sbj , then S = 0.

In the special case when d1 = N , the S array becomes a (N × N ) matrix. If bothsets | i 〉 and | a 〉 span the same Hilbert space, then S must have an inverse, and R

is related to R1 by a similarity transformation

M[1] = SMS−1. (3.17)

The two representations are said to be equivalent. It could also be that the tworepresentations have the same dimension, but live in different Hilbert spaces; inthat case S = 0. We conclude that a non-vanishing S implies either reducibility orsimilarity.

Now assume that R is irreducible. Any matrix S which satisfies

M(g)S = SM(g), ∀ g ∈ G, (3.18)

must be proportional to the unit matrix: if | i 〉 is an eigenket of S , then every | i(g) 〉is also an eigenket with the same eigenvalue, reducing S to a multiple of the unitmatrix (the eigenvalue cannot be zero). This is Schur’s second lemma.

Page 48: Pierre Ramond Group Theory a Physicists Survey 2010

36 Finite groups: representations

These two lemmas have important consequences. To start, consider Rα and Rβ ,two irreps of (dα × dα) matrices M[α](g) in Vα, and (dβ × dβ) matrices M[β](g)in Vβ . Construct the mapping S

S =∑

g

M[α](g)N M[β](g−1), (3.19)

where N is any (dα × dβ) array. It satisfies

M[β](g)S = SM[α](g). (3.20)

If the two representations are different, dα �= dβ , and by Schur’s first lemma, Snecessarily vanishes. If on the other hand, S �= 0, then one of the representationsmust be reducible.

Thus for any two different irreps, Rα and Rβ , an immediate consequence ofSchur’s first lemma is

1

n

∑g

M[α]i j (g)M[β]

pq (g−1) = 0, α �= β, (3.21)

so that for fixed i, j = 1, 2 . . . dα and p, q = 1, 2 . . . dβ , the matrix elements oftwo different irreps are orthogonal over an n-dimensional vector space.

Setting Rβ = Rα, we find

1

n

∑g

M[α]i j (g)M[α]

pq (g−1) = 1

δiqδ j p. (3.22)

The (properly normalized) matrix elements of an irrep form an orthonormal set.This equation can be verified by multiplying both sides by an arbitrary matrix t jp,and summing over j and p, reproducing the result of Schur’s second lemma. Theproper normalization is determined by summing over i and q.

Let us now introduce the group-averaged inner product(i , j

) ≡ 1

n

∑g

〈 i(g) | j (g) 〉, (3.23)

where the sum is over all elements of G. It is group-invariant by construction, inthe sense that for any element ga ,

( i(ga), j (ga) ) = 1

n

∑g

〈 i(gag) | j (gag) 〉

= 1

n

∑g′=ga g

〈 i(g′) | j (g′) 〉 = ( i, j ). (3.24)

Page 49: Pierre Ramond Group Theory a Physicists Survey 2010

3.2 Schur’s lemmas 37

Then for any two vectors u and v, we have

(M(g−1)v, u ) = ( v, M(g)u ), (3.25)

since the inner product is group-invariant. On the other hand, for any representa-tion,

(M†(g)v, u ) = ( v, M(g)u ), (3.26)

so that

M(g−1) = M†(g), M(g−1)M(g) = 1. (3.27)

It follows that all representations of finite groups are unitary with respect to thegroup-averaged inner product. This will not be true when we consider continuousgroups, where the group-averaged inner product cannot in general be defined.

We summarize all these results in a master equation for different unitary irreps:

1

n

∑g

M[α]i j (g)M [β]

qp (g) = 1

δαβδiqδ j p. (3.28)

This equation can be rewritten in terms of the character of an element g in arepresentation Rα, defined as the trace of its representation matrix

χ [α](g) ≡ Tr M[α](g). (3.29)

It is not a unique function of g; rather, the cyclic property of the trace

χ [α](gagg−1a ) = Tr

(M[α](ga)M[α](g)M[α](g−1

a )) = Tr M[α](g), (3.30)

shows that all elements of one class have the same character: character is a classattribute. In addition, for a unitary transformation,

χ [α](g−1) = Tr M[α] †(g) = χ [α] (g). (3.31)

Applied to characters, our master formula becomes

1

n

∑g

χ [α](g)χ [β] (g) = δαβ. (3.32)

In terms of ni the number of elements in the i th class, with character χ [α]i , the sum

over the group elements becomes

{χ [α], χ [β] } ≡ 1

n

nC∑i=1

ni χ[α]i χ

[β]i = δ αβ, (3.33)

Page 50: Pierre Ramond Group Theory a Physicists Survey 2010

38 Finite groups: representations

where nC is the number of classes. Assuming nR different irreps, α = 1, 2, . . . , nR ,this equation states that the nR vectors χ [α] form an orthonormal set in annC -dimensional vector space. This can only be if

nR ≤ nC : (3.34)

the number of different irreps of a group is no larger than the number of its classes.Another consequence of this equation is that for any irrep,{

χ [α], χ [α]} = 1. (3.35)

Consider a reducible representation that is the sum of rα irreps Rα, with characters

χ =∑α

rαχ[α] . (3.36)

The multiplicity of the Rα is easily computed from the scalar product

rα = {χ, χ [α]} , (3.37)

which holds for any character χ . Also

{χ, χ } =∑α β

rαrβ{χ [α], χ [β]} =

∑α

r2α, (3.38)

provides a simple test for irreducibility: if the scalar product of its characters isequal to one, the representation is irreducible.

These formulæ can be applied to the n-dimensional regular representation,which contains all the irreps Rα,

Rreg =nR∑α=1

rαRα, (3.39)

so that

n =nR∑α=1

rα dα, (3.40)

its characters are expressed as

χregi =

nR∑α=1

rαχ[α]i . (3.41)

In the regular representation, any group element shuffles the n-vector elementsinto different ones: they are represented by matrices which have zeroes on theirdiagonal. The only exception is the identity element. Therefore all the charactersof the regular representation vanish except for the identity

χreg1 = n, χ

regi = 0, i �= 1. (3.42)

Page 51: Pierre Ramond Group Theory a Physicists Survey 2010

3.2 Schur’s lemmas 39

As a consequence,

rα = {χ reg, χ [α]} = {

χreg1 , χ [α]} = χ

[α]1 . (3.43)

Since the unit element of any representation is represented by the unit matrix, itscharacter is the dimension of the representation

χ[α]1 = dα, (3.44)

so that

rα = dα. (3.45)

Hence we arrive at the following important result.

The multiplicity of any irrep in the regular representation is equal to its dimension.

A relation between the order of the group and the dimensions of its irreps ensues,

n =nR∑α=1

(dα)2 . (3.46)

This equation is quite useful in determining the various irreps of groups of loworder.

Given the class Ci made up of ni elements ( g(i)1 , g(i)

2 , . . . , g(i)ni

) obtained byconjugation, we define

Ai ≡ 1

ni

ni∑m=1

g(i)m , (3.47)

which is self-conjugate by construction,

Ai = Ai . (3.48)

Furthermore, the Ai form a closed algebra. Their product is of course a sum ofelements of the group, but it is a self-conjugate sum. Hence the elements mustcome in full classes. Their product is made up of Ak as well

Ai A j =∑

k

ci jk Ak, (3.49)

where the ci jk , sometimes called the class coefficients, depend on the group. In Rα,they are matrices

N [α](i) ≡ M[α](Ai ) = 1

ni

∑g∈Ci

M[α](g), (3.50)

Page 52: Pierre Ramond Group Theory a Physicists Survey 2010

40 Finite groups: representations

with

Tr N [α](i) = χ

[α]i . (3.51)

Since

M[α](g)N [α](i) M[α](g−1) = N [α]

(i) , (3.52)

the N [α](i) commute with all the group matrices: when Rα is irreducible, they are

multiples of the unit matrix in the representation

N [α](i) = χ

[α]i

χ[α]1

1. (3.53)

Since

N [α](i) N

[α]( j) =

∑k

ci jkN[α](k) , (3.54)

we arrive at

χ[α]i χ

[α]j = χ

[α]1

∑k

ci jkχ[α]k . (3.55)

We can rewrite the characters of the regular representation as

χregi =

nR∑α=1

rαχ[α]i =

nR∑α=1

dαχ[α]i =

nR∑α=1

χ[α]1 χ

[α]i . (3.56)

Hence we see that

nR∑α=1

χ[α]i χ

[α]j =

∑k

ci jkχregk . (3.57)

Since all characters of the regular representation vanish except for the identity, itfollows that

nR∑α=1

χ[α]i χ

[α]j = nci j1. (3.58)

Now let us consider Ci the class made of all the inverses of class Ci . In a unitaryrepresentation, their characters are related by complex conjugation,

χ[α]i

= χ[α]i . (3.59)

Thus if the character for a particular class is real, the class contains its own inverses;if it is complex, the class and its inverse class are different. The product Ai A j

Page 53: Pierre Ramond Group Theory a Physicists Survey 2010

3.3 The A4 character table 41

contains the unit element only if some member of the j class is the inverse of onein the i class, but this only happens if j = i , when we have

Ai Ai = 1

ni

e + · · ·· (3.60)

Thus we conclude that

ci j1 ={

0, j �= i1ni, j = i

. (3.61)

Putting it all together, we arrive at

nR∑α=1

χ[α]i χ

[α]j = n

niδi j . (3.62)

The nC vectors χi form an orthogonal basis in the nR-dimensional vector space,which implies that

nC ≤ nR. (3.63)

The number of classes is limited to the number of irreps, but we already saw thatthe number of irreps is limited by the number of classes. Hence we arrive at thefundamental result of representation theory.

The number of irreps of a finite group is equal to its number of classes.

Once we know the classes of a group, we can deduce its representations. Theinformation is usually displayed in a character table, which lists the values of thecharacters for the different representations. A note of caution: in rare occasions twodifferent groups (e.g. the quaternion group Q and the dihedral group D4) can havethe same character table. In the following, we show how to derive the charactertable for A4, and list the character tables of the low-order groups we have alreadyencountered.

3.3 The A4 character table

The alternating group A4 contains the even permutations on four letters, of theform (x x x) and (x x)(x x). A hasty conclusion would be that it has three classes,but there is a subtlety in the three-cycle: it contains elements that differ from oneanother by one transposition, such as (1 2 3) and (1 3 2), which cannot be mappedby A4 conjugation into one another. Hence the three-cycles break up into twoclasses which contain the inverses of one another. Their characters are related bycomplex conjugation.

Page 54: Pierre Ramond Group Theory a Physicists Survey 2010

42 Finite groups: representations

If a = (1 2)(3 4), and b = (1 2 3), the four classes are C1(e) which con-tains the unit element, C2(b, ab, aba, ba) with four elements, its inverse classC3(b2, ab2, b2a, ab2a) also with four elements, and C4(a, b2ab, bab2) with threeelements. It has four irreps, one of which is the trivial unit representation. The orderis low enough that we can easily find their dimensions di . Indeed, they must satisfy

12 = 1 + d22 + d2

3 + d24 , (3.64)

which has the unique solution d2 = d3 = 1, d4 = 3. Let us denote the four rep-resentations by 1, 11, 12, and 3. Their multiplicity in the regular representation isequal to their dimension, so that (we dispense with the arcane ⊗ and ⊕, and simplyuse × and + for the product and sum of representations)

Rreg(A4) = 1 + 11 + 12 + 3 + 3 + 3. (3.65)

Let χ [α] be the characters of the four irreps. We form a (4 × 4) character table. Itsrows are labeled by the characters, and its columns by the classes. The first columnis the unit matrix, the trace of which is the dimension of the irrep. The first rowis the (1 × 1) unit irrep where all characters are equal to one. We start with thefollowing.

A4 C1 4C2 4C3 3C4

χ [1] 1 1 1 1χ [11] 1 z z rχ [12] 1 u u sχ [3] 3 w w t

Since the class C4 is its own inverse, its characters r , s and t are real numbers.To find the characters in these representations, we can use the orthogonality ofcharacters, or that the weighted sum of the characters along any row (except forχ [1]) vanishes. The latter determines r , s, and t in terms of the real parts of z, uand w, but we need more information. It is encoded in the class coefficients ci jk

through

χ[α]i χ

[α]j = dα

∑k

ci jkχ[α]k . (3.66)

The class coefficients of A4 are easily worked out from the commutative multipli-cation table of the Ai matrices (eq. (3.49))

A4 A4 = 1

3A1 + 2

3A4, A2 A3 = 1

4A1 + 3

4A4,

A4 A2 = A3 A3 = A2; A4 A3 = A2 A2 = A3. (3.67)

Page 55: Pierre Ramond Group Theory a Physicists Survey 2010

3.3 The A4 character table 43

These tell us that for α = 1, 2, 3,

χ[α]3 χ

[α]3 = χ

[α]2 χ

[α]4 = χ

[α]2 , (3.68)

implying that

z z = zr = z, u u = us = u : (3.69)

so that z and u are either zero or else cubic roots of unity. The orthogonality of thecharacters with the unit representation leads to the vanishing of the weighted rowsum of χ [2]

1 + 4(z + z) + 3r = 0, (3.70)

together with 1 + z + z = 0, which yields r = 1, and z = e±2π i/3. We see that zcannot be zero; if it were, the class equation

0 = χ[2]2 χ

[2]3 = 1

4+ 3

[2]4 , (3.71)

would require r = −1/3, which would contradict the class relation

χ[2]4 χ

[2]4 = 1

3+ 2

[2]4 . (3.72)

The same method applied to the third row fixes

s = 1, u = z = e∓2π i/3, (3.73)

taking into account the orthogonality between the second and third rows. For thethree-dimensional irrep, the class relation,

χ[4]3 χ

[4]3 = χ

[4]2 χ

[4]4 = 3χ [4]

2 , (3.74)

yields

ww = w t = 3w, (3.75)

while

χ[4]4 χ

[4]4 = 3

(1

[4]1 + 2

[4]4

), (3.76)

Page 56: Pierre Ramond Group Theory a Physicists Survey 2010

44 Finite groups: representations

leads to two solutions, t = −1 and t = 3. The solution t = −1 yields byorthogonality w + w = 0. The second solution yields w + w = wz + wz =wz + wz = −12, which is not consistent unless w + w vanishes. Hence the onesolution t = −1, w = 0 completes the character table. The characters of 11 and 12

are complex conjugates of one another: these two irreps are said to be conjugatesas well, 12 = 11.

The three non-trivial classes are C [3]2 (b, ab, aba, ba), C [3]

3 (b2, ab2, b2a, ab2a),and C [2]

4 (a, b2ab, bab2), with the order of their elements in a square-bracketedsuperscript. The elements are written in term of the two generators with presen-tation < a, b | a2 = e, b3 = e, (ba)3 = e >. The A4 character table is then asfollows.

A4 C1 4C [3]2 4C [3]

3 3C [2]4

χ [1] 1 1 1 1χ [11] 1 e2π i/3 e4π i/3 1χ [11] 1 e4π i/3 e2π i/3 1χ [3] 3 0 0 −1

Since 3 is a real irrep, it is not hard to deduce the expression for its generators as(3 × 3) matrices:

3 : a =⎛⎝−1 0 0

0 1 00 0 −1

⎞⎠ ; b =⎛⎝0 1 0

0 0 11 0 0

⎞⎠ . (3.77)

3.4 Kronecker products

In order to study how representations combine, consider two irreps of G, Rα andRβ , represented on two Hilbert spaces spanned by dα and dβ bras and kets, 〈 i |α,i, j = 1, 2 . . . , dα, and 〈 s |β , s, t = 1, 2 . . . , dβ , respectively, that is

| i 〉α → | i(g) 〉α = M[α]i j (g)| j 〉α, (3.78)

| s 〉β → | s(g) 〉β = M[β]st (g)| t 〉β. (3.79)

We wish to study how G is represented on the product space spanned by

| i 〉α| s 〉β ≡ | A 〉, (3.80)

where A, B take on dαdβ values of the pairs {i, s}.

Page 57: Pierre Ramond Group Theory a Physicists Survey 2010

3.4 Kronecker products 45

We define a new matrix for the representation Rα × Rβ by

| A 〉 → | A(g) 〉 ≡ M[α×β]AB (g)| B 〉,

= M[α]i j (g)M[β]

st (g)| j 〉α| t 〉β. (3.81)

This new reducible representation, called the Kronecker product, is like any otherexpressible as a sum of irreps of G. We write the Clebsch–Gordan series

Rα × Rβ =∑γ

d(α, β | γ )Rγ , (3.82)

where the d(α, β | γ ) are to be determined. In terms of characters,

χ [α×β] =∑γ

d(α, β | γ )χ [γ ]. (3.83)

Now, trace over the AB to obtain the characters of the Kronecker product

χ [α×β] =∑

A

M[α×β]AA (g) =

∑i s

M[α]i i (g)M[β]

ss (g), (3.84)

that is for any class

χ [α×β] = χ [α] χ [β]. (3.85)

Using the orthogonality property of the characters, we obtain the coefficients

d(α, β | γ ) = {χ [α×β], χ [γ ] } = 1

n

nC∑i=1

niχ[α]i χ

[β]i χ

[γ ]i . (3.86)

This allows us to find the product of irreps from the character table. The productof the unit irrep 1 with any other irrep reproduces that irrep, using orthogonality ofthe characters. As an example, let us compute the Kronecker products of A4 irreps.From the character table, we find

d(3, 3 | 1) = d(3, 3 | 11) = d(3, 3 | 12) = 1; (3.87)

d(3, 3 | 3) = 33 + 3(−1)3

12= 2, (3.88)

whence

3 × 3 = 1 + 11 + 12 + 3 + 3. (3.89)

Similarly

3 × 11 = 3; 3 × 12 = 3; 11 × 12 = 1. (3.90)

We can separate the product of the same representation into its symmetric andantisymmetric parts, which must transform separately. Thus we see that we have

(3 × 3)sym = 1 + 11 + 12 + 3, (3 × 3)antisym = 3. (3.91)

Page 58: Pierre Ramond Group Theory a Physicists Survey 2010

46 Finite groups: representations

We want to find in detail the decomposition of the Kronecker product of two ofits representations r and s. In these representations, the group generators will bewritten as a[r], b[r], and a[s] b[s], assuming for simplicity that the group has twogenerators a and b (rank two). These act on the Hilbert spaces | i 〉r and | m 〉s,respectively. The Kronecker product r × s lives in the direct product Hilbert space| i 〉r | m 〉s: the group acts on this space with generators

A[r×s] = a[r]a[s], B[r×s] = b[r]b[s], (3.92)

in such a way that

A[r×s]| i 〉r | m 〉s = a[r]| i 〉r a[s]| m 〉s, (3.93)

and

B[r×s]| i 〉r | m 〉s = b[r]| i 〉r b[s]| m 〉s. (3.94)

Assuming that we know the action of the generators in the original representations,this very convenient formula allows us to explicitly decompose the direct productHilbert space into irreducible subspaces.

Consider for example the antisymmetric product of two A4 triplets. The productHilbert space contains the three antisymmetric states. Then by using the explicitform of a in the triplet, eq. (3.77), we find

A[(3×3)a ] (| 2 〉3 | 3 〉3′ − | 3 〉3 | 2 〉3′) = − (| 2 〉3 | 3 〉3′ − | 3 〉3 | 2 〉3′

)A[(3×3)a ] (| 3 〉3 | 1 〉3′ − | 1 〉3 | 3 〉3′

) = + (| 3 〉3 | 1 〉3′ − | 1 〉3 | 3 〉3′)

A[(3×3)a ] (| 1 〉3 | 2 〉3′ − | 2 〉3 | 1 〉3′) = − (| 1 〉3 | 2 〉3′ − | 2 〉3 | 1 〉3′

)so that A is diagonal with the same eigenvalues as in the triplet. The reader canverify that B also has the triplet form so that the antisymmetric product of twotriplets is a triplet, as derived earlier.

3.5 Real and complex representations

Consider an irrep r represented by non-singular matrices M[r](g), for any elementg of any finite group. It is obvious that the complex conjugates of these matrices

M [r](g) also form an irrep, r′, since they satisfy the same closure property. What

is the relation of this representation to r? Clearly, the characters of elements in thesame classes are related by complex conjugation

χ [r](g) = χ [r′](g). (3.95)

One can consider two possibilities. If r happens to have complex characters, r′ isa distinct representation with characters conjugate to those of the first; we call r′

Page 59: Pierre Ramond Group Theory a Physicists Survey 2010

3.5 Real and complex representations 47

the conjugate representation and denote it by r . In this case the two matrices Mand M cannot be related by a similarity transformation since they have differentcharacters.

If r has real characters, then the characters of r and r′ are the same, and theirrepresentation matrices must be related by a similarity transformation

M = SMS−1. (3.96)

Using unitarity, (M−1

)T = M, (3.97)

where the superscript T means transposition, we deduce that

S = MSMT . (3.98)

Taking the transpose and inverse of this equation, we arrive at

ST S−1 M = MST S−1. (3.99)

Since M represents an irreducible representation, STS−1 must be proportional tothe unit matrix, which implies that S is proportional to its transpose; hence it iseither symmetric or antisymmetric

ST = ±S. (3.100)

In both cases, r is the same (equivalent) to r′. However when S is symmetric, thesymmetric product of the representation contains a singlet, while in the second casethe singlet (invariant) is to be found in the antisymmetric product. Representationswith an antisymmetric S are called pseudoreal. In either case the invariant can bewritten in the form

Y T S X, (3.101)

where X and Y are column vectors acted upon by M[r].To summarize, representations are of three types: complex with complex char-

acters, real with real characters and with a symmetric quadratic invariant, andpseudoreal with an antisymmetric quadratic invariant.

Complex representations are obviously different from real and pseudoreal rep-resentations, while real and pseudoreal representations are harder to distinguish.Fortunately, there is a simple way to find out which is which in terms of characters.We can always express S in the form

S =∑

g

MT (g)V M(g), (3.102)

Page 60: Pierre Ramond Group Theory a Physicists Survey 2010

48 Finite groups: representations

in terms of an arbitrary matrix V , with the sum over all group elements. Indeed forany group element g′

MT (g′)SM(g′) =∑

g

MT (g′)MT (g)V M(g)M(g′), (3.103)

=∑

g′′MT (g′′)VM(g′′) = S. (3.104)

In terms of matrix elements, we have for the real and pseudoreal representations

Si j =∑

g

M(g)kiVkl M(g)l j = ±∑

g

M(g)k jVkl M(g)li . (3.105)

Hence for any k, l, i and j∑g

M(g)kiM(g)l j = ±∑

g

M(g)k jM(g)li . (3.106)

Sum independently over k j and il to obtain∑g

M(g)kiM(g)ik = ±∑

g

M(g)kkM(g)i i , (3.107)

that is ∑g

χ(g2) = ±∑

g

χ(g)χ(g). (3.108)

Since we are dealing with real representations, the right-hand side is just the orderof the group. When the representation is complex, S does not exist and the aboveexpression is zero. Hence our criterion

1

n

∑g

χ(g2) =

⎧⎪⎨⎪⎩+1 real

−1 pseudoreal

0 complex

. (3.109)

Using the character table of the quaternion group, one can show that its doubletrepresentation is pseudoreal, either by using its antisymmetric Kronecker productor by applying this criterion.

3.6 Embeddings

Let G be a group of order n, with irreps R[α], of dimension Dα and characters �[α],represented by matrices M[α](g). If K is an order-k subgroup of G, with irrepsT[a], of dimensions da and characters χ [a], we wish to study the embeddings of the

Page 61: Pierre Ramond Group Theory a Physicists Survey 2010

3.6 Embeddings 49

irreps of K into those of G. The set of matrices M[α](g) form an irrep of G, butspecialized to elements k in the subgroup

M[α](k), k ∈ K ⊂ G, (3.110)

they form a (Dα × Dα) reducible representation of K. Let N [a](k) be the (da × da)

matrices with characters χ [a] which represent K. The full reducibility propertyallows us, after the appropriate similarity transformation of the (Dα × Dα) matri-ces, to express them in block-diagonal form, each block being one of the N [a]s

M[α](k) =∑

a

f αa N [a](k), (3.111)

or more abstractly

R[α] =∑

a

f αa T[a]. (3.112)

The embedding coefficients f αa are positive integers, subject to the dimension

constraints,

Dα =∑

a

f αa da. (3.113)

We now show how they are determined from the class structures of G and of K. Theclasses Ci of G break into two categories, those that do not contain any element ofK, and those that do. In the latter case, a given class Ci of G will in general containelements of K which live in different classes im of K.

If the class Ci contains elements of K, we have

�[α]i =

∑a

f αa χ

[a]im

, (3.114)

where χ[a]im

is the K character of any of its im classes that appear in the Ci classof G. Note that this formula does not depend on m.

We first use the orthogonality of the G characters to extract

n

niδi j =

∑α

�[α]j �

[α]i =

∑a

(∑α

f αa �

[α]j

[a]im

. (3.115)

Since this applies to all the classes of K, we then use the orthogonality of itscharacters, where kim is the number of elements in im

, to find(n

k

)∑im

kim

niχ

[b]im

=∑α

f αb �

[α]i , (3.116)

where the sum is over the classes of K which are contained in the Ci class of G.

Page 62: Pierre Ramond Group Theory a Physicists Survey 2010

50 Finite groups: representations

If the class C⊥ with character �[α]⊥ has no element in the subgroup, a similar

reasoning leads to ∑α

f αb �

[α]⊥ = 0. (3.117)

Sums of characters with positive integer coefficients are called compound charac-ters. The length of a compound character

∑i

ni

n

(∑α

f αb �

[α]i

)2

(3.118)

is the number of (simple) characters it contains: a compound character of unitlength is a character. These formulæ can be used to find the embedding coefficients,and also determine the G characters from those of its subgroup K.

Example: S4 ⊃ A4

Let us find the characters of S4 from those of A4, as well as the embedding coeffi-cients. S4 has five classes. Two of them, C2 and C5 contain only odd permutationsand do not appear in A4. The other three split according to

C1 ⊃ 1, C3 ⊃ 4, C4 ⊃ 2 + 3, (3.119)

where 1,2,3,4 are the A4 classes. Hence∑α

f αb �

[α]2 = 0,

∑α

f αb �

[α]5 = 0. (3.120)

Applying to the remaining classes yields∑α

f αb �

[α]1 = 24

12χ

[b]1 , for C1, (3.121)

∑α

f αb �

[α]3 = 24

12χ

[b]4 , for C3, (3.122)

∑α

f αb �

[α]4 = 24

12

(4

[b]2 + 4

[b]3

), for C4. (3.123)

The next step is to evaluate the compound characters for the four A4 irreps. Theirvalues are summarized in the following table.

Page 63: Pierre Ramond Group Theory a Physicists Survey 2010

3.6 Embeddings 51

C1 6C2 3C3 8C4 6C5 length

1: 2 0 2 2 0 211: 2 0 2 −1 0 111: 2 0 2 −1 0 13: 6 0 −2 0 0 2

Here, we have used e2π i/3 + e4π i/3 = −1, and indicated their lengths. Since S4 hasa singlet representation with character �[1], it is easy to see that the first compoundcharacter contains the singlet once. This yields the first character by subtrac-tion, namely �[1′] = (1,−1, 1, 1,−1). This is the one-dimensional representationwhere the odd permutations are equal to minus one.

The second and third compound characters have unit length: they represent oneirrep with character �[2] = (2, 0, 2,−1, 0).

The fourth compound character of length two is made up of two characters. Onecan check that it is orthogonal to �[1], �[1′], and �[2]. It must therefore split intotwo characters each representing a three-dimensional irrep

�[31] = (3, 1,−1, 0,−1), �[32] = (3,−1,−1, 0, 1). (3.124)

The ambiguities in determining their values are fixed by orthogonality with theother characters. This leads us to the following S4 character table.

S4 C1 6C [2]2 3C [2]

3 8C [3]4 6C [4]

5

χ [1] 1 1 1 1 1χ [1′] 1 −1 1 1 −1χ [2] 2 0 2 −1 0χ [31] 3 1 −1 0 −1χ [32] 3 −1 −1 0 1

We can read off the embedding coefficients, and deduce the decomposition of theS4 irreps in terms of those of A4:

S4 ⊃ A4 : 1′ = 1; 2 = 11 + 11; 31 = 3; 32 = 3. (3.125)

We see that in going from A4 to S4, the A4 singlet splits into two one-dimensionalirreps of S4. The same happen to the A4 triplet which splits into two triplets ofS4. On the other hand, the complex A4 one-dimensional irrep and its conjugateassemble into one real S4 doublet.

Page 64: Pierre Ramond Group Theory a Physicists Survey 2010

52 Finite groups: representations

This is a general pattern: some irreps duplicate, and some assemble in goingfrom the subgroup to the group.

There is a systematic way to understand these patterns in terms of the auto-morphism group of A4. Some mappings of the group into itself, like conjugation,live inside the group; they are the inner automorphisms. Some groups haveautomorphisms which are not to be found inside the group, called outer automor-phisms. This is true for A4 (as we saw earlier), and in fact for any alternatinggroup.

Let r = (1 2) be the odd permutation of order two which permutes the first twoobjects, and generates Z2, the group of outer automorphisms of A4. It is not in A4,but maps it into itself by conjugation: ai → r air−1 with ai ∈ A4. This allows theconstruction of the group S4, which contains both even and odd permutations. Wedenote their relation symbolically (as in the ATLAS) by

S4 = A4 · 2, Z2 = S4/A4, (3.126)

with A4 as commutator subgroup of S4, and Z2 is the quotient group.

3.7 Zn character table

The cyclic groups of order n are generated by a single element a of order n, withpresentation

Zn : < a | an = e > .

All its elements a, a2, . . . , an−1, e commute with one another: all can be rep-resented by (1 × 1) matrices, that is (complex) numbers. All its n classes areone-dimensional since it is Abelian. The order of the element in each classdepends on n. If n is prime, all classes contain an element of order n, but ifn = p · q · · · is a product of integers, some classes have elements of orderp, q, etc. For this reason, we do not include the bracketed superscript forthe classes. As the presentation requires a to be any one of the nth root ofunity,

a = 1, e2iπ/n e2·2iπ/n, . . . , e(n−1)·2iπ/n,

each choice yields one of n different irreps, labeled by k = 0, 1, . . . , n − 1, inagreement with the n one-dimensional conjugacy classes. The first one is the trivialand (very) unfaithful singlet, with all elements equal to one. This information isencoded in the following Zn character table.

Page 65: Pierre Ramond Group Theory a Physicists Survey 2010

3.8 Dn character table 53

Zn C1 C2 C3 · · · · · · Cn

χ [1] 1 1 1 · · · · · · 1χ [11] 1 ε ε2 · · · · · · εn−1

χ [12] 1 ε2 ε4 · · · · · · ε2n−2

χ [13] 1 ε3 ε6 · · · · · · ε3n−3

· · · · · · · · · · · · · · · · ·χ [1n−1] 1 εn−1 ε2n−2 · · · · · · ε(n−1)2

Here, ε = e2iπ/n . Further, since ε∗ = 1/ε, we see that for even n two repre-sentations are real, and (n − 2)/2 are complex. For n odd = 2m + 1, only thesinglet is real with m complex representations (and their conjugates). In the regularrepresentation a is represented by the (n × n) matrix

a : M(a) =

⎛⎜⎜⎜⎜⎜⎜⎜⎝

0 1 0 0 · · · 00 0 1 0 · · · 0· · · · · · · ·· · · · · · · ·0 0 0 0 · · · 11 0 0 0 · · · 0

⎞⎟⎟⎟⎟⎟⎟⎟⎠,

which can be brought to diagonal form

T M(a) T −1 = Diag(1, ε(n−1), ε(n−2), · · · , ε

),

by

T = 1√n

⎛⎜⎜⎜⎜⎜⎜⎜⎝

1 1 1 · · · 11 ε ε2 · · · εn−1

1 ε2 ε4 · · · ε2n−2

· · · · · · ·· · · · · · ·1 εn−1 ε2n−2 · · · ε(n−1)2

⎞⎟⎟⎟⎟⎟⎟⎟⎠.

Finally, the Kronecker products of its representations are pretty easy to read offfrom the character table:

1i × 1j = 1k, k = i + j mod(n).

3.8 Dn character table

A slightly more complicated construction is that of the character tables of the dihe-dral groups. It is convenient to start from their expression as semi-direct productsof cyclic groups

Page 66: Pierre Ramond Group Theory a Physicists Survey 2010

54 Finite groups: representations

Dn = Zn � Z2,

with presentation < a, b | an = b2 = e ; bab−1 = a−1 >, where a and b gen-erate Zn and Z2, respectively. The 2n elements of Dn are then e, a, a2, . . . , an−1

and b, ba, ba2, . . . , ban−1. The class structure is different as to whether n is evenor odd.

For odd n, conjugation under b generates (n − 1)/2 classes:(a, a−1

),(a2, a−2

), . . . ,

(a(n−1)/2, a(1−n)/2

),

and conjugation under a, aba−1 = ba−2, yields one class(b, ba−2, ba−4, . . .

)of n elements: D2m+1 has (m + 2) classes, and of course as many irreps.

For even n, b-conjugation yields the same (n−2)/2 classes of the form (ak, a−k),plus one additional class of one element an/2. On the other hand, a-conjugationgenerates two classes, one with elements of the form baeven, the other baodd: D2m

has (m + 3) classes and as many irreps.To construct their character tables, consider the one-dimensional representations

for which a and b are commuting real or complex numbers. It follows that b = ±1,and bab−1 = a = a−1, so that we find both a2 = an = e.

For odd n, these have only one solution, a = 1: D2m+1 has two one-dimensionalirreps (a = 1, b = ±1).

Since b2 = e, the action of Z2 on Zn yields at most two elements of Zn (two-dimensional orbits), and we expect Dn to have two-dimensional irreps. The orderof the group formula

2n = 1 + 1 + (n − 1)

2(2)2

yields (n − 1)/2 two-dimensional irreps, in which a is represented by a reducible(2 × 2) matrix. To construct them, go to a basis where a is a diagonal matrix

a =(α 00 β

),

where α and β are nth roots of unity. Z2 is represented by its regular representation,but since b need not commute with a, it cannot be diagonal, leading to the unitarilyequivalent form

b =(

0 11 0

).

We see that b-conjugation of a requires β = α−1, yielding for all two-dimensionalrepresentations

Page 67: Pierre Ramond Group Theory a Physicists Survey 2010

3.8 Dn character table 55

a =(

e2π ik/n 00 e−2π ik/n

), χk(a) = 2ck ≡ 2 cos

(2πk

n

); b =

(0 11 0

),

with k = 1, 2, . . . , (n − 1)/2. The character table for n = 2m + 1 easily follows.

D2m+1 C1 2C2(a) 2C3(a2) · · · 2Cm+1(am) nCm+2(b)

χ [1] 1 1 1 · · · 1 1χ [11] 1 1 1 · · · 1 −1χ [21] 2 2c1 2c2 · · · 2cm 0χ [22] 2 2c2 2c4 · · · 2c2m 0· · · · · · · · · · · · · · · · ·χ [2m] 2 2cm 2c2m · · · 2c

m2 0

The Kronecker products of the two one-dimensional irreps 1, 11, and the mdoublets 2k, k = 1, 2, . . . ,m are rather easy to work out:

11 × 11 = 1, 11 × 2k = 2k,

2k × 2k = 1 + 11 + 22k,

2j × 2k = 2j−k + 2j+k j > k,

with the identification 2j = 22n−j for j = 1, 2, . . . , (n − 1).For even n = 2m, the presentation yields two possible solutions for the singlets,

a = ±1: D2m has four one-dimensional representations (a = ±1, b = ±1). Itfollows from the order of the group formula

2(2m) = 1 + 1 + 1 + 1 + (m − 1)(2)2,

that we have (m − 1) two-dimensional representations. The slightly more compli-cated character table reads as follows.

D2m C1 2C2(a) 2C3(a2) ·· 2Cm(am−1) Cm+1(am) mCm+2(b) mCm+3(ba)

χ [1] 1 1 1 ·· 1 1 1 1χ [11] 1 1 1 ·· 1 1 −1 −1χ [12] 1 −1 1 ·· (−1)m−1 (−1)m 1 −1χ [13] 1 −1 1 ·· (−1)m−1 (−1)m −1 1χ [21] 2 2c1 2c2 ·· 2cm−1 2cm 0 0χ [22] 2 2c2 2c4 ·· 2c2m−2 2c2m 0 0· · · · · · · · · · ·· · · · · · · · · · · · ·χ [2m−1] 2 2cm−1 2c2m−2 ·· 2c

(m−1)2 2cm(m−1) 0 0

Page 68: Pierre Ramond Group Theory a Physicists Survey 2010

56 Finite groups: representations

The doublet representations are of the same form as in the odd case. The Kroneckerproducts of the four one-dimensional 1 and 1r, r = 1, 2, 3 irreps and (m − 1)doublets 2k, k = 1, 2, . . . , (m − 1) are tediously worked with the results

1r × 1r = 1,

1r × 1s = 1t r, s, t = 1, 2, 3, r �= s �= t,

11 × 2k = 2k,

12 × 2k = 13 × 2k = 2|m−k|.

The Kronecker products of two different doublets satisfy more complicated rules:

2j × 2k = 2j−k + 2j+k, j > k �= (m − j)

2j × 2m−j = 12 + 13 + 2|m−2j|, j �= 2m.

The product of the same doublet is simpler

2k × 2k = 1 + 11 + 22k, k �= m/2,

For m even, the “half-way" doublet satisfies its own Kronecker,

2m/2 × 2m/2 = 1 + 11 + 12 + 13.

The maximal subgroup of D2m+1 is Z2m+1, which rotates the (2m+1) vertices. D2m

has two maximal subgroups, Z2m which rotates the 2m vertices of the polygons, butalso Dm which is the symmetry of the m-gon that is inscribed by m vertices of theoriginal 2m vertices (skipping one). The two ways to choose such an m-gon yieldtwo equivalent embeddings. Finally we note that the commutator subgroup of D2m

is Zm , that of D2m+1 is Z2m+1, indicating that the dihedral groups are not simple.

3.9 Q2n character table

The dicyclic groups Q2n of order 4n have the presentation < a, b | a2n = e, an =b2 ; bab−1 = a−1 >. They have four one-dimensional irreps: when the generatorscommute with one another, the presentation reduces to a2n = e, an = b2, a = a−1.Since a is its own inverse, and the (2n)th root of unity, it can only be ±1. Inaddition, b is a fourth root of unity, with four possible values, ±1,±i . The finalconstraint an = b2 amounts to (±1)n = b2. This leaves the following solutions:

– n even: four one-dimensional irreps, a = ±1, b = ±1;– n odd: four one-dimensional irreps: a = 1, b = ±1 and a = −1, b = ±i .

It is just as easy to construct two-dimensional irreps. To wit, choose a basiswhere a is diagonal

a =(ηk 00 ηl

), b =

(r st u

),

Page 69: Pierre Ramond Group Theory a Physicists Survey 2010

3.9 Q2n character table 57

where η2n = 1, k, l range over 1, 2, . . . , (2n−1), and r, s, t, u are to be determined.The presentation yields four equations

(1 − η2k) r = 0; (1 − η2l) u = 0; (1 − ηk+l) s = 0; (1 − ηk+l) t = 0.

If r, s, t, u are all different from zero, we are left with k = l = n and a becomes ±the unit matrix: this case reduces to one-dimensional irreps.

There is only one type of non-trivial solution to these equations (s, t �= 0, r =u = 0); the case (r, u �= 0, s = t = 0) yields only the previously obtainedone-dimensional irreps.

Setting (s, t �= 0, r = u = 0) requires k = −l, as well as

b2 = an → st = ηnk = η−nk = (−1)k,

so that the generators become

a =(ηk 00 η−k

), b =

(0 1

(−1)k 0

),

where we have set s = 1 with a harmless similarity transformation. For k = n, thisrepresentation reduces to two singlets. Also, the representations for k = (n + 1),(n + 2), . . . , (2n − 1) are the same as those for k = 1, 2, . . ., (n − 1), after a trivialrelabeling of the matrix elements. Hence we have (n − 1) independent doubletirreps. By checking the dimension of the regular representation,

4n = 1 + 1 + 1 + 1 + (n − 1)22,

we conclude that Q2n contains no more irreps. It follows that we have (n−1)+4 =(n + 3) different classes.

The character tables are different depending on whether n is even or odd. For nis even, it is given by the following.

Q2n C1 2C2(a) 2C3(a2) ·· 2Cn(an−1) Cn+1(an) nCn+2(b) nCn+3(ba)

χ [1] 1 1 1 ·· 1 1 1 1χ [11] 1 1 1 ·· 1 1 −1 −1χ [12] 1 −1 1 ·· (−1)n−1 (−1)n 1 −1χ [13] 1 −1 1 ·· (−1)n−1 (−1)n −1 1χ [21] 2 2c1/2 2c1 ·· 2c(n−1)/2 2cn/2 0 0χ [22] 2 2c1 2c2 ·· 2cn−1 2cn 0 0· · · · · · · · · · ·· · · · · · · · · · · · ·χ [2n−1] 2 2c(n−1)/2 2cn−1 ·· 2c

(n−1)2/22cn(n−1)/2 0 0

Page 70: Pierre Ramond Group Theory a Physicists Survey 2010

58 Finite groups: representations

It is the same as the character table of D2n , so that the Kronecker product of itsirreps is also the same.

When n is odd, the character table is slightly different.

Q2n C1 2C2(a) 2C3(a2) ·· 2Cn(an−1) Cn+1(an) nCn+2(b) nCn+3(ba)

χ [1] 1 1 1 ·· 1 1 1 1χ [11] 1 1 1 ·· 1 1 −1 −1χ [12] 1 −1 1 ·· (−1)n−1 (−1)n i −iχ [13] 1 −1 1 ·· (−1)n−1 (−1)n −i iχ [21] 2 2c1/2 2c1 ·· 2c(n−1)/2 2cn/2 0 0χ [22] 2 2c1 2c2 ·· 2cn−1 2cn 0 0· · · · · · · · · · ·· · · · · · · · · · · · ·χ [2n−1] 2 2c(n−1)/2 2cn−1 ·· 2c

(n−1)2/22cn(n−1)/2 0 0

The products of its irreps are given by

11 × 11 = 1, 12 × 12 = 13 × 13 = 11,

11 × 12 = 13, 11 × 13 = 12, 12 × 13 = 1,

11 × 2k = 2k, 12 × 2k = 13 × 2k = 2(n−k),

2k × 2k = 1 + 11 + 22k,

2j × 2k = 2j−k + 2j+k ( j + k) �= n, j > k,

2j × 2(n−j) = 12 + 13 + 2|2j−n|,

with the identification 2j = 22n−j for j = 1, 2, . . . , (n − 1).None of the dicyclic groups is simple, since Zn is the commutator subgroup

of Q2n .We are fortunate to be able to build these character tables and representation

in such a straightforward manner. In order to do the same for more complicatedgroups, new machinery and concepts need to be introduced.

3.10 Some semi-direct products

We have seen that the construction of the representations of the dihedral group wasfacilitated by its semi-direct product structure. In this section, we further elaborateon semi-direct products.

We begin by introducing the holomorph of a group as the semi-direct productof a group with its automorphism group (the group which maps its multiplicationtable to itself). Take the cyclic groups as an example. The holomorph of a cyclic

Page 71: Pierre Ramond Group Theory a Physicists Survey 2010

3.10 Some semi-direct products 59

group Hol(Zn) depends on n. When n = p is prime, Aut Zp = Zp−1, with mul-tiplication defined mod(p). For instance, the elements of Aut Z7 are the integers(1, 2, 3, 4, 5, 6); multiplication is mod(7), such that 5 ·4 = 20 = 6, etc. This yieldsthe family of semi-direct products

Hol(Z p) = Zp � Zp−1, (3.127)

of order p(p − 1). We use the convention that the group acted upon appears first.The second should be called the actor group.

When n = 2, 4, or n = pm, 2pm , with p �= 2 is prime, Aut Zn is also cyclic,with as many elements as the number of integers less than n with no integersin common (coprime): Aut Z6 is the cyclic group Z2, with two elements 1, 5,coprimes to 6. In general, the number of coprimes of n is given by Euler’s totientφ(n) function. Hence we can build the semi-direct product groups

Hol(Zn) = Zn � Zφ(n), n = 2, 4, pm, 2pm p prime �= 2. (3.128)

Low-lying examples are

Hol(Z5) = Z5 � Z4 : < a5 = b4 = e | b−1ab = a3 >, (3.129)

Hol(Z7) = Z7 � Z6 : < a7 = b6 = e | b−1ab = a5 > . (3.130)

In other cases, the automorphism group need not be cyclic: Aut Z8 is D2 = Z2 ×Z2, Klein’s Vierergruppe, yielding the order 32 holomorph

Hol(Z8) = Z8 � (Z2 × Z2) :< a8 = b2 = c2 = e | b−1ab = a−1 ; c−1ac = a5 > . (3.131)

In all cases, Aut Zn contains φ(n) elements, with the multiplication table definedmodulo n.

There are many other ways to build semi-direct products of two cyclic groups, asthe action of one cyclic group on the other may not be unique. Consider a specialcase: for any two integers m < n, let a, an = e generate Zn , and b, bm = e generateZm , and define the action of Zm on Zn by conjugation

b ab−1 = ar .

It follows that

b2 ab−2 = b ar b−1 = (b a b−1)r = ar ·r ,

so that eventually,

bm ab−m = a = arm,

Page 72: Pierre Ramond Group Theory a Physicists Survey 2010

60 Finite groups: representations

which requires the necessary condition

rm = 1 + l n,

where l is integer. This enables us to construct the semi-direct product of order nm

Zn �r Zm : < a, b | an = e, bm = e ; bab−1 = ar >, rm = 1 + ln, (3.132)

which may have several solutions. For example, this yields several groups oforder 32:

Z8 �5 Z4 : < a, b | a8 = e, b4 = e ; bab−1 = a5 >, 54 = 625 = 1 mod (8),

Z8 �−1 Z4 : < a, b | a8 = e, b4 = e ; bab−1 = a−1 >, (−1)4 = 1 mod (8),

Z8 �3 Z4 : < a, b | a8 = e, b4 = e ; bab−1 = a3 >, 34 = 81 = 1 mod (8),

showing different semi-direct product constructions, and

Z16 �7 Z2 : < a, b | a16 = e, b2 = e ; bab−1 = a7 >, 72 = 49 = 1 mod (16).

The infinite family of dihedral groups can be constructed in this way

Dn = Zn �−1 Z2, (3.133)

corresponding to r = −1, r2 = 1.Note that although the dicyclic groups Q2n are not in general semi-direct

products, there are special isomorphisms such as Q6 = Z3 �3 Z4.Consider semi-direct products of the form (Zr × Zs) � Zt , where r, s, t are

integers. The action of Zt on the (Zr ×Zs) may be more complicated, as it can acton either of the product groups or both. Let us take as an example the action of Z2

on (Z4 × Z2). Let a ∈ Z4, a4 = e, and b ∈ Z2, b2 = e with ab = ba generate thedirect product. The action of the first Z2 on a must send it into an element of order4 in Z4 × Z2. There is only one such element a3b as (a3b)4 = e. Hence we candefine the action as

Z2 : a → a3b, b → b,

or it can act on the Z2 in the direct product as

Z2 : a → a, b → a2b,

since (a2b)2 = e. If the actor group is generated by c, c2 = 1, these correspondrespectively to

c−1ac = a3b, c−1bc = b,

and

c−1ac = a, c−1bc = a2b.

Page 73: Pierre Ramond Group Theory a Physicists Survey 2010

3.11 Induced representations 61

It follows that there are two semi-direct products involving these groups. The rich-ness of the semi-direct product construction is evidenced by looking at the ordersixteen groups, where all but one of the nine non-Abelian groups can be expressedeither as a direct or semi-direct product

Q8,D8(= Z8 �−1 Z2),Q × Z2,D4 × Z2,Z8 �3 Z2,

Z8 �5 Z2,Z4 �3 Z4, (Z4 × Z2) � Z2, (Z4 × Z2) �′ Z2.

The last two groups correspond to the action of the actor group on the first andsecond group we just discussed.

3.11 Induced representations

This is a method which constructs representations of a group using the representa-tions of its subgroups. It is particularly well-suited for groups which are semi-directproducts. Let H be a subgroup of G of index N

G ⊃ H. (3.134)

Suppose we know a d-dimensional representation r of H, that is for every h ∈ H,

h : | i 〉 → | i ′ 〉 = M(h)[r]i ′ j | j 〉, (3.135)

acting on the d-dimensional Hilbert space H, spanned by | i 〉, i = 1, 2, . . . , d. Wewant to show how to construct a representation of G, starting from the known rep-resentation on H. It is called the induced representation of G by the r representationof H.

We begin by considering the coset made up of N points,

H ⊕ g1H ⊕ g2H ⊕ · · · ⊕ gN−1H, (3.136)

where gk are elements of G not in H. Consider the set of elements

G × H : ( g, | i 〉 ); g ∈ G, | i 〉 ∈ H. (3.137)

In order to establish a one-to-one correspondence with the coset, we assume thatwhenever the G group element is of the form

g = gkha, (3.138)

we make the identification(gkha, | i 〉) =

(gk, M(ha)

[r]i j | j 〉

), (3.139)

Page 74: Pierre Ramond Group Theory a Physicists Survey 2010

62 Finite groups: representations

so that both g and gh are equivalent in the sense that they differ only by reshufflingH. In this way, we obtain N copies of H, the Hilbert space of the r representationof H, one at each point of the coset, which we take to be

( e, H ), ( g1, H ), ( g2, H ), · · · , ( gN−1, H ). (3.140)

The action of G on this set is simply group multiplication

g : ( gk, | i 〉 ) → ( ggk, | i 〉 ), g ∈ G, k = 1, 2, · · · , N . (3.141)

Suppose that g = ha , the coset decomposition tells us that

hagk = glhb, (3.142)

where gl and hb are uniquely determined. Hence,

ha : (gk, | i 〉) → (

hagk, | i 〉) = (glhb, | i 〉) =

(gl, M(hb)

[r]i j | j 〉

). (3.143)

Similarly, when g = gl , the product gl gk is itself a group element and can berewritten uniquely as

gl gk = gmhb, (3.144)

for some gm and hb. It follows that

gl : ( gk, | i 〉 ) → ( gl gk, | i 〉 ) = ( gmhb, | i 〉 )=

(gm, M(hb)

[r]i j | j 〉

). (3.145)

We have shown that these N copies of the vector space are linearly mapped into oneanother under the action of G. The action of G is thus represented by a (d N × d N )

matrix. It may or may not be irreducible.In the special case when H is Abelian, its irreducible representations are one-

dimensional, and the induced representation on G by an irrep of H is in terms of(N × N ) matrices. This suggest that we think of the induced representation asmatrices in block form, (N × N ) matrix, whose elements are themselves (d × d)matrices.

For the more mathematically minded, we define N mapping (called sections)which single out N vectors in H:

fl : ( gk, | i 〉 ) −→⎧⎨⎩

| vk 〉, l = k,

0, l �= k.(3.146)

This defines a basis, and we can arrange these N vectors into a column, which isbeing acted on by the induced representation.

Page 75: Pierre Ramond Group Theory a Physicists Survey 2010

3.11 Induced representations 63

As an example, consider the group of order 3n2

�(3n2) = (Zn × Zn) � Z3, (3.147)

generated by a ∈ Z3, c ∈ Zn and d ∈ Zn , with the presentation

< a, c, d | a3 = cn = dn = e, cd = dc, aca−1 = c−1d−1, ada−1 = c > .

(3.148)Let us construct the representation induced by the one-dimensional representationof H = Zn × Zn , for which

c = ηk; d = ηl, ηn = 1, (3.149)

acting on complex numbers z. The coset contains just three points, and we considerthe action of G on

( e, z ), ( a, z ), ( a2, z ). (3.150)

Then

a( e, z ) = ( a, z ), a( a, z ) = ( a2, z ), a( a2, z ) = ( e, z ), (3.151)

so that a is represented by the (3 × 3) permutation matrix

a =⎛⎝0 1 0

0 0 11 0 0

⎞⎠ . (3.152)

Next, we consider the action of c

c( e, z ) = ( c, z ) = ( e, cz ) = ( e, ηk z ) = ηk( e, z ). (3.153)

Using

ca = ad; ca2 = a2c−1d−1, (3.154)

derived from the presentation, we find

c( a, z ) = ( ca, z ) = ( ad, z ) = ( a, dz ) = ( e, ηl z ) = ηl( e, z ), (3.155)

as well as

c( a2, z ) = ( ca2, z ) = ( a2c−1d−1, z ) = ( a2, c−1d−1z )

= ( e, ηk ηl z ) = η−k−l( a2, z ). (3.156)

Similarly, the action of d can be derived, using

da = c−1d−1; da2 = a2c. (3.157)

Page 76: Pierre Ramond Group Theory a Physicists Survey 2010

64 Finite groups: representations

The above show that c and d are represented by the diagonal matrices

c =⎛⎝ηk 0 0

0 ηl 00 0 η−k−l

⎞⎠ , d =⎛⎝ηl 0 0

0 η−k−l 00 0 ηk

⎞⎠ , (3.158)

completing the induced (3 × 3) representation.In general, the induced representation is reducible, but when one builds it out of

a subgroup that has a low index, one has a better chance of obtaining an irrep fromthis procedure. This is the case for this example.

Alas, it is not so easy to work out the character tables for more complicatedgroups. One runs out of tricks, and must rely on numerical methods. Fortunatelyfor us, character tables and other useful group properties have been tabulated inthe Atlas of Simple Groups for the simple groups, in book form (Oxford UniversityPress) and also on the internet: http://brauer.maths.qmul.ac.uk/Atlas.

A word of caution about the ATLAS’s dot notation. By A4 · 2, they mean thegroup S4 ⊃ A4, and derive specific matching rules for their irreps A5 → S4:1 → 1 + 11, 11 + 11 → 2, and 3 → 31 + 32.

On the other hand, the group 2 · A4 is the binary extension or double cover ofA4, to be discussed in detail later, but it does not contain A4 as a subgroup; ratherA4 is the quotient group of the binary tetrahedral group with Z2.

It is useful to tabulate group properties in a convenient form that general-izes the tables of Thomas and Wood. In Appendix 1, we gather in one placeimportant information (presentation, classes, character table, Kronecker products,embeddings, representations) for groups which appear ubiquitously in physicalapplications.

3.12 Invariants

In physics, invariance manifests itself when the Hamiltonian (or action in fieldtheory) can be expressed in terms of group invariants. Typically, the Hamilto-nian is a function of the dynamical variables which transform as some repre-sentation of the group. Mathematically, this amounts to constructing invariantsout of polynomials built out of group representations. Invariants appear when-ever the (multiple) Kronecker products of representations contains 1, the singletirrep.

For instance, the A4 Kronecker products show two quadratic invariants in theproduct of two irreps

[3 × 3]singlet, 11 × 11. (3.159)

Page 77: Pierre Ramond Group Theory a Physicists Survey 2010

3.12 Invariants 65

There are more ways to form cubic invariants. We can form cubic invariants in fourways [

(3 × 3)sym3 × 3

]singlet ,

[(3 × 3)antisym

3 × 3]

singlet,

[(3 × 3)sym

11× 11

]singlet

,[(3 × 3)sym

11× 11

]singlet

,

where the representation subscript denotes the projection of the product on thatrepresentation: 3 appears in both the symmetric and antisymmetric product of twotriplets. The second cubic invariant requires at least two triplets.

The same reasoning leads to four quartic invariants built out of triplets alone:[(3 × 3)sym

3 × (3 × 3)sym3

]singlet ,

[(3 × 3)antisym

3 × (3 × 3)sym3

]singlet

,

[(3 × 3)antisym

3 × (3 × 3)antisym3

]singlet

,[(3 × 3)sym

11× (3 × 3)sym

11

]singlet

,

but they are not all independent. This procedure, while very useful to identifyinvariants in the product of representations, produces too many redundancies.How do we find out how many of those are independent? Furthermore, theiractual construction requires detailed knowledge encoded in the Clebsch–Gordandecompositions.

The identification and construction of finite group invariants is clearly aformidable task, but it is tractable. The determination of invariants constructed outof several irreducible representations, can be somewhat alleviated by the fact thatall irreps of most groups can be generated by the Kronecker products of its fun-damental irrep. There are exceptions to this rule: in some cases, these productsgenerate different irreps always in the same combination, making it impossible todistinguish them by this construction, but in practice a good first step is to find allinvariants made out of the fundamental irrep.

Consider a finite group G of order N , with n irreducible representations ra, withr1 = 1 the singlet. We express the k-fold symmetric Kronecker product of anyirrep as a sum of irreps as

(ra × ra × · · · × ra)s︸ ︷︷ ︸k

=∑

b

N [k](rb; ra) rb. (3.160)

The integer coefficient N [k](rb; ra) denotes the number of rb irreps in the k-foldproduct of ra irreps. Consider Molien’s remarkable generating function

M(rb; ra; λ) = 1

N

n∑i=1

ni

χ[rb]i

det(

1 − λA[ra]i

) , (3.161)

Page 78: Pierre Ramond Group Theory a Physicists Survey 2010

66 Finite groups: representations

where i labels the class Ci of G, with ni elements. A[ra]i is any group element

in Ci expressed in the ra representation, and χ[rb]i are the characters of the rb

representation. Its power series

M(rb; ra; λ) =∞∑

k=0

N [k](rb; ra) λk, (3.162)

yields the desired coefficients. When we set rb to be the singlet irrep, the Molienfunction yields the number of possible invariants constructed out of one irreduciblerepresentation

M(1; r; λ) = 1

N

n∑i=1

ni

det(

1 − λA[r]i

) =∞∑

k=0

N [k] λk, (3.163)

where N [k] denotes the number of invariants of order k. What makes the Molienfunction particularly useful is that it can always be written in the form

M(1; r; λ) = (1 + ∑dk λ

k)

(1 − λa1)n1(1 − λa2)n2 · · · , (3.164)

where the numerator is a finite polynomial in λ, and the dk , ak and nk are pos-itive integers. The numerator yields dk invariants of order k, while expansion ofeach factor in the denominator generates nk invariants of order ak , etc. This infinitenumber of invariants can be expressed as products of a finite set of basic invariants,but the Molien expansion does not single them out. However, the subset of powersin the denominators ak which are smaller than those in the numerator determineinvariants of lesser order, and those must form the bulk of the basic invariants,but whenever they are comparable, the distinction becomes blurred. To makematters worse, invariants satisfy non-linear relations among themselves calledsyzygies.

In some cases, it can even happen that the invariants of order a1, a2, etc., fromthe denominator of the Molien function, do not satisfy syzygies among themselves.They are called free invariants. The powers in the numerator then refer to con-strained invariants which satisfy syzygies with the free invariants. Unfortunately,this neat distinction among invariants is of limited validity.

We apply this approach to the tetrahedral group A4, with four irreps, 1, 11, 11,

and 3. Using its character table, it is an easy matter to compute the Molien functionfor each. For the singlet irrep, we have of course

M(1; 1; λ) = 1

1 − λ, (3.165)

Page 79: Pierre Ramond Group Theory a Physicists Survey 2010

3.13 Coverings 67

corresponding to the one-dimensional trivial invariant. Applied to 11, the otherone-dimensional representations, we find

M(1; 11; λ) = M(1; 11; λ) = 1

1 − λ3, (3.166)

yielding one invariant of cubic order with the 11. Calling its one component z, itmeans that z3 is invariant, as one can easily check.

For the triplet, a straightforward computation of the Molien function yields

M(1; 3; λ) = 1 + λ6

(1 − λ2)(1 − λ3)(1 − λ4). (3.167)

We infer three free invariants, of order 2, 3, and 4, and one sixth-order constrainedinvariant. There is no easy way to construct these invariants, so we simply statetheir expression in terms of the three real coordinates xi , i = 1, 2, 3 spanning thetriplet irrep. The quadratic invariant is the length of the vector, as in SO(3)

〈2〉 = (x2

1 + x22 + x2

3

). (3.168)

The free cubic and quartic invariants are found to be

〈3〉 = x1x2x3, 〈4〉 = x41 + x4

2 + x43 , (3.169)

and the constrained sixth-order invariant is

〈6〉 = (x2

1 − x22

) (x2

2 − x23

) (x2

3 − x21

). (3.170)

Its square is related by one syzygy, to sums of polynomials in the free invariants:

4〈6〉2−2〈4〉3+108〈3〉4+〈2〉6+36〈4〉〈3〉2〈2〉−20〈3〉2〈2〉3+5〈4〉2〈2〉2−4〈4〉〈2〉4 = 0,(3.171)

which gives an idea of the level of complication.

3.13 Coverings

We have noted the close similarity between the dihedral and dicyclic groups, indeedtheir presentations are almost the same

Dn : < a, b | an = b2 = e ; bab−1 = a−1 >,

Q2n : < a, b | a2n = e ; an = b2 ; bab−1 = a−1 > .

In both, an and b2 are equal to one another, but in the dicyclic group, they are oforder two. Hence Q2n contains a Z2 subgroup generated by b2, called the Schurmultiplier. In symbols (following the ATLAS), we write

Page 80: Pierre Ramond Group Theory a Physicists Survey 2010

68 Finite groups: representations

Q2n = 2 · Dn,

and we say that Q2n is the double cover of Dn .This is a special case of a general construction. Let G be a finite group gen-

erated by a set of elements ai (letters), with presentation defined by a bunch ofrelations wα(a1, a2, . . . ) = e (words). We define a new group G ′ where the wordsare no longer equal to the unit element, and commute with the letters, that is,wα(a1, a2, . . . ) �= e, but such that they commute with the generators of G, thatis, wα ai = ai wα. The wα generate a subgroup C of G ′, which commutes with G.G ′ is called the covering group of G by a central subgroup C with G = G ′/C . It isclearly not simple, and the Schur multiplier is the intersection of C with its com-mutator subgroup. In the dicyclic group, there is only one w of order two. Dicyclicgroups have spinor representations while the dihedral groups does not. For physi-cists, this is a familiar situation: the spinor representations are obtained by goingfrom SO(3) to SU (2) where SU (2) is the cover of SO(3).

Page 81: Pierre Ramond Group Theory a Physicists Survey 2010

4

Hilbert spaces

Physics happens in Hilbert spaces. Physicists like to describe Hilbert spaces interms of their basis, in such a way that any vector can be represented as a lin-ear combination of basis vectors with complex coefficients. The number of basisvectors may be finite or infinite, with the basis described in terms of discrete orcontinuous variables defined over various intervals. All can be constructed out ofproducts of Hilbert spaces whose basis vectors are labeled by only one variable,continuous or discrete. In the continuous case, the range of the variable may beunbounded or bounded. We start with the simpler finite cases.

4.1 Finite Hilbert spaces

Since discrete groups are irreducibly represented by unitary matrices in finiteHilbert spaces, their study requires special attention. A D-dimensional Hilbertspace has a basis made up of D kets { | 1 〉, | 2 〉, . . . , | D − 1 〉, | D 〉 }. Its dualis made up of bras 〈 j |, j = 1, 2, . . . , D. They satisfy completeness andorthonormality, the hallmark of Hilbert spaces, that is

D∑j=1

| j 〉〈 j | = 1, 〈 j | k 〉 = δ j k . (4.1)

In this space, one can build D2 linear operators, of the form | j 〉〈 k |, whichgenerates | j 〉 from | k 〉. By completeness, one of them

D∑j=1

| j 〉〈 j |, (4.2)

reduces to the identity transformation, leaving (D2 − 1) non-trivial linear oper-ators. By taking appropriate linear combinations, all can be made into hermitianoperators:

69

Page 82: Pierre Ramond Group Theory a Physicists Survey 2010

70 Hilbert spaces(| k 〉 〈 j | + | j 〉 〈 k |), i(| k 〉 〈 j | − | j 〉 〈 k |), for j �= k, (4.3)

and the (D − 1) diagonal combinations which we can write in the physics way as(| 1 〉 〈 1 | − | 2 〉 〈 2 |), (| 1 〉 〈 1 | + | 2 〉 〈 2 | − 2 | 3 〉 〈 3 |),. . . . . .(| 1 〉 〈 1 | + | 2 〉 〈 2 | + · · · + | D − 1 〉 〈 D − 1 | − (D − 1)| D 〉〈 D |).

It might be inferred that a D-dimensional Hilbert space can sustain at least (D2−1)operators which can be used to generate an infinite number of linear transfor-mations. We shall see later that in some special cases, there can be even moresymmetry generators.

4.2 Fermi oscillators

A favorite way to construct finite Hilbert spaces is by means of n operators bi andtheir hermitian conjugates b†

i , i = 1, 2, . . . , n. They describe fermionic harmonicoscillators, which satisfy canonical anticommutation relations

{b i , b†j } = δ i

j , {b i , b j } = {b†i , b†

j } = 0, (4.4)

where the anticommutator is defined as

{A, B} ≡ A B + B A. (4.5)

We construct a Hilbert space from the vacuum state |� 〉, and its dual 〈� |, subjectto the n annihilation conditions

b i |� 〉 = 0, 〈� |b†i = 0. (4.6)

and normalization

〈� |� 〉 = 1. (4.7)

Using the anticommutation relations, we see that the n states

| i 〉 = b†i |� 〉, (4.8)

are normalized and orthogonal to the vacuum state

〈 i | j 〉 = δ ij , 〈 i |� 〉 = 0. (4.9)

The rest of the Hilbert space is spanned by the states

| [i1 i2 · · · ik] 〉 = b†i1

b†i2

· · · b†ik|� 〉, (4.10)

where k is limited to n since the square of any b†i is zero. In addition, they are anti-

symmetric under any i j ↔ il interchange, and thus in one-to-one correspondence

Page 83: Pierre Ramond Group Theory a Physicists Survey 2010

4.2 Fermi oscillators 71

with totally antisymmetric tensors (called forms by some). There are

(nk

)such

states which generate a Hilbert space with dimension 2n , given by the binomialformula

n∑k=0

(nk

)= 2n. (4.11)

This very rich finite-dimensional Hilbert space sustains many algebraic struc-tures, in particular continuous and finite groups. We postpone the discussion ofthe continuous groups, and focus on the finite groups.

One such group is the finite Dirac group (see Lomont [15]). Arrange thefermionic harmonic oscillators into hermitian combinations

γ j = b j + b†j , γ n+ j = i(b j − b†

j ), (4.12)

which we write in a unified way as γa , a = 1, 2, . . . , 2n. These obey the Dirac orClifford algebra

{ γ a, γ b } = 2δ ab. (4.13)

Now consider elements of the form [15]

±γl11 γ

l22 · · · γ l2n

2n , (4.14)

where lk = 0, 1, since (γ1)2 = (γ2)

2 = · · · = 1. There are 22n+1 elements of thiskind, including 1 and −1. Since the γa are their own inverses, all elements haveinverses as well; for example, (γ1γ4)

−1 = γ4γ1 = −γ1γ4. Closed under multipli-cation, they form a finite group of order 22n+1, called the Dirac group. For n = 2,its 32 elements are just

± 1, ±γa, ±γaγb, ±γaγbγc, ±γ1γ2γ3γ4, (4.15)

with a < b < c. Since for a �= b,

γ aγ bγ−1a = −γ b, (4.16)

we deduce that all group elements, except for 1 and −1, split into classes eachcontaining the element and its negative: (γa,−γa), (γaγb,−γaγb), a �= b, etc. TheDirac group has thus 2n + 1 classes, and as many irreducible representations.

Its only normal subgroup, Z2 = (1,−1), is also its commutator (derived) sub-group, so that it has 22n one-dimensional irreps. The dimension d of the extra irrepis deduced from the order formula

22n+1 = 22n + d2, (4.17)

so that d = 2n . In this representation, the γa are simply the well-known (2n × 2n)

Dirac matrices.

Page 84: Pierre Ramond Group Theory a Physicists Survey 2010

72 Hilbert spaces

Since its elements split into two types, those which square to 1, and thosewhich square to −1, the reality of the representation can easily be determined bycomputing

1

22n+1

∑g

χ(g2) = 1

22n+1

[n+χ(1) + n−χ(−1)

], (4.18)

where n± is the number of elements which square to ±1, and χ(±1) are thecharacters. We find

n+ = 22n + i n(n−1) 2n, n− = 22n+1 − n+ = 22 − i n(n−1) 2n, (4.19)

so that they add up to the order. Hence

1

22n+1

∑g

χ(g2) = i n(n−1). (4.20)

The representation is real whenever n is a multiple of four; otherwise it ispseudoreal.

4.3 Infinite Hilbert spaces

A Hilbert space is best described by its basis vectors, called kets. Consider firsta Hilbert space with an infinite-dimensional basis, which we choose to label withthe positive integers, | n 〉, where n = 0, 1, 2, 3, . . . ,∞. Any vector (ket) can bewritten as a linear combination

|� 〉 =∞∑

n=0

zn | n 〉, (4.21)

where the zn are arbitrary complex numbers. In order to associate numbers to thesekets, we need to introduce the dual vector space, which has a mirror basis madeof bras, 〈 n |, with n = 0, 1, 2, 3, . . .. An arbitrary bra can be written as a linearcombination of bras with arbitrary complex coefficients

〈� | =∞∑

n=0

wn 〈 n |, (4.22)

where wn are complex numbers. In particular, the bra associated with the state �

is given by

〈� | =∞∑

n=0

zn 〈 n |, (4.23)

Page 85: Pierre Ramond Group Theory a Physicists Survey 2010

4.3 Infinite Hilbert spaces 73

where the bar denotes complex conjugation. The vector space and its dual areneeded to map bras and kets into complex numbers, which “Di rac” calls bra kets:

〈� | |� 〉 → 〈� |� 〉 = complex number. (4.24)

It is particularly convenient to choose an orthonormal basis in the sense that

〈 n | m 〉 = δm n. (4.25)

Such a basis also satisfies the completeness relation

∞∑n=0

| n 〉〈 n | = 1. (4.26)

A vector space over the complex numbers with its dual space defined so as to havea complete orthonormal basis is called a Hilbert space. Then

〈� |� 〉 =∞∑

n=0

|zn|2, (4.27)

is the norm of the state vector represented by �; it is positive definite. In this spacedwell an infinite number of operators which can be written in the form

C =∑n,m

c nm | m 〉〈 n |, (4.28)

where the c nm are complex numbers. We define the hermitian operator

C† ≡∑n,m

c mn | n 〉〈 m |, (4.29)

in such a way that the action of C on a ket is the same as that of C† on itscorresponding bra. That is

〈� | C |� 〉 = 〈� |(

C |� 〉)

= 〈� |�′ 〉,=

(〈� | C

)|� 〉 = 〈� ′ |� 〉,

where

|�′ 〉 = C |� 〉, 〈� ′ | = 〈� | C. (4.30)

Then,

|� ′ 〉 = C† |� 〉. (4.31)

It follows that if C is a hermitian operator,

C = C† → c nm = c m

n : (4.32)

Page 86: Pierre Ramond Group Theory a Physicists Survey 2010

74 Hilbert spaces

any hermitian operator is a matrix which is equal to itself when its elements aretransposed and conjugated. We can view any operator as an infinite-dimensionalmatrix which acts on kets from the left to produce other kets, and on bras from theright to produce other bras. Of special interest are the operations that preserve thenorm of the states. Consider the transformation

|� 〉 → |� ′ 〉 = U |� 〉, (4.33)

such that

〈� ′ |� ′ 〉 = 〈� |� 〉. (4.34)

If this is true for any state �, it follows that

U† U = 1, → U−1 = U†. (4.35)

The operator U is said to be unitary. Now let us write it as

U = eM , (4.36)

where M is another operator. The condition means that

M† = −M, (4.37)

and M is an antihermitian operator, which can always be written as i times ahermitian operator.

While there is an infinite number of operators in an infinite-dimensional Hilbertspace, some particular ones have proven to be of great mathematical and physicalinterest, such as

a ≡∞∑

m=0

√m + 1 | m 〉〈 m + 1 |, (4.38)

which changes each basis vector to its lower sibling and annihilates the zero state

a | n 〉 = √n | n − 1 〉, a | 0 〉 = 0. (4.39)

With its hermitian conjugate,

a† ≡∞∑

m=0

√m + 1 | m + 1 〉〈 m |, (4.40)

it forms the simple commutation relation

[ a, a† ] ≡ a a† − a† a =∞∑

n=0

| n 〉〈 n | = 1. (4.41)

Page 87: Pierre Ramond Group Theory a Physicists Survey 2010

4.3 Infinite Hilbert spaces 75

It follows that any ket is viewed as the action of the creation operator a† on the“vacuum” state | 0 〉:

| n 〉 = (a†)n

√n! | 0 〉. (4.42)

It is well-known that we can change basis to a continuous variable. Namely,introduce the hermitian combinations

X ≡ 1√2(a† + a), P ≡ i√

2(a† − a), (4.43)

so that

[ X, P ] = i, (4.44)

is the Heisenberg algebra. With the proper units, X can be interpreted as a positionoperator whose eigenvalue denotes a coordinate which ranges from −∞ to +∞.We define the eigenket | x 〉 through

X | x 〉 = x | x 〉. (4.45)

They form a new basis which satisfies the normalization

〈 x | x ′ 〉 = δ(x − x ′), (4.46)

and completeness relation ∫ +∞

−∞dx | x 〉〈 x | = 1. (4.47)

The relation between the two bases is defined through

| x 〉 =∞∑

n=0

Kn(x) | n 〉, (4.48)

where

Kn(x) = 〈 n | x〉, (4.49)

is a function in the continuous variable x . It can be written in the form

Kn(x) = Hn(x) e−x2/2, (4.50)

where Hn is the Hermite polynomial of degree n. Completeness and orthogonalityof our basis translates in the same way for the Hn polynomials

Page 88: Pierre Ramond Group Theory a Physicists Survey 2010

76 Hilbert spaces∫ +∞

−∞dx e−x2

Hn(x)Hm(x) = δn m, (4.51)

+∞∑n=0

Hn(x) Hn(y) = e−(x2+y2)/2 δ(x − y). (4.52)

We conclude this section with two remarks and a caveat.

First remark. The physical relevance of operators in Hilbert space depends on theirstability under time evolution. In Quantum Mechanics, the state of a physical sys-tem at any time is represented by a state vector |� 〉t . At a later time, this vector hasevolved into another one |� 〉t ′ . Time evolution is regarded as an infinite numberof infinitesimal steps from t to t + δt . The Hamiltonian is the operator which takesthe state ket into its infinitesimal future

|� 〉t+δt ≈ (1 + i H δt)|� 〉t . (4.53)

The Hamiltonian H is a hermitian operator which generates time translations. Itsexponentiation generates a finite time evolution

ei H(t−t ′) |� 〉t ′ = |� 〉t , (4.54)

so that the operator

U(t − t ′) ≡ ei H(t−t ′), (4.55)

is unitary

U† U = U U† = 1. (4.56)

Second remark. In Quantum Mechanics, any Hilbert space operator can be viewedas generating some transformation; the Hamiltonian which generates time transla-tion, is often used to define a physical system. Let A and B be two operators, andconsider their commutator

[ A, B ].It has two different meanings: the infinitesimal change in the operator A due to atransformation generated by B, and that of B due to A. For instance if A = H ,our commutator means either the change in B under time evolution, or the changein H under the transformation generated by B. Thus if the commutator is zero itmeans that H is invariant under the B transformation and that B is invariant undertime evolution: it represents a constant of the motion. For example if H commuteswith the angular momentum operators (which generate space rotations), it meansthat the energy (eigenvalue of H ) is invariant under rotations, and that angularmomentum is a constant of the motion (generated by the Hamiltonian).

Page 89: Pierre Ramond Group Theory a Physicists Survey 2010

4.3 Infinite Hilbert spaces 77

Caveat. We described a Hilbert space with a simple basis labeled by an unboundedcontinuous variable. A product of such Hilbert spaces yields a space described bymany such coordinates, which can then serve as a label for a point in space. Butwhat if the coordinates are not globally defined, as happens when there are topo-logical obstructions? Then one has to use coordinate patches for different regions,with complicated matching conditions. What if there are “holes” in the sense of sin-gular points? How does one treat the innards of black holes? The Universe is, wethink, represented by a state in Hilbert space, and the determination of its Hilbertspace is at the heart of physics. Such knowledge for a particular physical system isa crucial step towards the understanding of its dynamics. With the success of theStandard Model of particle physics, one may even speculate that the dynamics maywell be uniquely determined by symmetries. Our purpose at hand, however, is notto tackle such deep questions but to learn how to represent transformation groupsin Hilbert spaces.

Page 90: Pierre Ramond Group Theory a Physicists Survey 2010

5

SU (2)

Continuous symmetry operations, e.g. rotation about some axis, can be viewedas analytic functions of its parameters, in this case a continuous angle, definedmodulo 2π . All can be understood, except for topological obstructions, as stem-ming from the repeated application of one operator which generates (in this case)infinitesimal rotations about that axis. Generators of rotations about different axesdo not in general commute; rather they form a Lie algebra, which closes undercommutation. This chapter is devoted to the study of the most elementary but non-trivial Lie algebra. In the process we will introduce many concepts that will be ofuse in the study of arbitrary Lie algebras. This algebra ubiquitously lies at the rootof many areas of physics, and mathematics.

5.1 Introduction

The smallest non-trivial Lie algebra is realized in a two-dimensional Hilbert space,in which roam three linear operators:

T + ≡ | 1 〉〈 2 |; T − ≡ | 2 〉〈 1 |; T 3 ≡ 12

(| 1 〉〈 1 | − | 2 〉〈 2 |

). (5.1)

They satisfy an algebra under commutation

[ T +, T − ] = 2 T 3; [ T 3, T ± ] = ± T ±. (5.2)

Define as before hermitian operators

T 1 ≡ 1

2

(T − + T +

); T 2 ≡ i

2

(T − − T +

). (5.3)

In terms of which

[ T A, T B ] = iεABC T C , (5.4)

78

Page 91: Pierre Ramond Group Theory a Physicists Survey 2010

5.1 Introduction 79

A, B,C = 1, 2, 3, and where εABC is the Levi–Civita symbol. An algebra closedunder commutation is called a Lie algebra provided it also satisfies the Jacobiidentity, in this case[

T A,[T B, T C

]] + cyclic permutations = 0. (5.5)

It is trivially satisfied by our construction.This is our first Lie algebra. It is so ubiquitous that it has many names: math-

ematicians call it A1, B1, C1, and physicists call it SU (2), SO(3), Sp(2). Themeaning of these names will become evident later. Any polynomial function of theelements of a Lie algebra that commutes with the Lie algebra is called a Casimiroperator. In this case there is only one; it is quadratic and given by

C2 ≡ (T 1

)2 + (T 2

)2 + (T 3)2; [C2, T A

] = 0. (5.6)

Our explicit construction represents this algebra in a two-dimensional Hilbertspace, our first representation. We say that the Lie algebra SU (2) has a two-dimensional irrep, written as 2. We note that the generators satisfy the additionalrelation (T ±)2 = 0, which is specific to the representation, and not to the algebra.

It is not too difficult to work out all the irreducible unitary representations of thisalgebra. There are many ways of doing it, but we will use a technique which willgeneralize naturally to more complicated algebras.

First some remarks. (1) In Quantum Mechanics, states are labeled by theeigenvalues of a maximal number of commuting operators. Our algebra has twosuch operators, the Casimir operator C2, and any one of its three elements T A,A = 1, 2, 3. Let us take C2 and T 3. The algebra will be represented on eigenstatesof these two operators. (2) The Casimir operator is positive definite, which meansthat the spectrum of any of the T A is bounded, for a finite value of C2. Roughlyspeaking, we expect the largest and lowest values of any of the T A to be like ±√

C2.So we look for states | c ; m 〉 which satisfy

C2 | c ; m 〉 = c | c ; m 〉, T 3 | c ; m 〉 = m | c ; m 〉, (5.7)

where c and m are real since these operators are hermitian. We can use theremaining operators in the algebra to create new states. It is easily seen fromthe commutation relations that T + raises the value of T 3 by one unit, and doesnot change the value of C2. Hence the state T + | c ; m 〉 must be proportional to| c ; m + 1 〉 (since there are no other labels), so we set

T + | c ; m 〉 = d (+)m | c ; m + 1 〉, (5.8)

where d (+)m is to be determined later. We can generate an infinite set of states with

T 3 eigenvalues m,m + 1,m + 2, . . ., by repeated application of T +, but, as we

Page 92: Pierre Ramond Group Theory a Physicists Survey 2010

80 SU (2)

have remarked, the spectrum of T 3 is bounded, and this sequence must stop: thereis a state where T 3 assumes its maximum value, j :

T + | c ; j 〉 = 0; T 3 | c ; j 〉 = j | c ; j 〉. (5.9)

This state is called the highest weight state of the representation. We can easilywork out the value of the Casimir operator on the highest weight state, using thecommutation relations

C2 | c ; j 〉 =((

T 3)2 + (

T + T − + T − T +) /2)

| c ; j 〉,=

((T 3

)2 + [ T +, T − ]/2)

| c ; j 〉,=

((T 3

)2 + T 3)

| c ; j 〉 = j ( j + 1) | c ; j 〉,

so that c = j ( j + 1). The maximum eigenvalue of T 3 determines the Casimiroperator. We thus relabel the kets by j instead of c: | c ; m 〉 → | j ; m 〉.

Similarly, we can generate states of lower T 3 values by repeated applicationof T −, but this process must terminate as well, yielding a lowest weight state thatis annihilated by T −

T − | j ; k 〉 = 0; T 3 | j ; k 〉 = k | j ; k 〉, (5.10)

where k is in the sequence j, j − 1, j − 2, . . .. However, this state has the samevalue of C2, and thus

C2 | j ; k 〉 =((

T 3)2 + T − T +/2

)| j ; k 〉,

=((

T 3)2 + [ T −, T + ]/2

)| j ; k 〉,

=((

T 3)2 − T 3

)| j ; k 〉 = k (k − 1) | j ; k 〉,

so that

k(k − 1) = j ( j + 1). (5.11)

This quadratic equation has two solutions, k = j + 1, which is impossible since kis always less than j , and this leaves the only possibility, k = − j .

With little work we have found that any irrep of this algebra contains (2 j + 1),so that 2 j must be an integer, with T3 eigenvalues ranging from j to − j in integersteps. Physicists call j the total angular momentum, while mathematicians call 2 jthe Dynkin label of the representation. The Lie algebra of SU (2) has thereforean infinite number of irreps, one for each integer value of 2 j . The value j = 0corresponds to the trivial singlet representation, j = 1/2 is the famous spinorrepresentation, j = 1 describes the three states of a vector, etc.

Page 93: Pierre Ramond Group Theory a Physicists Survey 2010

5.1 Introduction 81

The action of T − produces a state of lower T 3 value by one unit. Since there isno other label, it must be that it is proportional to | j ; m 〉:

T − | j ; m 〉 = d (−)m | j ; m − 1 〉. (5.12)

Using

[ T +, T − ] | j ; m 〉 = 2T 3 | j ; m 〉, (5.13)

we find that

d (+)

m−1 d (−)m − d (+)

m d (−)

m+1 = 2m, (5.14)

subject to d (+)j = d (−)

− j = 0. Since | j ; m 〉 and | j ; m ′ 〉 are eigenstates ofa hermitian operator, their inner product must vanish for m �= m ′. Furthermorewe can determine the constants d(±)

m so as to make them have unit norm. A littlecomputation shows that if

d (+)m = √

( j − m)( j + m + 1), d (−)m = √

( j + m)( j − m + 1), (5.15)

the states form an orthonormal set

〈 j ; m | j ; m ′ 〉 = δm,m′ . (5.16)

In addition, completeness is satisfied over the states of the representation, that is

m=+ j∑m=− j

| j ; m 〉〈 j ; m | = 1. (5.17)

It is useful to represent the Lie algebra in this representation as (2 j +1)× (2 j +1)matrices (

T A)

pq, (5.18)

where p, q run from 1, 2, . . . , 2 j + 1. We can choose to represent T 3 as a diagonalmatrix with entries j, j −1, j −2, . . . ,− j , in which case T +(T −) are upper(lower)triangular matrix (homework). The representation matrices are always traceless,since each is the commutator of two matrices, which is always traceless because ofthe cyclic property of the trace.

Two representations with “celebrity” status, are worthy of special mention.(1) The smallest one, with j = 1/2, in which the generators are represented interms of the three (2 × 2) Pauli spin matrices

T A = σ A

2, (5.19)

Page 94: Pierre Ramond Group Theory a Physicists Survey 2010

82 SU (2)

which satisfy

σ A σ B = δAB + i εABC σC; σ A ∗ = −σ 2 σ A σ 2, (5.20)

where star denotes charge conjugation.(2) The second celebrity representation is that which has as many states as ele-

ments in the Lie algebra. It is called the adjoint representation, with 2 j = 2, wherethe generators can be represented by (3 × 3) hermitian matrices

T 1 =⎛⎝0 0 0

0 0 i0 −i 0

⎞⎠ , T 2 =⎛⎝ 0 0 i

0 0 0−i 0 0

⎞⎠ , T 3 =⎛⎝ 0 i 0

−i 0 00 0 0

⎞⎠ . (5.21)

We can associate an infinite lattice with a Lie algebra, as an array of points,each representing a state in the totality of its irreps. In the SU (2) case the latticeis one-dimensional, with evenly spaced points. Each point corresponds to a par-ticular value of T 3, and represents a state that can appear in an infinite number ofrepresentations. If its axis is taken to be 2T 3, they sit at the integer points:

� � � � � � � � �

−4 −3 −2 −1 0 1 2 3 4

We can draw some lessons from the way we have constructed these representa-tions. (1) The SU (2) irreps can be identified by their highest weight state; (2) allother states of the irrep are generated by repeated application of the lowering oper-ator T −. SU (2) has only one label and one lowering operator, but for bigger Liealgebras, the highest weight state and its descendants will have several labels andmore lowering operators. SU (2) has one Casimir operator and one label; Lie alge-bras will have in general as many labels as Casimir operators. Their number iscalled the rank of the algebra, and it is equal to the number of commuting elementsof the Lie algebra. The SU (2) Lie algebra has rank one (hence the subscript in itsmathematical names).

5.2 Some representations

There is another way to generate the irreducible representations, dear to the heartof physicists. It is based on the fact that all representations can be generated bytaking direct products of the smallest representations, in this case the doublet 2.

Suppose we clone our Lie algebra, producing two identical copies, which wedenote by T A

(1) and T A(2), each satisfying the same Lie algebra, but commuting with

one another.

Page 95: Pierre Ramond Group Theory a Physicists Survey 2010

5.2 Some representations 83

[ T A(a), T B

(a) ] = i εABC T C(a), a = 1, 2 [ T A

(1), T B(2) ] = 0. (5.22)

Each acts on its own Hilbert space, spanned by | · · · 〉(1) and | · · · 〉(2). The sum ofthe generators generates the same algebra[

T A(1) + T A

(2), T B(1) + T B

(2)

] = i εABC(T C(1) + T C

(2)

), (5.23)

acting on the direct product states

| . . . 〉(1)| . . . 〉(2), (5.24)

which span a new Hilbert space, constructed out of the original two. Since theirsums satisfies the same commutation relations, one should be able to represent theiraction in terms of the representations we have previously derived. When SU (2) isidentified with three-dimensional rotations, physicists call this construction, theaddition of angular momentum.

Consider the (Kronecker) direct product of two arbitrary representations 2j + 1and 2k + 1. The highest weight direct product state, which we label only with theT 3 eigenvalue, since we know the range of its eigenvalues:

| j 〉(1) | k 〉(2), (5.25)

satisfies

T 3| j 〉(1) | k 〉(2) = (T 3(1) + T 3

(2)

) | j 〉(1) | k 〉(2) = ( j + k) | j 〉(1) | k 〉(2), (5.26)

and

T +| j 〉(1) | k 〉(2) = (T +(1) + T +

(2)

) | j 〉(1) | k 〉(2) = 0. (5.27)

It must be the highest weight state of the 2(j + k) + 1 representation. In order togenerate the rest of its states, we apply to it the sum of the two lowering opera-tors. The action of T − generates the linear combination (neglecting the numericalcoefficient in front of each ket)

| j − 1 〉(1) | k 〉(2) + | j 〉(1) | k − 1 〉(2). (5.28)

We form its orthogonal combination

| j − 1 〉(1) | k 〉(2) − | j 〉(1) | k − 1 〉(2), (5.29)

on which T 3 = j + k − 1, and check (homework) that it is annihilated by T +: it isthe highest weight of another representation, with highest weight 2(j + k − 1)+ 1.We repeat the process, and find new states which are linear combinations of thethree states

| j − 2 〉(1) | k 〉(2), | j − 1 〉(1) | k − 1 〉(2), | j 〉(1) | k − 2 〉(2). (5.30)

Page 96: Pierre Ramond Group Theory a Physicists Survey 2010

84 SU (2)

Two linear combinations are lower weight states of the two representations 2(j +k)+1 and 2(j+k−1)+1; the third is the highest weight state of a new 2(j+k−2)+1representation. The pattern becomes clear, we keep generating at each level onenew representation until we run out of states because we hit the lower bound of thetwo original representations. It is easy to see (homework) that we stop at the | j −k|representation, and thus

[2j + 1]× [2k + 1] = [2(j + k) + 1]+ [2(j + k − 1) + 1]+ · · ·+ [2(j − k) + 1],(5.31)

taking here j ≥ k. Hence by taking the product of the lowest representation wegenerate all the representations of the algebra. This is a very economical way toproceed as long as one does not have to deal with large representations which,thankfully is usually the case in physics.

To see how this works in detail, we give two examples. First take the directproduct of two spinor representations. Start from the direct product highest weightstate

| ↑ ↑ 〉 ≡ | 12 ,

12 〉(1)| 1

2 ,12 〉(2), (5.32)

and act on it with the sum of the lowering operators

T − ≡ T −(1) + T −

(2), (5.33)

to get the linear combination

| ↓ ↑ 〉 + | ↑ ↓ 〉 ≡ | 12 ,− 1

2 〉(1)| 12 ,

12 〉(2) + | 1

2 ,12 〉(1)| 1

2 ,− 12 〉(2), (5.34)

and do it again to obtain

| ↓ ↓ 〉 ≡ | 12 ,− 1

2 〉(1)| 12 ,− 1

2 〉(2). (5.35)

Further action of the lowering operator on this state yields zero. Hence thesethree states form the three-dimensional representation of the algebra. Note that thedirect product Hilbert space has four states, the fourth being the orthogonal linearcombination

| 12 ,− 1

2 〉(1)| 12 ,

12 〉(2) − | 1

2 ,12 〉(1)| 1

2 ,− 12 〉(2). (5.36)

It is easy to see that it is killed by either lowering or raising operators; it is there-fore a singlet state. We have just proven that two spin one-half form themselvessymmetrically into a spin one combination or antisymmetrically into a spin zerostate. We indicate this as

2 × 2 = 3 + 1, (5.37)

indicating here the representation by its dimensionality.

Page 97: Pierre Ramond Group Theory a Physicists Survey 2010

5.2 Some representations 85

For a second example, consider the Kronecker product of spin one-half with spinone representations. In our language, it is 2 × 3 which, according to what we havejust said,

2 × 3 = 4 + 2, (5.38)

and contains contains two representations, spin 3/2 and 1/2. To construct thesestates explicitly, starting from the highest weight states, we need the action of thelowering operators on the j = 1/2 and j = 1 representations

T −| 12 ,

12 〉 = | 1

2 ,− 12 〉; T −| 1, 1 〉 = √

2 | 1, 0 〉; T −| 1, 0 〉 = √2 | 1,−1 〉.

(5.39)

We first build the states of the spin 3/2 representation, starting from the highestweight of the constituent representations:

| 32 ,

32 〉 = | 1

2 ,12 〉 | 1, 1 〉. (5.40)

This state is uniquely determined, as we can see from its value of T 3. We generateall the remaining states of this representation by use of the sum of the loweringoperators. Using eq. (5.39), we have

| 32 ,

12 〉 = 1√

3T − | 3

2 ,32 〉,

=√

13 | 1

2 ,− 12 〉 | 1, 1 〉 +

√23 | 1

2 ,12 〉| 1, 0 〉. (5.41)

Then we continue the process

| 32 ,− 1

2 〉 = 12 T − | 3

2 ,12 〉,

= 1

2√

3

(2√

2 | 12 ,− 1

2 〉 | 1, 0 〉 + 2 | 12 ,

12 〉 | 1,−1 〉

),

=√

23 | 1

2 ,− 12 〉 | 1, 0 〉 +

√13 | 1

2 ,12 〉 | 1,−1 〉. (5.42)

Finally, we lower one more time to get the lowest weight state

| 32 ,− 3

2 〉 = 1√3

T − | 32 ,− 1

2 〉,

= | 12 ,− 1

2 〉 | 1,−1 〉.The remaining four states are then obtained as the combinations that are orthogonalto (5.41). In this case there is only one√

23 | 1

2 ,− 12 〉 | 1, 1 〉 −

√13 | 1

2 ,12 〉, (5.43)

Page 98: Pierre Ramond Group Theory a Physicists Survey 2010

86 SU (2)

and since it is annihilated by T +, it is the highest weight state of the spin one-halfrepresentation, that is

| 12 ,

12 〉 =

√23 | 1

2 ,− 12 〉 | 1, 1 〉 −

√13 | 1

2 ,12 〉. (5.44)

The coefficients which appear in these decompositions are called the Clebsch–Gordan coefficients, familiarly know as “Clebsches.” They are so important tophysical applications, that they have been tabulated. See for example the ParticleData Group (http://pdg.gov/).

There are countless ways to represent the SU (2) algebra. Our favorite is in termsof two fermionic harmonic oscillators. In this case, the Hilbert space contains justfour states

| 0 〉, b†1| 0 〉, b†

2| 0 〉, b†1b†

2| 0 〉. (5.45)

The three SU (2) generators are simply

T ji = b†

i b j − 1

ji b†

kbk, i, j = 1, 2. (5.46)

It is easily seen that the four states break up into two singlets and one doublet. Thisconstruction generalizes to an arbitrary number of harmonic oscillators.

5.3 From Lie algebras to Lie groups

What kind of group does this algebra generate? We have seen that a general uni-tary operator can be written as the exponential of i times a hermitian matrix. Thegenerators of the SU (2) Lie algebra are hermitian, so that any operator of the form

U(�θ) = ei θA T A, (5.47)

where the θA are three real parameters, is a unitary operator. These operators satisfythe group axioms. Each has an inverse, with �θ → −�θ . One, with �θ = 0, is the unitelement. Less obvious is their closure under multiplication, that is

U(�θ)U(�η) = U( �φ(�θ, �η)), (5.48)

as their associativity

U(�θ)(U(�η)U( �χ)) = (

U(�θ)U(�η))U( �χ). (5.49)

The proof of closure is based on the Baker–Hausdorf theorem which expresses theproduct of two exponentials as an exponential, with exponent in the Lie algebra.We do not prove it here, as the reader can convince him/herself by expanding both

Page 99: Pierre Ramond Group Theory a Physicists Survey 2010

5.3 From Lie algebras to Lie groups 87

sides to second order in the parameters. We can easily verify it for SU (2) in thej = 1/2 representation. A little mathematical exercise shows that

U(�θ) = exp

(i θ A σ A

2

)= cos(

θ

2) + i θ A σ A sin(

θ

2), (5.50)

where

θ A = θ A

θ; θ = √

θ Aθ A. (5.51)

In this form both closure and associativity are manifest: U(�θ) satisfies the groupaxioms. It is a continuous (Lie) group because the �θ are continuous parameters.

It describes a group transformation that is periodic only after a 4π rotation. A 2πtransformation results in a reversal in sign of the vector on which it acts. Viewedas an operator in a two-dimensional Hilbert space, this transformation leaves thequadratic form,

|z1|2 + |z2|2, (5.52)

invariant. This is the reason why the group of unitary transformation in two dimen-sions associated with the Lie algebra SU (2) is called SU (2). The ever-carefulmathematicians distinguish the algebra from the group by using lower-case let-ters for the algebra (su(2)), and upper-case letters for the group (SU (2)). The Usignifies that the transformations are unitary, and the S that these transformationshave unit determinant. Their general form is(

α β

−β α

), (5.53)

subject to

|α|2 + |β|2 = 1. (5.54)

Any group element can be represented as a point on the surface of a four-sphereS3, the group manifold of SU (2). A succession of transformations will trace acontinuous curve on S3. In particular any closed curve can be shrunk to a pointsince one cannot lasso a basketball, even in four dimensions! We say that the SU (2)group manifold is simply connected.

Not all group manifolds are simply connected. To see this, let us look at theexponential mapping of the same algebra, but in the j = 1 representation, wherethe SU (2) generators are represented by matrices(

T A)

BC= −iεABC . (5.55)

The group elements are expressed as (3 × 3) rotation matrices

RBC(�θ) = exp

(θ AεABC

), (5.56)

Page 100: Pierre Ramond Group Theory a Physicists Survey 2010

88 SU (2)

which generates rotations in space, acting on the three components of a vector

�v → �v′ = R(�θ) �v. (5.57)

This element generates the group of rotations in real three-dimensional space,called SO(3). It is the group of orthogonal transformations of unit determinantthat leave the length of a three-dimensional vector invariant

�v · �v = �v′ · �v′ → R(�θ)TR(�θ) = 1, (5.58)

where T means matrix transpose. By expanding the exponential, we find (home-work) that

RBC(�θ) = δBC cos θ + εBC AθA sin θ + θB θC(1 − cos θ), (5.59)

which describes a rotation by an angle θ about the axis θA. The group elementis then given by the unit vector θA and the angle θ . We can use this expressionto determine the SO(3) group manifold. It is a three-dimensional space with theangles θA labeling its orthogonal axes. The group element is the same for θ andθ + 2π , so that we limit its range to

−π ≤ θ ≤ π, (5.60)

and allow negative values for the angles. The SO(3) group manifold is the insideand the surface of a sphere of radius π , but since θ = π and θ = −π lead tothe same group element, antipodal points on the surface of the sphere are identifiedwith one another. Alternatively, a point in the manifold is a vector with components

ξ1 = θ

πsin η cosφ, ξ2 = θ

πsin η sinφ, ξ3 = θ

πcos η, (5.61)

subject to the constraint

(ξ1)2 + (ξ2)

2 + (ξ3)2 =

π

)2

, (5.62)

with |θ | ≤ π , with antipodal identification on the sphere |θ | = π . This describesthe SO(3) group manifold.

To investigate its topological properties, consider a rotation that starts at the ori-gin and does not make it to the surface of the sphere and comes back, forming aclosed curve on the manifold. That curve can be shrunk to a point without obstruc-tion. Now consider another path which starts at the origin and makes its way to thesurface of the sphere, and then reappears at the antipodal point before rejoining theorigin. This closed curve cannot be deformed to a point: SO(3) is multiply con-nected. Next, consider a curve which describes a path that pierces the sphere at twopoints. We leave it to the reader to show that this path can also be deformed to apoint. It means that SO(3) is doubly connected.

Page 101: Pierre Ramond Group Theory a Physicists Survey 2010

5.4 SU (2)→ SU (1,1) 89

It can be shown that our algebra generates yet another group, Sp(2), the group ofquaternionic transformations, a notion we will examine in detail when we considersymplectic algebras.

The same Lie algebra can generate groups with different group manifolds withdifferent topologies. When a Lie algebra generates several groups, the one whosemanifold is simply connected is called the universal covering group: SU (2) is theuniversal covering group of SO(3).

5.4 SU(2) → SU(1, 1)

The SU (2) Lie algebra has yet another interesting property: one can reverse thesign of two of its elements without changing the algebra. For example

T 1,2 → −T 1,2; T 3 → T 3. (5.63)

This operation is called an involutive automorphism of the algebra. It is an opera-tion which forms a finite group of two elements. Its eigenvalues are therefore ±1,and it splits the Lie algebra into even and odd subsets.

One can use this to obtain a closely related algebra by multiplying all oddelements by i . This generates another Lie algebra with three elements L A, withL1,2 ≡ iT 1,2, L3 ≡ T 3. Note that L1,2 are no longer hermitian. The new algebra is

[L1 , L2 ] = −i L3, [L2 , L3 ] = i L1; [L3 , L1 ] = i L2. (5.64)

It is still a Lie algebra because this replacement does not affect the Jacobi iden-tity. The minus sign in the first commutator has crucial consequences because theCasimir operator,

Q = (L1)2 + (L2)2 − (L3)2, [ Q, Li ] = 0, (5.65)

is no longer positive definite. This means that the spectrum of L3 need not bebounded: the natural realm of this Lie algebra is an infinite-dimensional Hilbertspace. What a difference a sign makes! This algebra is called a non-compact or realform of the SU (2) algebra. As you can see from the Casimir operator, the invariantquadratic form is not positive definite, but looks like the length of a vector in three-dimensional Minkowski space with two spatial and one time directions, hence thisalgebra is called SO(2, 1) (also SU (1, 1)) by physicists.

One can go further and absorb the i on the right-hand side of the commutators,and generate the algebra in terms of the real matrices (hence the name real form?)(

1 00 −1

),

(0 11 0

),

(0 −11 0

). (5.66)

Page 102: Pierre Ramond Group Theory a Physicists Survey 2010

90 SU (2)

Its representation theory is far more complicated than that of its compact coun-terpart. Its simplest unitary representations can be found in the infinite Hilbertspace generated by one bosonic harmonic oscillator. To see this, consider theoperators

L+ ≡ 1

2√

2a†a†; L− ≡ 1

2√

2a a = (L+)†. (5.67)

Compute their commutator

[L+, L− ] = −1

4(1 + 2a†a) ≡ −L3. (5.68)

The new operator itself has nice commutation relations with L±

[L3, L± ] = ±L±, (5.69)

completing the non-compact SO(2, 1) Lie algebra, as we can see by defining thehermitian generators through

L+ ≡ 1√2

(L1 + i L2

), L− ≡ 1√

2

(L1 − i L2

). (5.70)

In this representation the Casimir operator reduces to a c-number

Q = 3

16. (5.71)

The action of these operators is particularly simple. Starting from the vacuum state,| 0 〉, we note that

L− | 0 〉 = 0, (5.72)

whileL+ | 0 〉 = 1

2 | 2 〉, (5.73)

and then by repeated application of the raising operator L−, we generate an infinitetower of states

(L+)n | 0 〉 =√(2n)!

2√

2| 2n 〉. (5.74)

Each state is an eigenstate of L3,

L3 | 2n 〉 = (1 + 4n)

4| 2n 〉. (5.75)

Hence the infinite tower of states | 0 〉, | 2 〉, . . . , | 2n 〉, . . . forms a representationof the algebra. Furthermore since the operators are hermitian, this representationis unitary and manifestly irreducible. The L3 spectrum is bounded from below,starting at 1/4 and increasing forever in integer steps.

There is another tower of states, | 1 〉, | 3 〉, . . . , | 2n + 1 〉, . . . which are alsorelated to one another with the action of the generators of our Lie algebra. In this

Page 103: Pierre Ramond Group Theory a Physicists Survey 2010

5.4 SU (2)→ SU (1,1) 91

case, the L3 spectrum starts at 3/4. The harmonic oscillator space splits into tworepresentations of SO(2, 1) of even and occupation number states. They are calledsingleton representations since they involve only one harmonic oscillator. The one-dimensional harmonic oscillator has no particular symmetry of its own (except forparity x → −x), so where does this algebra fit in? The clue is that its Hamiltonian

H = a† a + 12 = 2L3, (5.76)

is itself a member of the algebra. The algebra relates states of different energy; suchan algebra is called a spectrum generating algebra; each representation contains allof its states of a given parity.

There are of course many other representations of this algebra. We can gen-erate as many finite-dimensional representations as we want, by starting from the(2 j +1)-dimensional hermitian representation matrices of the compact algebra andmultiplying any two generators by i . Of course none is unitary.

As we have seen, in the j = 1/2 representation, the three generators are

1

2

(0 1

−1 0

); 1

2

(1 00 −1

); i

2

(0 11 0

). (5.77)

The corresponding group elements generated by these (2 × 2) matrices are ofthe form

eM; M = 12

(iθ3 −θ1 + iθ2

−θ1 − iθ2 −iθ3

). (5.78)

They are not unitary, with matrices of the form

h =(α∗ β∗

β α

), (5.79)

of unit determinant

|α |2 − |β |2 = 1. (5.80)

These matrices generate a group called SU (1, 1), since it leaves the abovequadratic form (5.80) invariant. This representation is not unitary. The Casimiroperator expressed in terms of hermitian operators is not positive definite, whichmeans that unitary representations of this non-compact algebra have to be infinite-dimensional.

To illustrate the complexity of the unitary representations of non-compactgroups, we list without proof the many different types of unitary irreducible rep-resentations (irreps) of SO(2, 1) = SU (1, 1). We express the Casimir operatoras Q = − j ( j − 1). One finds, after much work, five types of unitary irreps.All are infinite-dimensional, the states within each irrep being labeled by l3, theeigenvalues of L3:

Page 104: Pierre Ramond Group Theory a Physicists Survey 2010

92 SU (2)

Continuous class

• Non-exceptional interval: j = − 12 − is, 0 ≤ s < ∞.

(i) Integral case, l3 = 0,±1,±2, . . ..(ii) Half-integral case: l3 = 0,± 1

2 ,± 32 , . . ..

They are realized on square-integrable functions on the unit circle f (ϕ), with innerproduct

( f, g ) = 1

∫ 2π

0dϕ f ∗(ϕ) g(ϕ),

U(h) f (ϕ) = |α∗ − β eiϕ |−1−2is f (�),

where

ei� = α eiϕ − β∗

α∗ − β e−iϕ.

• Exceptional interval: j = − 12 − σ, 0 < σ < 1

2 ; l3 = 0,±1,±2, . . ..

They describe functions on the circle such that

U(h) f (ϕ) = |α∗ − β eiϕ |−1−2σ f (�),

in a space with inner product

( f, g)σ = (σ + 1/2)

(2π)3/22σ(σ )

∫ 2π

0dϕ1

∫ 2π

0dϕ2

[1 − cos(ϕ1 − ϕ2)

]σ−1/2f ∗(ϕ1)g(ϕ2).

Discrete class

• j = 12 , 1, 3

2 , . . .; l3 = j, j + 1, j + 2, . . ..

• j = 12 , 1, 3

2 , . . .; l3 = − j,− j − 1,− j − 2, . . ..

• j = 14 ,

34 ; l3 = j, j + 1, j + 2, . . . (singleton representations).

They are realized on functions of a complex variable, f (z), defined on the open circle| z | < 1,

U(h) f (z) = (α∗ + iβz)2 j f

(αz − iβ∗

α∗ + iβz

)with inner product

( f, g ) j = −2 j − 1

π

∮d2z

[1 − |z|2

]−2 j−2f ∗(z) g(z).

In a sense the singletons are the simplest unitary irreps of this algebra, but theyare actually part of a representation of a super algebra which is an algebra closedunder both commutators and anticommutators.

Indeed the lowest states of the two singleton representations, | 0 〉 and | 1 〉, can berelated to one another by the application of the creation operator a†. Hence we can

Page 105: Pierre Ramond Group Theory a Physicists Survey 2010

5.5 Selected SU (2) applications 93

extend the Lie algebra to include the operators a and a† that relate the two singletonrepresentations. However, while their commutator is not in the Lie algebra, theiranticommutator

{a, a†} ≡ a a† + a† a = 1 + 2a† a = 4L3, (5.81)

is. We further check that

[ L+, a ] = − 1√2

a†, [ L3, a ] = − 12 a. (5.82)

Thus, by extending the Lie algebra operation to include both commutators and anti-commutators, we obtain a super-Lie algebra. Its singleton representation includesboth singleton representations of its Lie sub-algebra, and is the Hilbert space gener-ated by one harmonic oscillator. See how much one can do with one set of creationand annihilation operators! Could it be that any Hilbert space can be thought of asa representation of some algebraic structure?

5.5 Selected SU(2) applications

The smallest Lie algebra we have just studied and its related groups have manyapplications in physics (for particle physics, see the books by Cahn [1] andGeorgi [6]). In the following we have chosen a few examples which illustrate itsimportance.

5.5.1 The isotropic harmonic oscillator

The quantum-mechanical isotropic harmonic oscillator is treated in many books,and we simply state its main features. It is a physical system with time evolutiongenerated by the Hamiltonian

H = �p · �p2m

+ 12 k �x · �x, (5.83)

where �x and �p are the three-dimensional position and momentum vectors. It issolved by rewriting the Hamiltonian in terms of the three creation and annihilationoperators A†

i and Ai ,

H = hω( �A† · �A + 32), (5.84)

with

[ Ai , A†j ] = δi j , i, j = 1, 2, 3. (5.85)

Its eigenstates are

H | n1, n2, n3 〉 = hω(n1 + n2 + n3 + 32)| n1, n2, n3 〉, (5.86)

Page 106: Pierre Ramond Group Theory a Physicists Survey 2010

94 SU (2)

with

| n1, n2, n3 〉 =∏

i

(A†i )

ni

√ni ! | 0 〉. (5.87)

We want to analyze this spectrum using group theory. Its lowest state is the vacuumstate | 0 〉, annihilated by all three Ai . At the next energy level, one finds three states| 1, 0, 0 〉, | 0, 1, 0 〉 and | 0, 0, 1 〉, all with the same energy. Define three transitionoperators among these

P12| 1, 0, 0 〉 = | 0, 1, 0 〉 ; P13| 1, 0, 0 〉 = | 0, 0, 1 〉 ; P23| 0, 1, 0 〉 = | 0, 0, 1 〉.Since these states have the same energy, that means

[ H, Pi j ] = 0 (5.88)

acting on any of these three kets: the degeneracy of the first excited state tells us thatthere exist operators which commute with the Hamiltonian, at least when acting onthe first excited states. These operators generate symmetries of the Hamiltonian.The action of Ti j is to destroy A†

i and replace it with A†j , so they can be written as

hermitian operators of the form

Pi j = i(A†i A j − A†

j Ai ), (5.89)

which satisfy the SU (2) Lie algebra, and generate rotations in three dimensions.They are of course the angular momentum operators

L1 = P23 = x2 p3 − x3 p2, etc. (5.90)

They commute with the Hamiltonian and reflects its invariance under spacerotations. The commutator

[ Li , A†j ] = −iεi jk A†

k, (5.91)

denotes how the three creation operators transform under rotation: they transforminto one another like the components of a vector, but they are tensor operators.Hence they span the 3 representation of SU (2). Hence the N th level transforms asthe N -fold symmetric product of the vector(

3 × 3 × · · · × 3 × 3)

sym.

This enables us to decompose any excited level in terms of this product. As anexample take the second excited level. Since

3 × 3 = 5 + 1,

Page 107: Pierre Ramond Group Theory a Physicists Survey 2010

5.5 Selected SU (2) applications 95

it breaks up into the quadrupole representation and a singlet. Specifically thequadrupole is the second rank symmetric traceless tensor,[

A†i A†

j + A†i A†

j − 1

6δi j (

�A† · �A†)]| 0 〉,

while the singlet is just the dot product

( �A† · �A†)| 0 〉.We note that in this spectrum, only integer spin representations appear. Our analysisshows that we can think of all the states as symmetric tensors, with the tracesremoved by the δi j tensor. It is called an invariant tensor because it is mapped intoitself by an SO(3) transformation.

What about the rest of the representations, such as antisymmetric tensors andhalf-odd integer representations? Can we generate them with creation and annihi-lation operators? Of course: we can do anything with ladder operators! We havealready seen how to generate all representations by taking Kronecker productsof the two-component spinor representation. This suggests that we think of thetwo spinor states as being created by two creation operators, a†

α, α = 1, 2. Theoperators, simply given in terms of the Pauli matrices

T A = 1

2a†α (σ A)αβ aβ, (5.92)

satisfy the SU (2) commutation relations as long as

[ aα, a†β ] = δαβ. (5.93)

These operators (first constructed by J. Schwinger) are not to be confused withthose of the isotropic harmonic oscillator problem. They are SU (2) spinors, whilethose for the isotropic harmonic oscillator are SU (2) vectors. The Fock space gen-erated by arbitrary products of the spinor creation operators generate the SU (2)lattice, while that generated by the vector creation operators generate only a subsetof representations.

5.5.2 The Bohr atom

Bohr’s model of the hydrogen atom is a cornerstone of Quantum Mechanics. It isdescribed by the Hamiltonian of a particle of momentum �p, charge e and mass m,bound by an attractive Coulomb potential

H = �p · �p2m

− e2

r; r = √�x · �x . (5.94)

Page 108: Pierre Ramond Group Theory a Physicists Survey 2010

96 SU (2)

Its spectrum of negative energy bound states, familiar to physics students, isgiven by

En = −Ry

n2, (5.95)

where Ry is Rydberg’s constant and n is the principal quantum numbern = 1, 2, . . .. Each level is populated by states of angular momentum l =0, 1, 2, . . . , n − 1. The Bohr atom is manifestly rotationally invariant, meaningthat the angular momentum operators commute with the Bohr Hamiltonian

Li = εi jk x j pk, [ H, Li ] = 0. (5.96)

As we have seen in the previous example, this invariance implies that levels of thesame energy will come in sets of representations of the rotation algebra. Indeedthis happens, but the first excited level is made up of two SO(3) representations,the ones with l = 1 and l = 0. This implies that there is more symmetry in theproblem than meets the eye. In particular it hints at the existence of three operatorsthat map each of the l = 1 states into the l = 0 state. We infer the existence ofan operator which commutes with the Hamiltonian, and is a vector operator underSO(3).

Indeed, the existence of such a vector had been known for a long time (Laplace,Runge and Lenz) in the context of Kepler’s problem, which uses the same potential.Its classical form is

Aclassicali = εi jk p j Lk − m e2 xi

r. (5.97)

For use in quantum theory, we symmetrize on the momenta and angular momenta,to get its hermitian form

Ai = 12 εi jk (p j Lk + Lk p j ) − m e2 xi

r,

= εi jk p j Lk − m e2 xi

r− i pi .

It is not difficult to see that �A is orthogonal to �L�L · �A = �A · �L = 0. (5.98)

Also it is manifestly a vector operator, which can be verified by explicit computa-tion of its commutator with Li

[ Li , A j ] = i εi jk Ak . (5.99)

Finally, a careful computation yields

[ H, A j ] = 0, (5.100)

Page 109: Pierre Ramond Group Theory a Physicists Survey 2010

5.5 Selected SU (2) applications 97

which means that this vector is a constant of the motion, and that the Bohr Hamil-tonian is invariant under the transformations it generates. This extra invariance isnot at all manifest; it is a hidden symmetry.

It provides for a quick and elegant solution of the Bohr atom. A few monthsafter the creation of Heisenberg’s paper, Pauli was the first to solve the quan-tum mechanical Bohr atom. He used group theoretical methods, which we presentbelow.

We begin by computing the commutator of Ai with itself. The result is rathersimple

[ Ai , A j ] = i εi jk Lk (−2m H). (5.101)

Since the Bohr Hamiltonian describes a bound state system, we expect the eigen-values of H to be negative. Also, since H commutes with both �L and �A, we candefine without ambiguity the new operator

Ai ≡ Ai√−2m H, (5.102)

which satisfies the simpler commutation relation

[ Ai , A j ] = i εi jk Lk, (5.103)

closing the algebra. We note that the linear combinations

X (+)i ≡ 1

2(Li + Ai ), X (−)i ≡ 1

2(Li − Ai ), (5.104)

not only commute with one another

[ X (+)i , X (−)

j ] = 0, (5.105)

but also form two independent SU (2) Lie algebras

[ X (+)i , X (+)

j ] = i εi jk X (+)k ; [ X (−)

i , X (−)j ] = i εi jk X (−)

k . (5.106)

The invariance algebra of the Bohr atom is the direct product of two SU (2)algebras, SU (2) × SU (2). The two Casimir operators of these algebras must beconstants of the motion, but as we see from eq. (5.98), they are the same since

C (+)2 = 1

4

(Li + Ai

) (Li + Ai

)= 1

4

(Li − Ai

) (Li − Ai

)= C (−)

2 , (5.107)

that is

C (+)

2 = j1 ( j1 + 1) = C (−)

2 = j2 ( j2 + 1) ≡ j ( j + 1), (5.108)

where j = j1 = j2.The next step is to express the Hamiltonian in terms of the Casimir operators by

calculating the length of the LRL vector. This is a tricky calculation because we

Page 110: Pierre Ramond Group Theory a Physicists Survey 2010

98 SU (2)

are dealing with operators, so we give some of the details. We split the result intodirect and cross terms. The direct terms easily yield

Ai Ai

∣∣∣direct

= εi jk εimn p j pm Lk Ln − 2m e2

rεi jk xi p j Lk + m2 e4

= (p j p j ) Li Li − p j p j − 2m e2

rLi Li + m2e4.

The mass-independent cross terms are

−iεi jk(p j Lk pi + pi p j Lk) = 2pi pi , (5.109)

using

[ Lk, p j ] = iεk ji pi .

The first of the mass-dependent cross terms yields

−m e2εi jk

(p j Lk

xi

r+ xi

rp j , Lk

)= −2m e2

rLk Lk − 2im e2 pi

xi

r,

while the second gives

im e2(

pi

xi

r+ xi

rpi

).

Their sum is

−2m e2

rLk Lk − 3m e2

r− im e2

[pi xi ,

1

r

].

Evaluation of the commutator yields the final result

Ai Ai =(

�p · �p − 2m e2

r

) (Li Li + 1

)+ m2 e4 . (5.110)

We divide both sides by −2m H , and solve for H , to get

H = − m e4/2

�L · �L + �A · �A + 1= − m e4/2

4 C (+)

2 + 1. (5.111)

The relation between the Hamiltonian and the Casimir operator is explicit; in termsof j ,

H = − m e4

2(2 j + 1)2. (5.112)

This clearly shows that j is not the angular momentum but is related to the principalquantum number

n = 2 j + 1.

Page 111: Pierre Ramond Group Theory a Physicists Survey 2010

5.5 Selected SU (2) applications 99

The angular momentum is the sum of the two independent Lie algebras

Li = X (+)i + X (−)

i , (5.113)

so that finding the degeneracy of the levels with n = 2 j + 1 amounts to listingthe states in the direct product of (2j + 1) × (2j + 1), the classical problem as theaddition of two angular momenta with the same j , yielding states with angularmomentum l = 2 j = n − 1, 2 j − 1 = n − 2, . . . , 1, 0, which is the well-knownresult.

By recognizing the symmetries of Bohr’s hydrogen, and using his knowledge ofthe representations of SU (2), Pauli was the first (Zs. f. Phys. 36, 336 (1926)) toderive the spectrum of its Hamiltonian. Such was the power of Pauli and of grouptheory!

5.5.3 Isotopic spin

Nuclear forces appear to act equally on protons and neutrons, the constituents ofnuclei. This led W. Heisenberg to think of them as the two components of aniso doublet of a new SU (2) which he called isotopic spin. It cannot be an exactconcept since protons are charged and neutrons are not. Isotopic spin invarianceapplies only to nuclear forces, and it appears as a broken symmetry when the weakand electromagnetic forces are taken into account. The nuclear potential was soonunderstood as being caused by new elementary particles, the pions. They appear inthree charge states, π+, π0 and π−, suggesting that they could be assembled intoan isotriplet.

Fermi–Yang model

In this model (Phys. Rev. 76, 1739 (1949)), Fermi and Yang suggest that pions,the particles behind nuclear forces, might be nucleon–antinucleon bound states.Think of nucleons and antinucleons as constructed from Fermi oscillator states,that is

| p 〉 = b†1 | 0 〉, | n 〉 = b†

2 | 0 〉, (5.114)

for the nucleons, and

| p 〉 = b†1 | 0 〉, | n 〉 = b†

2 | 0 〉, (5.115)

for their antiparticles. The oscillators satisfy the usual anticommutation relations{bi , b†

j

}= δi j ,

{bi , b†

j

}= δi j , (5.116)

Page 112: Pierre Ramond Group Theory a Physicists Survey 2010

100 SU (2)

all other anticommutators vanishing. We construct the SU (2) × U (1) generators,

I j = 1

2b† σ j b − 1

2b† σ ∗

j b, j = 1, 2, 3, I0 = 1

2

(b† b − b† b

), (5.117)

where ∗ denotes conjugation, and the matrix indices have been suppressed. Fermiand Yang noted that the pions have the same quantum numbers as nucleon–antinucleon pairs, leading to

|π+ 〉 = b†1 b†

2 | 0 〉, |π0 〉 = 1

2

(b†

1 b†1 − b†

2 b†2

), | 0 〉, |π− 〉 = b†

2 b†1 | 0 〉,

(5.118)that is pions look like nucleon–antinucleons bound together; while erroneous asstated, their idea survives today, since pions contain quark–antiquark pairs. Thisis an early example of the use of group theory to the taxonomy of elementaryparticles. Soon we will generalize it to include strange particles.

The electric charge for the pion isotriplet is simply equal to the third componentof isospin, while for the nucleons, it is the linear combination

Q = I3 + I0, (5.119)

where I0 is equal to 1/2 for nucleons, −1/2 for antinucleons.

Wigner supermultiplet model

Wigner (Phys. Rev. 51, 106 (1939), and independently Stückelberg (Helv. Phys.Acta 11, 225 (1938)), extended isospin symmetry by combining it with spin. Thisleads to

SU (4) ⊃ SU (2)I × SU (2)spin, (5.120)

with the nucleons transforming both as spinors and isospinors,

| N 〉 ∼ 4 = ( 2, 2spin ), | N 〉 ∼ 4 = ( 2, 2spin ),

where the second labels the spin. The pions belong to the fifteen-dimensionaladjoint representation built from

4 × 4 = 15 + 1, (5.121)

with the result

15 = ( 3, 3spin ) + ( 1, 3spin ) + ( 3, 3spin ). (5.122)

It contains additional spin one states which are found in experiments: one isovector( 3, 3spin ), the ρ+, ρ0, ρ−, vector mesons, and one isoscalar vector meson, ω. Moremassive than the pions, they generate corrections to the nuclear potential whichprovide the short-range repulsion (because they are vectors) as required by nuclear

Page 113: Pierre Ramond Group Theory a Physicists Survey 2010

5.5 Selected SU (2) applications 101

data. They can also be viewed à la Fermi–Yang, as nucleon–antinucleon pairs.Today we know them to be quark–antiquark pairs.

Epilogue

We cannot close this brief survey of the uses of SU (2) in physics without mention-ing its use as the gauge group which underlies the modern theory of β decay. Thisdiscussion is relegated to the section where we discuss the groups of the StandardModel.

Page 114: Pierre Ramond Group Theory a Physicists Survey 2010

6

SU (3)

While many of the techniques introduced in the previous chapter carry over tothe study of more general Lie algebras (see for instance Gilmore [7]), there arenon-trivial generalizations which are best discussed in the context of the morecomplicated SU (3) algebra.

6.1 SU(3) algebra

Consider a Hilbert space and its dual spanned by three kets and three bras, whichform an orthonormal and complete basis,

〈 a | b 〉 = δab, a, b = 1, 2, 3;3∑

a=1

| a 〉〈 a | = 1. (6.1)

It is the natural habitat of nine operators

Xab ≡ | a 〉 〈 b |, a, b = 1, 2, 3; (6.2)

one combination is the identity operator by completeness, while the others obeythe commutation relations

[ Xab, Xc

d ] = δ cb Xa

d − δ ad Xc

b. (6.3)

Together with Xaa = 0, these commutation relations define the abstract descrip-

tion of the eight-dimensional Lie algebra known as SU (3) to physicists and A2 tomathematicians. Combinations of the form

(X · X)ab ≡ Xa

n Xnb, (X · X · X)a

b ≡ Xap X p

q Xqb, . . . (6.4)

transform the same way as the generators,

[ Xab, (X · · · X · · · X)c

d ] = δ cb (X · · · X · · · X)a

d − δ ad (X · · · X · · · X)c

b. (6.5)

102

Page 115: Pierre Ramond Group Theory a Physicists Survey 2010

6.1 SU (3) algebra 103

Hence their traces over a, b are invariants

[ Xab, (X · X)n

n ] = [ Xab, (X · X · X)n

n ] = · · · = 0, (6.6)

leading to an apparently infinite number of invariants. In fact there are only twoindependent invariants. To see this, construct the traceless antisymmetric tensor

A acbd = Xa

b Xcd − Xc

b Xad − Xa

d Xcb + Xc

d Xab − trace terms, (6.7)

with the trace terms determined by demanding

A anbn = A mn

mn = 0. (6.8)

Then by use of the Levi–Civita tensor on three indices, we write

Cnm = εmacε

nbd A acbd, A ac

bd = ε macε nbd Cnm . (6.9)

By taking traces we see that Cnm vanishes and thus so does Aac

bd . It is then a simplermatter by multiplying with Xn

m to deduce that

(X · X · X)ab = α δa

b + β Xab + γ (X · X)a

b, (6.10)

where α, β, γ are expressible in terms of traces. Hence all higher-order combina-tions can be reduced and SU (3) has only two Casimir invariants, one quadratic andone cubic

C2 = (X · X)aa, C3 = (X · X · X)a

a. (6.11)

Proceeding as for SU (2), we single out two “diagonal” operators,

T 3 ≡ 12

(| 1 〉 〈 1 | − | 2 〉 〈 2 |

),

T 8 ≡ 12√

3

(| 1 〉 〈 1 | + | 2 〉 〈 2 | − 2 | 3 〉 〈 3 |

), (6.12)

normalized in the same way so that

Tr (T 3 T 3) = Tr (T 8 T 8) = 12 ; Tr (T 3 T 8) = 0. (6.13)

These two operators clearly commute with one another, and their commutatorswith the remaining six operators are given by

[ T 3, X12 ] = X1

2; [ T 3, X13 ] = 1

2 X13; [ T 3, X2

3 ] = − 12 X2

3; (6.14)

[ T 8, X12 ] = 0; [ T 8, X1

3 ] =√

32 X1

3; [ T 8, X23 ] =

√3

2 X23, (6.15)

together with their hermitian conjugates, while

[ X12, X1 †

2 ] = 2 T 3, (6.16)

Page 116: Pierre Ramond Group Theory a Physicists Survey 2010

104 SU (3)

and finally

[ X13, X1 †

3 ] = T 3 + √3 T 8, [ X2

3, X2 †3 ] = −T 3 + √

3 T 8. (6.17)

This allows for a simple generalization of Pauli’s spin matrices to the (3 × 3)hermitian Gell-Mann matrices

λ1 =(

0 1 01 0 00 0 0

), λ2 =

(0 −i 0i 0 00 0 0

), λ3 =

(1 0 00 −1 00 0 0

)

λ4 =(

0 0 10 0 01 0 0

), λ5 =

(0 0 −i0 0 0i 0 0

), λ6 =

(0 0 00 0 10 1 0

)

λ7 =(

0 0 00 0 −i0 i 0

), λ8 = 1√

3

(1 0 00 1 00 0 −2

).

Now we study this algebra in a language that generalizes to arbitrary Lie algebras.Think of the two diagonal operators as generating a vector in a two-dimensionalspace, and introduce the notation Hi , i = 1, 2 with

H 1 ≡ T 3; H 2 ≡ T 8. (6.18)

Associate the remaining six “off-diagonal” elements with two-dimensional vectors

X12 ≡ E

β(1); X13 ≡ E

β(2); X23 ≡ E

β(3); (6.19)

where the components of the two-dimensional vectors β(a), a = 1, 2, 3, are definedby the commutators

[ Hi , Eβ(a) ] = β(a) i E

β(a) . (6.20)

It is easy to see that

β(1) = { 1, 0 }, β(2) = { 12 ,

√3

2 }, β(3) = { − 12 ,

√3

2 }, (6.21)

which expresses the vectors in the H-basis. The β(a) are called the roots of the Liealgebra. Since

Eβ(a)† = E−β(a) , (6.22)

their negatives are also root vectors. The three roots, living in a two-dimensionalspace, are not independent: they satisfy

β(2) = β(1) + β(3). (6.23)

Page 117: Pierre Ramond Group Theory a Physicists Survey 2010

6.1 SU (3) algebra 105

The commutator of the elements associated with a root with their hermitianconjugates yield linear combinations of the H elements, allowing us to write

[ Eβ(a) , E−β(a) ] = β

(a)i H i , (6.24)

thus defining the root components with lower indices. Evaluation of the commuta-tors show them to be simply related to the components with upper indices

β(a)

i ≡ gi j β(a) j , (6.25)

thus introducing the metric and its inverse

gi j = 2 δi j , gi j = 1

2δi j . (6.26)

Distinguishing upper and lower indices may seem like overkill at this stage, but itis warranted for general Lie algebras. This enables us to define a scalar product

(β(a), β(b) ) ≡ β(a)

i β(b) i = gi jβ

(a)iβ

(b)j . (6.27)

The orthogonality of the vectors along H 1 and H 2 with respect to this scalarproduct is the same as the vanishing of the trace of (T 3T 8).

The remaining non-zero commutators of this algebra are easily found to be

[ X12, X3

1 ] = −X32; [ X1

2, X23 ] = X1

3; [ X13, X3

2 ] = X12, (6.28)

and their hermitian conjugates. The commutator of any two root elements is eitherproportional to another root element, or zero. The criterion is that the root associ-ated with the commutator of two elements is the sum of the two roots associatedwith each generator in the commutator, that is

[ Eβ, Eβ ′ ] ∼ E

β+β ′ . (6.29)

This implies that the commutator

[ Hi , [ Eβ(a) , E−β(a) ] ],

must vanish, which is easily verified using the Jacobi identity.Next, consider a zero commutator, such as

[ X12, X1

3 ] = [ E(β(1)), E(β(2)) ] = 0. (6.30)

The sum of these two roots is

β(1) + β(2) = { 32 ,

√3

2 }, (6.31)

which is not a root vector: the rule seems to be that the commutator

[ Eβ, Eβ ′ ]

vanishes whenever the vector (β + β ′) is not a root.

Page 118: Pierre Ramond Group Theory a Physicists Survey 2010

106 SU (3)

One can plot the eight elements of the Lie algebra by their root vectors in theH 1 − H 2 plane, including the two zero-length roots associated with H 1 and H 2,yielding the following SU (3) root diagram.

� ���

��

��

In this simple example, we have introduced some important general concepts:to each element of the Lie algebra corresponds a root vector defined in a spaceendowed with a metric and scalar product. The dimension of that space is equalto the number of commuting generators (rank) of the algebra, two for SU (3). Theroots are not linearly independent, since their number is larger than the dimensionof the space they span.

While it is obviously desirable to describe Lie algebras in basis-independentterms, some bases are more useful than others, depending on the application.

For SU (3), bases are defined by the choice of two independent roots; below wegive several choices, and show in detail how they are related.

6.2 α-Basis

We begin by splitting the six non-zero roots into positive and negative ones. Theactual definition of which is positive and which is negative is arbitrary. Here wedefine a positive root as one with a positive H 2 component, and a positive H 1

component if its H 2 component is zero.Hence all three roots β(1), β(2), β(3) are positive. We take the dependent positive

root to be the one expressed as a linear combination of the other two with positivecoefficients. This leads to the unique choice

β(2) = β(1) + β(3). (6.32)

The remaining two positive roots, β(1) and β(3)are called the simple roots of theLie algebra. We define

Page 119: Pierre Ramond Group Theory a Physicists Survey 2010

6.3 ω-Basis 107

α1 ≡ β(1), α2 ≡ β(3), (6.33)

and express all roots in terms of their components along these two simple roots.This defines the α-basis; unlike the H -basis, it is not orthogonal. All properties ofthe algebra can be expressed in terms of the simple roots. Explicitly,

(α1, α1) = (α2, α2) = 2, (α1, α2) = −1. (6.34)

The two simple roots of SU (3) have equal length and lie at 120◦ from one another.This information can be summarized in several ways. First, through the Cartanmatrix with elements

Ai j ≡ 2(αi , α j )

(α j , α j ), (6.35)

which for SU (3) is (2 −1

−1 2

).

Second, through a two-dimensional diagram: draw a point for each simple root.Here the two points are the same indicating the roots have equal length. Connectthe points by a line when the roots are at 2π/3 from one another. This is the SU (3)Dynkin diagram.

� �This dumbbell diagram contains all the relevant information about the algebra, ifyou know how to read it, of course.

6.3 ω-Basis

A closely related basis is the ω-basis, spanned by two vectors ω1 and ω2. They aredefined such that the coefficients of the simple roots in the α-basis expressed in theω-basis, are the rows of the Cartan matrix, that is

α1 ≡ 2 ω1 − ω2,

α2 ≡ −ω1 + 2 ω2.

Their H -basis components are

ω1 = { 12 ,

12√

3}, ω2 = { 0, 1√

3}, (6.36)

showing that they have equal length, and make a π/3 angle, as shown below.

Page 120: Pierre Ramond Group Theory a Physicists Survey 2010

108 SU (3)

α1

ω1

ω2

α2

We can represent the elements of the Lie algebra in terms of their componentsin these two bases: the commuting elements H 1, H 2 have zero components in allbases. In the H -basis, the components of the six roots contain ugly square rootsand the like; in the α- and ω-bases, they become beautiful integers.

Roots: β(1) β(3) β(2) −β(1) −β(3) −β(2)

α-basis: (1,0) (0,1) (1,1) (-1,0) (0,-1) (-1,-1)ω-basis: (2,-1) (-1,2) (1,1) (-2,1) (1,-2) (-1,-1)

Together with the two (0, 0) states, they form the defining adjoint representation ofthe SU (3) Lie algebra.

6.4 α′-Basis

Since we have not justified our procedure nor claimed uniqueness in constructingthese bases, the reader’s confidence in the method may be frayed, lest we build yetanother set of bases.

Suppose that we had defined positive roots as those with a positive H 1 compo-nent, so that all three β(1), β(2), −β(3) are positive roots. We then follow the sameprocedure, using the same criterion to determine the dependent root: the root thatcan be expressed as a linear combination of the other two with positive coefficientsis the dependent one, in this case

β(1) = β(2) + [−β(3)]. (6.37)

We then take β(2) and −β(3) to be the simple roots, and define

α′1 ≡ β(2), α

′2 ≡ −β(3). (6.38)

Page 121: Pierre Ramond Group Theory a Physicists Survey 2010

6.4 α′-Basis 109

Call this new basis the α′-basis. The Cartan matrix is the same as before since

(α′1, α

′1) = (α

′2, α

′2) = 2, (α

′1, α

′2) = −1, (6.39)

as is the Dynkin diagram. The new ω′-basis is introduced in the same way, through

α′1 ≡ 2 ω

′1 − ω

′2, α

′2 ≡ −ω

′1 + 2 ω

′2. (6.40)

These have the H -basis components

ω′1 =

{12 ,

12√

3

}, ω

′2 =

{12 ,− 1

2√

3

}. (6.41)

The different primed basis vectors are displayed as follows.

�H 2

H 1

���

α′2

ω′2

ω′1

α′1

These diagrams look the same except for an overall rotation: ω1 = ω′1, but ω

′2 can

be obtained as the reflection of ω2 about the direction ω1. Such a reflection is calleda Weyl reflection. In both bases, the sliver contained inside and on the two ω vectorsis called the Weyl chamber.

These bases are widely used for building representations of the algebra. Theirconstruction for the SU (3) Lie algebra follows the same procedures as for SU (2).First we label the states by the eigenvalues of the maximal set of commuting gen-erators, two for SU (3). We set aside for the moment the more traditional labelsattached to the Casimir operators, since we saw that for SU (2), the range of eigen-values determined the Casimir. Here we have two operators, and thus two ranges,which suggests two Casimir operators. It is in fact true that there are as manyCasimir operators as the rank, the number of commuting generators. The first is ofcourse quadratic, but the second happens to be cubic, which is not obvious at all atthis stage.

Page 122: Pierre Ramond Group Theory a Physicists Survey 2010

110 SU (3)

6.5 The triplet representation

First we determine the representation of SU (3) in the three-dimensional Hilbertspace. Different bases correspond to different labeling schemes. We can label itsthree states by two numbers, the eigenvalues of �H ,

Hi | h1, h2 〉 = hi | h1, h2 〉,yielding the three states

| 12 ,

12√

3〉, | 0,− 1√

3〉, | − 1

2 ,1

2√

3〉.

Alternatively, we can express them in the α- and ω-bases as well, yielding for thesame three states

α-basis: | 23 ,

13 〉, | − 1

3 ,− 23 〉, | − 1

3 ,13 〉,

ω-basis: | 1, 0 〉, | 0,−1 〉, | −1, 1 〉.Note that the state labels are rational numbers in the α-basis, and integers in theω-basis. This feature generalizes to all Lie algebras.

To build representations, we seek the lowering operators. SU (3) has three low-ering operators, but one is linearly dependent since its root vector is a linearcombination of the simple roots (that is, is the commutator of the other two), sowe can express all lowering operators in terms of only two (same as the rank), eachassociated with a simple root. The application of a lowering operator associatedwith the simple root α1, say, will lower the state labels according to the value ofthat root:

E−α1| h1, h2 〉 ∼ | h1 − 1, h2 〉,

E−α2| h1, h2 〉 ∼ | h1 + 1

2 , h2 −√

32 〉.

We note that since the weights associated with the lowering (and raising) operatorshave integer components in the α- and ω-bases their application to the states willshift their labels by integers.

Therein lies the great advantage of the ω-basis: starting with a highest weightstate with integer labels in the ω-basis, we generate only states with (positiveand negative) integer labels. Furthermore, the two ω-basis raising or loweringoperators are associated with the vectors whose components are the rows of theCartan matrix, (2,−1) and (−1, 2), and their negatives, respectively. In the eight-dimensional adjoint representation, application of the raising operators to the state| 1, 1 〉 gives zero as they would generate either (3, 0) or (0, 3), neither of whichare in the representation. This enables us to infer that the highest weight states(annihilated by the raising operators) are those with either positive integers or zero

Page 123: Pierre Ramond Group Theory a Physicists Survey 2010

6.5 The triplet representation 111

ω-basis labels. As another example, the application of the lowering operators onthe highest weight state with ω-basis labels (1, 0), produces three integer-labeledstates

| 1, 0 〉 −→ | − 1, 1 〉 −→ | 0,−1 〉,which form the triplet representation. The third state is a lowest weight state, asits labels have no positive integers. Further application of the lowering operatorsgives zero. Similarly, we could have started with the state | 0, 1 〉 and generated theantitriplet representation. This illustrates the convenience of this basis.

The SU (3) group lattice is an infinite two-dimensional lattice. Its points arelabeled by two integers which indicate the projection along ω1 and ω2. Thus thepoint (5,−2) lies five units along ω1 and two units along −ω2. Each point corre-sponds to a possible state. The points with one zero entry lie at the edge of the Weylchamber. The points inside the Weyl chamber are labeled by two positive integers;those at the edge have one positive integer and one zero entry; all serve to labelirreducible representations. The fundamental domain of the SU (3) lattice is just anequilateral triangle.

Scholium. Conjugation

We can use this decomposition of the SU (3) Lie algebra in terms of its SO(3) sub-algebra to gain an insight in the operation of charge conjugation. As we have seen,the SU (2) and SU (3) algebras operate on complex vectors, and so it is natural tosee what conjugation does to them. Consider first the Pauli matrices, and define

σA ≡ −σ ∗A. (6.42)

It is easy to see that these matrices satisfy the same algebra as the Pauli matrices,and therefore represent the same SU (2) algebra. However, the signs of the chargesare reversed and, rather than containing the charges (+1/2,−1/2), they yield thecharges (−1/2,+1/2), but these are the same up to a flip. A specific way to seethis is to recall that

σ A ∗ = −σ 2 σ A σ 2,

so that the states of the tilde representation are obtained from those of the untildedone by the action of σ 2, which flips the two states. It follows that these two repre-sentations are equivalent. In this case even though there is conjugation, it does notyield any new representation. Wigner calls this case pseudoreal.

For SU (3) though, charge conjugation yields entirely new representations. Tosee this, do the same for the Gell-Mann matrices and define

λA ≡ −λ∗A. (6.43)

Page 124: Pierre Ramond Group Theory a Physicists Survey 2010

112 SU (3)

Since they are hermitian, conjugation is the same as transposition, and this opera-tion does not change the λ2,λ5 and λ7 which generate the SO(3) sub-algebra. Theremaining five matrices change sign under this operation. They transform as thequadrupole, and the commutation of any two of them yields an SO(3) generator,so that the tilde matrices satisfy the same algebra. Schematically,

[ L, L ] ∼ L; [ L, Q ] ∼ Q; [ Q, Q ] ∼ L. (6.44)

The tilde matrices therefore satisfy the SU (3) algebra. They act on a three-dimensional representation, but it is different from the one we have alreadyconstructed! To see this it suffices to look at the charges of its states, since theyare the negatives of the previous ones; in the ω-basis, they are

| − 1, 0 〉, | 0, 1 〉, | 1,−1 〉. (6.45)

Clearly the highest weight state is | 0, 1 〉. This entirely new representation is theconjugate of the triplet representation. It is denoted as 3. Its Dynkin label is (0 1).

In general, if any Lie algebra is satisfied by hermitian matrices T A, the sameLie algebra is satisfied by minus their transpose. This follows easily by taking thetranspose of the defining equation. It follows that these matrices live either in a newHilbert space, or in the same Hilbert space (as for SU (2)).

6.6 The Chevalley basis

In our previous discussion, the choice for H 1 and H 2 was motivated by physics.Mathematicians prefer to construct the algebra by generalizing the SU (2) structure.For SU (2) and SU (3) we found as many commuting generators as simple roots,a feature that generalizes to all Lie algebras. To describe SU (3), we define twocopies of SU (2) generators, hr , er and e−r , r = 1, 2 with commutation relations

[ hr , er ] = 2er , [ hr , e−r ] = −2e−r , [ er , e−r ] = hr . (6.46)

In addition, we have

[ h1, h2 ] = 0. (6.47)

In matrix form, these correspond to

h1 ∼⎛⎝1 0 0

0 −1 00 0 0

⎞⎠ , h2 ∼⎛⎝0 0 0

0 1 00 0 −1

⎞⎠ , (6.48)

to be compared with the Gell-Mann matrices λ3 and λ8 for H 1 and H 2. Hence itfollows that

h1 = 2H 1, h2 = √3H 2 − H 1. (6.49)

Page 125: Pierre Ramond Group Theory a Physicists Survey 2010

6.6 The Chevalley basis 113

The ei are identified with the elements of the simple roots

e1 = Eα1, e2 = Eα2

. (6.50)

We will see later that for some Lie algebras, a normalization factor will be required.In the α basis, this yields

[ h1, e2 ] = −e2, [ h2, e1 ] = −e1. (6.51)

Happily, we note that these commutators can be neatly written in terms of theCartan matrix

[ hr , es ] = Asr es . (6.52)

This feature generalizes to all Lie algebras. As we have seen the algebra containsmore elements; in this notation, they are obtained as commutators of the simpleroot elements. For SU (3), the only non-zero element is [ e1, e2 ] together with itscomplex conjugate. In terms of matrices, we have

e1 =⎛⎝0 1 0

0 0 00 0 0

⎞⎠ , e2 =⎛⎝0 0 0

0 0 10 0 0

⎞⎠ , (6.53)

while the third root is identified with

e3 ≡ [ e1, e2 ] =⎛⎝0 0 1

0 0 00 0 0

⎞⎠ . (6.54)

There are no other elements because all double commutators vanish

[ e1, [ e1, e2 ] ] = [ e2, [ e1, e2 ] ] = 0. (6.55)

The Serre presentation, to be discussed later, offers a systematic way to recast thesevanishings in terms of the Cartan matrix.

Note that, in this basis, all elements are represented by matrices with integerentries. This is an elementary example of the Chevalley basis, in which the algebrais represented by matrices with integer entries, and where all structure functions areintegers. It seems appropriate to denote the matrix with unit entry in the i j positionas ei j . Thus we have

e1 = e12, e2 = e23, e−1 = e21, e−2 = e32

as well as

h1 = e11 − e22, h2 = e22 − e33.

This shorthand notation is very useful.

Page 126: Pierre Ramond Group Theory a Physicists Survey 2010

114 SU (3)

6.7 SU(3) in physics

This group manifests itself in many different areas of particle physics, simplybecause Nature seems to like the number three! There are three families of ele-mentary fermions; there are three colors of quarks, and there are three quarks withmass significantly lower than the inverse diameter of a proton (with the help ofPlanck’s constant h and the speed of light c, mass and length are contragredient).

6.7.1 The isotropic harmonic oscillator redux

Let us take another look at the isotropic harmonic oscillator in the light of SU (3).We found its invariance group by examining the three-fold degeneracy of its firstlevel, with transitions between these states represented by operators which replaceda†

i with another a†j , and we suggested their form to be

Li j ≡ i(A†i A j − A†

j Ai ).

However, these transitions could have been generated just as well by the otherhermitian combinations

(A†i A j + A†

j Ai ),

and nobody asked about them! There are six such operators, and under SO(3) theytransform as the five-dimensional “quadrupole” representation represented by thesymmetric traceless tensor

Qi j = 1

2(A†

i A j + A†j Ai ) − 1

3δi j

�A† · �A, (6.56)

which satisfies

[ Li j , Qmn ] = iδ jm Qin + iδ jn Qim − iδim Q jn − iδin Q jm, (6.57)

while the trace is left invariant

Q0 = �A† · �A, [ Li j , Q0 ] = 0. (6.58)

Furthermore, their commutators

[ Qi j , Qmn ] = − i

4

(δim L jn + δ jm Lin + δin L jm + δ jn Lim

), (6.59)

and

[ Q0, Qmn ] = 0, (6.60)

Page 127: Pierre Ramond Group Theory a Physicists Survey 2010

6.7 SU (3) in physics 115

close into the SU (3) Lie algebra with its eight generators expressed in terms of itsSO(3) sub-algebra. We say that its adjoint representation splits in terms of SO(3)representations as

8 = 3 + 5 + 1.

It is easy to verify that all these generators commute with the Hamiltonian of theisotropic harmonic oscillator

[ H, Qmn ] = 0. (6.61)

We were too quick in our previous pass at identifying its invariance group; we seethat it is the much larger SU (3). It follows that all excited levels of this systemform representations of SU (3): since there are six states at the next level, theremust be a six-dimensional representation of SU (3), and so on.

6.7.2 The Elliott model

Think of the nucleus as made up of N nucleons moving independently in a har-monic oscillator potential. Each nucleus is located at x (a)

i with momentum p(a)i ,

where i = 1, 2, 3, and a = 1, 2, . . . N . We convert these in terms of creation andannihilation operators, and write the Hamiltonian as

H0 =N∑

a=1

3∑i=1

a† (a)i a(a)

i + 3N

2. (6.62)

It commutes with the nine generators

N∑a=1

a† (a)i a(a)

j , (6.63)

which form the SU (3) × U (1) algebra. At this stage the energy levels are d-folddegenerate, where d is the dimension of any SU (3) representation. This degen-eracy is broken when inter-nucleon interactions are taken into account. They aredescribed by one two-body potential which we approximate as

V (�x (a) − �x (b)) ≈ m2(�x (a) − �x (b))2 + λ (�x (a) − �x (b))4, (6.64)

with a harmonic potential and an anharmonic term. The Hamiltonian now becomes

H = H0 + c

2

3∑i, j=1

Qi j Qi j , (6.65)

Page 128: Pierre Ramond Group Theory a Physicists Survey 2010

116 SU (3)

where c is some constant and

Qi j =N∑

a=1

Q(a)i j , (6.66)

where

Q(a)i j = 1

2(x (a)

i x (a)i + p(a)

i p(a)i ), (6.67)

is the quadrupole moment of the ath nucleon. This is the Elliott model (J. P. Elliott,Proc. R. Soc. Lond. A245, 128 (1958)). Physically, the long-range interactiondepends on the shape: when one nucleon bulges in one direction and acquires alarge quadrupole moment, it interacts most with other nucleons which also have alarge quadrupole moment in the same direction. It also breaks the SU (3) × U (1)symmetry of the harmonic approximation.

As we have seen, these are SU (3) generators, so that this perturbation just splitsthe degeneracy among members of each SU (3) irrep, without generating mixingbetween different representations. The combination

C2 =3∑

i=1

Li Li + 1

2

3∑i, j=1

Qi j Qi j (6.68)

is nothing but the quadratic SU (3) Casimir operator. This allows us to rewrite theHamiltonian as

H = H0 + c

2

(C2 − L(L + 1)

), (6.69)

where L is the angular momentum. For a given SU (3) irrep, C2 depends on twointegers, and L occurs in a range of values L = 0, 1, 2, . . . , Lmax, where Lmax isthe value of the magnetic quantum number for the highest weight state, which hasthe largest deviation from a pure sphere. The d-fold degeneracy is thus split intoLmax + 1 levels.

Following Dyson [4], consider the magnesium nucleus Mg24, described by eightnucleons in six possible orbits (L = 0 and L = 2 unfilled shells) outside aclosed shell. In addition each nucleon has two possible spin values (±1/2), andtwo charges (zero and one), resulting in 6×4 = 24 possibilities. Hence the numberof configurations for the eight nucleons is simply

24!8! 14! ,

more than a mere million. Diagonalizing such a large Hamiltonian has led nuclearphysicists either to madness, or to treat the nuclear levels statistically and developthe theory of random matrices. This simple-minded model brings tremendous

Page 129: Pierre Ramond Group Theory a Physicists Survey 2010

6.7 SU (3) in physics 117

simplification to the problem. Comparing the measured energy levels with thosecalculated from the Elliott model, assuming only one SU (3) irrep, we find that70%–90% of the energy levels are correctly reproduced!

6.7.3 The Sakata model

With the advent of strange particles, the taxonomy of “elementary particles” hadto be extended. Sakata (Prog. Theor. Phys. 16, 686 (1956)) suggested that theisospin doublet be augmented by one new spin one-half neutral nucleon whichhe dubs � and its antiparticle. Such a particle had been discovered through itsdecay into a proton and a π+, leaving tracks bending in opposite ways, hence itsname.

In group-theoretical language, this amounts to generating the SU (3) × U (1)algebra through

Ti j = b†i b j − b†

j bi , i, j = 1, 2, 3, (6.70)

by including a third Fermi oscillator. The nucleons and antinucleons now form aSU (3) triplet and antitriplet, respectively. The electric charge resides in both SU (3)and U (1), as

Q = I3 + 1√3

I8 + 1

3N , (6.71)

where N is the nucleon number and I8 = (1, 1,−2)/(2√

3) is the second diagonalSU (3) generator for the triplet.

Pions are found to reside in the adjoint octet representation

3 × 3 = 8 + 1. (6.72)

In the decomposition

SU (3) ⊃ SU (2) × U (1), (6.73)

it contains

8 = 30 + 21 + 2−1 + 10, (6.74)

where the subscript denotes the U (1) hypercharge Y = 2√

3I8. The isodoublets

are the strange kaons, (K 0, K +) and (K −, K0), and the extra singlet is the η

meson. They are the pseudoscalar mesons. In picture form, this is represented asfollows.

Page 130: Pierre Ramond Group Theory a Physicists Survey 2010

118 SU (3)

�π−

�π+

�η

�π0

�K +

�K 0

�K

0

�K −

For the octet, the electric charge reduces to the Gell-Mann–Nishijima formula

Q = I3 + Y

2. (6.75)

The Sakata model correctly described the pseudoscalar mesons, but did not accountfor the plethora of baryons, the proton, the heavy unstable spin one-half particles,and spin 3/2 resonances that were being discovered in experiments, almost ona monthly basis at the time. Hence it serves as an early use of group theory tosystematize the taxonomy of elementary particles.

6.7.4 The Eightfold Way

It was Gell-Mann and Y. Ne’eman who independently arrived at the correct assign-ment for the baryons. In a nutshell, they assigned the proton and neutron to thefollowing SU (3) octet representation.

��−

��+

��

��0

�p

�n

��0

��−

Page 131: Pierre Ramond Group Theory a Physicists Survey 2010

6.7 SU (3) in physics 119

All eight have spin one-half, but differ by their masses and their “strangeness” S:the nucleons are not strange (S = 0), the � and � are a little strange (S = −1), andthe � more so (S = −2). All are stable with respect to strong decays (i.e. decayswith a characteristic decay time around 1024 seconds), but take much longer todecay eventually (weakly or electromagnetically) to the neutron.

The neutral � has a mass of 1115 MeV; The three � baryons center in massaround 1190 MeV. The charge decays either as �+ → p + π0 or �+ → n + π+

while �0 decays by photon emission: �0 → �+γ . The heavier “cascade” baryons�0 (1314 MeV) and �− (1321 MeV) decay mostly into � + π .

To account for these mass differences, the assumption was made that theHamiltonian which governs the strong interactions is made up of two pieces: thelargest, H0, is invariant under SU (3) and its contribution is to give the same massto all particles within a multiplet, since H0 commutes with the SU (3) generators.The smaller second piece generates the mass differences. Since the strong interac-tions preserve both isospin and strangeness, this second piece must commute withSU (2)I × U (1)Y , but not with the rest of the SU (3) generators. The first placeto look for an operator with these requirements is the octet representation whichcontains an isospin singlet with Y = 0.

In equations,

H = H0 + a T 33, (6.76)

where the a is some constant, and T cd is an operator which satisfies

[ Xab, T c

d ] = δ cb T a

d − δ ad T c

b. (6.77)

Other choices would have been tensor operators with multiple upper and lowerindices all aligned along the three direction, but this is the simplest.

Although T 33 is an infinite-dimensional operator, it only shuffles states within the

same representation, and it suffices to determine its matrix elements within one rep-resentation. As part of the strong Hamiltonian, the generator of time translations,particles belonging to different SU (3) representations will not mix.

We now show that T 33 in a given representation space can be written as a function

of the SU (3) generators. Consider any SU (3) irreducible representation, labeledby the kets | i 〉, i = 1, 2 . . . , d. We have completeness on the states of the samerepresentation,

1 =∑

i

| i 〉〈 i |, (6.78)

which allows us to write

T 33 =

∑i, j

| i 〉〈 i | T 33| j 〉〈 j |. (6.79)

Page 132: Pierre Ramond Group Theory a Physicists Survey 2010

120 SU (3)

But, we have seen that | i 〉 can be written as a polynomial in the (lowering)generators (X−) acting on a state of highest weight | hw 〉. Hence

T 33 =

∑i, j

〈 i | T 33| j 〉 fi (X−)| hw 〉〈 hw | f j (X+). (6.80)

Now comes Okubo’s trick: introduce P , the projection operator on the state ofhighest weight

P| hw 〉 = | hw 〉, P fi (X−)| hw 〉 = 0. (6.81)

Then

T 33 =

∑i, j

〈 i | T 33 | j 〉 fi (X−)P| hw 〉〈 hw | f j (X+), (6.82)

=∑i, j

〈 i | T 33| j 〉 fi (X−)P

∑k

| k 〉〈 k | f j (X+), (6.83)

=∑i, j

〈 i | T 33| j 〉 fi (X−)P f j (X+). (6.84)

Since the matrix elements 〈 i |T 33| j 〉 are just numbers and P can itself be written

in terms of the generators in the representation, we see that when restricted to arepresentation, the operator T 3

3 is a function of the generators.Hence the most general form is

T 33 = a0δ

33 + a1 X3

3 + a2 (X · X)33 + a3 (X · X · X)3

3 + · · · , (6.85)

which contains an infinite number of parameters, but use of eq. (6.10) reduces it to

T 33 = a0δ

33 + a1 X3

3 + a2 (X · X)33, (6.86)

which involves only three unknown representation-dependent constants. It is nowa simple matter to deduce that

X33 = Y, (X · X)3

3 = constant + 3

2Y + 1

4Y 2 − I (I + 1), (6.87)

leading to the final result

< H >= a + bY + c

(1

4Y 2 − I (I + 1)

), (6.88)

where a, b, c only depend on the representation.The application of this formula to the octet and decuplet representations is wildly

successful. This shift in the Hamiltonian is interpreted (in lowest order perturbation

Page 133: Pierre Ramond Group Theory a Physicists Survey 2010

6.7 SU (3) in physics 121

theory), as a mass shift for the baryons and a squared mass shift for the mesons. Inthe octet with four isospin multiplets, we expect one mass formula, given by

MN + M� = 1

2(3 M� + M� ). (6.89)

Applied to the neutral members of the isospin multiplets, it gives 2254.4 = 2269.7,accurate to less than 1%!

The same mass formula works well for the mass squared of the octet pseu-doscalar mesons

m2K + m2

K= 1

2(3 m2

η + m2π ), (6.90)

with (497)2 = (478)2, which is not as good as for the baryons.In the eightfold way, the spin 3/2 baryon resonances fall into the decuplet repre-

sentation, the symmetric tensor with three upper triplet indices, (its conjugate hasthree lower indices), that is

(3 × 3 × 3)sym = 10. (6.91)

The particle assignment is given as follows.

��∗−

��∗+

��∗0

��+

���0�−

��++

��∗0

��∗−

��−

The isospin 3/2 multiplet contains resonances found in π − N scattering, the threeother isospin multiplets contain strange resonances. We see that in this multipletisospin and strangeness are related I = 1+Y/2, so that we have two mass relations

M�− − M�∗ = M�∗ − M�∗ = M�∗ − M�. (6.92)

Page 134: Pierre Ramond Group Theory a Physicists Survey 2010

122 SU (3)

With (in MeV) M� = 1230, M�∗ = 1385, M�∗ = 1532, and M�− = 1672, theagreement is stunning. In fact this mass formula allowed Gell-Mann to successfullypredict both the existence and mass of the � baryon.

Many more details can be found in Gell-Mann and Ne’eman’s The EightfoldWay (Benjamin, 1964), as well as in S. Okubo’s Rochester Lectures on UnitarySymmetry.

Before leaving this section, we note that the electric charges of the particles in thedecuplet are the same as in the octet. Hence the Gell-Mann–Nishijima charge for-mula is now valid for all SU (3) representations, and the charge operator is a linearcombination of SU (3) generators. With such an assignment, the three “particles”in the triplet are the fractionally charged quarks, u, d, and s.

Page 135: Pierre Ramond Group Theory a Physicists Survey 2010

7

Classification of compact simple Lie algebras

Now that we have looked at several examples and established useful techniques, itis time to present the complete classification (see Humphreys [11]). We define Liealgebras as a finite collection of elements X A, A = 1, 2, . . . , N , that satisfies thefollowing axioms.

(i) It is closed under the antisymmetric operation of commutation

[ X A, X B ] = i f ABC XC . (7.1)

(ii) It obeys the Jacobi identity

[ [ X A, X B ], XC ] + cyclic permutations = 0. (7.2)

The coefficients f ABC are called the structure constants of the Lie algebra. They

are constrained to be antisymmetric under the interchange of A and B,

f ABC = − f B A

C ,

by the first axiom and to satisfy

f ABC f DC

F + f D AC f BC

F + f B DC f AC

F = 0, (7.3)

obtained from the Jacobi identity. We restrict ourselves to the cases where the ele-ments of the Lie algebra are hermitian. Then by taking the hermitian conjugate ofthe first axiom, it follows that the structure functions are real. Let us think of thestructure functions as the (N × N ) matrices

f ABC ≡ i (T A

adj)BC , (7.4)

acting on a N -dimensional real vector space, so that the Jacobi identity translatesinto

(T Aadj T D

adj)BF − (T D

adj T Aadj)

BF = i f AD

C (T Cadj)

BF ,

123

Page 136: Pierre Ramond Group Theory a Physicists Survey 2010

124 Classification of compact simple Lie algebras

after using the antisymmetry of the upper two indices. We recognize this equationas the (B F) matrix element of

[ T Aadj, T D

adj ] = i f ADC T C

adj.

This explicit representation of the Lie algebra in terms of (N × N ) matrices,on a real space with the same dimension as the algebra, is called the adjointrepresentation. The second-rank symmetric matrix

gAB ≡ − f ACD f B D

C = Tr (T Aadj T B

adj), (7.5)

is called the Cartan–Killing form. We restrict ourselves to the study of semi-simpleLie algebras for which its determinant does not vanish. In that case we can defineits inverse with lower indices, such that

gAB gBC = δAC . (7.6)

It acts like a metric, which can be used to lower and raise indices. In particular, wecan define the structure function with all three upper indices

f ABC ≡ f ABD gC D = − f AB

D f C EF f DF

E ,

which, after use of the Jacobi identity, can be rewritten as

f ABC = −iTr (T Aadj T B

adj T Cadj − T B

adj T Aadj T C

adj),

showing it to be manifestly antisymmetric in all three indices.

7.1 Classification

These axioms severely constrain the structure functions. In this section, we derivethe allowed Lie algebras. Consider a Lie algebra with r commuting generators Hi ,

[ Hi , H j ] = 0, (7.7)

with i, j = 1, 2, . . . , r , that is f i jC = 0. Since they mutually commute, the Hi

can be represented by diagonal matrices. In the adjoint representation, we set

−i f iab = (T i

adj)ab ≡ β i (a) δa

b, (7.8)

where a, b, . . . label the remaining (N − r) elements of the Lie algebra. Sincethe generators are hermitian, β i (a) are real, but otherwise undetermined. It followsfrom the basic commutation relations that

[ Hi , Xa ] = β i (a) δab Xb = β i (a) Xa. (7.9)

Page 137: Pierre Ramond Group Theory a Physicists Survey 2010

7.1 Classification 125

Every Xa is characterized by the r coefficients (β1(a), β2(a), . . . , βr (a)). It is avector in an r -dimensional vector space, β(a), called the root associated with theelement Xa . It is convenient to relabel these in terms of the roots:

Xa ≡ Eβ(a), (7.10)

assuming no two elements with the same root. This yields for any root vector β,

[ Hi , Eβ ] = β i Eβ . (7.11)

The hermitian conjugate equation is

[ Hi , E †β ] = −β i E †

β , (7.12)

so that the dagger belongs to the negative root and we set

E †β = E−β . (7.13)

Since the negative of each root is also a root, (N −r) is always even. A Lie algebraof rank r and dimension N contains (N − r) elements, each of which is describedby a non-zero vector in r -dimensional space, and its r commuting elements can beviewed as associated with r roots of zero length.

Consider the commutator [ Eβ, Eγ ], where β and γ are roots. Commute it withHi and use the Jacobi identity to find (if not equal to zero)

[ Hi , [ Eβ, Eγ ] ] = (β i + γ i ) [ Eβ, Eγ ],so that the commutator belongs to the root β + γ . Accordingly we write in allgenerality,

[ Eβ, Eγ ] = Nβ, γ Eβ+γ , (7.14)

where the yet undetermined constants satisfy

Nβ, γ = −Nγ ,β . (7.15)

But wait! There are only (N − r) roots, so that not all linear combinations of rootscan be roots. Since all elements of the Lie algebra are spoken for, it must be thatwhenever β + γ is not a root, the commutator vanishes, that is

Nβ, γ = 0, if β + γ not a root. (7.16)

By taking the hermitian conjugate we see that

Nα,β = −N ∗−β,−α = −N−β,−α, (7.17)

since the structure functions are real.

Page 138: Pierre Ramond Group Theory a Physicists Survey 2010

126 Classification of compact simple Lie algebras

If β + γ = 0, the commutator with Hi vanishes, and must therefore reduce to alinear combination of the Hi ,

[ Eβ, E−β ] ≡ βi H i , (7.18)

where βi are undetermined, not to be confused with β i , the components of the rootvector. To relate them, multiply this equation by H j , and take the trace. A fewmanipulations yield

β j Tr(

E−β Eβ

) = βi Tr(Hi H j

). (7.19)

The matrix

gi j ≡ Tr(Hi H j

)(7.20)

is the Cartan–Killing form of eq. (7.5). It is invertible for a semi-simple algebra,acts as a metric for the r -dimensional root space, relates covariant and contravariantcomponents of the root vectors, and is used to define a scalar product. The trace onthe left-hand side is arbitrarily set to one,

Tr(

E−β Eβ

) = 1, (7.21)

for each β. Thus

β j = βi gi j ; βi = gi j βj . (7.22)

This leads us to define the scalar product in the real r -dimensional root space as

(α, β ) = (β, α ) ≡ αi β j gi j = αi β j gi j . (7.23)

One can show that the length of all these vectors is positive definite.We are now in a position to determine the coefficients Nβ, γ , by clever repeated

use of the Jacobi identity. For three roots related by

α + β + γ = 0,

we easily see that

[ Eα, [ Eβ, Eγ ] ] = αi Nβ, γ Hi ,

Taking cyclic permutations, the Jacobi identity yields(αi Nβ, γ + βi Nγ ,α + γi Nα,β

)Hi = 0. (7.24)

Multiply by H j and take the trace, to get the constraint

αi Nβ, γ + β i Nγ ,α + γ i Nα,β = 0, (7.25)

which leads to the identities

Nα,β = Nβ,−α−β = N−α−β,α. (7.26)

Page 139: Pierre Ramond Group Theory a Physicists Survey 2010

7.1 Classification 127

A pair of integers p and q can be naturally associated to any two roots in thefollowing way: take Eα and commute it k times with Eβ , obtaining the α-chain ofroots

[ Eβ, [ Eβ, [ · · · , [ Eβ, Eα ] · · · ] ] ],which either yields an element belonging to the root α+k β, or else zero. Since theLie algebra contains a finite number of elements, there has to be a maximum valuekmax = q, after which further commutation yields zero, so that α + (q + 1)β is nota root. Similarly, we can repeatedly commute E−β with Eα, and there must be aninteger p such that α − (p + 1)β is not a root, with the vanishing of the multiplecommutator after (p + 1) steps.

Now we form the Jacobi identity for the three elements belonging to the rootsβ, −β, and α + k β, (with α �= β), and evaluate the commutators with Hi . Theresult is

βi (αi + k β i ) Eα+kβ = [ Eβ, [ E−β, Eα+kβ ] ] + [ E−β, [ Eα+kβ, Eβ ] ]. (7.27)

Evaluation of the double commutators yields

βi (αi + k β i ) = Nβ,α+(k−1)β N−β,α+k β + N−β,α+(k+1)β Nα+k β,β . (7.28)

Using (7.26), and the antisymmetry of the N symbols, it is rewritten as

βi (αi + k β i ) = F(k) − F(k − 1), (7.29)

where

F(k) = Nβ,α+k β N−β,−α−k β, (7.30)

and −p ≤ k ≤ q. Since α + (q + 1)β is not a root, it follows from (7.27) that

F(q) = 0, F(q − 1) = −βi (αi + q β i ). (7.31)

The recursion relation (7.29) is then easily solved

F(k) = (k − q) βi [αi + 1

2(k + q + 1)β i ]. (7.32)

At the lower end of the chain, the use of (7.26) yields

F(−p − 1) = 0, (7.33)

so that

(p + q + 1) [βiαi + 1

2(q − p)βiβ

i ] = 0. (7.34)

Since both p and q are positive integers, we deduce that

2βi αi

β j βj

= 2(β, α)

(β, β)= p − q ≡ n, (7.35)

Page 140: Pierre Ramond Group Theory a Physicists Survey 2010

128 Classification of compact simple Lie algebras

is always an integer. This equation, central to Lie algebra theory, has manyconsequences.

First, by setting k = 0 in (7.32), and using (7.17), we obtain a neat expressionfor the structure constants,

(Nα,β)2 = 1

2q (p + 1) (β, β ). (7.36)

It also means that

α′ = α − (p − q)β = α − 2βi αi

β j βj

β, (7.37)

is always a root. The component of α′ along β, is the negative of the projection ofα along the same direction, since

(α′, β ) = −(β, α ). (7.38)

This suggests a nice geometrical interpretation in root space: α′ is the Weyl reflec-tion of α about the hyperplane perpendicular to β. Weyl reflections generate rootsof the same length, since

(α′, α′ ) = (α, α ). (7.39)

A further consequence is a set of restrictive values for the angle between tworoots. To see this, note that we could have interchanged α and β, and arrived at aresult with the same structure, namely

2(α, β )

(α, α )= p′ − q ′ ≡ n′,

where p′ and q ′ are the relevant integers for the chain β + kα. Clearly n and n′

must have the same sign, since

n

n′ = (α, α )

(β, β ). (7.40)

In addition2(β, α )

(α, α )

2(β, α )

(β, β )= n n′, (7.41)

from which

(β, α )2 = (α, α ) (β, β )n n′

4. (7.42)

As a result, the angle �αβ between the two roots is restricted by

cos2 �αβ = n n′

4≤ 1. (7.43)

Page 141: Pierre Ramond Group Theory a Physicists Survey 2010

7.2 Simple roots 129

If cos2 �αβ = 1, the two roots are proportional to one another, but this runs intothe following theorem: if α and aα are roots, then a = ±1, but we assumed α �= β

from the onset. Hence, any two roots of a semi-simple Lie algebra must satisfy theconstraints in the following table:

n n′ � Length ratio

0 0 π/2 not fixed1 1 π/3 1

−1 −1 2π/3 11 2 π/4

√2

−1 −2 3π/4√

21 3 π/6

√3

−1 −3 5π/6√

3

so that always

4(β, α )2

(α, α )(β, β )= 0, 1, 2, 3, (7.44)

and since |n|, |n′| are at most three, there are no chains with five roots or more.

7.2 Simple roots

The expressions derived above apply to any two arbitrary roots α and β. However,the non-zero (N − r) roots live in an r -dimensional space, implying many lineardependencies. For SU (3), we had split them into an equal number of positive andnegative roots. Although our notion of positivity was basis-dependent, the finalresults were unaffected.

Repeating this procedure, let us order the components of any root along H 1, H 2,. . . , Hr . A root is then said to be positive if its first non-zero component along theordered axes is positive. This is fine since the negative of every non-zero root is alsoa root. Not all (N − r)/2 positive roots are independent. We introduce r linearlyindependent simple roots, αi , i = 1, 2, . . . , r , which are positive roots with theproperty that none can be expressed as linear combination of other simple rootswith positive coefficients.

We now apply the results of the preceding section to chains of simple roots. Letαi and α j be two simple roots, and form their difference αi − α j , which is eitherpositive, negative or zero. If positive, or negative, we see that the simple rootssatisfy αi = (αi − α j ) + α j , or α j = (α j − αi ) + αi , contrary to their definition.Hence the difference of two simple roots must be zero.

Page 142: Pierre Ramond Group Theory a Physicists Survey 2010

130 Classification of compact simple Lie algebras

It follows that any chain constructed out of two simple roots has p = 0, sothat (αi , α j ) ≤ 0, and the angle between simple roots has only three possiblevalues, 2π/3, 3π/4, and 5π/6. This leads to only three chains built out of simpleroots

� �αi αi +α j

� �αi αi +α j �αi +2α j

� �αi αi +α j �αi +2α j �αi +3α j

for q = 1, 2, 3, respectively. Only the first root is simple. For all three chains, wefind

(αi + α j , αi + α j )

(αi , αi )= 1

q, (7.45)

for a chain starting at the simple root αi and ending at αi + q α j .It is convenient to normalize each element of the Lie algebra as

eα ≡√

2

(αi , αi )Eα, (7.46)

and write those corresponding to the simple roots as ei , i = 1, 2 . . . , r . Then

[ ei , e−i ] = 2

(αi , αi )(αi ) j H j , (7.47)

where (αi ) j is the j th component of αi . In terms of

hi = 2

(αi , αi )(αi ) j H j , (7.48)

we have

[ ei , e−i ] = hi , (7.49)

and

[ hi , ei ] = 2 ei . (7.50)

Page 143: Pierre Ramond Group Theory a Physicists Survey 2010

7.3 Rank-two algebras 131

Thus each of the r sets ei , e−i and hi , generates an SU (2) algebra. However, theirelements do not commute with one another:

[ hi , e j ] = 2(αi , α j )

(α j , α j )e j , (7.51)

= A ji e j , (7.52)

written in terms of the (r × r) Cartan matrix

A ji ≡ 2(αi , α j )

(α j , α j ), (7.53)

with negative integer off-diagonal elements, −1, −2, and −3. We also have

[ eα, eβ ] = N ′α,β eα+β, (7.54)

with

N ′α,β ≡

√2(α + β, α + β )

(α, α ) (α, β )Nα,β . (7.55)

Now, specialize this expression to the commutator of two simple roots, for whichp = 0; use of eqs. (7.36) and (7.45) then yields the simple result

[ ei , e j ] = ± eαi +α j. (7.56)

What of the commutator of elements which are not simple roots? It turns out thatfor any two roots α and β, eq. (7.45) generalizes to

(α + β, α + β )

(α, α )= p + 1

q, (7.57)

which leads to

[ eα, eβ ] = ± (p + 1) eα+β . (7.58)

This is the Chevalley basis, where all structure constants are integers. The signambiguities can be determined with more work (see R. W. Carter, Simple Groupsof Lie Type (John Wiley & Sons Ltd, London)). This is the Chevalley basis whichwe alluded to in the context of SU (3); it is central to the construction of a plethoraof finite simple groups.

7.3 Rank-two algebras

With the restrictions just derived, it is straightforward to build the root space ofrank-two Lie algebras. The two zero roots are situated at the origin of the two-dimensional root space. The procedure is to align the first simple root along thehorizontal positive axis. Its mirror image about the vertical negative represents its

Page 144: Pierre Ramond Group Theory a Physicists Survey 2010

132 Classification of compact simple Lie algebras

negative root. We then draw the second simple root at the prescribed length andangle. The rest of the roots are generated by Weyl reflections. Depending on theorientation of the second simple root, we have the following possibilities.

(i) Draw the second root at π/2 with arbitrary length. The Weyl reflections in this caseproduce the negative roots, and we obtain the following root diagram.

��

����

The two simple roots are indicated by the dark circles. Since Weyl reflections do notrelate the roots on the vertical with those on the horizontal, we recognize two SU (2)algebras, with their roots along the horizontal and vertical axes, respectively. This isthe root diagram of the semi-simple algebra SU (2) × SU (2).

(ii) The second root is drawn at 2π/3 from the first with the same length as the first.Draw their negatives, and start the Weyl reflections about the various perpendiculars.We generate three non-zero roots and their negatives, yielding the following SU (3)root diagram.

Note that it contains on its horizontal axis the three roots of the SU (2) root spaces aswell as a singlet root and two pairs of roots, one above, one below. This simply saysthat the SU (3) adjoint representation contains an SU (2) triplet (its adjoint), a singletand two doublets. This diagram can be rotated by π/3 without affecting its structure.

(iii) The second simple root is now at an angle of 3π/4 to the first, with length 1/√

2 that ofthe first. Add their negatives, and proceed by Weyl reflections. Reflect the small rootabout the vertical axis, then about the horizontal axis, then about the vertical axis. This

Page 145: Pierre Ramond Group Theory a Physicists Survey 2010

7.3 Rank-two algebras 133

yields four small roots. Further Weyl reflections generate no new roots. Now reflectthe large root about the hyperplane perpendicular to the small roots. This yields alarge root on the vertical axis. This process yields two more large roots on the axes.Hence we have obtained by Weyl reflection eight non-zero roots. This corresponds,adding the two zero-length roots, to a ten-dimensional algebra, called B2 or C2. If wehad started with the small root on the horizontal axis, we would have obtained thesame root structure, rotated by π/4.

�� ��

��

We recognize that this root diagram contains the diagram of two independent SU (2).It also contains roots that look like doublets under both SU (2)s. Hence we infer thatthis algebra contains the six elements of SU (2) × SU (2) and four elements whichtransform as a doublet-doublet, that is a doublet under each!

(iv) Finally, draw the second root at an angle of 5π/6 to the first. Take it to be the shorterone, with length (3)−1/2 times that of the first. Draw their negatives, and proceed tobuild the root diagram by a judicious succession of Weyl reflections. With two zeroroots, the result is the 14-dimensional Lie algebra G2.

���

���

� �

��

��

Page 146: Pierre Ramond Group Theory a Physicists Survey 2010

134 Classification of compact simple Lie algebras

We recognize that the long roots form the SU (3) root diagram, so that G2 containsSU (3), but we also discern two SU (2)s, so that it also contains SU (2) × SU (2).

This exhausts all possible rank-two semi-simple Lie algebras. All have roots ofat most two different non-zero lengths. This is a general feature: the three possiblelength ratios, 1 : 1, 1 : √

2, 1 : √3 imply that an algebra with all three lengths

would contain roots in the ratio√

2 : √3, contrary to our general results.

7.4 Dynkin diagrams

For higher-rank algebras, it is clearly not practical to describe the roots by draw-ing them in r -dimensional space. Fortunately, all the relevant information can bedisplayed in two-dimensional patterns for simple Lie algebras of any rank.

A Dynkin diagram is an assembly of r dots, one for each simple (positive) root.When the algebra contains roots of different lengths, short roots are indicated byan open dot, long roots by a filled dot.

It is convenient to work with the largest roots normalized to unit length

( αi , αi ) = 1, i = 1, 2, . . . , r.

As we have seen, the angle between two simple roots can take on four possiblevalues. The rules for their construction are as follows.

(i) Dots representing roots which satisfy

4( αi , α j )2 = 0, 1, 2, 3, i �= j,

are connected by zero, one, two or three lines, respectively.In the latter two cases, they must have different lengths, as indicated below.

� � � � � �These are the Dynkin diagrams of the rank-two algebras we just constructed.

(ii) Removing dots (roots) from a Dynkin diagram generates Dynkin diagram(s) associatedwith lesser-rank Lie algebra(s) (simple or semi-simple).

These simple rules are sufficient to derive the Dynkin diagrams of all simple Liealgebras.

Consider the vector

α =r∑i

αi .

Page 147: Pierre Ramond Group Theory a Physicists Survey 2010

7.4 Dynkin diagrams 135

Its norm

(α, α ) =r∑i

( αi , αi ) + 2r∑

i< j

( αi , α j ),

is strictly positive, that is

2r∑

i< j

( αi , α j ) <

r∑i

( αi , αi ) = r,

and the largest possible number of connected pairs by one line is (r − 1).

• The number of connected pairs of roots is at most (r − 1).Assume a Dynkin diagram with a closed cycle. Remove all dots outside the cycle,

generating by rule (ii) a Dynkin diagram with s roots on the cycle, and s connectedpairs, describing a rank s algebra! But we just proved that the number of connectedpairs must be less than the rank. Hence we reach the following conclusion.

• There are no Dynkin diagrams with closed cycles.Let α be a root connected to n roots β i , i = 1, 2, . . . , n, so that

( α, α ) = ( β i , β i ) = 1, ( α, β i ) < 0.

The absence of closed cycles implies

( β j , β i ) = 0, i �= j.

The vector orthogonal to all β i s

α −n∑i

( α, β i ) β i ,

has a positive norm, yielding the inequality

n∑i

( α, β i )2 < 1,

but the left-hand side is simply one-quarter the number of lines emanating from the rootα. Hence

number of lines = 4n∑i

( α, β i )2 < 4.

It implies the following.

Page 148: Pierre Ramond Group Theory a Physicists Survey 2010

136 Classification of compact simple Lie algebras

• No more than three lines can originate from one root.This limits the allowed multiple connectors to four types.

� � � � � �

� �� ����

���

�When two roots are connected by three lines, they cannot be further connected to otherroots. Hence this last diagram stands by itself as the Dynkin diagram of the rank twoalgebra G2

G2� �with Cartan matrix

AG2=

(2 −3

−1 2

).

Consider a Dynkin diagram with a linear chain of roots, {β1, β2, . . . , βn, }, connectedto one another by one line, that is

( β i , β j ) = −1

2, i �= j = 1, 2, . . . , n.

Remove these n roots from the diagram, and replace them with one root described bythe unit vector

β =n∑i

β i .

Let α be a root that was connected to the chain in the original Dynkin diagram; it waslinked to at most one of βk of the chain, otherwise there would have been a cycle, that is

( α, β ) = ( α, βk ),

and we see that β has the same projection along its connected root as the original rootβk , and thus satisfies all the axioms to be a root of the same length as those in theoriginal chain.

Page 149: Pierre Ramond Group Theory a Physicists Survey 2010

7.4 Dynkin diagrams 137

• Replacing a linear chain of roots by one root generates a Dynkin diagram.As a result, there is no Dynkin diagram with a linear chain connecting to two differentroots, each connected by two lines to the rest of the diagram; otherwise we could shrinkthe chain and put the two pairs together, resulting in a contradiction of a root with fourlines coming out of it.

Consider a Dynkin diagram with two roots αn and βm connected by two lines, so that

( αn, βm )2 = 1/2.

Each is connected with its own linear chain, one of length n with roots αi , the other oflength m, with roots β p. For i < n, p < m,

( αi , αi+1 ) = ( β p, β p+1 ) = −1

2, ( αi , β p ) = 0.

The vectors

α =n∑

j=1

j α j , β =m∑

p=1

p β p,

have norms

(α, α ) =n∑j

j2 −n−1∑

i

j ( j + 1) = n(n + 1)

2,

(β, β ) =m∑p

p2 −m−1∑

p

p(p + 1) = m(m + 1)

2.

Since the roots in the two chains are orthogonal except for those at the end, we have

(α, β )2 = m2n2 ( αn, βm )2 = m2n2

2,

which, by Schwartz’s inequality, must be less than the product of their norm,

m2n2

2<

n m (n + 1) (m + 1)

4,

that is

(n − 1)(m − 1) < 2.

This equation has three solutions.– When n = m = 2, we find F4, the rank-four algebra with the following Dynkin

diagram.

F4

� �

Page 150: Pierre Ramond Group Theory a Physicists Survey 2010

138 Classification of compact simple Lie algebras

– When n = 1, and m arbitrary, there is only one linear chain connected to the longroot. This generates the Lie algebra Br or SO2r+1 with the following Dynkin dia-gram.

Br

�– With m = 1, and n arbitrary, the chain is connected to the short root. The resulting

Lie algebra is called Cr , or Sp2r with the following Dynkin diagram.

Cr

� � � � The two algebras obviously have the same Dynkin diagram when r = 2, as we havepreviously noted.

Now consider a Dynkin diagram with a root connected to three linear chains oflengths m, n, p with roots αi , β j , and γ k . The three chain vectors

α =n−1∑

i

i αi , β =m−1∑

j

j β j , γ =p−1∑

k

k γ k,

have norms

(α, α ) = n(n − 1)

2, (β, β ) = m(m − 1)

2, ( γ , γ ) = p(p − 1)

2.

Let δ be the root that is connected by one line to αn−1, βm−1 and γ p−1. Now,

( δ, α )2 = (n − 1)2( δ, αn−1 )2 = (n − 1)2

4,

so that the angle between δ and α is restricted to be

cos2 �δ α

= 1

2− 1

2n.

Similarly for the other chain vectors,

cos2 �δ β

= 1

2− 1

2m, cos2 �

δ γ= 1

2− 1

2p.

The vector perpendicular to all three normalized chain vectors α, β and γ is

δ − ( δ, α ) α − ( δ, β ) β − ( δ, γ ) γ ,

Page 151: Pierre Ramond Group Theory a Physicists Survey 2010

7.4 Dynkin diagrams 139

and the positivity of its norm requires

( δ, α )2 + ( δ, β )2 + ( δ, γ )2 < 1.

Putting it all together, the three integers m, n, p are restricted by the inequality

3

2− 1

2n− 1

2m− 1

2p< 1,

or

1

n+ 1

m+ 1

p> 1. (7.59)

This inequality has four types of solutions, one where two of the integers have the samevalue and three others where all assume different values.– One solution leaps to the eye: if m = p = 2, this equation has solutions for arbitrary

n, leading to the following Dynkin diagram of rank n + 2 algebras, Dn+2, SO2n+4 tophysicists.

Dn+2

���

���

It is symmetric under the interchange of the two antennas at the end.

– When p = 2, the inequality reduces to the very famous Diophantine equation

1

n+ 1

m>

1

2. (7.60)

It has a remarkable geometrical meaning: let n be the number of m-polygons thatmeet at any vertex. Each m-gon comes with an angle π − 2π/m, since the sum of allthe angles they make at the vertex must be less than 2π , we get the inequality

n

(π − 2π

m

)< 2π,

which is the same as eq. (7.60), and its solutions are in one-to-one correspondencewith the platonic solids.

– The tetrahedron case, with m = n = 2, yields the following Dynkin diagram of therank-six exceptional algebra E6.

Page 152: Pierre Ramond Group Theory a Physicists Survey 2010

140 Classification of compact simple Lie algebras

E6

� � � � �

– The octahedron and cube correspond to n = 3, m = 4, resulting in the followingDynkin diagram of the rank-seven exceptional algebra E7.

E7

� � � � � �

– Finally the icosahedron, n = 5, m = 3 yields the rank-eight exceptional Lie algebraE8, with the following Dynkin diagram.

E8

� � � � � � �

Lastly, roots of equal length can always be connected by one line, producing theDynkin diagram of Ar or SUr+1.

Ar

This exhausts all possible Dynkin diagrams: Lie algebras come as members of thefour infinite families An , Bn , Cn , and Dn , n = 1, 2 . . ., or one of the five exceptionalalgebras, G2, F4, E6, E7, and E8.

7.5 Orthonormal bases

The information included in Dynkin diagrams and/or Cartan matrices can alsobe summarized by expressing the roots in terms of an orthonormal basis in theroot space. Introduce an orthonormal basis, traditionally denoted by l vectors ni ,satisfying

( ni , n j ) = δi j . (7.61)

The roots of all the simple Lie algebras are then expressed as linear combinations.

Page 153: Pierre Ramond Group Theory a Physicists Survey 2010

7.5 Orthonormal bases 141

• An = SU (n + 1). In this case the roots all have the same length with the simple rootssatisfying

(αi , αi ) = 2, (αi , αi+1 ) = −1, i = 1, 2, . . . , n,

which suggest that we define (n + 1) orthonormal vectors ei for i, j = 1, 2, . . . , n + 1,and set

simple roots: ni − ni+1.

All the roots lie in the plane orthogonal to the vector∑

ni . There are (n2 − n − 1)non-zero roots, half of them positive

roots: (ni − n j ), i �= j;positive roots: (ni − n j ), i < j.

We define (for later use) the Weyl vector as half the sum of all positive roots

ρ ≡ 1

2

∑pos roots

α = (n1 − nn+1).

• Bn = SO(2n + 1). The roots all have the same length but one which is shorter. Theycan be written in terms of n orthonormal unit vectors ni , i = 1, 2, . . . , n with

roots:

{±ni ,

±(ni ± n j ),(7.62)

simple roots: nn, ni−1 − ni ,

with the short root of unit length, the long roots of length√

2. An easy way to get thisform is to start from the linear chain of n−1 long roots, which are given by the A-series,to which we add one root of unit length that is perpendicular to all but the last one in thechain, and that is nn .

• Cn = Sp(2n). This case is the same as the previous one except that the short and longerroots are interchanged, resulting in

roots:

{±2 ni ,

±(ni ± n j ),(7.63)

simple roots: 2 nn, ni−1 − ni , i = 2, . . . , n,

where the long roots are of length√

2, while the short root has unit length.• Dn = SO(2n). We have n orthonormal vectors ni , i = 1, 2, . . . , n, with

roots: ±(ni ± n j );simple roots: nn−1 + nn, ni−1 − ni .

This set of simple roots is obtained from the linear chain of the A-series, to whichone must add a root of the same length that is perpendicular to all but the next to lastroot in the chain, and that is nn−1 + nn .

Page 154: Pierre Ramond Group Theory a Physicists Survey 2010

142 Classification of compact simple Lie algebras

• G2. Its roots are best expressed in terms of three unit vectors ni , i = 1, 2, 3, with

roots:

{±(ni − n j ),

±(2ni − n j − nk), i �= j �= k;simple roots: (n1 − n2),−(2 n1 − n2 − n3).

• F4. In this case we need only four unit vectors to express the roots

roots:

⎧⎪⎪⎨⎪⎪⎩±ni ,

± 12 (n1 ± n2 ± n3 ± n4),

±(ni ± n j ), i �= j = 1, 2, 3, 4;simple roots: (n2 − n3), (n3 − n4), n4,

12 (n1 − n2 − n3 − n4).

In the above, we recognize the simple roots of B3, to which we have added the shortroot (n1 − n2 − n3 − n4)/2 that is perpendicular to all but n4.

• E8. It is easy to express its simple roots, if we notice that by deleting one root at theright end of its Dynkin diagram, we obtain that of D7, for which we already know howthe simple roots are expressed. All we have to do is to add one of the same length thatis perpendicular to all but the last one, n6 − n7 in our notation. The result is

simple roots: n6 + n7, ni−1 − ni , i = 2, . . . , 7; 12

(n8 + n7 −

6∑i=1

ni

).

• E7. This one is obtained simply from the previous case by chopping off the root at theleft of the Dynkin, resulting in

simple roots: n6 + n7, ni−1 − ni , i = 3, . . . , 7; 12

(n8 + n7 −

6∑i=1

ni

).

These are still in an eight-dimensional root space of E8. If one wants to express the rootsin seven dimensions, the last root is given by n8/

√2 + (n7 − ∑6

i=2 ni )/2, which doesnot have rational coefficients. Hence it is best to leave its orthonormal basis in eightdimensions.

• E6. Take out one more root to get

simple roots: n6 + n7, ni−1 − ni , i = 4, . . . , 7; 12

(n8 + n7 −

6∑i=1

ni

).

As in E7, we need to be in eight dimensions to express its last root with rationalcoefficients.

Page 155: Pierre Ramond Group Theory a Physicists Survey 2010

8

Lie algebras: representation theory

Now that we have determined all simple and semi-simple Lie algebras, the nextstep is to represent them in Hilbert spaces, where connections with physics may beestablished. The technique for building irreducible representations is rather simpleto state: identify a highest weight state and then generate its other states by therepeated application of lowering operators until the process terminates, and yourun out of steam. We have seen that this process is very efficient for SU (2) whichhas only one lowering operator. However, a rank-r Lie algebra has r independentnon-commuting annihilation operators, and the process becomes rather unwieldy,and a systematic approach must be adopted.

8.1 Representation basics

We work in the Dynkin basis where all states are labeled by r integers (positive,negative or zero), denoted by the ket

| λ 〉,

where λ is a vector whose components are the Dynkin labels of the state

λ = ( a1 a2 . . . an ). (8.1)

To each simple root, associate a lowering operator T (i)− which, together with its

hermitian conjugate, generates an SU (2) algebra

[ T (i)+ , T (i)

− ] = T (i)3 , i = 1, 2 . . . , r. (8.2)

We have

T (i)3 | λ 〉 = ai | λ 〉. (8.3)

143

Page 156: Pierre Ramond Group Theory a Physicists Survey 2010

144 Lie algebras: representation theory

The action of the lowering operators on this state yields another state withlabel λ − αi

T (i)− | λ 〉 ∼ | λ − αi 〉. (8.4)

When all ai are positive, the state is said to be inside the Weyl chamber of thealgebra’s lattice, or on its wall if any of the entries are zero. A state inside or on thewalls of the Weyl chamber is said to be dominant; denote it a capitalized vector

= ( a1 a2 . . . an ), ai ≥ 0.

It is annihilated by the r raising operators

T (i)+ | 〉 = 0, i = 1, 2 . . . , r, (8.5)

since the commutator of two raising operators is also a raising operator. Any domi-nant weight generates a representation of the algebra by repeated application of thelowering operators. For this reason, it is called the highest weight state of the rep-resentation. Representations of the Lie algebra are in one-to-one correspondencewith the states inside or on the walls of the Weyl chamber. This is in contrast withthe physics way to label representations in terms of the Casimir operators of thealgebra.

Consider a highest weight state with entry ai . We can subtract the simple rootαi ai times, by applying T (i)

− = E(−αi ) as many times. The simple roots areexpressed in Dynkinese by the rows of the Cartan matrix.

The number of applications of the annihilation operators to generate all states ofa representations is called the level or height of the representation. It is given by thedot product of the Dynkin indices of the highest weight state with the level vectorR = (R1, R2, . . . , Rr ), whose components are given by twice the sum of the rowsof the inverse Cartan matrix

Ri = 2∑

j

A−1i j . (8.6)

The fundamental representations are those which are labeled by the Dynkin labels( 0p 1 0n−p−1 ), where p = 1, 2, . . . n, and 0p is shorthand for p zero entries.

8.2 A3 fundamentals

In this section, we use this technique to build the fundamental representations ofA3. There are three fundamental representations, since its Dynkin diagram is alinear chain with three nodes. Its three simple roots, expressed in the Dynkin basis,are the rows of the Cartan matrix (1 stands for −1)

α1 = ( 2 1 0 ), α2 = ( 1 2 1 ), α3 = ( 0 1 2 ).

Page 157: Pierre Ramond Group Theory a Physicists Survey 2010

8.2 A3 fundamentals 145

From its Cartan matrix, we find the level vector

RSU (4) = (3, 4, 3), (8.7)

whose entries give the levels of the fundamental representations.The highest weight state of the first fundamental representation is ( 1 0 0 ). With

only one positive entry in the first spot, the only way to lower it is by appli-cation of the operator associated with α1. This produces the state labeled by( 1 1 0 ) = ( 1 0 0 ) − ( 2 1 0 ). This new state has a positive entry in the secondspot: we apply T (2)

− which subtracts the second root α2. The result is a new state( 0 1 1 ) = ( 1 1 0 ) − ( 1 2 1 ), with positive unit entry in the third spot. Then weapply T (3)

− which subtracts the third simple root to get ( 0 0 1 ) = ( 0 1 1 )− ( 0 1 2 ).We are done because this state has no positive entries: application of any loweringoperator gives zero. We have generated the four-dimensional representation of A3.We observe that it has three levels, as expected.

Let us apply the same procedure to ( 0 1 0 ). We first subtract α2 to get ( 1 1 1 ).We now have positive entries in the first and third spots, so we need to subtract withα1 and α3, to produce two states. Then, as shown in the figure, the next level ofsubtractions produces the same state, so that the diagram is spindle-shaped. This isa general feature of Lie algebra representations. At the fourth level we finally arriveat a state without positive entries and the representation terminates. The height ofthis representation is four and it contains six states.

The highest weight state of the third fundamental is ( 0 0 1 ). The first level isobtained by subtracting α3, yielding ( 0 1 1 ). Subtraction by α2 produces ( 1 1 0 );finally subtraction by α1 yields the lowest weight state ( 1 0 0 ); as expected it hasthree levels.

These constructions are summarized by the following diagrams.

SU(4) fundamental representations

( 1 0 0 )

−α1

−α2

−α3

( 1 1 0 )

( 0 1 1 )

( 0 0 1 )

( 0 1 0 )

� −α2

( 1 1 1 )

��� ��� −α3−α1

������ −α1−α3

( 1 0 1 ) ( 1 0 1 )

( 1 1 1 )

−α2�( 0 1 0 )

( 0 0 1 )

−α3

−α2

−α1

( 0 1 1 )

( 1 1 0 )

( 1 0 0 )

Page 158: Pierre Ramond Group Theory a Physicists Survey 2010

146 Lie algebras: representation theory

We see that the states in (0 0 1) have opposite Dynkin labels from those in (1 0 0).These representations are complex conjugates of one another, a consequence ofthe Dynkin diagram’s reflection symmetry. Also the weights in the (0 1 0) are allaccompanied by their negatives: this representation is self-conjugate, as expectedfrom the Dynkin diagram.

Another method, more familiar to some physicists, is to represent the quadru-plet (1 0 0) by a complex vector with upper index, the sextet (0 1 0) by a complexantisymmetric tensor with two upper indices, and the antiquadruplet (0 0 1) with acomplex antisymmetric tensor with three upper indices

(1 0 0) ∼ T a, (0 1 0) ∼ T ab, (0 0 1) ∼ T abc.

We could have gone the other way, associating to the antiquadruplet a complexvector with one lower index, with the sextet represented by a tensor with two lowerindices, and the quadruplet with a tensor with three lower indices

(0 0 1) ∼ Ta, (0 1 0) ∼ Tab, (1 0 0) ∼ Tabc.

This dual tensor description of the same representation is resolved by using theLevi–Civita tensor; for SU (4), its four upper(lower) indices enable to trade onelower(upper) index for three upper(lower) indices and vice-versa. It allows us towrite

Ta = 1

3! ε abcd T bcd . (8.8)

This is called Poincaré duality. The sextet which sits in the middle of theDynkin diagram is represented in two ways by a complex two-form (twelve realparameters), with upper or lower indices related by complex conjugation

T ab = Tab. (8.9)

The same duality relation

T ab = 1

4! εabcd Tcd, (8.10)

cuts the number of real parameters in half to six, the required number needed todescribe the real sextet.

Page 159: Pierre Ramond Group Theory a Physicists Survey 2010

8.2 A3 fundamentals 147

This pattern readily generalizes to all SU (n + 1) algebras, for which the basicrepresentations correspond to the antisymmetric tensor products of its fundamen-tal (n + 1)-plet. One moves along the Dynkin diagram by adding an index andantisymmetrizing. Half way through the Dynkin diagram, one can use the totallyantisymmetric Levi–Civita tensor with (n + 1) indices

εa1a2...an+1,

to lower the indices.The dimensions of the fundamental representations of An are summarized in the

following table.

Dynkin Symbol Dimension

( 1 0 · · · · 0 ) T a(

n + 11

)

( 0 1 · · · · 0 ) T [a1 a2](

n + 12

)

( 0 · · 1 · · 0 ) T [a1 a2...ak ](

n + 1k

)

( 0 0 · · · 1 0 ) T[a1 a2](

n + 12

)

( 0 0 · · · 0 1 ) Ta

(n + 1

1

)

In this table we have used duality in the last two entries. Each of these representa-tions corresponds to one node in the Dynkin diagram.

The adjoint representation (1 0 1) is real but is not a fundamental representation.Since the highest weight has two positive entries, there are two states at the next

level. These produce two weights each of which have two positive entries, resultingagain in two paths to the next level. There, redundancy occurs and there are onlythree states at this level, each of which has only one positive entry. The next levelhas three states, each of which has zero Dynkin labels; they correspond to therank of the algebra. Then the process continues in reverse, ending with the loweststate ( 1 0 1 ). This produces fifteen states, arranged in a self-conjugate spindle-shaped arrangements, with the conjugate weights appearing in the bottom half ofthe spindle.

Page 160: Pierre Ramond Group Theory a Physicists Survey 2010

148 Lie algebras: representation theory

SU(4) adjoint representation

( 1 0 1 )

��� ��� −α3−α1

( 1 1 1 ) ( 1 1 1 )

��� ��� ��� ���−α2−α2 −α3 −α1

( 0 1 2 ) ( 1 2 1 ) ( 2 1 0 )

−α2 −α1−α3 � � �

( 0 0 0 ) ( 0 0 0 ) ( 0 0 0 )

−α3 −α2 −α1� � �

( 0 1 2 ) ( 1 2 1 ) ( 2 1 0 )

������ ��� ���−α2−α2 −α3 −α1

( 1 1 1 ) ( 1 1 1 )

( 1 0 1 )

��� ���−α3−α1

The positive roots lie in the upper half of the spindle; they are easily identifiedas (α1 + α2 + α3), (α2 + α3), (α1 + α2), and the three simple roots. Half the sumof all positive roots, the Weyl vector, is then

ρ = 1

2(3α1 + 4α2 + 3α3).

Does this ring a bell? Yes you noticed it, the numerical factors are the same asin the level vector. This turns out to be a feature of the Lie algebras An , Dn , andE6,7,8 which have a symmetric Cartan matrix, with roots of the same length. Forthe others, it is not true. The Weyl vector looks particularly simple in the Dynkinbasis

ρ = ω1 + ω2 + ω3 = ( 1 1 1 ). (8.11)

This is a general feature: the components of the Weyl vector of any simple Liealgebra are all equal to one in the Dynkin basis.

The height of the adjoint representation, ladj (six for SU (4)), is related to theCoxeter number of the algebra

h = 1

2ladj + 1, (8.12)

Page 161: Pierre Ramond Group Theory a Physicists Survey 2010

8.3 The Weyl group 149

which relates D, the dimension of the algebra, to its rank r ,

D = r (h + 1). (8.13)

8.3 The Weyl group

We have already seen that if α and β are roots, then

α ′ = α − 2βi αi

β j βj

β, (8.14)

is always a root. The root α ′ is the reflection of the root α about the hyperplaneperpendicular to the direction of β, with the same length, and its component alongβ is the negative of the projection of α along the same direction:

(α ′, α ′ ) = (α, α ), (α ′, β ) = −(α, β ). (8.15)

Repeated applications of these reflections generate a finite group. A simple way tosee this is by associating a basic reflection ri with each simple root, so there are asmany such reflections as the rank. For example, in SU (3), the action of the basicreflections about the simple roots are

r1 α1 = −α1, r1 α2 = α1 + α2,

r2 α2 = −α2, r2 α1 = α1 + α2. (8.16)

These can be elegantly written in terms of the Cartan matrix

ri α j = α j − A jiαi . (8.17)

Reflections are elements of order two which generate all other positive and negativenon-zero roots of SU (3) through the three operations

r2 r1, r1 r2, r1 r2 r1.

All other operations are redundant, for example one checks that r2 r1 r2 is thesame as r1 r2 r1. Together with the identity, these six operations generate a six-dimensional finite group, called the Weyl group. The Weyl group of SU (4) is S3,the permutation group on three objects. We can also express the action of thesereflections in the ω-basis

r1 ω2 = ω2, r1 ω1 = ω2 − ω1,

r2 ω1 = ω1, r2 ω2 = ω1 − ω2. (8.18)

Each rank n Lie algebra has its own Weyl group, generated by n basic reflectionsri , i = 1, 2, . . .. Their action on the simple roots are elegantly expressed in termsof the Cartan matrix

Page 162: Pierre Ramond Group Theory a Physicists Survey 2010

150 Lie algebras: representation theory

ri α j ={

−αi j = i

α j − A ji αi j �= i, (8.19)

with no sum over i . The action of the Weyl group on the Weyl chamber generatesthe full group lattice. We note that the action of the basic reflections on the Weylvector is simply to subtract their root

ri ρ = ρ − αi , i = 1, 2, . . . , n. (8.20)

The Weyl groups of the simple Lie algebras are all finite, and they have an irre-ducible representation of dimension equal to the rank of the group. Below we listthe Weyl groups for the simple Lie algebras.

Algebra Weyl group Order

SU (n + 1) Sn+1 (n + 1)!SO(2n + 1) Sn � (Z2 × · · · × Z2)︸ ︷︷ ︸

n

2nn!

Sp(2n) Sn � (Z2 × · · · × Z2)︸ ︷︷ ︸n

2nn!

SO(2n) Sn � (Z2 × · · · × Z2)︸ ︷︷ ︸n−1

2n−1n!

G2 D6 = S3 × Z2 12

F4 S3 × S4 × Z2 × Z2 × Z2 27 · 32

E6 U4(2) · 2 27 · 34 · 5

E7 Sp6(2) · 2 210 · 34 · 5 · 7

E8 2 · O+8 (2) · 2 214 · 35 · 52 · 7

The Weyl groups of the classical algebras are easy to understand. For instance,the An = SU (n+1) Weyl group permutes the n vectors ei of its orthonormal basis.For SO(2n + 1) and Sp(2n) their Weyl groups also permute their n orthonormalbasis elements ei , but also include the n reflections ei → ±ei . For SO(2n), thereflections are restricted further with one less parity.

For G2, S3 permutes its three orthonormal vectors ei , and Z2 changes their sign.The Weyl group of E6 is almost simple, since it has only one normal subgroup,

Z2. The quotient group is the simple U4(2), the simple group of unitary (4 × 4)matrices over the Galois field G F(2).

The same applies for E7’s Weyl group, with a Z2 normal subgroup and thesimple Sp6(2) of (6 × 6) symplectic matrices over G F(2).

Page 163: Pierre Ramond Group Theory a Physicists Survey 2010

8.4 Orthogonal Lie algebras 151

For E8, the Weyl group has a more complicated structure. It is the double coverof O+

8 (2) ·2, which has a Z2 normal subgroup with the simple O+8 (2) ·2 as quotient

group (see Chapter 9 for notation).Weyl groups are examples of the more general Coxeter groups, defined by the

presentation

< r1, r2, . . . rn | (rir j )mi j = 1 >,

where the symmetric matrix elements mi j are either positive integers or ∞

mi j ={

1 i = j

≥ 2 i �= j,

and mi j = ∞ if there is no such relation between ri and r j .Finally, we note that irreducible representations can be described in terms of

Weyl orbits, the collection of states obtained by repeated application of Weyl trans-formations. Their size is limited by the number of elements of the Weyl group:SU (2) Weyl orbits are at most two-dimensional since its Weyl group is the two-element Z2. Its Weyl orbits contain the two states | a 〉 and | − a 〉, where a is theDynkin weight, while the lone one-dimensional orbit is | 0 〉.

SU (4) Weyl orbits come in different sizes: 1, 4, 6, 12, and 24, the most allowedby its Weyl group S4, and there can be many different orbits of the same size.Irreducible representations can then be viewed as a collection of Weyl orbits; forexample, 4 and 4 each contains one different size-four orbit, and the 6 is made up ofa single orbit. On the other hand, 15 splits up into three orbits of unit length (for thethree zero weights), and one twelve-dimensional orbit. For more details, the readershould consult Tables of Dominant Weight Multiplicities For Representations ofSimple Lie Algebras, M. R. Bremner, R. V. Moody, and J. Patera (Marcel Dekker,Inc., New York and Basel).

8.4 Orthogonal Lie algebras

We have noted from the Dynkin diagrams several peculiar relations between uni-tary and orthogonal algebras. In particular, SU (2) has the same algebra as SO(3),as do SU (4) and SO(6).

In order to understand these types of relations, let us take a closer look at algebraswhich generate rotations in space. Start from an N -dimensional Hilbert space inwhich the largest algebra is generated by the (N 2 − 1) Gell-Mann matrices. Somematrices have purely imaginary entries; in the original construction they are λ2, λ5,λ7, λ10, λ12, etc., one for each off-diagonal entry in the defining (N × N ) matrix.There are N (N − 1)/2 such entries. When exponentiated into group elements ofthe form

Page 164: Pierre Ramond Group Theory a Physicists Survey 2010

152 Lie algebras: representation theory

exp(

iωaλa

2

),

they form real matrices. As we have seen, the SU (N ) generators act on complexnumbers zi , i = 1, 2, . . . , N . On the other hand the group elements generated bythe matrices with imaginary entries act independently on the real and imaginaryparts of these complex numbers. While the SU (N ) algebra preserves the norm ofa complex N -dimensional vector, these matrices form a Lie algebra SO(N ) whichpreserves the norm of a real N -dimensional vector: they generate rotations.

From a geometrical point of view, we can think of rotations as transformationswhich rotate the i th axis into the j th axis. Denote their generators by Xi j . Clearly

Xi j = −X ji . (8.21)

Their commutator with another rotation is easy to visualize; if the i−j plane is notconnected to the k−l plane, they must commute

[ Xi j , Xkl ] = 0, (8.22)

as long as all indices are different. If the two planes have a common axis, they nolonger commute but their commutator generates another rotation, viz.

[ Xi j , X jk ] = i Xik . (8.23)

When acting on a real space with real coordinates xi , i = 1, 2, . . . , N , they aregiven by

Xi j = i

(xi

∂ x j

− x j

∂ xi

). (8.24)

Suppose N = 2n is even. It is easy to see that the n generators

X12, X34, . . . , X N−3 N−2, X N−1 N ,

commute with one another since they generate rotations in planes with no axis incommon. Thus SO(2n) has rank n.

The rotation generators in (2n + 1) dimensions are the same as the SO(2n)generators plus those of the form X2n+1 i for i = 1, 2, . . . , 2n. This does not addany generator to the previous commuting set, so that the rank of SO(2n +1) is alson. We have seen that the Dynkin diagrams of the orthogonal groups in even andodd dimensions are different, although their tensor structures are much the same,except for the range of their indices. Their differences become evident in the studyof their spinor representations. The best known example is the two-dimensionalrepresentation of SO(3).

Page 165: Pierre Ramond Group Theory a Physicists Survey 2010

8.5 Spinor representations 153

8.5 Spinor representations

The link between the two are the (2 × 2) Pauli matrices (found by F. Klein fiftyyears earlier). They satisfy

σa σb = δab + iεabcσc,

from which one can obtain the SO(3) Lie algebra and the anticommutator Cliffordalgebra

{ σa, σb } = 2 δab. (8.25)

Also, the product of any two, say σ1 and σ2 generates the third one

σ3 = −i σ1σ2. (8.26)

In the geometrical notation, its generators are X12, X23, X31, which suggests therewriting

X12 = − i

2σ1 σ2, X23 = − i

2σ2 σ3, X31 = − i

2σ3 σ1. (8.27)

To derive their commutation relations, it suffices to know that the Pauli matricessatisfy a Clifford algebra. This gives a recipe for constructing generators of orthog-onal groups from Clifford algebras by expressing the SO(n) generators as bilinearproducts of Clifford elements. Building these representations of orthogonal alge-bras is reduced to the construction of all possible Clifford algebras and the spaceson which they act.

It is convenient to use the previously introduced fermionic harmonic oscillatoroperators. Recall they generate a finite-dimensional Hilbert space spanned by statesof the form

b†i1

b†i2

· · · b†ik|�〉 ≡ | [i1i2 · · · ik] 〉.

They can be thought of as complex totally antisymmetric tensors, and we know ongeneral grounds that such tensors form representations of SU (n). Indeed it is nottoo hard to see that the quadratic operators

T ij = b†

j b i − 1

nδ i

j b†k b k, T = 1

2n

(b k b†

k − b†k bk

),

form the SU (n) and U (1) compact Lie algebras, respectively, as

[ T ij , T k

l ] = δ il T k

j − δ kj T i

l, [ T, T kl ] = 0.

Except for the vacuum state and the totally antisymmetric tensor with n indices,the states of the Hilbert space are thus the fundamental representations of SU (n),one for each node in the Dynkin diagram.

Page 166: Pierre Ramond Group Theory a Physicists Survey 2010

154 Lie algebras: representation theory

Consider the remaining quadratic combinations

A i j = b i b j , A i j = b†i b†

j ,

which transform as antisymmetric two-forms of SU (n),

[ T ij , A kl ] = δ l

j A ik − δ kj A il,

and their conjugates. More computation shows that they close on the SU (n)×U (1)generators

[ A i j , A kl ] = δ ik T j

l − δj

k T il − δ i

l T jk + δ

jl T i

k − 2(δ ik δ

jl − δ

jk δ i

l ) T .

Taken together, these quadratic operators generate the larger SO(2n) Lie algebra,with the embedding

SO(2n) ⊃ SU (n) × U (1),

defined by the decomposition of its real vector representation

2n = n + n,

as can be verified by forming the adjoint decomposition

2n(2n − 1)2

= 1 + (n2 − 1) + n(n − 1)2

+ n(n − 1)2

.

8.5.1 SO(2n) spinors

All states of the Hilbert space have a U (1) charge, starting with the vacuum state,

T |� 〉 = |� 〉,with unit charge, and ending with the totally antisymmetric state made up of noscillators

T | [i1i2 · · · in] 〉 = (1 + n)| [i1i2 · · · in] 〉.We can build representations of SO(2n) by repeated application of the double-raising operator Ai j on the vacuum state. This action generates all states with aneven number of oscillators. A similar action starting on the one-oscillator state gen-erates a second representation made up of states with an odd number of oscillators.These are the two spinor representations of SO(2n).

Let us consider some simple examples. To build the spinors of SO(6), westart from the SU (3) Dynkin diagram to which we append the vacuum stateand the totally antisymmetric state (open dots). We construct the two spinors byhopscotching across the Dynkin, starting with the vacuum state

Page 167: Pierre Ramond Group Theory a Physicists Survey 2010

8.5 Spinor representations 155

� � � � � � � � �� � or the one-oscillator state

� � � � � � � � � � �yielding the two spinor representations in terms of SU (3) representations

1 + 3, 3 + 1.

They are evidently complex representations and also conjugate to one another,nothing but the 4 and 4 of SU (4) ∼ SO(6), respectively.

Our second example is SO(8). The first hopscotching

� � � � � � � � � � � � � � � � � � � � �yields one real spinor representation with eight components, which contains thefollowing SU (4) representations

1 + 6 + 1.

The second hopscotching

� � � � � � � � � � � �yields the second spinor representation made up of

4 + 4,

which is also eight-dimensional, real, and clearly different from the first one.Judging from these examples, we infer that the structure of the spinor represen-

tations of SO(2n) differ as to when n is even or odd. When n is even, the twospinor representations are complex and conjugates to one another; when n is odd,the two spinors are real and different from one another. In either case, both have

Page 168: Pierre Ramond Group Theory a Physicists Survey 2010

156 Lie algebras: representation theory

the same dimensions. The two spinor representations of SO(2n) correspond to therabbit ears of its Dynkin diagram.

Dn� � � ���

��

��

For orthogonal algebras in even dimensions, the first (n−2) fundamental represen-tations are antisymmetric tensors, and the last two are the spinor representations.

Dynkin Tensor Dimension

( 1 0 · · · 0 ) T a(

2n1

)( 0 1 · · · 0 ) T [a1 a2]

(2n2

)( 0 · · · 1 · · · 0 ) T [a1 a2...ak ]

(2nk

)( 0 0 · · · 1 0 0 ) T [a1 a2... an−2]

(2n

n − 3

)( 0 0 · · · 1 0 ) χ 2(n−1)

( 0 0 · · · 0 1 ) χ ′ 2(n−1)

When n is even, the two spinor representations are complex and conjugates to oneanother; when n is odd, the two spinors are real and inequivalent.

8.5.2 SO(2n + 1) spinors

On the other hand, the spinor structure of SO(2n−1) is very different as its Dynkindiagrams leads us to expect only one real spinor representation. This is exactly whathappens for SO(3) where we already found a (pseudoreal) spinor doublet. Thiscan be seen by constructing the SO(2n + 1) generators in terms of the fermionicoscillators.

We begin with the expression of the SO(2n) generators in terms of n fermionicoscillators. Then we add one more b and b†. According to the previous section,this would enable us to construct the generators of SO(2n + 2), but here we areinterested in its SO(2n + 1) subgroup. In the embedding

Page 169: Pierre Ramond Group Theory a Physicists Survey 2010

8.5 Spinor representations 157

SO(2n + 1) ⊃ SO(2n), (2n + 1) = 2n + 1,

the coset SO(2n + 1)/SO(2n) is just the 2n representation of SO(2n). In terms ofSU (n) it decomposes as n and n. Our extra fermionic oscillator allows us to writeseveral quadratic expressions with the right SU (n) transformation properties. Thecommutator of the combinations

n : ai = bi (b + b†), n : ai = b†i (b

† + b).

closes on the SO(2n) × U (1) algebra

[ ai , a j ] = 2 T ij + 1

nδi

j T .

They are the coset generators of SO(2n + 1)/SO(2n). These explicit expressionsenable us to compute the spinor representation.

By adding one oscillator, we have doubled the size of the Hilbert space, whichis now spanned by

| [i1i2 · · · ik] 〉, b†| [i1i2 · · · ik] 〉.

Each branch transforms by itself under the action of the SO(2n) generators, but theai and ai coset generators make transitions between them. To be explicit, considertheir action on the vacuum state

ai |� 〉 = b†i b†|� 〉.

Further action yields

a j ai |� 〉 = −b†j b

†i |� 〉,

and so on. This produces a crisscross pattern of states. It is manifestly self-conjugate, and yields half the states in the Hilbert space. We could have startedwith the action of the coset operators on the state b†|� 〉. The result would havebeen the same, that is a self-conjugate representation with half the state. Both repre-sentations are equivalent, and each is the real spinor representation of SO(2n + 1).Hence the fundamental representations of SO(2n+1) differ from those of SO(2n)by having only one real spinor representation (corresponding to the one dot of itsDynkin).

Page 170: Pierre Ramond Group Theory a Physicists Survey 2010

158 Lie algebras: representation theory

Dynkin Symbol Dimension

( 1 0 · · · 0 ) T a(

2n + 11

)( 0 1 · · · 0 ) T [a1 a2]

(2n + 1

2

)( 0 · · · 1 · · · 0 ) T [a1 a2...ak ]

(2n + 1

k

)( 0 0 · · · 1 0 ) T [a1 a2]

(2n + 1n − 1

)( 0 0 · · · 0 1 ) χ 2n

The lone spinor representation is of course real.Let us now use this method to construct the spinor representation of SO(9).

We start from two copies of the SU (4) Dynkin diagram, augmented by the twosinglets. The leftmost dots represent the vacuum state. The diagonal dots representthe action of ai , and link the states of the real spinor representation (not shown isthe other equivalent way to construct the same representation).

� � � � � � �� � � �� � � �

� � � �

1 4 6 4 1 � � �

1 4 6 4 1

This expresses the real 16-dimensional SO(9) spinor representation in terms of itsSU (4) subgroup

16 = 1 + 4 + 6 + 4 + 1.

These graphical constructions show that there is only one real fundamental spinorrepresentation in odd numbers of dimensions. As we saw, in even dimensions thetwo spinors can be either real and inequivalent, or else complex conjugates of oneanother.

We summarize these properties in a table of the spinor representations of thelowest orthogonal algebras.

Page 171: Pierre Ramond Group Theory a Physicists Survey 2010

8.5 Spinor representations 159

Algebra Spinor representation Type

SO(3) 2 Pseudoreal

SO(3) × SO(3) (2, 1 ), (1, 2 ) Pseudoreal

SO(5) 4 Real

SO(6) 4, 4 Complex

SO(7) 8 Real

SO(8) 8, 8′ Real

SO(9) 16 Real

SO(10) 16, 16 Complex

SO(11) 32 Real

SO(12) 32, 32′ Real

8.5.3 Clifford algebra construction

A more familiar way to build spinor representations is through the explicit con-struction of Clifford algebras. In the following we proceed in this way by usingan iterative procedure which builds larger Clifford algebras starting from directproducts of Pauli matrices. The direct product of matrices M and N , is the matrixM ⊗ N , such that matrix multiplication satisfies(

M ⊗ N) (

P ⊗ Q)

= (M P) ⊗ (N Q). (8.28)

For example we can construct (4 × 4) matrices in this way, such as

σ1 ⊗ σ3 =(

0 σ3

σ3 0

), σ2 ⊗ σ3 =

(0 −i σ3

i σ3 0

), (8.29)

and the like. Let us apply this construction to Clifford algebras.

• Start with the (2 × 2) matrices

1 = σ1, 2 = σ2, 3 = (−i)σ1σ2 = σ3. (8.30)

• By multiplying these by σ1 and adding one along σ2, we obtain the four (4×4) Cliffordmatrices

1 = σ1 ⊗ σ1, 2 = σ1 ⊗ σ2,

3 = σ1 ⊗ σ3, 4 = σ2 ⊗ 1, (8.31)

and their product

5 = 1234 = σ3 ⊗ 1 =

⎛⎜⎜⎝1 0 0 00 1 0 00 0 −1 00 0 0 −1

⎞⎟⎟⎠ , (8.32)

Page 172: Pierre Ramond Group Theory a Physicists Survey 2010

160 Lie algebras: representation theory

which is a diagonal matrix in this representation. Note that in this construction theClifford matrices are all hermitian. We take their bilinear product to build the SO(4)generators

Xi j = − i

2i j = 1

21 ⊗ σk; i, j, k = 1, 2, 3, cyclic,

X4k = − i

24 k = −1

2σ3 ⊗ σk . (8.33)

The combinations

L±k = 1

2εi jk Xi j ± Xk = 1

2(1 ∓ σ3) ⊗ σk, (8.34)

commute with one another and form two SU (2) Lie algebras. In block form

L+k =

(0 00 σk

), L−

k =(σk 00 0

). (8.35)

The SO(4) algebra is semi-simple, the sum of two SU (2)s. We can also define theoperators

P± ≡ 1

2(1 ± 5), (8.36)

which satisfy the projection operator relations

P± P± = P±, P± P∓ = 0, (8.37)

so that they split the four-dimensional space into two parts. Since they clearly commutewith the SO(4) generators, the four-dimensional spinor space on which the generatorsact is reducible.

However, the full four-dimensional space is needed to realize the SO(5) generators,since we need to add

X45 = − i

24 5, Xi5 = − i

2i 5, (8.38)

showing that SO(5) has a four-dimensional irreducible representation.• We repeat the procedure to build (8 × 8) Clifford matrices, by multiplying the five

previous matrices by σ1, and adding one along σ2

1 = σ1 ⊗ σ1 ⊗ σ1, 2 = σ1 ⊗ σ1 ⊗ σ2,

3 = σ1 ⊗ σ1 ⊗ σ3, 4 = σ1 ⊗ σ2 ⊗ 1,5 = σ1 ⊗ σ3 ⊗ 1, 6 = σ2 ⊗ 1 ⊗ 1, (8.39)

and one more by taking their product

7 = (−i) 1 2 3 4 5 6 = σ3 ⊗ 1 ⊗ 1. (8.40)

This means that we can represent both SO(6) and SO(7) algebras in this eight-dimensional space, using the projection operators

P± ≡ 1

2(1 ± 7), (8.41)

Page 173: Pierre Ramond Group Theory a Physicists Survey 2010

8.5 Spinor representations 161

to separate the eight-dimensional spinor space into two four-dimensional spaces. Indeed,the fifteen SO(6) generators

Xi j = − i

2i j , i �= j = 1, 2, . . . , 6, (8.42)

are 8 × 8 matrices which split (in this representation) into two 4 × 4 block-diagonalmatrices, since 7 is diagonal. Thus, SO(6) acts independently on two four-componentspinors, and because SO(6) has the same algebra as SU (4), they must correspond toits four-dimensional complex representation and its conjugate. Write the infinitesimalchanges in the form

i

2ωA

(λA 00 λA

) (ψ

χ

), (8.43)

where A = 1, 2, . . . , 15, ψ and χ are four-component spinors, and of course λA are theusual 4 × 4 Gell-Mann matrices. The hatted matrices are closely related, as it is onlyrotations containing 6 that act on ψ and χ with different sign. For instance the diagonalgenerator corresponding to ω3 is actually the same on both spinors

1

2

(X12 + X34

)=

(λ3 00 λ3

). (8.44)

However, the generator corresponding to λ8 is actually different⎧⎪⎨⎪⎩λ8 = 1

2√

3

(X34 − X12 + 2X56

), on ψ

λ8 = 12√

3

(X34 − X12 − 2X56

), on χ.

(8.45)

Some computation shows that

λA = −C λ∗A C−1 = C λA C−1, (8.46)

where the tilde matrix is the representation matrix on the 4 representation, and the (4×4)charge conjugation matrix is

C =

⎛⎜⎜⎝0 −1 0 01 0 0 00 0 0 10 0 −1 0

⎞⎟⎟⎠ = i σ3 ⊗ σ2, (8.47)

or as an eight-dimensional matrix,

C = i

2(1 − σ3) ⊗ σ3 ⊗ σ2. (8.48)

Is there a direct way to see how conjugation arises? Our Clifford matrices are man-ifestly hermitian, but not symmetric nor real. Note that the Clifford algebra is equallysatisfied by the transposed and complex conjugated matrices

±(i

)T, ±

(i

)∗.

Page 174: Pierre Ramond Group Theory a Physicists Survey 2010

162 Lie algebras: representation theory

Similarity transformations on Clifford matrices

i → S i S−1 (8.49)

preserve the Clifford algebra, and they can be used to change representations; in ourcase we have opted for a representation where 2n+1 is diagonal. This also means theexistence of a similarity transformation which transposes or takes the complex conjugateof the matrix.

Conjugation of spinors is tricky. We have seen that the SO(3) spinor representationis pseudoreal in the sense that it has a charge conjugation operation

C = σ2, (8.50)

which is a member of the algebra, so that the 2 and 2 are the same representation.

These examples suggest a recursive construction of Clifford algebras in higherdimensions. Given the 2n−1 Clifford (2n−1×2n−1) matrices, i , we build (2n ×2n)

Clifford matrices as follows

i = σ1 ⊗ i , i = 1, 2, . . . , 2n − 1.2n = σ2 ⊗ 1,

2n+1 = (−i)n 1 2 . . . 2n = σ3 ⊗ 1. (8.51)

These (2n × 2n) matrices, a , satisfy the (2n + 1)-dimensional Clifford algebra

{a, b } = 2 δab, a, b = 1, 2, . . . , 2n + 1. (8.52)

The SO(2n + 1) generators can be written as (2n × 2n) matrices

Xab = − i

2a b, a �= b. (8.53)

By forming the projection operators out of 2n+1, we split the 2n spinor intotwo 2n−1 spinors that transform according to SO(2n). Hence we see in this waythat there are two spinor representations in even dimensions and only one in odddimensions.

The product of two SO(3) spinors yields a triplet which describes a vectorin three-space, and a singlet. Mathematically, the outer product of two two-component spinors ψ, χ is a (2 × 2) matrix, which can be written as a linearcombination of the three Pauli matrices and the unit matrix

ψ χ T = 1 v0 + σi vi , (8.54)

where the coefficients are obtained by taking the appropriate traces; for example

Tr(ψ χ T σ j ) = Tr(σi σ j ) vi = 2 v j .

Page 175: Pierre Ramond Group Theory a Physicists Survey 2010

8.5 Spinor representations 163

This is called the Fierz expansion of spinor products. Let us apply it to the spinorrepresentations of SO(6), where

4 × 4 = 10 + 6, 4 × 4 = 15 + 1, 4 × 4 = 10 + 6. (8.55)

Representations appearing on the right-hand side are easily identified as SO(6)tensors: 6 is the vector, 15 is the second-rank antisymmetric tensor. The third-rankantisymmetric tensor Ti jk has twenty real components, while only 10 and 10 are inthe product. The real combination is of course their sum which has twenty compo-nents. The way to understand this is to note that in SO(6), there is the totallyantisymmetric Levi–Civita tensor on six indices, εi jklmn , which can be used togenerate self- and antiself-dual third rank tensors

T ±i jk = 1

2

(Ti jk ± Ti jk

), Ti jk ≡ 1

3! εi jklmn Tlmn. (8.56)

Similarly, we can use the Levi–Civita symbol to rewrite a four-component anti-symmetric tensor in terms of a second rank tensor.

The numbers check since we can express the 8 · 8 = 26 components break up as

26 = (1 + 1)6 =6∑1

(6i

)= 1 + 6 + 15 + 20 + 15 + 6 + 1,

each being interpreted as rank p antisymmetric tensors (called p-forms by mathe-maticians: a scalar is a zero-form, a vector is a one-form, etc.). Their construction isstraightforward in terms of products of different anticommuting Clifford matrices

i1i2...in= i1

i2· · · in

, i1 �= i2 �= · · · �= in. (8.57)

Their antisymmetry is obvious since all the matrices anticommute with oneanother. They are also traceless

Tr(i1i2...in

)= 0, (8.58)

as one can verify from our explicit construction. The Fierz expansion expresses theouter product of two spinors as a linear combination of all these matrices

� �T =6∑

j=1

i1i2...i jvi1i2...i j

. (8.59)

The coefficients are obtained by taking the appropriate traces. For example

�T i j � = 8 vi j , (8.60)

where we have used

Tr(i j kl

)= 8 (δikδ jl − δ jk δil). (8.61)

Page 176: Pierre Ramond Group Theory a Physicists Survey 2010

164 Lie algebras: representation theory

Cn ∼ Sp(2n)

Symplectic algebras are similar to the orthogonal ones except that symmetriza-tion and antisymmetrization are interchanged. The fundamental representationsmove along the Dynkin diagram; all, including the nth node, are represented bysymmetric tensors.

Dynkin Tensor Dimension

( 1 0 · · · 0 ) T a(

2n1

)( 0 1 · · · 0 ) T (a1 a2)

(2n2

)−

(2n0

)( 0 · · · 1 · · · 0 ) T (a1 a2...ak )

(2n + k − 1

k

)( 0 0 · · · 1 0 ) T (a1...an−1)

(3n − 2n − 1

)( 0 0 · · · 0 1 ) T (a1...an)

(2nn

)−

(2n

n − 2

)

8.6 Casimir invariants and Dynkin indices

We have seen that irreducible representations of simple Lie algebras of rank n arecharacterized by n (positive or zero) integers, the Dynkin indices of its highestweight state. There are other ways to specify these representations, in terms of noperators constructed out of the generators, called Casimir operators, which com-mute with all generators of the algebra. Lie algebras always have one quadraticCasimir element, but the degree of the higher order Casimir invariants depend onthe algebra. If we denote the generators of the algebra by the (dr ×dr ) matrices T A

in the r representation, the quadratic Casimir is defined to be

C (2)r = T A T A, (8.62)

where we have omitted the (dr × dr ) unit matrix on the right-hand side.In the case of SU (n), these invariants can be constructed out of traces of prod-

ucts of (n × n) matrices. Since the determinant of such an (n × n) matrix can beexpressed in terms of the trace of powers of the matrix up to the nth power, it fol-lows that we have only (n − 1) independent invariants of degree 1, . . . , n. Sincethe trace of the matrix vanishes for an irrep, SU (n) has (n − 1) Casimir invariants

Page 177: Pierre Ramond Group Theory a Physicists Survey 2010

8.6 Casimir invariants and Dynkin indices 165

which are polynomials in its generators of order 2, . . . , n. To see how this works,consider the algebra A2 = SU (3), where we have only two quantities to consider

Tr(

T A T B), Tr

(T A T B T C

).

Since there is only one invariant tensor with two adjoint indices, we set

Tr(

T A T B)

≡ I (2)r δAB,

where I (2)r is the quadratic Dynkin index of the r representation.

The trace of the cube can be split into an antisymmetric third-rank tensor, whichis nothing but the structure function f ABC , while the totally symmetric part iswritten as

Tr(

T A T B T C)

Sym= d ABC I (3)

r ,

yields I (3)r , the cubic Dynkin index. Here, d ABC is a symmetric third-rank tensor

made out of the structure functions. Supposing that T A are matrices in the adjointrepresentation, a straightforward computation yields

d ABC = f AM N f B N P f C P M ,

normalized so that I (3) = 1 for the adjoint representation. The cubic invariant existsonly for those algebras with the property that the symmetric product of its adjointcontains the adjoint. This is true only for the SU (n) (n ≥ 3) algebras.

To relate the quadratic Dynkin index to the quadratic Casimir operator, take thetrace over A, B

I (2)r δAA = Tr

(T A T A

)= C (2)

r Tr(1) = C (2)r dr, (8.63)

so that

I (2)r = dr

DC (2)

r , (8.64)

where D is the dimension of the algebra. A similar computation expresses the thirddegree Dynkin index in terms of the cubic Casimir operator.

By taking the trace of arbitrary products of representation matrices,

C A1···Ak ≡ Tr(

T A1 T A2 · · · T Ak

),

we generate quantities which are expressed as invariant kth-degree Dynkin indicesmultiplied by the appropriate invariant tensors.

In general a rank n Lie algebra has n independent invariants which are poly-nomials in its generators. For the simple Lie algebras, the orders of the invariantpolynomials are as follows.

Page 178: Pierre Ramond Group Theory a Physicists Survey 2010

166 Lie algebras: representation theory

An : 2, 3, . . . , n + 1

Bn : 2, 4, 6, . . . , 2n − 1

Cn : 2, 4, 6, . . . , 2n − 1

Dn : 2, 4, 6, . . . , 2n − 4, n − 2

G2 : 2, 6

F4 : 2, 5, 8, 12

E6 : 2, 5, 6, 8, 9, 12

E7 : 2, 6, 8, 10, 12, 14, 18

E8 : 2, 8, 12, 14, 18, 20, 24, 30

The irreducible representations of most Lie algebras can be further character-ized by a congruence number, with the important property that it is additive underKronecker products, modulo an integer specified for each Lie algebra.

Also, we can define the Chang number of a representation which is equal to itsdimension mod(h + 1), where h is the Coxeter number (sum of the marks).

It is very useful to develop a graphical representation for the T A matrices inarbitrary representations, inspired by Feynman diagram techniques. Following P.Cvitanovic (Bird Tracks, Lie’s and Exceptional Groups, Princeton University Press,2008), we assign to each state of an arbitrary d-dimensional representation a solid“propagator” line, with labels a, b = 1, 2, . . . , d. Special status is given to theadjoint representation in the form of a dotted line.

In this language, we associate a “propagator line” to the Kronecker δab and δAB .

δab δAB

Closing the lines is the same as summing over the indices, so that a solid circle isd , the dimension of the representation.

���

= d

The representation matrices in the d-dimensional representations are depicted asfollows, respectively.

= T Aab

Page 179: Pierre Ramond Group Theory a Physicists Survey 2010

8.6 Casimir invariants and Dynkin indices 167

Tracelessness of these matrices is then shown as follows.

���

= 0

The second Dynkin index of the representation r is then represented by thefollowing diagram.

���r

This makes it easy to develop a way to generate the quadratic Dynkin indices forany representation, using the Kronecker products. Label the reducible representa-tion r × s by two solid lines,

s

r

and compute its quadratic Dynkin index, the sum of the quadratic indices ofthe irreducible representations in r × s. Since the representation matrices aretraceless, they connect to the lines in pairs, leading to the sum of the followingdiagrams.

r

s

���

����

+

r

s

���

����

It follows that

I (2)r×s = ds I (2)

r + dr I (2)s ,

which gives us an easy way to compute its value for low-lying representations.When r = s, we can even specialize the rule diagrammatically by considering thesymmetric and antisymmetric products in terms of the lines

±����������������

Page 180: Pierre Ramond Group Theory a Physicists Survey 2010

168 Lie algebras: representation theory

and derive, just by counting diagrams, the following equation

I (2)(r×r)s,a

= (dr ± 2) I (2)r .

A similar reasoning can be applied to the cubic Dynkin index which arises in thefollowing diagram.

���

r

Using the same techniques we outlined for the quadratic Dynkin index, we derivethe same composition law

I (3)r×s = ds I (3)

r + dr I (3)s ,

but the counting of possible diagrams yields for the symmetric and antisymmetricproducts a slightly different result

I (3)(r×r)s,a

= (dr ± 4) I (3)r .

The composition laws for the higher-order indices are complicated by the fact thattwo and more adjoint lines can hook to the same solid line, and we do not list themhere.

The numerical values of the Dynkin indices for low-lying representations ofchoice Lie groups can be found in Appendix 2.

8.7 Embeddings

It is kind of obvious that small Lie algebras can live inside bigger ones: rotationsin N dimensions contain rotations in (N − 1) dimensions, so we would write

SO(N ) ⊃ SO(N − 1).

This is an example of embedding. How can we systematize all embeddings insidean algebra of rank n? That is the subject of this section.

The analysis is based on two remarks. If a Lie algebra H lives inside anotherG ⊃ H, then

• the adjoint representation of G contains the adjoint representation of H;• the smallest representation of G is the sum of non-trivial representation(s) of H.

Page 181: Pierre Ramond Group Theory a Physicists Survey 2010

8.7 Embeddings 169

This suggests a two-step method to catalog all embeddings: express the smallestirrep of the bigger algebra in terms of representation(s) of the smaller one; thentake its Kronecker product to form the adjoint of the larger algebra and check if itcontains the adjoint of the smaller one; if it does we have an embedding. We beginwith several simple examples.

First example: inside SU (3)

This case is easy because there is only one simple algebra smaller than SU (3):SU (2) ∼ SO(3) of dimension three and one semi-simple algebra, SU (2)× SU (2)with dimension six. Consider the defining fundamental 3. First we note that there isnot enough room for SU (2)×SU (2) in the triplet, since its smallest representation,(2, 1) + (1, 2), or (2, 2) is four-dimensional. This leaves two ways to express thetriplet in terms of representations of the SU (2) ∼ SO(3) algebras

3 = 2 + 1; 3 = 3.

Start with the first embedding 3 = 2 + 1, and form the adjoint

3 × 3 = (2 + 1) × (2 + 1),‖ ‖

8 + 1 = 3 + 2 + 2 + 1 + 1.

On the left-hand side are SU (3) representations, and on the right-hand side areSU (2) representations. Taking out a singlet on both sides, the SU (3) adjoint is

8 = 3 + 2 + 2 + 1.

We recognize the triplet as the SU (2) adjoint, so this is a bona fide embedding, butthere is more since the adjoint octet also contains a singlet, which generates a U (1)Abelian algebra. Hence we really have

SU (3) ⊃ SU (2) × U (1).

The U (1) means that we can assign a charge to the members of the fundamen-tal triplet, as long as they sum to zero. We indicate the value of the charge by asubscript, and choose an arbitrary normalization. The first embedding now reads

3 = 2−1 + 12,

so that the charges in the adjoint are the sum of those in the fundamental

8 = 30 + 2+3 + 2−3 + 10.

The charge of the triplet is zero as it should be; otherwise the U (1) would notcommute with SU (2).

Page 182: Pierre Ramond Group Theory a Physicists Survey 2010

170 Lie algebras: representation theory

Before leaving this embedding, we note that there are three ways to choose theSU (2) singlet inside the triplet, leading to three equivalent embeddings, corre-sponding to the Weyl symmetry of the SU (3) root diagram. These embeddingsoccur in elementary particle physics with SU (2) as isospin and SU (3) as the clas-sification algebra of mesons and baryons, the eight-fold way of Gell-Mann andNe’eman. The other SU (2) are called U - and V -spins.

Now for the second embedding 3 = 3, we form the adjoint

3 × 3 = 3 × 3,‖ ‖

8 + 1 = 3 + 5 + 1,

so that

8 = 3 + 5.

This time there are no singlets, and thus no U (1). The embedding is just

SU (3) ⊃ SO(3),

written with SO(3) rather than SU (2) since only tensor representations appear. Wesaw that the Elliott model uses this embedding, where SO(3) are space rotations,and SU (3) classifies nuclear levels.

Second example: inside SO(8)

This is much less trivial because SO(8) has rank 4 and also displays triality; itsfundamental can be any one of the three real octets 8i , where i stands for vector(v), and the two real spinors (s, s ′). It does not matter which we start from, so wetake the vector. The 28 adjoint is to be found in its antisymmetric product

(8v × 8v)a = 28.

We consider all possible embeddings by listing the ways in which the vector rep-resentation can be expressed in terms of representations of algebras with rankequal or lower to 4. Simple algebras with representations smaller than eightare SU (2) ∼ SO(3) ∼ Sp(2), SU (3), SU (4) ∼ SO(6), SO(5) ∼ Sp(4),SO(7), Sp(6), and G2. SU (5) does not appear in the list although its fundamen-tal quintet is small enough, but since the 8v is real, complex representations ofsub-algebras must appear together with their conjugate. We can also have directproducts of these algebras if their representations fit in the vector octet. We listthe possibilities in terms of the partitions of eight into the largest integers, that is8 = 8, 7 + 1, 6 + 2, 5 + 3, 4 + 4, and identify these integers as possible repre-sentations of the sub-algebras. Then we form the adjoint and express it in terms ofrepresentations of the (assumed) sub-algebras.

Page 183: Pierre Ramond Group Theory a Physicists Survey 2010

8.7 Embeddings 171

(1) 8 = 8. There two possible sub-algebras with real octet representations, the SU (3)adjoint octet,

SO(8) ⊃ SU (3)

8v = 8; 28 = 8 + 10 + 10,

and the SO(7) spinor.

SO(8) ⊃ SO(7)

8v = 8 ; 28 = 21 + 7.

These embeddings of SU (3) and SO(7) inside SO(8) are peculiar as they seem torely on numerical coincidences, the equality of the dimensions of the SU (3) adjointand the SO(7) spinor to the SO(8) vector representation.

(2) 8 = 7 + 1. Since SO(7) has a seven-dimensional representation, this case leads to

SO(8) ⊃ SO(7)

8v = 7 + 1; 28 = 21 + 7.

Since the adjoint representation is the same as in the previous case, they must be thesame embedding. This is explained by SO(8) triality; after triality, one is the spinor,the other is the vector embedding.

(3) 8 = 7 + 1. The seven can be interpreted as the 7 of the exceptional group G2, leadingto a perfectly good embedding with

SO(8) ⊃ G2

8v = 7 + 1; 28 = 14 + 7 + 7.

By comparing with the previous case, we see that it also defines the embedding of G2

inside SO(7). This is not a maximal embedding as SO(7) is in between SO(8) andG2 (SO(8) ⊃ SO(7) × G2) .

(4) 8 = 6 + 2. We write the vector in the suggestive form

8v = 6 + 2 = ( 6, 1 ) + ( 1, 2 )

since it looks like SO(6) and SU (2). The adjoint is easily worked out

28 = ( 15, 1 ) + ( 1, 1 ) + ( 6, 2 )

but we see that, while it includes the SU (4) adjoint, it does not contain the SU (2)adjoint: there is no SU (2) although there is a U (1) factor corresponding to the singlet.This embedding is then

SO(8) ⊃ SU (4) × U (1) ∼ SO(6) × SO(2).

Putting the U (1) charges, we get

8v = 60 + 1+1 + 1−1; 28 = 150 + 10 + 6+1 + 6−1.

Page 184: Pierre Ramond Group Theory a Physicists Survey 2010

172 Lie algebras: representation theory

(5) 8 = 5 + 3. This looks like

8v = ( 5, 1 ) + ( 1, 3 ),

which yields the adjoint

28 = ( 10, 1 ) + ( 1, 3 ) + ( 5, 3 )

showing that we are dealing with

SO(8) ⊃ SO(5) × SO(3) ∼ Sp(4) × Sp(2).

(6) 8 = 4 + 4. The possible embedding

8v = ( 4, 2 )

can be interpreted either in terms of SO(4) × SU (2) or Sp(4) × Sp(2). To see which,we form the adjoint (the (a)sym subscript stands for the (anti)symmetrized product),

28 = ( [4 × 4]sym, [2 × 2]a ) + ( [4 × 4]a, [2 × 2]sym ),

= ( [4 × 4]sym, 1 ) + ( [4 × 4]a, 3 ).

The adjoint of each sub-algebra must be a singlet of the other; but the SO(4) adjointis in the antisymmetric product which transforms as an SU (2) triplet, and so cannotbe SO(4). On the other hand, the Sp(4) adjoint lies in the symmetric product of itsfundamental and the symmetric product contains a singlet. Hence we conclude that theembedding is

SO(8) ⊃ Sp(4) × Sp(2) ∼ SO(5) × SO(3).

The adjoint representation is the same as that previously obtained; this is another exam-ple of triality as one embedding started from the vector, and the other (by triality) fromthe spinor. These two embeddings are equivalent.

(7) 8 = 4 + 4. The decomposition

8v = ( 4, 1 ) + ( 1, 4 )

implies either Sp(4) × Sp(4) or SO(4) × SO(4). From the adjoint

28 = ( [4 × 4]a, 1 ) + ( 1, [4 × 4]a ) + ( 4, 4 ),

it is obvious that the embedding is

SO(8) ⊃ SO(4) × SO(4),

since the Sp(4) adjoint lies in the symmetric product of its fundamental, while theSO(4) adjoint is to be found in the antisymmetric product.

(8) 8 = 4 + 4. This can also be interpreted as

8v = 4 + 4,

Page 185: Pierre Ramond Group Theory a Physicists Survey 2010

8.7 Embeddings 173

leading to

28 = 15 + 1 + 6 + 6.

It contains SU (4) and a U (1) factor, as indicated by the singlet. Hence the embedding

SO(8) ⊃ SU (4) × U (1)

8v = 4+ 12

+ 4− 12; 28 = 150 + 10 + 6+1 + 6−1.

We have chosen the normalization of the U (1) charges so as to get the same for thepreviously worked out (case (4)) embedding of SU (4). This is yet another example oftriality at work.

(9) 8 = 4 + 4. This can be interpreted in terms of SO(4) for the first four and two copiesof one doublet for an additional SU (2). We set

8v = ( 2, 2, 1 )0 + ( 1, 1, 2 )+1 + ( 1, 1, 2 )−1,

where we have anticipated a U (1) charge since there are two copies of the doublet.The adjoint is then

28 = ( 3, 1, 1 )0 + ( 1, 3, 1 )0 + ( 1, 1, 3 )0 + ( 1, 1, 1 )0 + ( 2, 2, 2 )1 + ( 2, 2, 2 )−1.

It contains three SU (2)s and a U (1) factor, as indicated by the singlet. This embeddingis therefore

SO(8) ⊃ SU (2) × SU (2) × SU (2) × U (1).

Finally we note that the possible decomposition

8v = ( 2, 2, 2 ),

which looks like SU (2) × SU (2) × SU (2), is not an embedding: when we form theadjoint,

28 = ( 1, 1, 1 ) + ( 3, 3, 1 ) + ( 1, 3, 3 ) + ( 3, 1, 3 ),

we do not find any adjoint SU (2) (triplet). Rather, the SU (2) adjoints are found inthe symmetric product of two real octets, which means that this decomposition can beviewed as an embedding of the symplectic algebra Sp(8)

Sp(8) ⊃ SU (2) × SU (2) × SU (2) .

Page 186: Pierre Ramond Group Theory a Physicists Survey 2010

174 Lie algebras: representation theory

We can summarize the embeddings we have found in terms of rank, as shownbelow.

SO(8) maximal embeddings Remarks

SU (2) × SU (2) × SU (2) × SU (2) Semi-simple; rank 4SU (4) × U (1) Not semi-simple; rank 3SU (2) × SU (2) × SU (2) × U (1) Not semi-simple; rank 3SO(5) × SO(3) Semi-simple; rank 3SO(7) Simple; rank 3SU (3) Simple; rank 2

We have derived the largest embeddings into SU (3) and SO(8) to provide exam-ples and motivation for our method. Is there a systematic way to understand theseembeddings, particularly in terms of Dynkin diagrams?

We note that by removing one dot from the SO(8) Dynkin, we produce eitherSU (4) or SU (2)× SU (2)× SU (2); both of which appear on our list with an extraU (1) factor. The same thing happens for the embeddings into SU (3) as removingone dot yields SU (2) which is embedded with a U (1) factor as well. This leads usto the following rule.

• Start with the Dynkin diagram of an algebra G of rank n. Remove a dot from it, yieldingthe Dynkin of a smaller algebra H of rank n − 1. This results in the regular maximalnon-semi-simple embedding with a U (1) factor

G ⊃ H × U (1).

This accounts for two of the SO(8) embeddings, but we also found one embeddingof products of simple algebras without loss of rank. To account for this embedding,Dynkin invented the extended Dynkin diagram which has one more node, and thenremoved a node from it to obtain the Dynkin diagram of the sub-algebra! If theextra node is the negative of the highest root of the original algebra, the differenceof that root with any other simple root is not a root, so that the removal of any nodefrom the extended Dynkin diagram produces a Dynkin diagram of the same rankas the original algebra. Hence the second rule:

• Remove a dot from the extended Dynkin diagram of an algebra G of rank n to generatethe Dynkin diagram of an algebra H of the same rank, resulting in the regular maximalsemi-simple embedding

G ⊃ H.

Page 187: Pierre Ramond Group Theory a Physicists Survey 2010

8.7 Embeddings 175

Let us see how this works for SO(8) by building its extended Dynkin diagram.We start from the SO(8) Cartan matrix and its inverse

AD4=

⎛⎜⎜⎝2 −1 0 0

−1 2 −1 −10 −1 2 00 −1 0 2

⎞⎟⎟⎠ ; A−1D4

= 1

2

⎛⎜⎜⎝2 2 1 12 4 2 21 2 2 11 2 1 2

⎞⎟⎟⎠ .

In this numbering, the adjoint representation ( 0 1 0 0 ), corresponds to the secondrow of these matrices. It follows that the highest root, the highest weight of theadjoint representation, is read from the inverse Cartan matrix

θ = α1 + 2α2 + α3 + α4.

It is straightforward to compute its inner product with the simple roots to find

( θ , α1 ) = 0; ( θ , α2 ) = 1,

( θ , α3 ) = 0; ( θ , α4 ) = 0, (8.65)

so that θ has the same length as the other simple roots and is connected by one lineto the second simple root. The negative of the highest root satisfies all the axiomsto be part of a Dynkin diagram. This yields the SO(8) extended Dynkin diagram.

Extended SO(8)

⊗ �

This diagram shows manifest quadrality, and a higher degree of symmetry. We canassociate with it a (5 × 5) matrix

A(1)D4

=

⎛⎜⎜⎜⎜⎝2 0 −1 0 00 2 −1 0 0

−1 −1 2 −1 −10 0 −1 2 00 0 −1 0 2

⎞⎟⎟⎟⎟⎠ , (8.66)

which satisfies all the properties to be a Cartan matrix, except that its determinantvanishes

det A(1)D4

= 0,

Page 188: Pierre Ramond Group Theory a Physicists Survey 2010

176 Lie algebras: representation theory

so that it has one zero eigenvalue. We will see later how this matrix can be usedto generate infinite-dimensional Kac–Moody algebras, but for the present we aremore interested in using it to generate embeddings.

How do we know which dot to remove to obtain, say the maximal subgroup?There is a simple way to label each dot. The extended matrix has vanishing determi-nant, and one eigenvector with zero eigenvalue. Its components ci are the co-marks,which satisfy ∑

A(1)i j c j = 0.

The marks ci are the components of the zero eigenvector of the transposed matrix∑A(1)

j i c j = 0.

In this case, all roots are of equal length, and the marks and co-marks are the same,and equal to (1, 1, 2, 1, 1).

One dot is singled out by this labeling. Removing the central dot from theextended diagram yields four unconnected nodes corresponding to SU (2) ×SU (2) × SU (2) × SU (2), while removing any one of the outer nodes brings usback to the original diagram.

These two rules account for the so-called regular maximal embeddings. Thereare some notable exceptions where this rule with the extended Dynkin diagramdoes not yield maximal sub-algebras, all having to do with the exceptional algebrasF4, E7 and E8.

All other maximal embeddings are dubbed special. The three embeddings ofSO(5) × SO(3), SO(7), and SU (3) are special. They seem to pop up because ofapparent numerical coincidences relating dimensions of their representations. Forinstance, the embedding

SO(2n) ⊃ SO(2n − 1)

is special because it relies on the numerical equality between the dimension ofeither the SO(2n) spinor or the SO(2n − 1) spinor. The SU (3) embedding is alsospecial to SO(8) because of the numerical equality with the SU (3) adjoint. Finally,

SO(2n) ⊃ Sp(4) × SO(2n − 5)

is also a special embedding, because of the dual role of Sp(4) as a symplectic andorthogonal algebra. It is clear that the special embeddings have to be worked outon a case by case basis, although there are also some general patterns as we havejust seen.

Page 189: Pierre Ramond Group Theory a Physicists Survey 2010

8.7 Embeddings 177

As a final example, we construct the G2 extended Dynkin. The G2 Cartan matrixand its inverse are

AG2=

(2 −3

−1 2

); A−1

G2=

(2 31 2

).

In this numbering, the long root is in the first Dynkin label, so the adjoint ( 1 0 )

corresponds to the first row. Hence the highest root

θ = 2α1 + 3α2.

Since

( θ , α1 ) = 1; ( θ , α2 ) = 0,

the negative highest root is a long root connected to the α1 node by one line. Thisleads to the extended Dynkin diagram for G2.

Extended G2

⊗ � �From it we obtain the extended Cartan matrix

A(1)G2

=⎛⎝ 2 −1 0

−1 2 −30 −1 2

⎞⎠ ,

whose determinant vanishes. The marks are (1, 2, 3), and the co-marks are differentand equal to (1, 2, 1). The co-marks single out one node, and its removal producesthe maximal embedding. Removal of the short node yields the SU (3) Dynkin, withembedding G2 ⊃ SU (3):

7 = 3 + 3 + 1; 14 = 8 + 3 + 3.

Removal of the middle node yields the Dynkin diagram of SU (2) × SU (2), that isthe embedding G2 ⊃ SU (2) × SU (2)

7 = (2, 2) + (1, 3) ; 14 = (1, 3) + (3, 1) + (2, 4) .

Removal of the extended node yields of course the G2 Dynkin, resulting in thetrivial embedding. As we noted for SO(8), the extended Dynkin yields only themaximal embedding without loss of rank. There is for example the anomalousembedding G2 ⊃ SU (2), with 7 = 7.

The extended Dynkin diagrams can be constructed for any of the simple Lie alge-bras. In each case, the resulting extended Cartan matrix has one zero eigenvalue,since one row is linearly dependent.

Page 190: Pierre Ramond Group Theory a Physicists Survey 2010

178 Lie algebras: representation theory

8.8 Oscillator representations

The representation theory of Lie algebras is best understood in terms of the definingrepresentation of the algebra, which generates all others by direct product. Forexample, the three fundamental representations of SU (4), (1 0 0) ≡ 4, (0 1 0) ≡ 6,and (0 0 1) ≡ 4 are generated by 4 since the other two can be obtained by theantisymmetric Kronecker products

(4 × 4)a = 6, (4 × 4 × 4)a = 4,

so that 4 is the defining representation.The antisymmetrization means that we need three copies of the fundamental

weights to generate all others. Note that, in SO(6) language, the spinor repre-sentation is defining, not the vector: one can build a vector out of two spinorsbut not vice versa. This is a general feature of orthogonal groups, and for all Liealgebras, the defining representation is not necessarily the one with the smallestdimension, as the spinor is larger than the vector for SO(n) with n ≥ 9. Thedefining representation resides at the end of the Dynkin diagram.

We have already seen how Schwinger constructed all the SU (2) representations,starting with creation and annihilation operators that transform as the 2 of SU (2).His construction can be generalized, but with some important distinctions.

We start with four creation and annihilation operators, A†a and Ab, which satisfy

the usual commutation relations

[ A b, A†a ] = δ b

a, a, b = 1, 2, 3, 4. (8.67)

The SU (4) generators are

X ba = i A†

a A b, (8.68)

so that the A†a transform like the defining quartet representation. The highest

weight state of the quartet is then expressed as

A†1 | 0 〉,

and repeated application of this operator on the vacuum state(A†

1

)a | 0 〉,

generate all irreducible representations of type (a 0 0).So far this is Schwinger’s construction, but one type of harmonic oscillator is

not sufficient to generate the whole SU (4) lattice, which contains states of theform (0 b 0), or (0 0 c), obtained by antisymmetrizing the defining representation.

Page 191: Pierre Ramond Group Theory a Physicists Survey 2010

8.8 Oscillator representations 179

Two more copies of the quartet oscillators are required to generate all irreduciblerepresentations. For example, the 6 is made up of states of the form(

A(1)†a A(2)†

b − A(1)†b A(2)†

a

)| 0 〉,

while the 4 is made up of three different oscillators

εabcd A(1)†a A(2)†

b A(3)†c | 0 〉,

using the Levi–Civita symbol. The Hilbert space generated by these three setsof oscillators contains all the SU (4) irreducible representations, although it is noeasy task to separate them out. Alternatively, one could have introduced one set ofoscillators for each of the fundamental representations with the same result; eitherway, the resulting Hilbert space is large and redundant, and to be useful it must beorganized into irreducible blocks of the algebra.

This oscillator representation of the group lattice can be generalized for anysimple Lie algebra (T. Fulton, J. Phys. A: Math. Gen. 18, 2863 (1985)).

A non-trivial example is the construction of all representations of the exceptionalgroup F4. We need three sets of oscillators transforming as 26. Each copy of 26oscillators is decomposed in terms of the SO(9) subgroup as one singlet A[κ]

0 , ninevectors A[κ]

i , i = 1, . . . , 9 and sixteen SO(9) spinors B[κ]a , a = 1, . . . , 16, and

their hermitian conjugates, and where κ = 1, 2, 3. They satisfy the commutationrelations of ordinary harmonic oscillators

[ A[κ]i , A[κ ′] †

j ] = δi j δκ κ ′

, [ A[κ]0 , A[κ ′] †

0 ] = δκκ′, [ B[κ]

a , B[κ ′] †b ] = δab δ

κκ ′,

and commute with one another. The generators Ti j and Ta

Ti j = −i3∑

κ=1

{(A[κ]†

i A[κ]j − A[κ]†

j A[κ]i

)+ 1

2B[κ]† γi j B[κ]

},

Ta = − i

2

3∑κ=1

{(γi )

ab(

A[κ]†i B[κ]

b − B[κ]†b A[κ]

i

)− √

3(

B[κ]†a A[κ]

0 − A[κ]†0 B[κ]

a

)},

satisfy the F4 algebra,

[ Ti j , Tkl ] = −i (δ jk Til + δil Tjk − δik Tjl − δ jl Tik),

[ Ti j , Ta ] = i

2(γi j )ab Tb,

[ Ta, Tb ] = i

2(γi j )ab Ti j .

The structure constants are given by

fi j ab = fab i j = 1

2(γi j )ab.

Page 192: Pierre Ramond Group Theory a Physicists Survey 2010

180 Lie algebras: representation theory

The reader is encouraged to verify the commutation relations, with the use of theidentity

3 δacδdb + (γ i )ac (γ i )db − (a ↔ b) = 1

4(γ i j )ab (γ i j )cd,

satisfied by the (16 × 16) Dirac matrices. Similar constructions can be found forthe exceptional groups, through the use of their orthogonal subgroups.

8.9 Verma modules

We have stated that irreducible representations of simple groups are character-ized by their highest weight state, the state annihilated by all creation (raising)operators, and also have outlined their construction through the action of loweringoperators. Physicists, who usually deal with low-lying representations, do not con-cern themselves with the construction of arbitrarily large representations. In thissection, we introduce some mathematical machinery which is useful in buildingthem.

While informal, the following presentation relies on the book by James E.Humphreys [11], and S. Sternberg’s unpublished notes on Lie Algebras.

Algebras contain at least as many lowering operators as their rank, and the statesobtained by their repeated application on the highest weight state form a highlyredundant set, from which it is not a priori obvious how to single out the irreduciblerepresentations. Yet, this is the vehicle used by mathematicians.

A useful concept in this extraction is the Verma module Verm(): the infinitenumber of states obtained by repeated application lowering operators to a state ofhighest weight (living in or on the walls of the Weyl chamber) | 〉.

To see how it works, consider the SU (2) Verma module associated with thehighest weight state with Dynkin label (twice the magnetic quantum number) | 2 〉.Since SU (2) has only one lowering operator T−, the full Verma module is gener-ated by its repeated application on | 2 〉. In this way we obtain the infinite set ofstates

Verm(2) : { | 2 〉, | 0 〉, | − 2 〉, | − 4 〉, · · · }, (8.69)

where

| 2 − 2k 〉 = (T−)k | 2 〉. (8.70)

We are free to apply the raising operator to any one of these, and we find from thealgebra that the state | − 4 〉 is annihilated by T+

T+| − 4 〉 = T+(T−)3| 2 〉 = 0. (8.71)

Page 193: Pierre Ramond Group Theory a Physicists Survey 2010

8.9 Verma modules 181

Thus (T−)3| 2 〉 is a highest weight state. The first three states form the three-dimensional irrep, and we have decomposed the Verma module into an irreduciblerepresentation and another Verma module.

Verm(2) = {| 2 〉, | 0 〉, | − 2 〉 } ⊕ Verm(−4). (8.72)

We can then continue the decomposition of Verm(−4). It does not contain anymore highest weight states, since

T+| 2 − 2k 〉 = [ T+, (T−)k ]| 2 〉 = k(3 − k)| 1 〉, (8.73)

vanishes only for k = 0, 3. Thus, if we take away the first three states, we obtainanother Verma module with a state of highest weight with a negative Dynkin index,which contains no other state of highest weight. What do the states | 2 〉 and | − 4 〉have in common?

This is a general feature: a Verma module with highest weight state in or on thewalls of the Weyl chamber contains one finite-dimensional representation, and therest assemble into infinite-dimensional Verma modules, each of which starts witha state outside the Weyl chamber.

In general things are not so simple: consider a rank r algebra, and form its Vermamodule associated with the highest weight state | 〉 = | a1 a2 · · · ar 〉 in or on thewalls of the Weyl chamber (all ai positive or zero), that is

T (i)+ | 〉 = 0, i = 1, 2, . . . , r. (8.74)

The Verma module Verm(�) is the collection of states[T (i1)−

]n1[T (i2)−

]n j · · ·[T (in)−

]nr |� 〉, (8.75)

for all values of im = 1, 2, . . . , r , and arbitrary integers n j . Among this infinite setof states lie the irreducible representation labelled by , and more. The systematicdecomposition of the Verma module is what makes this concept useful.

Since the effect of T (i)− is to subtract the simple root αi , the Verma module

consists of states of the form

| μ 〉 = | −∑

ni αi 〉, (8.76)

where the ni are either zero or positive integers. In addition, since the loweringoperators associated with the simple roots do not operate in any preassigned order,their commutator produces all the lowering operators associated with the positiveroots. For instance, in SU (4), the three lowering operators generate by commuta-tions three more lowerers. As a result, there are many different ways to obtain aparticular weight. The multiplicity of a weight μ in Verm(), N(μ), is equal tothe number of ways − μ can be expressed as a sum of the positive roots withpositive coefficients.

Page 194: Pierre Ramond Group Theory a Physicists Survey 2010

182 Lie algebras: representation theory

In order to determine this multiplicity, it is convenient to introduce an operatorwhich does this counting. First, let us associate to any state | μ 〉, an operator e(μ)

which satisfies an exponential-like multiplication rule

e(λ) · e(μ) = e(λ + μ), (8.77)

for any two weights λ and μ. Introduce the Kostant counter

C ≡∏

pos. roots

1

1 − e(α), (8.78)

where the product is over the positive roots. For SU (3) it is the product of threeterms

1

1 − e(−α1)

1

1 − e(−α2)

1

1 − e(−α1 − α2). (8.79)

For a Lie algebra of rank r and dimension D, it will be the product of (N − r)/2terms. Expand each product as a power series

1

1 − e(−α1)= 1 + e(−α1) + e(−2α1) + · · · , (8.80)

using the multiplication rule (8.77) and then multiply them together. A little algebrayields

C = 1 +∑

μ

N (μ) e(μ), (8.81)

where the sum is over the weights of the form

μ =∑

ni αi , (8.82)

with ni positive integers or zero, and N (μ) simply counts the sets of sums overpositive roots that yields μ. In SU (3),

C = 1 + e(−α1) + e(−α2) + 2 e(−α1 − α2) + · · · , (8.83)

since there are two linear combinations of positive roots that yields that combi-nation: one is by adding two roots α1 and α2 and the other is the positive rootα1 + α2, and C is indeed a counting operator. It is closely related to the Weylfunction

Q ≡∏

pos. roots

(e(α/2) − e(−α/2)

), (8.84)

Page 195: Pierre Ramond Group Theory a Physicists Survey 2010

8.9 Verma modules 183

since

Q =∏

pos. roots

e(α/2)(

1 − e(−α))

= e(ρ)∏

pos. roots

(1 − e(−α)

)= e(ρ) C−1, (8.85)

where ρ is the Weyl vector, half the sum of the positive roots, which we introducedearlier.

The formal character over the Verma module

ChVerm() ≡∑

μ

N(μ) e(μ), (8.86)

with the sum over all the weights in the module, is closely related to the counter

ChVerm() = C e(). (8.87)

We also have

QChVerm() = QC e() = e( + ρ). (8.88)

Further understanding is gained by splitting up Verma modules into irreduciblecomponents. We have seen that our sample SU (2) Verma module contains a stateannihilated by creation operators. In order to identify such states in a generalVerm(�), consider the quadratic Casimir operator

C2 ≡∑

A

T Ar T A

r ,

which commutes with all elements of the Lie algebra,

[ C2, T Ar ] = 0. (8.89)

We rewrite it in the root notation as

C2 ≡ gi j H i H j +∑

roots α

Eα E−α, (8.90)

where the sum is over all non-zero roots. Its value on a highest weight state | 〉 isparticularly simple. Split the sum over positive and negative roots, and use for thepositive roots

Eα | 〉 = 0,

Page 196: Pierre Ramond Group Theory a Physicists Survey 2010

184 Lie algebras: representation theory

to find

C2 | 〉 =(

gi j H i H j +∑

pos. roots

Eα E−α

)| 〉,

=(

gi j H i H j +∑

pos. roots

[ Eα, E−α ])| 〉. (8.91)

Using the commutation relations, we obtain

C2 | 〉 =(

gi j H i H j +∑

pos. roots

αi H i)| 〉, (8.92)

which can be rewritten as

C2 | 〉 =[( + ρ, + ρ ) − ( ρ, ρ )

]| 〉, (8.93)

by completing the squares, and using the inner product defined by the Killing form.Since the Casimir operator commutes with all generators, its value is the same onany state of the Verma module.

C2 | μ 〉 =[( + ρ, + ρ ) − ( ρ, ρ )

]| μ 〉, (8.94)

for | μ 〉 in Verm(). Furthermore, if μk is a highest weight state in Verm(), theabove reasoning shows that also

C2 | μk 〉 =[(μk + ρ, μk + ρ ) − ( ρ, ρ )

]| μk 〉, (8.95)

so that the highest weight states in Verm() are those which satisfy

( + ρ, + ρ ) = (μk + ρ, μk + ρ ). (8.96)

Since they have the same length, this suggests that + ρ and μk + ρ are relatedby Weyl transformations; in fact they are. Using our sample SU (2) Verma module,with ρ = 1, the two highest weight states yield | 2 + 1 〉 = | 3 〉 and | − 4 + 1 〉 =| − 3 〉, which are indeed on the same Weyl orbit.

This singles out a finite set of weight states μk , k = 1, 2 . . . n, each standing atthe top of an irreducible module, Irr(μk), which contains no other state annihilatedby the creation operators. Again in our SU (2) example, there are two irreduciblemodules; one is finite-dimensional, with its highest weight in the Weyl chamber,the other is infinite-dimensional with its highest weight outside the Weyl chamber.

Verma modules split up as finite sums of irreducible modules; all are infinite-dimensional except for Irr(), which is the finite-dimensional irrep of the Liealgebra.

The next step is to define a Verma module for each μk , and express it in termsof irreducible modules contained within. To that effect, we order the weights μk

inside Verm() so that μ1 is the “largest” weight obtained by subtracting positive

Page 197: Pierre Ramond Group Theory a Physicists Survey 2010

8.9 Verma modules 185

roots, μ2 the second largest, etc. This enables us to write a relation among theVerma and irreducible characters

ChVerm(μk)= ChIrr(μk)

+∑j<k

a[μk ]j ChIrr(μj)

, (8.97)

where the character of the irreducible module, labeled by the highest weight μk , isgiven by

ChIrr(μk)=

∑λ

n[k](λ) e(λ), (8.98)

with the sum over the n[k](λ) states of weight λ in the irreducible module.Alternatively, we can invert this equation and express the character of the irre-

ducible representations in terms of the character of the Verma module of highestweights which satisfy eq. (8.96). In so doing, we should be able to write the char-acter of any irreducible representation in terms of those of the Verma modulesassociated with the weights which satisfy eq. (8.96)

ChIrr() = ChVerm() +∑k=1

a[]k ChVerm(μk)

, (8.99)

where of course μk < . Therefore, using eq. (8.88),

QChIrr() = e( + ρ) +∑k=1

a[]k e(μk + ρ). (8.100)

Further simplification is achieved by considering the effect of a Weyl transforma-tion on this equation, as the character is invariant under Weyl transformations sinceirreducible transformations are a collection of weights on Weyl orbits.

In addition, the Weyl function transforms simply under the Weyl group: a simplereflection ri will permute positive roots into other positive roots, except for thesimple root αi which goes into its negative, that is

ri :[e(αi/2) − e(−αi/2)

]→ −

[e(αi/2) − e(−αi/2)

],

and all others in the product get permuted into one another. It follows that

ri : Q → −Q, (8.101)

so that for any element w of the Weyl group

w: Q → (−1)l(w) Q, (8.102)

where l(w) is the number of reflections in the Weyl group element. Hence

QChIrr() = (−1)l(w)

(e(w[ + ρ]) +

∑k

a[]k e(w[μk + ρ])

), (8.103)

Page 198: Pierre Ramond Group Theory a Physicists Survey 2010

186 Lie algebras: representation theory

where w[μk + ρ] is the Weyl image of μk + ρ. Thus, for any w ∈ W ,

QChIrr() = e( + ρ) +∑

a[]k e(μk + ρ),

= (−1)l(w)(

e(w[ + ρ]) +∑

a[]k e(w[μk + ρ])

), (8.104)

which holds for all elements of the Weyl group. This implies that e(μk + ρ) ∼e(w[ + ρ)]) for some Weyl group element. It follows that

QChIrr() =∑w∈W

(−1)l(w) e(w[ + ρ]), (8.105)

as long as that all weights of the form | μk + ρ 〉 are on the Weyl orbit of + ρ.To see that it is indeed the case, assume the existence of a state, say | μm + ρ 〉,

which is not on the orbit of +ρ. By repeated application of Weyl transformations,we arrive at a new state in the Weyl chamber or on its wall. This would imply thatboth this state and | + ρ 〉 live in the Weyl chamber.

It is easy to show that with the same length, and differing by a sum of positiveroots with positive coefficients, they must be equal to one another. Let δ = −μm

be their difference, and write

0 = ( + ρ, + ρ ) − (μm + ρ, μm + ρ ),

= ( + ρ, δ ) + ( δ, μm + ρ ),

≥ ( + ρ, δ ) ≥ 0,

since it is assumed to be in the Weyl chamber ( δ, μm + ρ ) ≥ 0. Hence, it canonly be zero and μm = 0. The two states are the same and eq. (8.105) is exact. For = 0, it reduces to

Q =∑w∈W

(−1)l(w) e(w[ρ]), (8.106)

leading us to the Weyl character formula

ChIrr() =∑

w (−1)w e(w[ + ρ])∑w (−1)w e(w[ρ]) . (8.107)

This formula appears to be quite unwieldy: for SU (4) the sum is over the 4! = 24different Weyl elements, but for SU (2), it is easy to describe. We have only onesimple positive root α, and its Weyl group of one reflection, Z2, which changes itssign. All weight vectors are of the form | nα/2 〉, n = 0,±1,±2, . . . , and the Weylvector is ρ = 1. Hence∑

w

(−1)w e(w[ρ]) = e(α/2) − e(−α/2), (8.108)

Page 199: Pierre Ramond Group Theory a Physicists Survey 2010

8.9 Verma modules 187

so that

1

e(α/2) − e(−α/2)= 1

e(α/2)

(1 +

∑p

e(−α)

). (8.109)

The numerator is given by∑w

(−1)w e(w[ + ρ]) = e((n + 1)α/2) − e(−(n + 1)α/2), (8.110)

where

| 〉 = | n

2α 〉. (8.111)

Now if we divide one by the other, we obtain

n∑p=0

e((n − 2p)

2α), (8.112)

which contains (n + 1) terms in the representation n (n = 2 j). This ratio of char-acters exactly reproduces the states in the SU (2) irreducible representation labeledby = n.

8.9.1 Weyl dimension formula

From the Weyl character formula to the dimensions of irreducible representationsis but a small step, since the dimension is simply the character with the operatorse(μ) replaced by one. Thus for any weight μ, we define

(e(λ)

)≡ exp

{(μ, λ ) t

}, (8.113)

where t is some real parameter. Write the Weyl formula for convenience as

ChIrr() = A( + ρ)

A(ρ), (8.114)

with

A(μ) ≡∑w∈W

(−1)l(w) e(w[μ]). (8.115)

Then

Eμ(A(ν)) =∑w∈W

(−1)l(w) exp{(μ, w[ν] ) t

}, (8.116)

and observe that, since Weyl transformations do not change the length,

(μ, w[ν] ) = ( w−1[μ], ν ),

Page 200: Pierre Ramond Group Theory a Physicists Survey 2010

188 Lie algebras: representation theory

after summing over w′ = w−1, we obtain

Eμ(A(ν)) =∑w∈W

(−1)l(w) exp{( ν, w[μ] ) t

},

= Eν(A(μ)). (8.117)

Hence,

Eρ(A(λ)) = Eλ(A(ρ)) = Eλ

(∏[e(α/2) − e(−α/2)]

),

=∏[

exp{(λ,α/2)t

}− exp

{− (λ,α/2)t

}]. (8.118)

Now expand about t = 0 to get

Eρ(A(λ)) ≈ t N∏

pos. roots

(λ,α) + · · · , (8.119)

where N is the number of positive roots and the dots denote higher powers of t . Itfollows that

(A( + ρ)

A(ρ)

)=

∏pos. roots

( + ρ,α)

(ρ,α)+ O(t). (8.120)

We set t = 0 and note that the left-hand side is the dimension of the representation,the character of the irreducible representation at e = 1. Hence, Weyl’s dimensionformula:

DimIrr() =∏α>0

(1 + (,α)

(ρ,α)

). (8.121)

This yields a convenient way to get the dimension as a polynomial in the Dynkinlabels of the highest weight state. In SU (4), with six positive roots, the dimensionof the representation ( a1 a2 a3 ) is a sixth-order polynomial

(1 + a1) (1 + a2) (1 + a3)

(1 + a1 + a2

2

)(1 + a2 + a3

2

)(1 + a1 + a2 + a3

3

),

corresponding to the six positive roots in the α-basis

( 1 0 0 ) ( 0 1 0 ) ( 0 0 1 ) ( 1 1 0 ) ( 0 1 1 ) ( 1 1 1 ).

8.9.2 Verma basis

It is possible to construct directly the linearly independent states of irreduciblerepresentations, using the Verma module construction.

We state the results without proof and refer the reader to S.-P. Ling, R. Moody,M. Nicolescu, and J. Patera, J. Math. Phys. 27, 666 (1986).

Page 201: Pierre Ramond Group Theory a Physicists Survey 2010

8.9 Verma modules 189

For SU (3), the (un-normalized) states in the Verma basis are of the form[T (1)

−]n3

[T (2)

−]n2

[T (1)

−]n1 | a1 a2 〉, (8.122)

with

0 ≤ n1 ≤ a1,

0 ≤ n2 ≤ a2 + n1,

0 ≤ n3 ≤ min[a2, n2].For G2, the (un-normalized) states are given by products of the two lowering

operators associated with the simple roots[T (2)

−]n6

[T (1)

−]n5

[T (2)

−]n4

[T (1)

−]n3

[T (2)

−]n2

[T (1)

−]n1 | a1 a2 〉, (8.123)

with

0 ≤ n1 ≤ a1,

0 ≤ n2 ≤ a2 + 3n1,

0 ≤ n3 ≤ min[(a2 + 2n2)/3, n2],0 ≤ n4 ≤ min[(a2 + 3n3)/2, 2n3],0 ≤ n5 ≤ min[(a2 + n4)/3, n4/2],0 ≤ n6 ≤ min[a2, n5].

While clumsy to write, these constructions are very easy to implement oncomputers, and as such are quite useful.

Page 202: Pierre Ramond Group Theory a Physicists Survey 2010

9

Finite groups: the road to simplicity

Earlier, we saw that finite groups can be taken apart through the compositionseries, in terms of simple finite groups, that is groups without normal subgroups.Remarkably the infinitude of simple groups is amenable to a complete classifi-cation. Indeed, most simple groups can be understood as finite elements of Liegroups, with parameters belonging to finite Galois fields. Their construction relieson the Chevalley basis of the Lie algebra, as well as on the topology of its Dynkindiagram. The remaining simple groups do not follow this pattern; they are the mag-nificent 26 sporadic groups. A singular achievement of modern mathematics wasto show this classification to be complete.

So far, this beautiful subject has found but a few applications in physics. Wefeel nevertheless that physicists should acquaint themselves with its beauty. In thismostly descriptive chapter we introduce the necessary notions from number theory,and outline the construction of the Chevalley groups as well as that of some spo-radic groups. We begin by presenting the two smallest non-Abelian simple finitegroups.

A5 is simple

The 60 even permutations of the alternating group A5 are 3-ply or triplytransitive. We can use this fact to prove something startling about alternatinggroups.

By definition, all even permutations are generated by the product of two transpo-sitions, which can be reduced to three-cycles or the product of three-cycles. Indeed,the product of two transpositions is either

• the unit element if they are the same (a b)(a b) = e,• a three-cycle if they have one element in common: (a b)(b c) = (a b c),• the product of two three-cycles if they have no elements in common: (a b)(c d) =

(a b)(c a)(c a)(c d) = (a b c)(c a d).

190

Page 203: Pierre Ramond Group Theory a Physicists Survey 2010

Finite groups: the road to simplicity 191

It follows that all even permutations are generated by three-cycles (any five-cyclecan be written as the product of two three-cycles: (a b c d e) = (a b c)(c d e)). Wecan use this fact and the (n − 2)-ply transitivity of An to prove that all alternatinggroups are simple as long as n ≥ 5. Below we carry the proof for the case n = 5,but the generalization to larger n is easy.

Assume that A5 has a normal subgroup B, with elements of the form ( x x x x x ),( x x x ), or ( x x )( x x ).

Suppose first that B contains the five-cycle b = (1 2 3 4 5). Consider the elementt = (3 4 5), and form its conjugate

b = tbt−1 = (1 2 4 5 3). (9.1)

Since B is assumed to be normal, b and its inverse are also in B, as is their product

b(b)−1 = (1 4 3) ∈ B. (9.2)

If B contains the five-cycle, it must also contain a three-cycle.Then suppose that b = (1 2)(3 4). Let t = (1 2 5). Form

b = tbt−1 = (2 5)(3 4), (9.3)

and

bb = (1 5 2), (9.4)

which must be the normal subgroup: any normal subgroup of A5 must contain atleast one three-cycle.

Now for l’envoi: assume that B contains the three-cycle (1 2 3). Since A5

is 3-ply transitive, any other three-cycle is related to it by conjugation, forexample

(345) = t (1 2 3)t−1, t = (1 3 5), (9.5)

and (3 4 5) must belong to the normal B. Since A5 is generated by all three-cycles,it must be that B = A5, and A5 is simple. The proof carries for arbitrary n alongthe same lines: first one shows that any normal subgroup of An must containthree-cycles (the case n = 5 is a bit easier to analyze), then use the triple tran-sitivity of An . Note that the proof does not carry for n = 4 which is at mostdoubly transitive, and in fact A4 is not simple. We have identified two infinitefamilies of simple groups, Zp, p prime, and An , n ≥ 5. But there are manymore!

Note that A5 has only one singlet representation; this is a consequence of thefact that it is a perfect group, equal to its derived subgroup.

Page 204: Pierre Ramond Group Theory a Physicists Survey 2010

192 Finite groups: the road to simplicity

A5 C1 12C2 12C3 15C4 20C5

χ [1] 1 1 1 1 1

χ [31] 3 12 (1 + √

5) 12 (1 − √

5) −1 0

χ [32] 3 12 (1 − √

5) 12 (1 + √

5) −1 0

χ [4] 4 −1 −1 0 1

χ [5] 5 0 0 1 −1

It is the symmetry group of the icosahedron, the platonic solid with 12 ver-tices, 20 triangular faces, and 30 edges. Its classes describe rotations of the vertices(12C2, 12C3), edges (15C4) , and faces (20C5) into themselves.

A closely related structure, the truncated icosahedron, is found in Nature,notably in the form of a football (soccer ball), Fullerene, and virus coating. Atruncated icosahedron is simply an icosahedron where each vertex is chopped offand replaced by a pentagon.

Fullerene, C60 is a molecule with 60 carbon atoms. Its surface has 5 · (12) = 60vertices. Its 20 + 12 = 32 faces, split into 20 hexagons, and 12 pentagons whichdo not touch one another. The carbons are linked by two types of chemical bonds:30 “double bonds” edge two hexagons, and 90 slightly shorter “single bonds” edgea pentagon and a hexagon, making up the 90 edges required by Euler’s formula.Each carbon has one double bond and two single bonds.

A5 = PSL2(5) (see later) acts on pairs of vertices in C60, but does not takeinto account its intricate bond structure. Remarkably, by upgrading to PSL2(11),the bond structure can be encoded as well, as shown by B. Kostant (Notices of theAmerican Mathematical Society, 42, number 9, 959 (1995)). PSL2(11), the order660 projective group of (2 × 2) matrices over G F11 (see later), contains the icosa-hedral group, and a conjugacy class of 60 elements. Crucial to this identification isthe fact that its transformations can be written as z h, where z ∈ Z11, and h ∈ A5,as noted by Galois, the night before his death.

9.1 Matrices over Galois fields

In this section, we study groups as matrices whose elements are members of finitefields, called Galois fields. This approach has not yet proven useful for physicalapplications, but it is by far the most efficient way to present the landscape of finitegroups. We begin by reminding the reader of some basic mathematical facts aboutfinite fields.

Page 205: Pierre Ramond Group Theory a Physicists Survey 2010

9.1 Matrices over Galois fields 193

Scholium. Rings and fields

A set of elements with two operations, addition + and multiplication �. A ringcontains the zero element so that it is an Abelian group under addition. Under mul-tiplication, a ring contains the unit element, and satisfies closure and associativity,but its elements need not have multiplicative inverses, so it is not a group undermultiplication. The product law is not necessarily commutative. Octonions do notform a ring, since their multiplication is not associative.

A field is just a commutative ring, i.e. one where the multiplication law iscommutative, and where each element has a unique inverse. Hence it forms anAbelian group under both addition and multiplication. It must contain a zero anda unit element, denoted by 0 and 1, respectively. The real and complex numbers,including zero and one form a field. Quaternions do not form a field because theirmultiplication law is not commutative: they form only a ring.

Scholium. Galois fields

A Galois field is a finite set of elements which satisfy the field axioms. The simplestexample is the p elements 0, 1, 2, . . . , p − 1, where p is a prime number. Takeboth addition and multiplication to be mod(p). They form the Galois field G F(p).An example will suffice to illustrate its properties.

The Galois field G F(7) has seven elements, 0, 1, 2, 3, 4, 5, 6. Their inverses areobtained by modular addition:

−1 ∼ 6 : 1 + 6 = 0 mod(7), . . . ,−6 ∼ 1 : 6 + 1 = 0 mod(7). (9.6)

This makes perfect sense under ordinary multiplication mod(7), as 1 = (−1)2 =(6)2 = 36 = 1 + 35 = 1 mod(7). Galois fields have two elements which are rootsof one, except for the simplest Galois field G F(2) with elements 0, 1, where 1 isits own inverse.

One might think that any set of integers with mod(n) addition and multiplicationforms a field, but that is not so. It suffices to look at the mod(6) case: the element2 clearly does not have an integer inverse: ? · 2 = 1 mod(6) has not an integersolution. To find out the allowed values of n, let us start from G F(p) (p prime),with elements u0, u1, . . . , u p−1, and build bigger finite fields. Assume a finite fieldwith q elements, where q > p, which contains G F(p). Since q > p, it has atleast one element v not in G F(p). The p new elements uiv, are in the field, butnot in G F(p) (since v is not). Together with ui , we can form p2 elements, thesums

ui + u j � v. (9.7)

If q = p2, we have constructed a new field of order p2. Otherwise, q > p2, allow-ing us to repeat the procedure: pick a new element v′, and generate p3 elements,

Page 206: Pierre Ramond Group Theory a Physicists Survey 2010

194 Finite groups: the road to simplicity

and so on until we run out of new elements, with q = pn . We have shown howto build a Galois field of order pn , which contains the elements of G F(p). Thisprocedure is general: finite fields are necessarily of order pn . The prime p is calledthe characteristic of the field.

Having established its existence, we have yet to find its multiplication table. ForG F(pn) to be a field, it must be an Abelian group of order (pn − 1) under itsmultiplication rule. Hence the order of any of its elements must divide (pn − 1),that is

(x � x � · · · � x)︸ ︷︷ ︸(pn−1)

= x (pn−1)� = 1, (9.8)

under the yet to be defined multiplication rule for the field (this does not meanthat x is a complex root of unity!). This equation simply says that the elements ofG F(pn), u0, u1, . . . , ua, . . . , u pn−1, are the roots of the polynomial

p(x) = x pn

� − x = (x − u0) � (x − u1) � (x − u2) � · · · � (x − u pn−1). (9.9)

To find the multiplication rule, Galois started with a polynomial of degree n overG F(p) (he did not call it that, of course), that is

p(x) = v0xn + v1xn−1 + · · · + vpn , (9.10)

where va ∈ G F(p). There are p2 such polynomials. Only a subset of these poly-nomials, πa(x), have no roots which belong to G F(p). They are called irreduciblepolynomials.

The multiplication rule of the elements of G F(pn) are determined by settingone of these irreducible polynomials to zero, and it does not matter which wechoose!

Some simple examples: n = 2, p = 2. To find the quadratic irreducible poly-nomials, we form all second order polynomials with roots over G F(2): (x − 1) �(x − 1) = x2 − 2x + 1 = x + 1 and , x � (x − 1) = x2 − 1 = x2 + 1 mod(2).Comparing with all possible second-order polynomials over G F(2), we see thatthere is only one irreducible polynomial

π(x) = x2 + x + 1, (9.11)

leading to the multiplication rules: x � x = x2 = −x − 1 = x + 1 mod(π). Wehenceforth dispense with the �, and build the elements of G F(4) by successiveadditions {0, 1, x, x + 1}. Believe it or not, all are roots of unity under multiplica-tion. For instance, x3 = xx2 = x(x + 1) = x2 + x = 2x + 1 = 1, all mod(2). Thisleads to the multiplication table for G F(4)

Page 207: Pierre Ramond Group Theory a Physicists Survey 2010

9.1 Matrices over Galois fields 195

1 x x + 1x x + 1 1

x + 1 1 x (9.12)

where of course the zero element does not appear.A less trivial example with several possible irreducible polynomials is G F(9).

Start from the integral elements of G F(3), 0, 1, 2, to which we add polynomialsobtained by successive addition, leading to the nine (three integral plus four non-integral) elements

G F(9) : {0, 1, 2, x, x + 1, x + 2, 2x, 2x + 1, 2x + 2} . (9.13)

For convenience, we label them by their number, a = 1, 2 . . . , 8. Their multi-plication is determined from any irreducible polynomial. The reducible quadraticpolynomials are (x − 1)2, (x − 2)2, (x − 1)(x − 2), x(x − 1), x(x − 2), x2 to beevaluated mod(3). This leaves three possible irreducible polynomials

π1(x) = x2 + 1, π2(x) = x2 + 2x + 2, π3(x) = x2 + x + 2. (9.14)

It is simplest to use π1(x) to find the following multiplication table, that isx2 = −1 ∼ 2 mod(π1).

1 2 3 4 5 6 7 82 1 6 8 7 3 5 43 6 2 5 8 6 4 74 8 5 6 1 7 2 35 7 8 1 3 4 6 26 3 1 7 4 2 8 57 5 4 2 6 8 3 18 4 7 3 2 5 1 6

The reader can verify that the other choices for irreducible polynomials, π2 andπ3, lead to the same multiplication table. Note the interesting fact that one of itselements 4, cannot be written as a square.

Finally, we note that a Galois field G F(pn), with pn elements, p prime, sustainsan important map into itself, called the Frobenius map. It is defined as

u → u′ = u p, (9.15)

and it clearly has order n, since (symbolically)

(u′)′n = u pn = u. (9.16)

Page 208: Pierre Ramond Group Theory a Physicists Survey 2010

196 Finite groups: the road to simplicity

For special values of n, the Frobenius map plays a crucial role in the constructionof some infinite families of simple finite groups.

We are now in a position to study the property of matrices whose elements areelements of finite Galois fields.

Matrices over finite fields

As long as they have inverses, matrices over the real and complex number fieldsgenerate infinite Lie groups, since they satisfy all the group axioms. Hence we canexpect that non-singular matrices over finite fields also form (finite) groups.

We can immediately consider the general linear group of (n × n) matrices overthe Galois field G F(q), where q is a power of a prime. It is denoted by GLn(q).These matrices form finite groups which we wish to catalog. An instructiveexample is GL2(7), of (2 × 2) matrices over the Galois field G F(7):

GL2(7):(

a bc d

), a, b, c, d ∈ G F(7), (9.17)

with

(ad − bc) �= 0 mod(7) (9.18)

required because these matrices must have inverses to form a group. To determinehow many such matrices there are, look at the first row: a and b can take on sevenvalues each, but a = b = 0 is forbidden by the determinant condition. Hence ithas 72 − 1 possible entries. The second row c, d also has 72 possible values, butfor each assignment of the first row, there are seven choices which give vanishingdeterminant, when both elements are zero and when the second row is a multiple ofthe first. Hence the second row can have 72 − 7 = 42 assignments. Hence, GL2(7)contains 48 · 42 = 2016 matrices.

We then specialize to the special linear group, the set of matrices in GL2(7) withunit determinant:

SL2(7):(

a bc d

), a, b, c, d ∈ G F(7); (ad − bc) = 1 mod(7). (9.19)

Since the determinant can have one of six values, we see that it has one sixth asmany elements as GL2(7): SL2(7) is a group with 2016/6 = 336 elements.

We can think of these matrices as acting on two-vectors(xy

)→

(x ′

y′

)(a bc d

)(xy

), (9.20)

and define projective transformations

u → u′ = au + b

cu + d, u = x

y, (9.21)

Page 209: Pierre Ramond Group Theory a Physicists Survey 2010

9.1 Matrices over Galois fields 197

where u is called the homogeneous coordinate. These are generated by a smallergroup, the projective special linear group, PSL2(7). It is smaller because wecan multiply the first row and the second row by the same element of G F(7)without altering the transformation. In order to stay within the unit determinantrequirement, the element must square to one. Hence, in PSL2(7), the mappings(

1 00 1

)(a bc d

),

(6 00 6

)(a bc d

), (9.22)

lead to the same projective map, since(6 00 6

)(6 00 6

)=

(1 00 1

)mod(7). (9.23)

In PSL2(7), these two elements are taken to be the same, resulting in 336/2 = 168elements. Technically speaking, these two diagonal matrices form the center ofSL2(7).

This construction can be generalized to any (n×n) matrices over any Galois fieldG F(pm). We have then GLn(pm), SLn(pm), and PSLn(pm). There are specialcases where simplifications arise. Clearly GLn(2) = SLn(2), since the determinantcan only have one non-zero value, and it is also the same as PSLn(2), since inG F(2), 1 = −1.

All the projective groups are simple, except for PSL2(2). Sometimes you willsee the notation PSLn(pm) = Ln(pm). By counting their number of elements, wecan see the isomorphisms

PSL2(2) ∼ S3; PSL2(3) ∼ A4. (9.24)

The smallest simple group is also of this type, as

PSL2(4) ∼ PSL2(5) ∼ A5. (9.25)

By specializing the matrices of GLn(pm), one can find several infinite families ofsimple groups. These are the groups of Lie type, whose construction will have toawait our discussion of Dynkin diagrams.

9.1.1 PSL2(7)

PSL2(7) is the second smallest simple group. It is not an alternating group, and itis a perfect group.

To start, we easily see that PSL2(7) contains only three diagonal matrices

e =(

1 00 1

), b =

(3 00 5

), b2 =

(−5 00 −3

)=

(5 00 3

), (9.26)

Page 210: Pierre Ramond Group Theory a Physicists Survey 2010

198 Finite groups: the road to simplicity

using mod(7) addition and multiplication, and remembering that projective trans-formations are the same as their negatives. Thus b generates a Z3 subgroup. Wecan identify another subgroup. The matrix

a =(

1 10 1

), (9.27)

is easily seen to generate a Z7 subgroup; indeed, since

ak =(

1 k0 1

), k ∈ G F(7), (9.28)

it follows that a7 = e. Now let us form the product

b−1ab =(

5 00 3

)(1 10 1

)(3 00 5

)=

(1 40 1

)= a4. (9.29)

It therefore has a semi-direct product structure, so that

Z7 � Z3 ⊂ PSL2(7). (9.30)

This subgroup of order 21 is sometimes called the Frobenius group.We can recognize yet another subgroup. Adjoin to the above Z3 subgroup

generated by b, the matrix

c =(

0 16 0

), c2 = e, (9.31)

and we see that

cbc−1 =(

2 00 4

)=

(5 00 3

)= b2, (9.32)

which indicates a semi-direct product Z3 � Z2 = S3, the permutation group onthree letters. A convenient presentation of PSL2(7) in terms of the generators ofZ3 and Z7 is given by < a, b | a7 = e, b3 = e, (ba)3 = (ba4)4 = e >.

9.1.2 A doubly transitive group

Consider a set of seven points, imaginatively labeled as 1, 2, 3, 4, 5, 6, 7. One coulduse any label or names, for instance (Bashful, Doc, Dopey, Grumpy, Happy, Sleepy,Sneezy), or (E, G, A, L , O, I, S). We wish to construct the group formed by thosepermutations that permute the seven pillars (of wisdom).⎡⎣ 1

24

⎤⎦ ⎡⎣ 235

⎤⎦ ⎡⎣ 346

⎤⎦ ⎡⎣ 457

⎤⎦ ⎡⎣ 561

⎤⎦ ⎡⎣ 672

⎤⎦ ⎡⎣ 713

⎤⎦

Page 211: Pierre Ramond Group Theory a Physicists Survey 2010

9.1 Matrices over Galois fields 199

This group is not triply transitive since it does not contain transformations that mapall sets of three points (for instance [1 2 4] does not map into [1 2 6]). We will seethat it is doubly transitive.

We start by determining its subgroup H that leaves the first column invariant. Itcontains permutations that leave the three points in the first pillar invariant: 1 → 1,2 → 2, 4 → 4. Inspection of the other six columns yields only four elements,depending on the transformation of the 3 element:

(3)(5)(6)(7), (35)(67), (36)(57), (37)(56), (9.33)

yielding a subgroup of order four. With three elements of order two, it is theAbelian dihedral group D2 = Z2 × Z2. Next, we can also transpose 2 and 4,a = (24) without changing the first column. This produces 8 = 2·4 transformationsthat leave the 1 element invariant

H2 : D2 + D2 a = D4. (9.34)

The only other transformations that leave the first column invariant are in the three-cycle (1 2 4). One of these is the permutation of order three

b = (124)(3)(567), b3 = e, (9.35)

which we use to find H through the coset construction

H = H2 + H2b + H2b2. (9.36)

This exhausts all transformations that leave the first column invariant. We see thatH contains the 3 · 8 = 24 permutations of S4.

Finally, we consider the transformation that maps the columns into one another,such as the order seven permutation

c = (1234567), c7 = e, (9.37)

which maps the first column into the second one. Similarly, c2 maps the first col-umn into the third one, and so on. This produces the group we are seeking throughthe coset decomposition

G = H + Hc + · · · + Hc6. (9.38)

This is a group of order 7 · 24 = 168. Geometrically, we can think of the sevenpillars as lines, leading to the interpretation of a projective geometry with sevenpoints and seven lines, such that two points are only on one line. It follows that thisgroup is doubly transitive.

Page 212: Pierre Ramond Group Theory a Physicists Survey 2010

200 Finite groups: the road to simplicity

We can summarize this construction by means of the following diagram.

��

��

��

��

��

��

������������

��

��

��

1 2 4

3

5

7 6

This picture is useful to identify the subgroup that leaves the line [1 2 4] invariant.We can permute the remaining points without changing the picture: the three trans-positions (6 7), (5 7), and (3 7) do not alter the picture but relabel the points whilekeeping the base of the triangle invariant. For example, (6 7) simply interchangesthe two outer edges of the triangle. All three generate the permutations on the fourpoints [3, 5, 6, 7]: the subgroup that leaves one line invariant is nothing but S4.

This geometry is an example of a Steiner system, in this case S(2, 3, 7): sevenpoints assembled in sets of three, such that any two points belong only to one three-set. In general, finite projective planes can be defined in terms of Steiner systems asfollows: let p be a prime number, and consider a geometry with p2 + p + 1 points,such that each line will contain p + 1 points, with two points lying on only oneline. These form the Steiner systems S(2, p + 1, p2 + p + 1). Ours is the simplestnon-trivial case with p = 2.

Sylow’s theorems tell us that PSL2(7) has elements of order 2, 3, 4 and 7. Theyare conveniently displayed in the presentation < a, b | a2 = e, b3 = e, (ab)7 = e,([a, b])4 = e >, where [a, b] = a−1b−1ab.

The generators have simple expressions in terms of (2×2) matrices over G F(7)

a =(

0 −11 0

), b =

(0 −11 1

). (9.39)

Note that a belongs to one class with 42 elements of order 2, while b gives rise toa second class with 56 elements of order 3. The element

[a, b] =(

1 11 2

)(9.40)

Page 213: Pierre Ramond Group Theory a Physicists Survey 2010

9.2 Chevalley groups 201

generates a class of 21 elements of order 4. Finally, the two elements of order 7

ab =(

1 10 1

), abb =

(1 01 1

), (9.41)

generate two classes, each of 24 elements of order 7. It follows that there are sixirreps. Only one of these is one-dimensional. To see this, note that in such irrep, aand b are simply commuting numbers, so that

a2 = 1, b3 = 1, (ab)7 = ab = 1, (9.42)

with unique solution a = b = 1. The rest of the character table is given as follows.

PSL2(7) C1 21C [2]2 (a) 56C [3]

3 (b) 42C [4]4 ([a, b]) 24C [7]

5 (ab) 24C [7]6 (ab2)

χ [1] 1 1 1 1 1 1

χ [3] 3 −1 0 1 12

(−1 + i

√7)

12

(−1 − i

√7)

χ [3] 3 −1 0 1 12

(−1 − i

√7)

12

(−1 + i

√7)

χ [6] 6 2 0 0 −1 −1

χ [7] 7 −1 1 −1 0 0

χ [8] 8 0 −1 0 1 1

This group is one of an infinite family of finite simple groups, and there are manyothers. Thankfully, there is a systematic way to construct all finite simple groups.This will be the subject of the next section.

9.2 Chevalley groups

We have seen that there exists a basis for any Lie algebra where all the struc-ture functions and matrix elements are integers, the Chevalley basis. Chevalleyobserved that a Lie group element in this basis with parameters belonging to a finiteGalois field G F(q) will generate a finite group. They are the Chevalley groups,most of which are simple. For details the reader is referred to Roger Carter’s classicbook [3].

• An(q), generated by elements of the form

Ui (u) ≡ e u ei , U−i (u) ≡ e u e−i , (9.43)

Page 214: Pierre Ramond Group Theory a Physicists Survey 2010

202 Finite groups: the road to simplicity

where ei are the generators associated with the simple roots in the Chevalley basis ofAn , i = 1, 2, . . . , n, and where u is an element of the Galois field, G F(q), with uq = 1,q being any power of a prime.

Let us see how this works for n = 2, where we use the (3 × 3) representation of thetwo nilpotent matrices. The group elements are generated by the matrices⎛⎝1 u 0

0 1 00 0 1

⎞⎠ ,

⎛⎝1 0 00 1 u0 0 1

⎞⎠ , (9.44)

as well as ⎛⎝1 0 0u 1 00 0 1

⎞⎠ ,

⎛⎝1 0 00 1 00 u 1

⎞⎠ . (9.45)

All have unit determinant, and generate PSLn+1(q). Most are simple with fewexceptions: PSL2(5) ∼ PSL2(4) ∼ A5, PSL2(3) ∼ A4, and PSL2(2) ∼ S3.

• Bn(q), the classical orthogonal finite groups O2n+1(q), also known as P�2n+1(q),which leave the quadratic form

( u, u ) = u20 + u2

1 + · · · + u22n,

invariant, where the ui belong to G F(q).• Cn(q), the projective groups PSp2n(q), which leave the antisymmetric symplectic

quadratic form

( u, v ) =n∑

i=1

(ui vi+n − vi ui+n)

invariant. All but PSp2(2) ∼ S3, PSp2(3) ∼ A3, and PSp4(2) ∼ S6, are simple.• Dn(q) are classical finite groups of the orthogonal type. Here the situation is a little

bit more complicated because we are dealing with finite fields. Two different quadraticinvariants can be defined for finite orthogonal transformations. This is because someelements of finite fields may not have a square root, as we have seen.

Consider a symmetric quadratic form (u, v), where u, v are m-dimensional vec-tors over G F(q), q odd. Suppose that we have managed to find a basis in terms ofm orthogonal vectors ei , that is

( ei , e j ) = 0, i �= j. (9.46)

So far so good. If the field elements (ei , ei ) are squares, a trivial redefinition normalizesthem to the unit element. If this is true for all basis vectors, we arrive at the familiarquadratic invariant (with a + superscript)

(u, u )+ =2n∑

i=1

u2i = u2

1 + · · · + u22n,

Page 215: Pierre Ramond Group Theory a Physicists Survey 2010

9.2 Chevalley groups 203

where

u =2n∑

i=1

ui ei. (9.47)

However, suppose there is one vector ek for which

( ek, ek ) = v,

where v is not a square element of the finite field. Then we have to define a new quadraticform

( u, u )− =2n∑

i=1,�=k

u2i + v u2

k .

The case with two vectors whose norm is not a square, is reduced by judicious lineartransformations to the plus case. Hence there are only two distinct possibilities.

The Chevalley groups Dn(q) are the orthogonal groups defined in terms of the +quadratic form. They are called O+

2n(q), or P�+2n(q). There are interesting isomor-

phisms, O+4 (q) = PSL2(q) × PSL2(q), O+

6 (q) = PSL4(q).When q is even, a whole new discussion is needed.

The remaining Chevalley groups are constructed from the exceptional groups,yielding G2(q), F4(q), E6(q), E7(q), and E8(q).

Twisted Chevalley groups

The Chevalley construction can be further refined to produce more groups. Wehave seen that the Dynkin diagrams of the A − D − E Lie algebras have a paritywhich maps the Dynkin diagram into itself. We also noted the Frobenius map ofthe elements of the Galois field into itself; in particular when q = p2, it is also aparity, splitting the elements into even and odd parts.

The combined operation of both order two maps, one on the Dynkin diagram, theother on the Galois field, maps the Chevalley groups An(p2), Dn(p2), and E6(p2)

into themselves.The beautiful thing is that each contains a subgroup, the twisted Chevalley

groups, which are invariant under the combined mapping. If we split the Galoisfield elements into even and odd elements u±, and the generators as e±, even andodd under Dynkin parity, these groups are generated by

U+i (u+) = e u+ e+

i , U−i (u−) = e u− e−

i , (9.48)

and their conjugates.Finally, to match the order three Z3 automorphism of the SO(8) Dynkin dia-

gram, we consider a Galois field G F(p3) which has an order-three Frobeniusautomorphism. In this way we obtain the following.

Page 216: Pierre Ramond Group Theory a Physicists Survey 2010

204 Finite groups: the road to simplicity

• 2 An(p2) ≡ Un+1(p2) = PSUn+1(p2). Defined for n ≥ 2, all but 2 A2(22) are simple.

They are the classical unitary groups, whose elements can be viewed as matrices act-ing on (n + 1)-dimensional vectors with elements in the Galois field, and which leaveinvariant the quadratic form

n+1∑j=1

u j u j , (9.49)

where u j = u pj is the Frobenius conjugate.

For 2 A2(p2), the Dynkin automorphism is a Weyl reflection about the line whichlinks the origin to the (1, 1) state. Thus we can readily identify the even (3×3) matricesin the Chevalley basis ⎛⎝0 1 0

1 0 00 0 0

⎞⎠ ,

⎛⎝0 0 00 0 10 1 0

⎞⎠ . (9.50)

Exponentiating with even elements of the Galois field yields the desired group elements,as well as the exponentiation of the odd Lie algebra elements⎛⎝0 1 0

1 0 00 0 0

⎞⎠ ,

⎛⎝0 0 00 0 10 1 0

⎞⎠ , (9.51)

with the odd Galois elements.Most are simple, except for U2(2) = S3, U2(3) = A3, U3(2) = (Z3 × Z3) � Q8

• 2 Dn(p2) ≡ O−2n(q) = P�−

2n(q). These are the finite classical orthogonal group definedfor n ≥ 4 (n = 3 reduces to A3(p2)), which leaves the quadratic form

( u, u )− =2n∑

i=1,�=k

u2i + v u2

k,

invariant. All are simple, with notable isomorphisms, O−4 (q) = PSL2(q2), O−

6 (q) =U4(q).

• 2 E6(p2) are all simple.• 2 D4(p3) are simple as well.

Suzuki and Ree groups

Their construction is much sneakier. It starts with the remark that the Dynkin dia-grams of B2, G2 and F4 with simple roots of different lengths also have a (morecomplicated) automorphism: flip the Dynkin diagram and interchange the shortand long roots.

Looking at the root diagrams of B2 and G2, take the two simple roots and drawthe line that bisects the angle between them, shown here for B2.

Page 217: Pierre Ramond Group Theory a Physicists Survey 2010

9.3 A fleeting glimpse at the sporadic groups 205

��

��

��

A reflection of the large root about the bisectrix (indicated by an arrow) yields aroot of the same length, but in the direction of the shorter root, and vice versa. Thisreflection can be equally well applied to the other roots. The net result is simplya rotation of the Dynkin diagram about the origin by π/4 for B2 and π/6 for G2.Doing it again reproduces the original diagram, so this is an automorphism of ordertwo. In the case of F4, it is harder to draw but we can define a three-dimensionalplane that bisects the four simple roots, and the reflections about this plane producethe same result.

As in the previous construction of twisted Chevalley procedure, this Dynkinsymmetry must be matched with that of Galois fields which contain a specifiedFrobenius map. For B2 and F4, the Galois field must have characteristic two, andcontain a map that is of the form the Frobenius map must satisfy

u → u′ = uθ , 2 θ2 = 1.

The Galois field can only be G F(22l+1), for any integer l.For G2, the Galois field must be of characteristic three, and contain a similar map

with 3 θ2 = 1, which restricts it to G F(32n+1). Only then are the group propertiessatisfied. Hence we have three more families

2 B2

(22n+1

), 2 F4

(22n+1

), 2G2

(32n+1

).

Most are simple with the exceptions

2 B2(2) = Z5 � Z4,2G2(3) = PSL2(8) � Z3,

2 F4(2) = T · 2,

where T is the Tits group, which is simple as well.This almost completes the list of simple finite group. However, there are twenty-

six more simple groups which defy all classification: the sporadic groups, to whichwe now turn.

9.3 A fleeting glimpse at the sporadic groups

Emile Léonard Mathieu discovered in 1860 (and 1873) five groups of multiplytransitive permutations. Known today as Mathieu groups, they make up the firstgeneration of sporadic groups, the twenty-six simple non-Abelian finite groupswhich are neither alternating nor Chevalley groups.

Page 218: Pierre Ramond Group Theory a Physicists Survey 2010

206 Finite groups: the road to simplicity

They are, in order of order:

• M11 group of 24 · 32 · 5 · 11 quadruply transitive permutations on eleven objects,• M12 group of 26 · 33 · 5 · 11 quintuply transitive permutations on twelve objects,• M22 with 27 · 32 · 5 · 7 · 11 triply transitive permutations of twenty-two objects,• M23 group of 27 · 32 · 5 · 7 · 11 · 23 quadruply transitive permutations on twenty-three

objects, and• M24 with 210 ·33 ·5 ·7 ·11 ·23 quintuply transitive permutations on twenty-four objects.

They show up as automorphism groups of Steiner systems, as shown in the tablebelow.

M11 M12 M22 M23 M24

S(4, 5, 11) S(5, 6, 12) S(3, 6, 22) S(4, 7, 23) S(5, 8, 24)

These groups are quite extraordinary, but with the help of Galois fields, theyare generated by relatively simple transformations. Complicated groups can begenerated by simple operations; for instance the transformations

x → x ′ = x + 1, x → x ′ = −x5 − 2x2,

with x ∈ G F(7), generate our old friend PSL2(7).One can adjoin to Galois fields one extra point, called the point at infinity, in

order to define inversion as a group operation. Then M12 is generated by the threetransformations

x → x ′ = x + 1, x → x ′ = 4x2 − 3x7, x → x ′ = −1/x,

where x ∈ G F(11). Its triply transitive subgroup M11 is generated by

x → x ′ = x + 1, x → x ′ = 7x4 − 6x9, x → x ′ = −1/x .

In the same vein, M24, the largest Mathieu group is also generated quite“simply” by

x → x ′ = x + 1, x → x ′ = −3x15 + 4x4, x → x ′ = −1/x,

where x ∈ G F(23). It appears as the automorphism group of the perfectbinary Golay code (for definitions and delightful reading, see T. M. Thompson,From Error Correcting Codes Through Sphere Packings to Simple Groups, CarusMathematical Monographs, MAA, 1983).

It is well beyond the scope of this book to go beyond these.Seven more sporadic groups make up the second generation; all are relevant in

the context of the automorphism group of the Leech lattice, an infinite lattice ofpoints in twenty-four dimensional Euclidean space. It yields the densest packingof spheres in twenty-four dimensions, where one sphere touches 196 560 others!

Page 219: Pierre Ramond Group Theory a Physicists Survey 2010

9.3 A fleeting glimpse at the sporadic groups 207

The Leech lattice can be defined in terms of the light-like vector in 26 (1 + 25)-dimensional discrete Minkowski space-time. The totally unique light-like vectorwith integer components

(70, 0, 1, 2, . . . , 24),

serves to define the points of the Leech lattice by orthogonality. One cannot helpbut notice that this is the same (continuous) space where the quantum bosonicstring roams.

The remaining fifteen sporadic groups make up the third generation. All areconnected to the largest of them all, the Monster. It is beyond the scope of thisbook to go into any more details. These very unique mathematical objects havenot yet proven very useful in physics. Yet, given Nature’s attraction to unusualmathematical structures, it is easy to think they will some day assume physicalrelevance.

Page 220: Pierre Ramond Group Theory a Physicists Survey 2010

10

Beyond Lie algebras

10.1 Serre presentation

We have seen that Dynkin diagrams contain the information necessary to recon-struct the simple Lie algebras. Since that information is encoded in the Cartanmatrix, there must be a way to build the algebra directly from it. This procedure,we alluded to earlier, is called the Serre presentation of the Lie algebra.

Let us review the properties of the Cartan matrices of the simple Lie algebras.The Cartan matrix for a rank r Lie algebra is a (r ×r) matrix A with integer entriesarranged such that:

• its diagonal elements are equal to 2;• its off-diagonal elements are negative integers or zero;• if one of its off-diagonal elements is zero, so is its transpose.

Our explicit constructions suggest three more properties:

• it can always be brought to symmetric form S by multiplication by a diagonal matrix Dwith positive entries: S = DA;

• it cannot be written in block-diagonal form by rearranging its rows and columns;• A has only positive eigenvalues.

For SU (2), the Cartan matrix is just a number, with only one “diagonal” generatorH1, and one root α1 with corresponding operators Eα1 and its conjugate E−α1 , with

[ H1, E±α1 ] = ±2E±α1,

defined in such a way that the numerical factor on the right-hand side is the Cartanmatrix element. With this normalization,

[ Eα1, E−α1 ] = H1,

which completely defines the SU (2) algebra.

208

Page 221: Pierre Ramond Group Theory a Physicists Survey 2010

10.1 Serre presentation 209

Recall that for SU (3), we have the (2 × 2) Cartan matrix

ASU (3) =(

2 −1−1 2

),

two diagonal generators Hi , and two simple roots, αi i = 1, 2, with correspondinggenerators E±αi . With the normalization

[ Hi , E±α j ] = ±A ji E±α j ,

we obtain

[ Eαi , E−α j ] ={

Hi i = j,

0 i �= j.

So far these yield two SU (2) sub-algebras, which accounts for only six out of itseight elements. The extra generators are to be found in the commutator

[ Eα1, Eα2 ],and its conjugate. This exhausts the number of SU (3) generators. Hence, any otherpossible generator, written as multiple commutators between the Eαi , must vanish,that is

[ Eα1, [ Eα1, Eα2 ] ] = [ Eα2, [ Eα2, Eα1 ] ] = 0.

Serre cleverly rewrites these relations in terms of the off-diagonal elements of theCartan matrix as [

Ad(Eαi

]1−A ji Eα j = 0, i �= j,

since A12 = A21 = −1, and where the adjoint operation Ad(E)k is just the k-foldcommutator

[Ad(E)]k X = [ E, [ E, · · · [ E, X ] · · · ] ]︸ ︷︷ ︸k

.

A similar expression holds for the conjugate equation. This is called the Serrepresentation, which can be shown to completely define the Lie algebra in terms ofits Cartan matrix and its generators Eαi . To simplify notation, we let Eαi ≡ Ei ,E−αi ≡ Ei , i = 1, 2, · · ·, r .

To instill faith in this method, let us build the exceptional G2 algebra from itsCartan matrix

AG2=

(2 −3

−1 2

).

Like SU (3), it has two generators Ei , i = 1, 2, but now we can form many moreby commutation. Indeed, the Serre relations now read (A21 = −1)

[ E2, [ E2, E1 ] ] = 0,

Page 222: Pierre Ramond Group Theory a Physicists Survey 2010

210 Beyond Lie algebras

and (A12 = −3)

[ E1, [ E1, [ E1, [ E1, E2 ] ] ] ] = 0.

These leave room to construct more generators by commutation. First we have

[ E2, E1 ],the next in line triple commutator being forbidden by the first Serre relation. Fromthe second Serre relation, we can form

[ E1, [ E1, E2 ] ], [ E1, [ E1, [ E1, E2 ] ] ].Finally we have the four-fold commutator

[ E2, [ E1, [ E1, E2 ] ] ].Together with E1,2, these account for six generators. Together with their conjugatesand Hi , they form the fourteen-dimensional G2 algebra, and the reader can verify,through the use of the Jacobi identities, that all other commutators vanish.

10.2 Affine Kac–Moody algebras

In our analysis of embeddings, we came across extended Dynkin diagrams, used todescribe maximal embeddings without loss of rank. They are obtained by addinga node associated with the negative of the highest root to the original Dynkindiagram. Since the highest root is a linear combination of the simple roots, thecorresponding extended Cartan matrix has vanishing determinant, and one zeroeigenvalue, all other eigenvalues being positive.

Let us apply the Serre presentation to the singular (2×2) SU (2) extended Cartanmatrix

A(1)SU (2) =

(2 −2

−2 2

),

and analyze in detail the algebraic structure it produces. By convention, we labelthe extra node with a zero subscript, associated with the root α0.

The Serre-generated algebra starts with H0, H1, E0, E0, E1, and E1, andcommutators

[ H0, H1 ] = [ E1, E0 ] = 0,

[ H0, E0 ] = 2E0, [ H0, E1 ] = −2E1,

[ H1, E0 ] = −2E0, [ H1, E1 ] = 2E1,

[ E0, E0 ] = H0, [ E1, E1 ] = H1.

Page 223: Pierre Ramond Group Theory a Physicists Survey 2010

10.2 Affine Kac–Moody algebras 211

The Serre relations are now

[Ad(E0)]3 E1 = 0, [Ad(E1)]3 E0 = 0.

We first note that one linear combination of the Cartan sub-algebra commuteswith all the members of the algebra since

[ H0 + H1, Ei ] = 0.

This is easily traced to the vanishing of the determinant of the extended Cartanmatrix: these algebras have a non-trivial center, and this linear combination is justa c-number, written as

2k = H0 + H1.

Unlike Lie algebras, repeated commutations yield an infinite set of new elements.Start with

[ E1, E0 ],associated with the root

δ ≡ (α0 + α1).

It is a peculiar root of zero length,

(δ, δ ) = (α0, α0 ) + (α0, α1 ) + (α1, α0 ) + (α1, α1 ) = 0

and is orthogonal to the simple root of the algebra

(δ, α1 ) = (α0, α1 ) + (α1, α1 ) = 0.

The new elements

[ E0, [ E0, E1 ]], [ E1, [ E1, E0 ]],are associated with the roots

(2α0 + α1) = δ − α1, (2α1 + α0) = δ + α1.

On the other hand, the vanishing of the commutator [ E0, E1 ] tells us thatα0 − α1 = δ − 2α1 is not a root. However, 2δ is also a root since the element

[ E1, [ E0, [ E0, E1 ]]],does not vanish. We can then commute it with E1 to obtain a new root 2δ +α1, andso on. In this way, we generate infinite towers of roots of the form

±δ, ±2δ, · · · ; (±α1 ± δ), (±α1 ± 2δ), · · · ,yielding the infinite-dimensional affine Kac–Moody algebra.

Page 224: Pierre Ramond Group Theory a Physicists Survey 2010

212 Beyond Lie algebras

This proliferation of roots is rather frightening, but it can be expressed diagram-matically in a two-dimensional root diagram: setting α1 as a unit vector along thehorizontal, and δ along the vertical. We then have an infinite tower of roots asshown below.

���

���

������������

���������

α−α

−δ

δ

−α−δ α−δ

α+δ−α+δ

It is nothing but an infinite stack of SU (2) weight diagrams, each differing by δ

in the vertical direction. Invariance of this weight diagram under vertical transla-tions explains the mathematical term affine, which is associated with structures thatreproduce under translations.

The Weyl group generalizes to the affine case in an obvious way: it is generatedby the Weyl transformations on each line, that is w(α1) = −α1, and translationsfrom one stack to the next, say α1 → α1 +δ. The latter generate an infinite Abeliangroup.

This pattern generalizes to all affine Kac–Moody algebras: the same weight dia-gram as for the finite sub-algebra, translated ad infinitum by translations in thedirection of the null root. The Weyl group has the corresponding structure.

The complicated-looking commutation relations can be cast elegantly by stick-ing close to those of the underlying SU (2) algebra. We start from the familiaridentification of its generators through

E1 = (T 1

0 + iT 20

) ≡ T (1+i2)0 , E1 = (

T 10 − iT 2

0

) ≡ T (1−i2)0 ,

H1

2= T 3

0 ,

where we have added a zero subscript. They lead to the usual SU (2) commutationrelations [

T a0 , T b

0

] = iεabc T c0 ,

a, b, c = 1, 2, 3. We then organize the new elements in terms of their trans-formation properties with T 3

0 . Since the new element associated with the specialzero-length root δ,

[ E1, E0 ],

Page 225: Pierre Ramond Group Theory a Physicists Survey 2010

10.2 Affine Kac–Moody algebras 213

commutes with T 30 , we define it as

T 31 ≡ 1

2[ E1, E0 ], T 3 †

1 ≡ T 3−1.

This suggests the definitions

E0 ≡ T (1−i2)1 , E0 ≡ T (1+i2)

−1 ,

so as to satisfy [T (1+i2)

0 , T (1−i2)1

] = 2T 31 .

Note that the operators T a1 and T a

−1 are no longer real but are hermitian conjugatesof one another. It follows that[

T (1+i2)−1 , T (1−i2)

1

] = 2T 30 − 2k.

The new element

T (1+i2)1 ≡ 1

2[ E1, [ E1, E0 ]]

yields, after a modest amount of Jacobi gymnastics,[T (1+i2)

1 , T (1−i2)0

] = 2T 31 .

More Jacobi acrobatics produces the suggestive commutator[T (1+i2)

1 , T (1−i2)−1

] = 2T 30 + 2k

(note the change in sign of the c-number k), where

T (1−i2)−1 ≡ 1

2

[E0,

[E0, E1

]],

is the hermitian conjugate. All other elements of the algebra are obtained bycommutation. For instance define the element with root 2δ through[

T (1+i2)1 , T (1−i2)

1

] = 1

2[ E0, [ E1, [ E0, E1 ]]] ≡ 2T 3

2 ,

and so on. The full algebra is surprisingly simple to state in terms of SU (2)-likecommutation relations,[

T am , T b

n

] = iεabc T cm+n + m kδabδm+n,0, m, n = 0,±1,±2, . . ..

The undetermined constant k is at the heart of Kac–Moody algebras. It is useful toadjoin to this algebra an operator L0 which counts the subscript, that is

[ L0, T an ] = n T a

n .

Page 226: Pierre Ramond Group Theory a Physicists Survey 2010

214 Beyond Lie algebras

Before leaving SU (2), we note the existence of another Cartan-like matrix (neg-ative integer off-diagonal entries and diagonal elements equal to 2) with vanishingdeterminant

A(2)1 ≡

(2 −1

−4 2

).

The Serre presentation applies equally well, and generates a new infinite-dimensional algebra, the twisted affine Kac–Moody algebra.

Coming back to the general case, the identification of the generators appliesequally well to an arbitrary rank r Lie algebra. The elements belonging to the rootδ are now the r -fold commutators involving one E0, and we identify them with T j

1

where j runs over the r labels of the Cartan sub-algebra. One only needs to identifythe generators with subscript equal to one in order to generate the full algebra.

The null root is defined in terms of the marks as

δ =r∑

i=0

ci αi ,

so as to insure that it has zero length, and is orthogonal to all the other roots of theunderlying Lie algebra,

( δ, αi ) = 0, i = 1, 2, . . . , r.

The procedure is of course more algebraically tedious, but the result is to producethe general commutation relations[

T am , T b

n

] = i f abc T cm+n + m kδabδm+n,0, m, n = 0,±1,±2, . . . ,

obtained by simply replacing εabc with a general structure function f abc!

Representation theory

The representation theory of affine Kac–Moody algebras proceeds in much thesame way as for Lie algebras, except that all are infinite-dimensional. This shortsection only skims the surface of this rich subject. The interested reader is referredto S. Kass, R. V. Moody, J. Patera, and R. Slansky [13], and Fuchs [5].

We use affine SU (2) as an example, where the states are labeled by two integers(a0 a1). When both a0, a1 ≥ 0 the state lies inside or on the walls of the Weylchamber, and it can be used to label an irreducible representation. The rest of thestates in the representation are obtained by subtracting the positive roots from thehighest weight state. In the Dynkin basis, the positive roots are given by the row ofthe extended Cartan matrix,

α0: (2, −2), α1: (−2, 2).

Page 227: Pierre Ramond Group Theory a Physicists Survey 2010

10.2 Affine Kac–Moody algebras 215

We note that the linear combination with the co-marks

c0 a0 + c1 a1,

is invariant under subtraction by the roots because the co-marks are the componentsof the zero eigenvector of the Cartan matrix. It is called the level of the represen-tation. In this case it is just l = a0 + a1, and it can serve as a partial label for therepresentation.

We can subtract α0 a0-times from the highest weight, and α1 a1-times. Thisresults in a zig-zag pattern for the infinite number of states in the representation.The zig (zag) direction denotes subtraction by α0(α1). To keep track of these states,weights are assigned a null weight which tells how many α0 subtractions it is awayfrom the highest weight. These weights assemble in an infinite number of SU (2)representations.

Explicit examples of these representations are found in the physics literature.When k = 0, this algebra is rather trivial to realize: introduce an infinite number

of oscillators, each of which transform as SU (2) doublets,[am, a†

n

] = δm,n, m, n = 0,±1,±2, . . . ,

and define the generalized Schwinger representation

T am = 1

2

∞∑j=−∞

a†j+m σ a a j . (10.1)

It is easily seen to satisfy the Kac–Moody algebra with k = 0. We also have

L0 =∞∑

j=−∞j a†

j a j . (10.2)

Oscillator representations with k �= 0 are less obvious, but c-numbers genericallyoccur in commutators of quadratic combinations when normal ordering is invoked(: :→ annihilation operators on the right),[

a†1 a†

2, a1 a2

]= −a†

1 a1 − a2 a†2 = −a†

1 a1 − a†2 a2 − 1,

= − :(

a†1 a1 + a†

2 a2

): −1.

The Serre presentation is used to generate yet more general algebras with morecomplicated Cartan matrices. When S has one negative eigenvalue, the rest beingpositive, it leads to the Hyperbolic Algebras. Some have found application in thedescription of physics near a black hole singularity (M. Henneaux, D. Persson, andP. Spindel, Living Rev. Rel. 11, 1 (2008))!

Page 228: Pierre Ramond Group Theory a Physicists Survey 2010

216 Beyond Lie algebras

There are also connections with graph theory. The matrix elements of a matrixof the form A = DS must satisfy certain relations among themselves. For a (3×3)the matrix condition is

a12a23a31 = a13a32a21.

This equation has a natural graphical explanation. Consider a graph with threepoints, with ai j denoting the transformation from point i to point j . This equationtranslates into the equivalence of the transformation 1 → 2 → 3 and its inverse3 → 2 → 1.

10.3 Super algebras

Super algebras are an extension of Lie algebras which contain two types ofelements, even elements which satisfy commutation relations and contain Lie alge-bras, and odd elements which transform as representations of the Lie sub-algebras,and have the important property that their anticommutators close into the even ele-ments. This last condition is necessary but not sufficient, as a suitably generalizedJacobi identity must also hold.

To see how this works, consider SU (2). We know that its doublet representationsatisfies the first criterion

2 × 2 = 3,

and since this is a strict equality, the graded Jacobi identity must follow. Specifi-cally, we denote the SU (2) generators as T A with A = 1, 2, 3, and the complexdoublet by qα, with α = 1, 2. Clearly

[ T A, qα ] = 1

2(σ A)αβ qβ. (10.3)

Rather than taking the complex conjugate, we form the combination

qα = (σ2)αβ qβ, (10.4)

which also transforms as a doublet

[ T A, qα ] = 1

2(σ A)αβ qβ. (10.5)

Next, we form the anticommutator

{ qα, qβ } = (σ Aσ2)α β T A, (10.6)

and we verify that the Jacobi identities

[ T A, { qα, qβ } ] + { qα, [ qβ, T A ] } + { qβ, [ qα, T A ] } = 0, (10.7)

Page 229: Pierre Ramond Group Theory a Physicists Survey 2010

10.3 Super algebras 217

and

[ qα, { qβ, qγ } ] + [ qβ, { qγ , qα } ] + [ qγ , { qα, qβ } ] = 0, (10.8)

hold. Similarly, the anticommutator

{ qα, qβ } = (σ Aσ2

)α β

T A, (10.9)

also satisfies the graded Jacobi identities. Together with

[ T A, T B ] = iεABC T C , (10.10)

these define the smallest super algebra, made up of T A, qα and its conjugate. Thisis the simplest case of a symplectic super algebra.

All Lie super algebras can be constructed in the following manner: begin witha Lie algebra or direct products of Lie algebras with the property that its funda-mental representation are such that their symmetric product, corresponding to theanticommutator, belong to the adjoint representation of the Lie algebra. Using ournewly acquired knowledge of the properties of the low-lying representations of Liealgebras, it is straightforward to list the possibilities.

Since the symmetric product of the fundamental 2n of Sp(2n) is exactly itsadjoint representation

2n × 2n = n(2n + 1),

and the graded Jacobi identities are automatically satisfied. This infinite family ofsuper algebras is denoted by O Sp(2n), and the smallest one is O Sp(2), which wehave just explicitly constructed.

The only other simple Lie algebra with the property that the symmetric productof its fundamental contains the adjoint is the exceptional group E7. Its fundamental56 is pseudoreal and satisfies

56 × 56 = 133 + 1463,

but the Jacobi identities are not satisfied: whenever the symmetric product containsother representations, the validity of the Jacobi identities is not automatic, andneeds to be checked. With this caveat in mind, we simply list the simple superalgebras, following the work of V. Kac.

Much of the machinery we have developed for Lie algebras applies to the superalgebras, by a careful distinction of even and odd elements. Assign to any elementof a Lie super algebra X an index x , which is equal to zero if X is an even element,and one if it is an odd element. This enables us to introduce the elegant notationfor the graded commutators

{ X, Y ] ≡ X Y − (−1)xy Y X, (10.11)

Page 230: Pierre Ramond Group Theory a Physicists Survey 2010

218 Beyond Lie algebras

in terms of which the graded Jacobi identity becomes

{ X, { Y, Z ] ] + { Y, { Z , X ] ] + { Z , { X, Y ] ] = 0. (10.12)

The Lie super algebras are constructed from as many O Sp(2) as the rank, just asLie algebras of rank r are generated by r copies of non-commuting SU (2) algebras.The explicit construction is beyond the scope of this book, and we refer the readerto V. Kac’s book [12] for detailed constructions.

The simple super algebras are as follows.

• SU (n | m). The even elements form the Lie algebra SU (n) × SU (m) × U (1). The oddelements transform as (n, m) and its complex conjugate. Hence its content

(1, 1) +(

1, n2 − 1)

+(

n2 − 1, 1)

+ (n, m) f + (n, m) f .

The f subscript applies to the odd elements. The full super algebra can be convenientlywritten in terms of (m + n) × (n + m) hermitian matrices of the form(

SU (n) (n, m) f

(n, m) f SU (m)

)+

(n 00 m

).

The diagonal matrix generates U (1). The graded commutator of these matrices is suchthat the off-block-diagonal odd elements obey anticommutation relations while theblock-diagonal elements satisfy commutation relations. These rules single out the super-trace, defined as the trace of the upper diagonal block minus the trace of the lowerdiagonal block. Thus when n = m, the supertrace of the diagonal matrix vanishes, andthe U (1) decouples from the algebra. The resulting super algebra, SSU (n | n), containsonly SU (n)×SU (n) in its Lie part. Its non-compact form SSU (2, 2 | 4) is the symmetrygroup of the N = 4 super Yang–Mills theory.

• O Sp(n | 2m). The Lie part is SO(m) × Sp(2m), and the odd elements transform asthe real (n, 2m) f . This is the generalization of the smallest Lie super algebra dis-cussed earlier. It is possible because orthogonal and symplectic have opposite symmetryproperties: the antisymmetric (symmetric) product of the fundamental of the orthogonal(symplectic) vector representation contains the adjoint representation of the algebra.

• A[n − 1]. The Lie part is SU (n). The odd elements are the symmetric traceless (n × n)matrices which transform as the adjoint of SU (n). Recall that SU (n) algebras con-tain (for n ≥ 3) the adjoint in both the symmetric (d-coupling) and antisymmetric( f -coupling since it is a Lie algebra) product of its adjoint representation.

• P(n − 1) and P(n − 1). These super algebras also contain SU (n) as their Lie part. Theodd elements are different; for P(n − 1), they are the symmetric second rank tensorS(ab), and the second rank antisymmetric tensor A[ab]. This is a complex super algebra,and its conjugate P(n − 1) contains the symmetric S(ab), and the antisymmetric A[ab].These require some getting used to. For n = 3, the odd part is the sum of 6 and 3. Hencemany of the anticommutators vanish since only 6 × 3 contains the adjoint 8.

Page 231: Pierre Ramond Group Theory a Physicists Survey 2010

10.3 Super algebras 219

• F[4]. Its even Lie part is SO(7)×SU (2). Its sixteen complex odd elements transform as(8, 2). The 8 here is the real SO(7) spinor representation. It has the same pseudorealityas the SU (2) doublet. Now

(8, 2) ×s (8, 2) = (8 ×s 8, 2 ×s 2

) + (8 ×a 8, 2 ×a 2

),

where ×s,a denotes the symmetric and antisymmetric products. Since 8 ×s,a 8 containthe SO(7) singlet and adjoint 21 respectively, the above product contains the sum of theadjoints, as required.

• G[3]. The Lie part is G2 × SU (2), under which the odd elements transform as (7, 2).It has similar structure as F[4], since 7 ×s,a 7 respectively contain the G2 singlet andadjoint 14. It is the only super algebra with an exceptional group in its Lie part.

• D(2 | 1 ; a). The even elements are SU (2) × SU (2) × SU (2). The odd elements aredoublets under each of these, that is they transform as (2, 2, 2 ). In order to explainthe parameter a, introduce spinor indices α, β, γ for each SU (2). The odd elements arethen represented as χα β γ , such that under each SU (2), generated by R A, S A and T A

respectively, [R A, χα β γ

]= 1

2

(σ A

)αα′ χα′ β γ ,

[S A, χα β γ

]= 1

2

(σ A

)ββ ′ χα β ′ γ ,

[T A, χα β γ

]= 1

2

(σ A

)γ γ ′ χα β γ ′ .

The anticommutator contains three terms, by symmetrizing on one SU (2) and antisym-metrizing on the other two. Specifically{

χα β γ , χα′ β ′ γ ′}

= a1 (σ2)αα′(σ2)ββ ′(σ Aσ2

)γ γ ′ T A

+ a2 (σ2)αα′(σ2)γ γ ′(σ Aσ2

)ββ ′ S A

+ a3 (σ2)γ γ ′(σ2)ββ ′(σ Aσ2

)αα′ R A,

with a1 + a2 + a3 = 0, all from the graded Jacobi identity. We fix a1 = 1 for normaliza-tion, which leaves us with a2 = a and a3 = −1 + a, and one undetermined parameter.For the special choices a = 1/2, 2,−1, this algebra reduces to O Sp(4 | 2).

In addition to these super algebras, one finds the so-called hyperclassical simplesuper algebras. They also have even and odd elements, but their even elementsdo not form Lie algebras, with some exceptions. They are constructed out ofGrassmann numbers θi and their derivatives ∂i , i, j = 1, 2, . . . , n, which satisfyanticommutation relations {

θi , θ j

} = {∂i , ∂ j

} = 0, (10.13)

Page 232: Pierre Ramond Group Theory a Physicists Survey 2010

220 Beyond Lie algebras

with

{ ∂i , θ j } = δi j . (10.14)

There are three infinite families of hyperalgebras.

• S(n) and S(n). The elements of S(n) are of the form

S(n) : ∂iPn∂ j + (i ↔ j),

where Pn is the most general polynomial in the Grassmann variables; it terminatesat θ1θ2 · · · θn because of the anticommutation relations. The even (odd) parts con-tain even (odd) products of Grassmann numbers and derivatives. It is easy to see thatS(3) = P(2). However, for larger n we get new super algebras. For instance, S(4)contains only representation of SU (4),

S(4) : 15 + 10 + (4 + 20) f ,

and the even part is not a Lie algebra although it contains the SU (4) adjoint.• W (n) and W (n). These contain elements of the form

W (n) : Pn∂i .

We can see that while W (2) = SU (2 | 1), we have

W (3) : 8 + 3 + (3 + 3 + 6) f ,

which contains the SU (3) adjoint and the complex 3 in its even part.• H(n). Its elements are of the form

H(n)n∑

i=1

∂iPn∂i .

For low values of n, these are trivial. However, we have the isomorphism H(4) =SSU (2 | 2). When n > 4, its even part contains the SO(n) generators togetherwith antisymmetric tensors with an even number of indices. Its odd elements are theantisymmetric tensors with an odd number of indices.

It is beyond the scope of this book to construct their representations.

Page 233: Pierre Ramond Group Theory a Physicists Survey 2010

11

The groups of the Standard Model

The symmetries of a physical system manifest themselves through its conserva-tion laws; they are encoded in the Hamiltonian, the generator of time transla-tions. In Quantum Mechanics, the hermitian operators which commute with theHamiltonian generate the symmetries of the system.

Similar considerations apply to local Quantum Field Theory. There, the mainingredient is the Dirac–Feynman path integral taken over the exponential of theaction functional. In local relativistic field theory, the action is the space-time inte-gral of the Lagrange density, itself a function of fields, local in space-time, whichrepresent the basic excitations of the system. In theories of fundamental interac-tions, they correspond to the elementary particles. Through E. Noether’s theorem,the conservation laws are encoded in the symmetries of the action (or Lagrangian,assuming proper boundary conditions at infinity).

Physicists have identified four different forces in Nature. The force of gravity, theelectromagnetic, weak, and strong forces. All four are described by actions whichdisplay stunningly similar mathematical structures, so much so that the weak andelectromagnetic forces have been experimentally shown to stem from the sametheory. Speculations of further syntheses abound, unifying all three forces exceptgravity into a Grand Unified Theory, or even including gravity in Superstring or MTheories!

In his remarkable 1939 James Scott lecture, Dirac speaks of the mathematicalquality of Nature and even advocates a principle of mathematical beauty in seekingphysical theories!

The so-called Standard Model of Elementary Particle Physics describes the fun-damental interactions of Nature as we know it. It is not a complete theory as it onlyaccounts for the gravitational force in the large, and fails in its quantum treatment.On the other hand, the other three forces are successfully described in the quantumrealm.

221

Page 234: Pierre Ramond Group Theory a Physicists Survey 2010

222 The groups of the Standard Model

The Standard Model symmetries can be split into two categories; the first typeof symmetry acts directly on space-time; it is encapsulated into a set of transforma-tions called the Poincaré group; the second type, which acts directly on the chargesof the elementary particles, is called gauge symmetries; they are connected with theubiquitous Lie algebras, SU (3), SU (2), and U (1) through the Yang–Mills con-struction. Electric charge conservation is traced to a local gauge symmetry whichacts on the fields with space-time dependent parameters. On the other hand, inthe absence of gravity, the space-time transformations are global, the same for allspace-time points. These invariances are still not sufficient to totally determine themultiplicity and interactions of the fields. There is a mysterious triplication of ele-mentary particles into three families which awaits explanation, pointing perhaps toan hitherto unknown family symmetry.

For reasons that lie beyond the scope of this book, it turns out not to be possibleto unify space-time and gauge symmetries under the aegis of one mathemati-cal group, limiting their union to a direct product structure: the Standard ModelLagrangian is a polynomial of local fields, invariant under the direct product of thePoincaré group with the three gauge symmetries. Group theory comes in throughthe construction of invariants under all these symmetries.

11.1 Space-time symmetries

We first discuss symmetries which act on space and time. H. Lorentz had foundMaxwell’s equations to be invariant under six transformations, today called theLorentz group, the non-compact group SO(3, 1). Einstein boldly posited that theinvariances of Maxwell’s equations applied to Mechanics as well, thereby replacingNewton’s laws with his own theory of Special Relativity. Theories which satisfyspecial relativity have a Lagrangian that is invariant under space-time rotations;invariance under space rotations translates into angular momentum conservation.

Noether’s theorem associates time translation invariance with energy conser-vation, and invariance under space translations with momentum conservation.Poincaré obtained the ten-parameter Poincaré group by adding constant space andtime translations, thus incorporating, à la Noether, conservation of momentum andenergy.

Five years later, Bateman found the invariance group of the free Maxwell equa-tions to be the larger fifteen-parameter conformal group, SO(4, 2), which containsthe Poincaré group as a subgroup. The conformal group is not a symmetry ofNature as we perceive it, since it contains no dimensionful parameters, such asthe masses of elementary particles, the quark confinement scale, Newton’s con-stant (Planck mass), to name a few. If it were to apply at all to the real world, itwould be in the particular limit where all parameters with the dimension of mass

Page 235: Pierre Ramond Group Theory a Physicists Survey 2010

11.1 Space-time symmetries 223

are set to zero. Since much physics is to be gained from the study of Maxwell’sequations, we discuss both the Poincaré and conformal groups.

11.1.1 The Lorentz and Poincaré groups

The Lorentz group SO(3, 1) contains six antisymmetric generators Mμν whichsatisfy the commutation relations

[ Mμν, Mαβ ] = i(ημα Mνβ − ηνα Mμβ + ηνβ Mμα − ημβ Mνα), (11.1)

where the Greek indices run over 0, 1, 2, 3, and with ημν = ημν = diag(−1, 1, 1, 1). They act on the space-time coordinates xμ as

xμ → aμνxν, (11.2)

where aμν = −aνμ are six infinitesimal constant rotation parameters which leavethe quadratic form

ημνxμxν = −(x0)2 + (x1)2 + (x2)2 + (x3)2,

invariant. We also have

[ Mμν, xσ ] = i(ημσ xν − ηνσ xμ). (11.3)

Since the quadratic invariant is hyperbolic, these transformations cannot be repre-sented with hermitian generators acting in a finite-dimensional space; all unitaryrepresentations of non-compact groups are infinite-dimensional.

In our study of compact Lie algebras, we have noted the isomorphism SO(4) ∼SO(3) × SO(3). It translates to the non-compact case by defining

Mi j = εi jk J k; M0i = K i , (11.4)

where i, j, k run over the three space indices. Then we see that

[ J i , J j ] = iεi jk J k, [ J i , K j ] = iεi jk K k, (11.5)

while

[ K i , K j ] = −iεi jk J k . (11.6)

The all-important minus sign shows that SO(3, 1) splits into two sets of commut-ing generators,

1

2(J i + i K i ),

1

2(J i − i K i ), (11.7)

each of which satisfies the SO(3) commutation relations. Neither is hermitian,and that is the difference with the compact case. Under the parity operation

Page 236: Pierre Ramond Group Theory a Physicists Survey 2010

224 The groups of the Standard Model

P : xi → −xi , we see that J i → −J i , and K i → −K i , so that the two SU (2) gointo one another; the same is true under complex conjugation C : i → −i .

We can use this decomposition to obtain the representations of the Lorentzgroup. First represent the Lorentz generators in terms of (2 × 2) Pauli matrices as

J i = 1

2σ i , K i = i

2σ i , (11.8)

which act on a two-dimensional complex spinor field ψL . The simplest representa-tion is then

ψL → ψ ′L = e

i2 σ

i (ωi +iνi ) ψL , (11.9)

where ωi (νi ) are the rotation (boost) parameters. It is called a (left-handed)Weyl spinor, and we describe it as (2, 1). The right-handed Weyl spinor (1, 2),transforms as

ψR → ψ ′R = e

i2 σ

i (ωi −iνi ) ψR. (11.10)

For those who like indices, spinors which transform as a left-handed Weyl spinorare denoted with a subscript α, while those transforming as right-handed spinorsare given dotted indices, α

ψα ∼ (2, 1); ψα ∼ (1, 2).

Clearly, while the space rotations are unitarily represented, the boosts are not.Left- and right-handed spinors can be related since the reader will verify thatσ 2 ψ∗

R transforms as a left-handed spinor: all spinor representations can be writ-ten in terms of left-handed Weyl spinors. In this notation, real representations areof the form (n, n): four-vectors are (2, 2), corresponding to the (2 × 2) matrixrepresentation

xμ → X =(

x0 + x3 x1 − i x2

x1 + i x2 x0 − x3

), (11.11)

(with matrix elements written as Xαα) whose determinant is the invariant quadraticform. The electric and magnetic fields assemble as �E + i �B ∼ (3, 1), while �E −i �B ∼ (1, 3). Since we are interested in the action of these generators on localfields, we introduce the operators conjugate to xμ

[ xμ, pν ] = iημν, (11.12)

and separate the Lorentz generators into orbital and spin parts

Mμν = xμ pν − xν pμ + Sμν. (11.13)

The spin parts Sμν do not act on the space-time coordinates, but satisfy the samecommutation relations as Mμν . For left-handed Weyl spinors, we would have Si j =εi jkσ k/2, S0i = iεi jkσ k/2.

Page 237: Pierre Ramond Group Theory a Physicists Survey 2010

11.1 Space-time symmetries 225

The Poincaré group adds to the Lorentz group Pμ, the space-time translationgenerators (momenta) of

xμ → xμ + aμ, (11.14)

where aμ is a constant. They transform as a four-vector,

[ Mμν, Pσ ] = i(ημσ Pν − ηνσ Pμ), (11.15)

and commute among themselves,

[ Pμ, Pν ] = 0. (11.16)

We can always take

Pμ = −i∂

∂xμ

≡ pμ, (11.17)

where xμ = ημνxν . We see that the Poincaré group is the semi-direct product of theLorentz group with the Abelian translations. Its representations are characterizedby the values of its two Casimir operators, PμPμ, and WμWμ, where Wμ is thePauli–Lubanski vector

Wμ = 1

2εμνρσ Pν Mρσ , (11.18)

and where εμνρσ is the totally antisymmetric Levi–Civita symbol (ε0123 = 1). Theysatisfy the commutation relations

[ Wμ, Wν ] = −iεμνρσ PρW σ , (11.19)

and are orthogonal to the momenta

PμWμ = 0.

The representations of the Poincaré group fall into three distinct types, sincePμPμ can be negative, zero, or positive.

• PμPμ = −m2

With m real, the four-momentum is a time-like vector. Accordingly, we can byLorentz transformations go to the rest frame, where

P0 = m, Pi = 0.

Evaluation of the Pauli–Lubanski vector in the rest frame yields

W 0 = 0, W i = −mSi , Si = 1

2εi jk S jk, i, j, k = 1, 2, 3.

Page 238: Pierre Ramond Group Theory a Physicists Survey 2010

226 The groups of the Standard Model

Si are the three generators of SO(3),

[ Si , S j ] = iεi jk Sk,

whose quadratic Casimir operator is s(s + 1), with s = 0, 1/2, 1, 3/2, . . .. Theserepresentations with

PμPμ = −m2, Wμ Wμ = m2 s(s + 1), (11.20)

describe particles of mass m and spin s. We can further express the generators bysingling out different ways in which the physical system evolves.

The most familiar one is the Newton–Wigner representation, to which we nowturn our attention.

Newton–Wigner representation

The idea is to represent the generators on an initial surface in space-time. We iden-tify x0 with physical time, so that the initial surface is the slice of space-time x0 =constant. From the general representation eq. (11.13), the Pauli–Lubanski vector isgiven by

W0 = 1

2εi jk pi S jk, Wi = −εi jk

(1

2p0S jk − p j S0k

). (11.21)

The Lorentz generators split into space rotations Mi j which do not affect the ini-tial surface, and the boosts M0i which do; they will have a more complicatedexpression on the initial surface. To find it, we set

S0i = A εi jk p j Sk, (11.22)

and obtain, after a little bit of algebra ,

WμWμ = [(p0)2 + (2Ap0 + A2 pi pi )p j p j ]Sk Sk + (pi Si )2(2Ap0 + A2 p j p j +1).

(11.23)

Comparison with eq. (11.20) requires

2Ap0 + A2 pi pi + 1 = 0,

that is,

A = 1

p0 ± m, (11.24)

which yields S0i in terms of the other generators.In these expressions, x0 is to be identified with physical time, and the generators

are expressed at an initial fixed value of x0. This means that its conjugate variablep0 = −i∂0 is no longer well defined: it must be expressed in terms of the remaining

Page 239: Pierre Ramond Group Theory a Physicists Survey 2010

11.1 Space-time symmetries 227

variables. This is achieved through the method of Lagrange multipliers. FollowingP. A. M. Dirac (Rev. Mod. Phys. 29, 312 (1949)), we set

P0 = p0 + λ(pμ pμ + m2), (11.25)

and choose λ so as to eliminate p0. The result is

P0 = ±√

pi pi + m2, (11.26)

corresponding to positive and negative energy solutions. At x0 = 0, we obtain theNewton–Wigner representation of the Poincaré group generators

P0 = ±√

m2 − ∂ i∂ i , pi = −i∂ i , (11.27)

Mi j = i(x j∂ i − xi∂ j ) + εi jk Sk, M0i = −1

2{xi , P0} − i

εi jk∂ j Sk

P0 ± m, (11.28)

accounting for the fact that xi and P0 no longer commute. The generators on theinitial surface x0 = 0, split into kinematical generators, pi and Mi j , and dynamicalgenerators, the Hamiltonian P0 and the boosts M0i , which move the system awayfrom the initial surface. The Poincaré group therefore acts on Hilbert states labeledby kets of the form

| pi ; m, s, s3 >,

which are interpreted as particles of mass m, spin s, with three-momenta pi andmagnetic quantum number s3. Their (positive or negative) energy is given by themass-shell condition (11.26) in terms of their mass and three-momenta.

Light-cone representation

Dirac introduced the front form, a different way of expressing dynamical evolution,known today as the light-cone. The initial surface is light-like, xμxμ = 0, labeledby constant values of

x+ ≡ 1√2(x0 + x3). (11.29)

For fixed x+, its conjugate variable

p− = 1√2(p0 − p3), [ x+, p− ] = −i, (11.30)

is no longer well-defined, and must be expressed in terms of the remainingvariables, that is x−, p+,

x− = 1√2(x0 − x3), p+ = 1√

2(p0 + p3), [ x−, p+ ] = −i, (11.31)

Page 240: Pierre Ramond Group Theory a Physicists Survey 2010

228 The groups of the Standard Model

and the transverse variables xa, pb, which satisfy

[ xa , pb ] = i δab, a, b = 1, 2. (11.32)

There are seven kinematical Poincaré generators, p+, pa , Mab, M+a , and M+−.The latter is kinematical only at x+ = 0, since it contains x+ p−.

There are only three dynamical Poincaré generators, the light-cone Hamilto-nian P− and the two boosts M−a . Using Dirac’s method of Lagrange multipliers,we find

P− = pa pa + m2

2p+ . (11.33)

The remaining three translation generators P+ and Pi are kinematical. We expectthe kinematical generators to retain a simple form, for example

Mab = xa pb − xb pa + Sab (11.34)

generates the transverse plane rotation. Finding the expression for S+a , S+−, andS−a take a bit more doing.

Let us write the momentum vector in matrix form

P =(

p0 + p3 p1 − i p2

p1 + i p2 p0 − p3

), (11.35)

with det P = m2. In the rest frame of the particle this matrix reduces to m timesthe unit matrix. It is easy to see that

P = U(

m 00 m

)U†, (11.36)

where

U =⎛⎜⎝

√√2p+m 0

(p1+i p2)√m

√2p+

√m√2p+

⎞⎟⎠ (11.37)

is the matrix that transforms the rest frame into the light-cone frame. We can applyit to the Pauli–Lubanski vector as well,

W =(

W 0 + W 3 W 1 − iW 2

W 1 + iW 2 W 0 − W 3

)= −m U

(S12 S23 − i S31

S23 + i S31 −S12

)U†,

(11.38)using its form in the rest frame. In light-cone coordinates, the components of thePauli–Lubanski vector are

W + = −p+S12 + εab pa S+b, W − = p−S12 − εab pa S−b, (11.39)

Page 241: Pierre Ramond Group Theory a Physicists Survey 2010

11.1 Space-time symmetries 229

and

W a = εab(p−S+b + pb S−+ − p+S−b), (11.40)

where ε12 = −ε21 = 1. By comparing (11.38) with (11.39) for W +, we findS+a = 0, so that

M+a = x+ pa − xa p+. (11.41)

A similar procedure for the other components requires S+− = 0, and thus

M+− = −x− p+ + x+ p− = −x− p+, (11.42)

at x+ = 0. We can evaluate S−a from comparing the entries in W1 + iW2 on bothsides as well as W −, yielding

S−a = i

p+ (mSa + i pb Sab). (11.43)

We set T a = mSa , keeping the commutation relation,

[ T 1, T 2 ] = im2 S12. (11.44)

This yields the light-cone representation of the Poincaré generators, first derivedby Bacry and Chang (Ann. Phys. (N.Y.) 47, 407 (1968)),

P− = pa pa + m2

2p+ , P+ = p+, Pa = pa, (11.45)

M+a = −xa p+, M+− = −x− p+, Mab = xa pb − xb pa + Sab,

M−a = x− pa − 1

2{xa, P−} + i

p+ (T a + i pb Sab). (11.46)

In these coordinates, the Poincaré group acts Hilbert states labeled by the kets

| p+, pa; m, s, s3 >,

representing massive states with light-cone momenta p+, pa , and spin s withp− obeying the mass-shell condition (11.33).

This light-cone representation is very useful in other contexts; it readily gen-eralizes to any number of transverse dimensions, for which we have the samerepresentation, except that the indices a, b run over an arbitrary number of values,and thus

[ T a, T b ] = im2 Sab. (11.47)

It arises in the quantization of the Nambu–Goto string by J. Goldstone, P. Goddard,C. Rebbi and C. B. Thorn (Nucl. Phys. B56, 109 (1973)). For the string, the oper-ators m2 and Sab are quadratic combinations of an infinite number of harmonic

Page 242: Pierre Ramond Group Theory a Physicists Survey 2010

230 The groups of the Standard Model

oscillators operators, while T a are cubic polynomials in the same harmonic oscil-lators. Owing to normal ordering, this commutation relation is satisfied only whena, b range over twenty-four values: relativistic strings live only in a twenty-sixdimensional space-time.

• PμPμ = 0.

This case can be viewed in terms of the previous representation in the limit m → 0.The Pauli–Lubanski vector is now light-like. Since both Pμ and Wμ are light-likeand orthogonal to one another, they must be proportional. In the light frame where

P− = P1 = P2 = 0, W − = W 1 = W 2 = 0, W + = P+ S12, (11.48)

so that

W +

P+ = S12 ≡ λ, (11.49)

where the value of the SO(2) generator is the helicity. We can use the same rep-resentation for the operators, the difference being that we can no longer divideby m.

In the light-cone representation, the T a operators transform as a transversevector,

[ S12, T 1 ] = T 2, [ S12, T 2 ] = −T 1, (11.50)

and commute with one another

[ T 1, T 2 ] = 0. (11.51)

This leads to two possibilities.

– T a = 0. The ket labels are now just the momenta and the helicity λ, describing masslessparticles

| p+, pa; m = 0, λ >, (11.52)

of fixed helicity. The values of the helicity are restricted, not by the algebra, but by thegroup. Exponentiation of the transverse rotations must be single-valued, which requiresthe helicity to be ± half-odd integers. This case represents a massless particle with asingle value of helicity, λ = 0,±1/2,±1, . . ..

From the point of view of group theory, massive representations can be thought of as asum of massless representations; for instance the massive spin one vector representationconsists of three states with magnetic component −1, 0, 1, each of which occurs as amassless representation.

Page 243: Pierre Ramond Group Theory a Physicists Survey 2010

11.1 Space-time symmetries 231

– T a �= 0. In this case, T a are simply the c-number components of a transverse vector.The states on which the Poincaré algebra is realized must now be labeled by the maximalnumber of commuting operators, that is the transverse vector ξa

T a | p+, pa; ξa >= ξa | p+, pa; ξa > . (11.53)

The representations are labeled by the length of the transverse vector ξa , since the Pauli–Lubanski vector is now space-like with

WμWμ = ξaξa . (11.54)

Helicity is no longer a good quantum number as the light-cone boosts which containT i change the helicity by one unit. Hence the representation is made up of an infinitenumber of helicities, spaced by one unit. We have therefore two distinct representa-tions, depending on whether the helicities are integer or half-odd integer. They are called“continuous spin representations” by Wigner in his original work. These have no simpleinterpretations in terms of particles, since they describe a massless object with an infinitenumber of integer-spaced helicities.

Finally, let us mention that the Poincaré group can be extended to a super algebra, theSuperPoincaré or Supersymmetry algebra. Yu. A. Golfand and E. P. Likthman (JETPLett. 13, 323 (1971)) included new generators that transform as spinors such that theiranticommutator yields the translation generators

{ Qα, Qα } = (σμ)αα Pμ, (11.55)

and commute with the translations

[ Qα, Pμ ] = [ Qα, Pμ ] = 0. (11.56)

It follows that PμPμ is still an invariant and all of its representations can be discussedin terms of its value.• Its m = 0 representation simply consists of two massless Poincaré algebra represen-

tations separated by half a unit of helicity, denoted as (λ, λ + 12 ).

• The m �= 0 representations are assembled out of massless representations, as in thePoincaré case. For instance a massive spin one-half particle is made up as ( 1

2 , 0), and(0,− 1

2 ), so that it is accompanied by a complex scalar with the same mass. A massivespin one particle with three degrees of freedom belongs to (1, 1

2 )+( 12 ,−1)+( 1

2 , 0)+(0,− 1

2 ), and its supermultiplet contains two spinors and one complex scalar.• The two continuous spin representations, one with integer helicity, the other with

half-odd integer, neatly assemble into one massless supermultiplet.

11.1.2 The conformal group

The conformal group is the invariance group of the free Maxwell’s equations. Itis generated by a simple algebra, and is the simplest extension of the Poincarégroup. While this group is of great theoretical interest, it is not a symmetry group

Page 244: Pierre Ramond Group Theory a Physicists Survey 2010

232 The groups of the Standard Model

of Nature as we know it. There are many ways to characterize it; for our purposes,we define it as the group of transformations which leaves the equation of a lightray invariant,

xμ → xμ′; ημνxμxν = ημνxμ′xν ′, (11.57)

It means that the light-cone analysis of the Poincaré algebra can be easilyimplemented.

The conformal algebra contains, in addition to the ten Poincaré generators, fivenew transformations, the dilatation

xμ → s xμ, (11.58)

and four conformal transformations

xμ → xμ − cμ (xνxν)

1 − 2cνxν + (cνcν)(xρxρ)

, (11.59)

with constant s and cμ. The conformal transformations can be thought of as aninversion

xμ → xμ

(xνxν), (11.60)

sandwiched between two translations. In space-time, the new generators are thedilatation

D = 1

2(xμ pμ + pμxμ) + d, (11.61)

and the conformal transformations

K μ = −2xν Mνμ + (xνxν)pμ + 2(d − 2i)xμ + κμ. (11.62)

They do not act linearly on space-time. In the above, d and κμ are the non-orbitalparts. Assuming that they satisfy the same commutators as the orbital parts, we find

[ D, Pμ ] = i Pμ, [ D, K μ ] = −i K μ, (11.63)

and

[ Pμ, K ν ] = 2i Mμν − 2i ημν D, [ K μ, K ν ] = 0. (11.64)

These commutation relation can be rewritten in a pleasing way by identifying

D = M56, K μ = (M5μ − M6μ), Pμ = (M5μ + M6μ), (11.65)

which are seen to satisfy the SO(4, 2) Lie algebra,

[ M AB, MC D ] = i(ηAC M B D − ηBC M AD + ηB D M AC − ηAD M BC), (11.66)

with A, B, · · · = 0, 1, 2, 3, 5, 6, and ηAB = ηAB is a diagonal matrix with entries(−1, 1, 1, 1, 1,−1).

Page 245: Pierre Ramond Group Theory a Physicists Survey 2010

11.1 Space-time symmetries 233

The conformal group generators leave the quadratic form ηAB ζ A ζ B invariant,where ζ A are the six homogeneous coordinates on which they act linearly. It alsohas a cubic and a quartic Casimir operator, one more than the Poincaré group. Onthe light-cone

ηAB ζ A ζ B = 0, (11.67)

the homogeneous coordinates are related to the space-time variables by stereo-graphic projections

xμ =√

2

(ζ 5 + ζ 6)ζμ, ημν xμ xν = 2

(ζ 6 − ζ 5)

(ζ 6 + ζ 5). (11.68)

We note that PμPμ is no longer invariant; its eigenvalues are either zero orcontinuous. When it is zero, the particle interpretation of the Poincaré group isretained.

The m = 0 representations of the conformal group can be constructed usingDirac’s front form since x2 = 0 is conformally invariant, but there are other waysto do it: we note the especially elegant representation “witten” (Commun. Math.Phys. 252, 189 (2004)) in terms of Penrose’s twistors, the classical (boson-like)coordinates which transform as left- and/or right-handed Weyl spinors,

λα ∼ ( 2, 1 ), μα ∼ ( 1, 2 ). (11.69)

In this language, indices can be raised by means of the antisymmetric symbols εαβ

and εαβ ,

λα = εαβ λβ, μα = εαβ μβ . (11.70)

We also introduce the derivative operators which satisfy

∂λα

λβ = δ αβ ,

∂μα

μβ = δ α

β. (11.71)

The momenta are given in matrix form by

Pαα = iλα

∂μα. (11.72)

Its determinant clearly vanishes, and m = 0. The conformal generators are simply

Kαα = iμα

∂λα, (11.73)

with the dilatation in the form

D = i

2

(λα

∂λα

− μα

∂μα

). (11.74)

Page 246: Pierre Ramond Group Theory a Physicists Survey 2010

234 The groups of the Standard Model

The Lorentz generators are written in symmetric bispinor form as

Mαβ = i

2

(λα

∂λβ+ λβ

∂λα

), M

αβ= i

2

(μα

∂μβ+ μ

β

∂μα

). (11.75)

Thus while the conformal group acts non-linearly on space-time, its action ontwistor space is linear.

A world with only massless particles, while in obvious contradiction with facts,may well be an interesting theoretical abstraction, because in the Standard Modelmasses result from spontaneous breaking of symmetries. For instance, the continu-ous spin representations of the Poincaré group are absent in both Nature and in therepresentations of the conformal group.

We close this section by determining the non-relativistic limit of the Poincaréand conformal groups. In that limit, the speed of light c is taken to infinity. Itenters the generators through x0 = c t , and the non-relativistic limit is obtained bybuilding operators that survive it.

As c → ∞, both Mi j and pi retain their original form, while the Hamiltonianand boosts reduce to

c P0 → i∂

∂t,

1

cM0i → −i t

∂xi. (11.76)

Their action on xi (t) is then

xi (t) → xi (t) + ai j x j (t) + ai + vi t, (11.77)

where ai j , ai and vi are constant. It is the group that leaves the equation

d2xi (t)

d t2= 0, (11.78)

invariant, in accordance with Newton’s law for a free particle.A similar analysis of conformal transformation shows that the conformal

generators become

1

cK0 → i t2 ∂

∂t− 2i t x i ∂

∂xi,

1

c2Ki → −i t2 ∂

∂xi. (11.79)

The action of these generators on the trajectory xi (t) shows that it is quadratic intime. This means that conformal non-relativistic physics is invariant under

xi (t) → xi (t) + ai t2, (11.80)

ai is a constant acceleration. The equation that replaces Newton’s is of third order,

d3xi (t)

d t3= 0. (11.81)

Page 247: Pierre Ramond Group Theory a Physicists Survey 2010

11.2 Beyond space-time symmetries 235

In such a theory, one could not distinguish a particle at rest from one with a constantacceleration: protons uniformly accelerated around a ring would not radiate, muchto the delight of accelerator physicists, but alas it is not so: conformal theories donot reproduce Nature.

11.2 Beyond space-time symmetries

It is natural to seek a unified description of space-time and internal symmetries,such as SU (2)-isospin and flavor SU (3). We have already seen that by combiningisospin and spin, Wigner and Stückelberg constructed the supermultiplet theorywhich could be used to catalog particles of different spin. With the advent of theflavor SU (3) classification, F. Gürsey and L. Radicati (Phys. Rev. Letters 13, 173(1964)) and B. Sakita (Phys. Rev. 136, B1756 (1964)) independently proposed tounify the eight-fold way with spin. This is akin to finding a group which naturallycontains SU (3) and SU (2). The natural candidate is SU (6) with embedding,

SU (6) ⊃ SU (3) × SU (2), 6 = ( 3, 2). (11.82)

Hence the 35-dimensional adjoint deduced from 6 × 6, yields

35 = (8, 1 ) + ( 1, 3 ) + ( 8, 3 ). (11.83)

For physical identification, we are guided by the Fermi–Yang model, and recog-nize the pseudoscalar octet, and nine spin one vector mesons, which decomposeunder SU (3) as an octet and a singlet. There are indeed two isoscalar vectormesons, the ω and φ particles, while the isotriplet vector is identified with theρ-mesons, and the isodoublets correspond to the “K -star” vector resonances, anda linear combination of the ω and φ vector mesons in the octet, leading to thefollowing.

�ρ−

�ρ+

�(ω, φ)

�ρ0

�K ∗+

�K ∗0

�K ∗0

�K ∗−

Page 248: Pierre Ramond Group Theory a Physicists Survey 2010

236 The groups of the Standard Model

So far so good. When this classification is generalized to the baryons, the spin one-half octet, and the spin three-half decuplet combine into one SU (6) representation,the 56 with decomposition

56 = ( 8, 2 ) + ( 10, 4 ), (11.84)

where we have labeled the spin representations by their dimensions. This unifiesthe baryon resonances in the octet and decuplet into one representation. The baryonrepresentation is generated by the symmetric product of three fundamental sextets

56 = (6 × 6 × 6)sym . (11.85)

Symmetries can be used to relate the magnetic moments of the different particlesin the baryon octet. Classically, the magnetic moment of a charge distribution isrelated to the electromagnetic current �J by the formula

�μ = 1

2c�r × �J . (11.86)

In the quantum realm, this identification still applies, albeit altered by the spin ofthe particles. Hence the symmetry properties of the electromagnetic current implyphysical consequences for the magnetic moments. Recall that the electric charge isgiven by the Gell-Mann–Nishijima formula (6.75)

Q = I3 + Y

2,

where I3 is the third component of isospin and Y is the isospin singlet in SU (3).This implies that the electromagnetic current is the sum of two parts; one that isan isospin singlet, and one that transforms as an isotriplet. The same applies to themagnetic moment. For SU (2), this means, simply based on the symmetries, that(Wigner–Eckhardt theorem)

μ = 〈 I, I3 | (a + b I3) | I, I3 〉,where the hat denotes the operator, and a, b are undetermined constants. Appliedto the isotriplet of �+, �0 and �−, this formula implies

μ�± = a ± b

2, μ

�0 = a,

which yields one relation among physical quantities.

μ�+ + μ�− = 2μ�0 . (11.87)

To derive similar formulas for flavor SU (3), we need to determine the trans-formation properties of the electric current. To that effect, we note that theelectric charge commutes with an SU (2) subgroup of SU (3). Dubbed U -spin by

Page 249: Pierre Ramond Group Theory a Physicists Survey 2010

11.2 Beyond space-time symmetries 237

S. Meshkov, C. A. Levinson, and H. J. Lipkin (Phys. Rev. Letters 10, 361 (1963)),its multiplets can be read off from the oblique lines in the octet diagram, whichcontains two U -spin doublets

(�−, �− ), ( p, �+ ),

and a U -spin triplet and singlet

( n, T ≡ 1

2(√

3�0 − �0), �0 ), S ≡ 1

2(�0 + √

3�0).

We already learned how to compute the matrix elements of tensor operators in theoctet representation. In particular the part of the Hamiltonian which breaks SU (3)is along the X3

3 direction, and was given by eq. (6.88)

< X33 >octet= a + bY + c

(1

4Y 2 − I (I + 1)

).

This formula can be taken over, by simply changing from hypercharge to chargeand I -spin to U -spin, with

< X11 >octet= a + bQ + c

(1

4Q2 − U (U + 1)

).

The electromagnetic current is the traceless part since the electric charge is con-tained within SU (3). The reader might verify that on the octet, we are left with

< Jelec >octet= β Q + γ

(1 + 1

4Q2 − U (U + 1)

), (11.88)

written in terms of only two undetermined constants. Many relations among thebaryon’s magnetic moments follow from

μ∗ = 〈 ∗ | b Q + c

(1 + 1

4Q2 − U (U + 1)

)| ∗ 〉. (11.89)

As an example, we form

μ�0 = c 〈�0 | 1 − U (U + 1) |�0 〉, (11.90)

and by expressing the physical particles in terms of the U -triplet and U -singlet,

|�0 〉 = 1

2(√

3| S 〉 − | T 〉), |�0 〉 = 1

2(| S 〉 + √

3| T 〉),we find

μ�0 = c

2. (11.91)

Page 250: Pierre Ramond Group Theory a Physicists Survey 2010

238 The groups of the Standard Model

A similar reasoning leads to

μ�0 = −c

2, μ

�0→�0 = −c√

3

8,

where the latter is the transition magnetic moment. The eight plus one (transition)magnetic moments of the octet baryons are determined in terms of the proton andthe neutron magnetic moments, for instance

μ�− = −μp − μn.

With SU (6), these relations can be improved even further. The magnetic momentcoupling of a fermion is given by

�† �σ � · �B, (11.92)

where �B is the magnetic field, and � is the Dirac field representing the fermion (itsspinor indices are not shown).

In SU (6), the baryons are represented by a Dirac field which is a symmetricthird-rank tensor 56, �(abc), where a, b, c = 1, 2 . . . , 6. Each SU (6) index is reallya pair of SU (3) indices A, and spin indices α: a = (A, α), b = (B, β), etc., whereA, B, . . . = 1, 2, 3, and α, β, . . . = 1, 2. In index language, the decomposition(11.84) reads

�(abc) = �((A,α)(B,β)(C,γ )) (11.93)

= ψαβγ TABC +∑sym

(εαβψγ εAB D T D

C

). (11.94)

The first term contains ψαβγ the spin three-half Dirac field, and TABC a symmetrictensor that transforms as the 10 of SU (3). The second term describes the baryonoctet, with the spin one-half Dirac field ψα, and a tensor octet T D

C . Its form isdetermined by the only invariant tensors in spin space εαβ = −εβα, and the Levi–Civita tensor in triplet space. Finally, the sum is required to symmetrize over thethree pairs of indices.

The magnetic moments of the baryons in the octet can thus be determined interms of one of them. Indeed, after substituting this expression for the baryon octetin (11.92), M. A. B. Bég, B. W. Lee, and A. Pais (Phys. Rev. Letters 13, 514 (1964))find

2μp + 3μn = 0, (11.95)

corresponding to b = c. This relation is in spectacular agreement with experiment!There are many other applications of SU (6); to name a few: improving the SU (3)mass relations by relating pseudoscalar to vector octet masses, in particular

m2K − m2

π = m2K ∗ − m2

ρ,

Page 251: Pierre Ramond Group Theory a Physicists Survey 2010

11.2 Beyond space-time symmetries 239

as well as finding rules for the baryon decuplet masses, and also vector meson ω−φ

mixing. We leave to the reader the joy of rediscovering these remarkable relations.

11.2.1 Color and the quark model

Identifying the baryons with the totally symmetric 56 is at odds with the quarkmodel. In the eight-fold way, the “elementary” particles belong to the SU (3)octet and decuplet. Group-theoretically, these representations can be constructedas products of the fundamental representations. The octet appears in both

8 ∈ 3 × 3, 8 ∈ (3 × 3 × 3)mixed, (11.96)

while the decuplet is the three-fold symmetric product

10 = (3 × 3 × 3)sym . (11.97)

This led M. Gell-Mann and G. Zweig to speculate independently that the eight-fold way particles behave as if they are made up of some more elementary entitiestransforming as triplets of flavor SU (3); they are the quarks of Gell-Mann (Zweig’saces), the nucleons as three-quark composites and the mesons as quark–antiquarkcomposites. Scattering experiments later confirmed the quark hypothesis.

In the light of SU (6), the quark model is problematical: quarks are fermions,since they belong to the 6: flavor SU (3) triplets with spin one-half. Accordingto the Pauli exclusion principle (spin-statistics theorem in local quantum field the-ory), states containing several identical fermions must be antisymmetric under theirinterchange, and the three-quark combination should appear in the antisymmetricproduct of three SU (6) sextets; yet it lies in the symmetric product.

The road to a solution for this mystery was found when W. Greenberg (Phys.Rev. Letters 13, 598 (1964)) proposed a way to retain the spin-statistics connection:quarks have a hitherto unknown quantum number, which he couched in the lan-guage of parastatistics of order three, where the quark creation operator is the sumof three different creation operators, and the antisymmetry occurs over the three“colors,” as they were later called. The rest of the triple product is symmetric,hence the nucleons and their resonances are in the 56.

The next step was taken by M. Y. Han and Y. Nambu (Phys. Rev. 139, B1006(1965)), who boldly assumed that this new quantum number indicates that thestrong force can be understood as a gauge theory based on color SU (3), and thatthe color force was somehow caused by its eight vector particles. The observedparticles are color singlets under this new SU (3), but their hypothesis was shorton dynamical details. This new theory is quantum chromodynamics (QCD). ColorSU (3) should not be confused with flavor SU (3) of the eight-fold way.

Page 252: Pierre Ramond Group Theory a Physicists Survey 2010

240 The groups of the Standard Model

11.3 Invariant Lagrangians

In field theory, Lagrange densities L are real polynomials built out of local fields.In fundamental theories of Nature, these polynomials are invariant under somesymmetries, such as Poincaré and gauge symmetries. We can apply group theoryto their construction.

If we use fields local in space-time, translation invariance is assured as long asthe coordinates xμ do not appear, except as arguments of local fields, but translationinvariance does allow the derivative operator ∂μ ∼ (2, 2) acting on the fields.

• Lagrangians constructed out of scalar fields ϕ(x) are easiest to manufacture.Any function V (ϕ(x)) is Poincaré invariant. Use of an even number of derivativeoperators leads to an infinity of invariants, the simplest of which ∂μϕ(x) ∂μϕ(x),has two derivatives; its variation yields a second-order differential equation, themassless Klein–Gordon equation. Invariants with more derivatives lead to higher-order equations of motions. A spinless particle is represented by a scalar field withLagrangian

L = 1

2∂μϕ(x) ∂μϕ(x) − V (ϕ(x)). (11.98)

Since the action is dimensionless (in units of h), ϕ has mass (length) dimension1(−1). For a free particle of mass m, V = m2ϕ2/2.

The kinetic invariant of N scalar fields ϕa , a = 1, 2 . . . N ,

1

2

N∑a=1

∂μϕa(x) ∂μϕa(x), (11.99)

is a quadratic form, manifestly invariant under a global SO(N ) transformation

δ ϕa(x) = ωabϕb(x), ωab = −ωba, (11.100)

where ωab do not depend on x .

• Next we consider invariants built out of one Weyl spinor, represented by acomplex two-component local Grassmann field ψα(xμ) ∼ (2, 1); its conjugateis ψα ∼ (1, 2). Group theory points to an immediate quadratic invariant in theantisymmetric product of the Weyl spinor with itself

(2, 1) × (2, 1) = (1, 1)a + (3, 1)s . (11.101)

In terms of the fields, εαβψα(x)ψβ(x), and its conjugate εαβ ψα(x)ψβ(x). In fieldtheory, the spin-statistics connection is assured by taking fermion fields to beGrassmann, so this invariant does not vanish; when added to its conjugate to makeit real, it is interpreted as the Majorana mass of the Weyl spinor.

Page 253: Pierre Ramond Group Theory a Physicists Survey 2010

11.3 Invariant Lagrangians 241

The product of the Weyl fermion with its conjugate

(2, 1) × (1, 2) = (2, 2), (11.102)

transforms as a Lorentz vector, but we can make it Lorentz-invariant by using thederivative operator, obtaining

ψ†α (x) (σ

μ)αα ∂μ ψα(x), (11.103)

with the matrix σμ = (1, σ i ) written in terms of Pauli matrices. There are two otherinvariants of this kind; one, with the derivative operator acting on both spinors, isa surface term which does not affect the equations of motion, the other with thederivative operator acting on the conjugate spinor is needed for a real invariant.The equation derived from this invariant is the massless Dirac equation for a left-handed Weyl spinor. From the kinetic term, we note that the mass dimension of ψis 3/2; thus the Majorana term must be multiplied with a parameter with dimensionof mass; hence its name.

We note that since ψ is a complex field, the kinetic term is invariant under theglobal U (1) phase transformation

δ ψ(x) = iω ψ(x), δ ψ(x) = −iω ψ(x), (11.104)

where ω is independent of x .The Dirac kinetic term of N Weyl fermions ψa(x) is the quadratic sum

N∑a=1

ψ†a α(x) (σ

μ)αα ∂μ ψa α(x), (11.105)

manifestly invariant under the same global U (1) phase transformation on all thefields, and global SU (N ) transformations

δ ψa(x) = i

2ωA (λA)b

a ψb(x), (11.106)

where λA are the Gell-Mann matrices, and the (N 2 − 1) parameters ωA do notdepend on x . The Majorana mass terms assemble into

mab εαβ ψa α(x)ψb β(x), (11.107)

where mab is a symmetric mass matrix. This term is no longer invariant underSU (N ) × U (1).

If the theory contains scalar fields, ϕn(x), the Majorana invariant can be turnedinto interaction terms, called Yukawa couplings,

Yn ab ϕn(x) ψa α(x) εαβ ψb β(x), (11.108)

where Yn ab are dimensionless coupling parameters.

Page 254: Pierre Ramond Group Theory a Physicists Survey 2010

242 The groups of the Standard Model

• The treatment of theories with local four-vector fields Aμ(x) which transform as(2, 2) is more complicated. Naively, such a field contains four degrees of freedom,one for each component. Yet, the group theory of the Poincaré group indicatesthat a massless spin one particle contains one helicity state. However, this state iscomplex, and when we add its conjugate, we obtain two states of opposite helicity±1. This is consistent with two types of photons, corresponding to left- and right-circularly polarized states.

The description of a spin one massless particle in terms of a four-vector fieldmust account for this overcounting. The answer is to be found in Maxwell’s theoryof electromagnetism, where the overcounting riddle is solved by gauge invariance.Maxwell’s theory is a gauge theory associated with a U (1) phase transformation.

The vector field is introduced by generalizing the derivative to a covariantderivative, constructed as to ensure that the global symmetry of the kinetic termis upgraded to a local symmetry where the transformation parameters depend onthe space-time coordinates xμ. Taking as an example the kinetic term for a Weylspinor, we generalize it to

ψ†α (x) (σ

μ)αα Dμ ψα(x), (11.109)

where

Dμ = ∂μ + i Aμ(x), (11.110)

is the covariant derivative. In order to preserve the symmetry (11.104), it mustsatisfy under the gauge transformation

Dμ → eiω(x) Dμ e−iω(x), ψ(x) → eiω(x) ψ(x). (11.111)

For this to happen, the vector field must transform inhomogeneously as

Aμ(x) → eiω(x) Aμ(x) e−iω(x) − ieiω(x)(∂μ e−iω(x)

). (11.112)

The covariant derivative clearly induces a gauge coupling of the vector field withthe fermion bilinear current. From the covariant derivative, we construct the fieldstrength

Fμν = i [Dμ, Dν ], (11.113)

which is invariant under the gauge transformation. The Maxwell kinetic termensues

1

4g2Fμν Fμν, (11.114)

where g is the coupling strength. In Maxwell’s theory, the electron is described bytwo Weyl spinors e(x) and e(x), with equal gauge couplings to the vector field,resulting in the QED Lagrangian

Page 255: Pierre Ramond Group Theory a Physicists Survey 2010

11.4 Non-Abelian gauge theories 243

LQE D = 1

4g2Fμν Fμν +e†(x) σμ Dμ e(x)+ e†(x) σμ Dμ e(x)+me(x) e(x)+c.c.,

(11.115)suppressing all spinor indices. It is invariant under the local U (1), and we haveadded the Dirac mass term, accounting for the electron mass.

11.4 Non-Abelian gauge theories

C. N. Yang and R. L. Mills (Phys. Rev. 96, 191 (1954)) generalized the QEDconstruction to gauge transformation generated by arbitrary Lie algebras. Our dis-cussion is limited to the basic ideas, and the interested reader is urged to consultthe many excellent physics and mathematics treatises dedicated to this rich subject.

Start with the kinetic term for N Weyl fermions, assembled into one �. We haveseen that it is invariant under a global SU (N ) transformation

�(x) = U �(x), (11.116)

and we wish to extend the invariance when U = U(x) depends on x , andU(x)U†(x) = 1. Following the Abelian case, we construct a covariant derivativesuch that

�(x) → U(x)�(x), Dμ �(x) → U(x)Dμ �(x). (11.117)

This requires the covariant derivative to be a matrix with the transformation

Dμ → U(x)Dμ U†(x). (11.118)

We introduce (N 2 − 1) vector fields through(Dμ

)b

b= ∂μ δb

a + i

2AC(x) (λC)b

a. (11.119)

In terms of

Aμ(x) ≡ 1

2AC(x) (λC)b

a, (11.120)

the gauge transformation reads

Aμ(x) → U(x)Aμ(x)U†(x) − i U(x)(∂μ U†(x)

). (11.121)

The field strength

Fμν = −i [Dμ, Dν ], (11.122)

= ∂μ Aν − ∂ν Aμ + i[ Aν, Aν ] (11.123)

is a matrix that transforms covariantly

Fμν → U(x)Fμν U†(x). (11.124)

Page 256: Pierre Ramond Group Theory a Physicists Survey 2010

244 The groups of the Standard Model

The Yang–Mills Lagrangian for the vector fields is then

LY M = 1

4g2Tr

(Fμν Fμν

), (11.125)

tracing over the Lie algebra. Although our construction was for SU (N ) transfor-mations, it applies to any other Lie algebra. We can abstract from this constructionthat Yang–Mills theories contain the following.

• Self-interacting massless spin one gauge bosons, one for each parameter of the Liealgebra. Gauge bosons can also interact with matter, if present. All gauge interactionsdepends on one coupling “constant.” Matter consists of the following.

• Spin one-half particles, which transform as some representation of the Lie algebra, and• Spin zero particles, also transforming as some representation of the Lie algebra.

Before describing the Standard Model in group-theoretical terms, two remarks arein order. One is that the restriction of matter to spins no higher than one-half comesfrom the additional requirement of renormalizabilty, not from any group-theoreticalconsiderations.

Secondly, specifying a gauge theory and its matter content does not fully deter-mine the theory. Fermions and scalars can have interactions outside the realm ofgauge interactions: Yukawa couplings and self-couplings of the scalars. Renormal-izability of the theory requires the mass dimension of these interactions to be atmost four.

In fact, the Standard Model must have Yukawa couplings as necessary to accountfor the masses of the elementary particles, and weak decays, as well as scalar self-interactions to generate masses for some of its vector bosons.

11.5 The Standard Model

The Standard Model of particle physics of strong, weak and electromagnetic forcesis described by three gauge theories SU (3) × SU (2) × U (1), linked through theirmatter content.

Quantum Chromodynamics (QCD) is the color SU (3) gauge theory with eightmassless vector bosons called gluons. Color matter consists of six flavors of quarks,three with charge 2/3, the up quark u, the charmed quark c, and the top quark t ,together with three charge −1/3 quarks, the down quark d, the strange quark s,and the bottom quark b. The gluons coupling to the quarks conserves parity.

The weak and electromagnetic forces are described by two gauge theoriesassociated with SU (2) and U (1).

In our notation, all fermions are left-handed Weyl spinors under the Lorentzgroup. This means that it takes two Weyl spinors to define a quark or a charged

Page 257: Pierre Ramond Group Theory a Physicists Survey 2010

11.5 The Standard Model 245

lepton. For instance the left-handed up quark is described by u, which transforms asa color triplet 3c, while u, transforming as 3c of color, is the conjugate right-handedup quark.

It is convenient to label all fermions by their transformation properties underthe three gauge groups SU (2) × SU (3)c × U (1)Y with their hypercharge Y asa subscript, determined from the Gell-Mann–Nishijima relation for the electriccharge

Q = I3 + Y

2.

All spin one-half elementary particles split into two types: leptons, which arecolor singlets,

leptons:(νe

e

)∼ ( 2, 1c )−1; e ∼ ( 1, 1c )2,

and quarks, which have color,

quarks:(

ud

)∼ ( 2, 3c ) 1

3; u ∼ ( 1, 3c )− 4

3; d ∼ ( 1, 3c ) 1

3.

Stable matter is made up of these particles.This pattern is triplicated with two more exact copies, with the same gauge prop-

erties. There are two heavier charged leptons, the muon and tau leptons with theirassociated neutrinos,(

νμμ

),

(νττ

)∼ ( 2, 1c )−1; μ, τ ∼ ( 1, 1c )2,

as well as the heavier charm, strange, top and bottom quarks(cs

),

(tb

)∼ ( 2, 3c ) 1

3; c, t ∼ ( 1, 3c )− 4

3; s, b ∼ ( 1, 3c ) 1

3.

These elementary constituents have similar group properties, although with widelydifferent masses. The top quark weighs by itself almost as much as a hafnium atom!The fermions of the heavier families decay into the lightest one, and were presentin Boltzmann abundance in the early Universe. It is beyond the scope of this bookto go into any more details.

The Standard Model also contains a spinless weak doublet and color singlet ∼( 2, 1c )−1. Its self-interactions are engineered by hand to produce the spontaneousbreakdown of the SU (2)×U (1) gauge symmetry to Maxwell’s U (1). The SU (3)×U (1) gauge symmetries are unbroken, with nine massless gauge bosons, the photonand the eight gluons. The other gauge bosons are massive, to account for the finiterange of the weak interactions.

Page 258: Pierre Ramond Group Theory a Physicists Survey 2010

246 The groups of the Standard Model

In theories with spontaneous breaking, the Lagrangian is invariant under thegauge symmetries, but its lowest energy configuration is not. This is done by addingto the theory spinless fields; the lowest energy configuration of the theory is at theminimum of their potential. Careful crafting of this potential is necessary to ensurebreaking of the symmetries. The symmetry breaking generates a mass for the gaugebosons associated with the broken symmetry generators, the Higgs mechanism.

While fixing the potential in such a way appears to be an unsatisfactory adhoc feature of the Standard Model, it also generates the masses for the fermions.The Higgs mechanism leaves behind a massive scalar particle, the Higgs, whosecoupling to matter is proportional to the mass of the particles with which it cou-ples. The Higgs still awaits discovery, but since it cannot be arbitrarily heavy, itsexistence provides an experimental test of the validity of this idea.

11.6 Grand Unification

As its name indicates, the strong force between quarks is strong, so much sothat quarks are forever prisoners inside nucleons. No free quarks have ever beenobserved. However, the QCD coupling “constant” varies with the scale at which itis measured, in such a way that at short distances, the strong interaction betweenquarks becomes weaker and weaker, a phenomenon called asymptotic freedom.Thus, it could even be that, at some very short distance, quarks may not be toodissimilar from leptons. Jogesh C. Pati and Abdus Salam (Phys. Rev. D8, 1240(1973)) were the first to suggest that quarks and leptons could be unified: quarkcolor stems from a broken SU (4), whose quartet representation contains one quarkand one lepton, through the embedding

SU (4) ⊃ SU (3)c × U (1); 4 = 3 + 1, (11.126)

where the U (1) charge is identified with baryon number (1/3 for quarks) minuslepton number.

SU(5)

In a different realization of this idea, H. Georgi and S. L. Glashow (Phys. Rev.Letters 32, 438 (1974)) gather all gauge symmetries of the Standard Model in onegauge group, SU (5),

SU (5) ⊃ SU (2) × SU (3)c × U (1), (11.127)

with embedding defined by the decomposition of its fundamental quintet

5 = ( 2, 1c )1 + ( 1, 3c )− 13, 5 = ( 2, 1c )−1 + ( 1, 3c ) 1

3, (11.128)

Page 259: Pierre Ramond Group Theory a Physicists Survey 2010

11.6 Grand Unification 247

as the SU (2) doublet is pseudoreal, and the subscript denotes the U (1) charge.The other fundamental 10 and the adjoint 24 representations are obtained fromKronecker products (see Appendix 2 for more details)

5 × 5 = 10antisym + 15sym, 5 × 5 = 24 + 1. (11.129)

Multiplying out the representations in (11.128), we obtain

10 = ( 2, 3c ) 13+ ( 1, 3c )− 4

3+ ( 1, 1c )2, (11.130)

as well as the adjoint representation

24 = ( 3, 1c )0 + ( 1, 8c )0 + ( 1, 1c )0 + ( 2, 3c ) 23+ ( 2, 3c )− 2

3. (11.131)

We can now interpret these group-theoretical results.The elementary particles of each family fit in two representations

5:[(

νe

e

), d

]; 10:

[(ud

), u, e

]. (11.132)

Remarkably, the hypercharge is in SU (5), from which the third integral charges ofthe quarks arise naturally (a similar relation was determined in the Standard Modelby requiring that it be free from anomalies). This pattern is repeated for the othertwo families.

The gauge bosons of an SU (5) Yang–Mills theory reside in its adjoint repre-sentation. Comparing with (11.131), we recognize the eight gluons ( 1, 8c ), thehypercharge ( 1, 1c ), and the weak bosons ( 3, 1c ) of the Standard Model.

SU (5) predicts new forces beyond the Standard Model, mediated by the othertwelve gauge bosons. As their charges indicate, they cause transitions betweenleptons and the anti-down quark when acting on the 5, as well as quark–antiquarktransitions when acting on the 10. They are called leptoquark gauge bosons. Ifmassless, they would cause immediate proton decay. The experimental limit on theproton lifetime (� 1033 years) requires these gauge bosons to be very massive, sothat the SU (5) symmetry, while extraordinary for taxonomy, must be broken atdistances very much smaller than a nuclear radius.

This conclusion is quantitatively in accord with the strength of the couplings. Forleptons and quarks to unify, the strength of their couplings must be similar. Usingthe experimentally determined values of the three gauge couplings of the Stan-dard Model, as boundary condition, their renormalization group running implies apossible merging at scales which are also extraordinarily small.

The Achilles’ heel of grand unified theories is that the symmetries they implymust be broken. In analogy with the Standard Model, the symmetry breaking isassumed to be spontaneous, that it is a property of the lowest energy state. Thisrequires an elaborate machinery. To this day, there are no credible ways to achieve

Page 260: Pierre Ramond Group Theory a Physicists Survey 2010

248 The groups of the Standard Model

such breakings of symmetry. On the other hand, the elementary fermions fit beau-tifully into representations of these gauge group, and the evolution of the couplingstowards a single value at shorter scales gives credence to the idea of a single gaugegroup.

SO(10)

The Pati–Salam SU (4) can be extended to SO(10) (H. Fritzsch and P. Minkowski,Ann. Phys. 93, 193 (1975) and independently H. Georgi in Proceedings of theAmerican Institute of Physics 23 ed. C. E. Carlson, 575 (1974)) through theembedding

SO(10) ⊃ SU (4) × SU (2)L × SU (2)R. (11.133)

The first SU (2)L is the one in the Standard Model, while the second describesthree gauge bosons which cause transitions among right-handed particles. Underthis embedding (see Appendix 2), the SO(10) spinor representation is given by

16 = ( 4, 2, 1 ) + ( 4, 1, 2 ). (11.134)

After interpreting leptons as the fourth color, we recognize the particles of theStandard Model, and a new lepton. It is the right-handed partner of the neutrino.This is easiest seen through the embedding

SO(10) ⊂ SU (5) × U (1), (11.135)

with

16 = 53 + 10−1 + 1−5. (11.136)

In SU (5), the electric charge is within, and the new particle has no color, no weaknor electric charge; it can rightly be interpreted as the right-handed partner of theneutrino in each family. Its Majorana mass bilinear transforms under SU (4) ×SU (2) × SU (2) as

( 10, 1, 3 ) + ( 6, 1, 1 ), (11.137)

and is commensurate with the scale of the breaking of Pati–Salam’s SU (4), whichis expected to be very large. Such a large Majorana mass suggests a natural mech-anism, the see-saw mechanism, for generating tiny neutrino masses. These lead toneutrino oscillations of the type observed in the recent past. It provides yet anothercircumstantial evidence for the idea of grand unification.

Its adjoint representation

45 = 240 + 10 + 104 + 10−4, (11.138)

shows more gauge bosons than in SU (5), all of which cause interactions notyet observed in Nature. Breaking the larger SO(10) symmetry presents a chal-lenge; the mechanism by which it occurs remains a mystery to be solved by futuregenerations.

Page 261: Pierre Ramond Group Theory a Physicists Survey 2010

11.7 Possible family symmetries 249

E6

The next rung on the ladder of gauge symmetries

SU (3) × U (1) ⊂ SU (3) × SU (2) × U (1) ⊂ SU (5) ⊂ SO(10) ⊂ ?,

is the exceptional group E6. Its gauge theory (F. Gürsey, P. Ramond, and P. Sikivie,Phys. Lett. B60, 177 (1976)) uses the embedding

E6 ⊃ SO(10) × U (1), (11.139)

with each family described by

27 = 161 + 10−2 + 14. (11.140)

Here the 10 is the vector representation of SO(10), which breaks under SU (5) asthe vector-like pair 5 +5, with masses expected to be out of present experimentalrange. This sequence is familiar to mathematicians, in terms of plucking dots outof Dynkin diagrams

SU (3)×U (1) ⊂ SU (3)× SU (2)×U (1) ⊂ SU (5) ⊂ SO(10) ⊂ E6 ⊂ E7 ⊂ E8

Amazingly, E8 × E8 appears as the gauge symmetry of the Heterotic String model.Nature does seem to like exceptional structures!

11.7 Possible family symmetries

The patterns of elementary fermions appear in triplicate, pointing to a three-foldsymmetry. It is as mysterious today as it was when the first heavy lepton, the muon,was discovered (e.g. I. I. Rabi’s famous remark “who ordered that?”).

The impulse has been to assemble the three families in a family group, with two-and three-dimensional representations. This group may be continuous or finite. Ifcontinuous, it means either SU (2) or SU (3), as Lie algebra with such representa-tions; if finite, the family group has to be a finite subgroup of SU (2), SO(3), orSU (3). The recent discovery of large neutrino mixing angles has prompted interestin exploring finite family groups. In the following, we list the finite groups withsuch representations. While this connection is pure speculation, it allows for theintroduction of a number of finite groups that may be of physical interest.

11.7.1 Finite SU(2) and SO(3) subgroups

All finite groups with two-dimensional representations were identified as earlyas 1876 by Felix Klein, and summarized by G. A. Miller, H. F. Blichfeldt, andL. E. Dickson [16]. Consider the transformations

Page 262: Pierre Ramond Group Theory a Physicists Survey 2010

250 The groups of the Standard Model

Z =(

z1

z2

)→ M Z , (11.141)

where M is a unitary matrix with unit determinant. Klein constructed threequadratic forms,

�r = Z† �σ Z , (11.142)

where �σ are the Pauli matrices. These transformations act on the vector �r asrotations about the origin. These are five types of such rotations:

• rotations of order n about one axis that generate the finite Abelian group Zn ;• rotations with two rotation axes that generate the dihedral group Dn ;• rotations that map the tetrahedron into itself form T = A4, the tetrahedral group;• rotations that map the cube into itself form O = S4, the octahedral group;• rotations that map the icosahedron (dodecahedron) into itself form I = A5, the

icosahedral group.

Recall that the dihedral groups have only singlet and real doublet representations(n > 2). T , O and I are the symmetry groups of the platonic solids. Since they areSO(3) subgroups, all have real triplet representations, but no real two-dimensionalrepresentations.

Each can be thought of as the quotient group of its binary-equivalent groupwith at least one two-dimensional “spinor” representation and twice the numberof elements (double cover). In ATLAS notation, these are

2 · Dn, 2 · A4, 2 · S4, 2 · A5. (11.143)

The first is just Q2n , the binary dihedral or dicyclic group introduced earlier.Since A4 = PSL2(3), the binary tetrahedral group is SL2(3), obtained by drop-

ping the projective restriction, with 24 elements. Its presentation is closely relatedto that of A4

PSL2(3) : < a, b, | a2 = b3 = (ab)3 = e >,

SL2(3) : < a, b, | a2 = b3 = (ab)3 >.

In the binary group, a2 is not the identity, but the negative of the identity, and it isthus of order four: A4, with no element of order four, is not a subgroup, rather it isits largest quotient group

A4 = SL2(3)/Z2. (11.144)

The same reasoning applies to I = A5 = PSL2(5), and the binary icosahedralgroup is SL2(5) with 120 elements, dropping the projective requirement. No longersimple, it has a normal subgroup of order two.

Page 263: Pierre Ramond Group Theory a Physicists Survey 2010

11.7 Possible family symmetries 251

Its presentation is of the same form as for A5

PSL2(5) : < a, b | a2 = b3 = (ab)5 = e >,

SL2(5) : < a, b | a2 = b3 = (ab)5 > .

This generates (at least) one two-dimensional spinor representation to those of A5.In that representation, the generators are written in terms of η, η5 = 1, as

ab =(−η2 0

0 −η3

), b = 1√

5

(η4 − η2 η − 11 − η4 η − η3

). (11.145)

SL2(5) has in fact four spinor-like representations, 2s, 2′s, 4s, and 6s. The character

table is given by the following, where ϕ = (1 + √5)/2, ϕ = (−1 + √

5)/2.

SL2(5) C1 12C [5]2 12C [5]

3 1C [2]4 12C [10]

5 12C [10]6 20C [3]

7 20C [6]8 30C [4]

9

χ [1] 1 1 1 1 1 1 1 1 1

χ [31] 3 ϕ −ϕ 3 −ϕ ϕ 0 0 −1

χ [32] 3 −ϕ ϕ 3 ϕ −ϕ 0 0 −1

χ [4] 4 −1 −1 4 −1 −1 1 1 0

χ [5] 5 0 0 5 0 0 −1 −1 0

χ [2s] 2 −ϕ ϕ −2 −ϕ ϕ −1 1 0

χ [2′s] 2 ϕ −ϕ −2 ϕ −ϕ −1 1 0

χ [4s] 4 −1 −1 −4 −1 −1 1 0 −1

χ [6s] 6 1 1 −6 −1 −1 0 0 0

The Kronecker products of its representations follows an intriguing pattern,known as the McKay correspondence. Start from the singlet irrep and multiplyit with the spinor irrep 2s, to get of course

1 × 2s = 2s.

We associate one dot with 1, and one for 2s, and connect them by a line. Next wegenerate new representations by continuing to take products with the spinor:

2s × 2s = 1 + 31, 2s × 31 = 2s + 4s,

and

2s × 4s = 5 + 31, 2s × 5 = 6s + 4s,

Page 264: Pierre Ramond Group Theory a Physicists Survey 2010

252 The groups of the Standard Model

creating four more dots for 31, 4s, 5, and 6s, connecting the new irrep with ahorizontal line as we go. The next product contains three irreps

2s × 6s = 5 + 32 + 4,

so we branch out two lines to the new representations. Now we have

2s × 4 = 6s + 2′s,

generating one new representation. However, the two chains now terminate becausefurther multiplication does not generate any new irrep

2s × 2′s = 4, 2s × 32 = 6s.

Remarkably, McKay’s graphical rules generate the following extended E8 Dynkindiagram with each of its dots associated with an irrep of the binary icosahedralgroup.

32

1 2s 31 4s 5 6s 4 2′s

This pattern holds true for the binary extensions of Zn , Dn , T , O, in McKay cor-respondence with the extended Dynkin diagrams of SU (n + 1), SO(2n), E6 andE7, respectively.

11.7.2 Finite SU(3) subgroups

The finite subgroups of SU (3) are described in Miller et al. [16], and discussedextensively by W. M. Fairbanks, T. Fulton, and W. H. Klink (J. Math. Phys. 5, 1038(1964)).

While not all finite subgroups of SU (3) have complex triplet representations,they find two infinite families and two simple groups with triplet representations:

• �(3n2) = (Zn × Zn) � Z3

• �(6n2) = (Zn × Zn) � S3

• A5 (see Appendix 1)• PSL2(7) of order 168 (see Appendix 1).

Four SU (3) subgroups have no triplet representations:

• Hessian group: �(216) = (Z3 × Z3) � SL2(3)• �(72) = (Z3 × Z3) � Q ⊂ �(216)

Page 265: Pierre Ramond Group Theory a Physicists Survey 2010

11.7 Possible family symmetries 253

• �(36) = (Z3 × Z3) � Z4 ⊂ �(72)• A6, the order 360 alternating group.

In the SU (2) case, we used Z2, the center of SO(3), to construct the double coverof its subgroups without spinor representations. Here, the center of SU (3) is Z3,allowing us to construct the triple cover of its subgroups without triplets. Thisgenerates four groups with complex triplet representations, and triple the numberof elements:

3 · �(216), 3 · �(72), 3 · �(366), 3 · A6,

shown here in ATLAS language.

Page 266: Pierre Ramond Group Theory a Physicists Survey 2010

12

Exceptional structures

Nature (and humans) seems to like those special mathematical structures whichhave unusual properties. For instance, the platonic solids have held a special placein both human imagination and group theory, where they are associated with excep-tional groups. For that reason, it is interesting to study these groups, using theirunique underlying algebraic structures.

12.1 Hurwitz algebras

Consider a complex number and its conjugate

z = x + iy, z = x − iy,

where x and y are real numbers. Its norm is defined as

N (z) ≡ √z z =

√x2 + y2.

The square of two complex numbers z, w is also a complex number, whose normwe can define. Trivially,

N (z w) = √z wzw = N (z) N (w), (12.1)

and the norm of the product is the product of the norms. This is also true of realnumbers for which the norm is the number itself. Can we generalize to numberswith more imaginary units? It turns out that there are only two more algebras whichshare this property. They are called Hurwitz algebras: the real numbers R, the com-plex numbers C, the quaternions Q with three imaginary units, and the octonions(Cayley numbers) O, with seven imaginary units. All have norms such that thenorm of the product of any two members is the product of their norms.

254

Page 267: Pierre Ramond Group Theory a Physicists Survey 2010

12.1 Hurwitz algebras 255

Quaternions

We have already encountered quaternions in the context of finite groups. A quater-nion q and its conjugate q are defined as

q ≡ x0 + e1 x1 + e2 x2 + e3 x3, q = x0 − e1 x1 − e2 x2 − e3 x3, (12.2)

where xi , i = 0, 1, 2, 3 are real numbers, and ea , a = 1, 2, 3 are three imaginaryunits

e1 e1 = e2 e2 = e3 e3 = −1. (12.3)

They must satisfy

ea eb = εabc ec, a �= b = 1, 2, 3, (12.4)

in order to share the same norm property as complex numbers for any twoquaternions q and q ′,

N (q q ′) = N (q) N (q ′), (12.5)

where the norm is defined in the usual manner

N (q) ≡ √q q =

√x2

0 + x21 + x2

2 + x23 . (12.6)

These are Hamilton’s quaternions, the natural generalization of complex numbers,and a unit quaternion represents a point on the unit four-sphere.

Let η, χ be pure imaginary quaternions, η = −η, χ = −χ , and consider thetransformation

q → q ′ = eη q eχ , (12.7)

and take its conjugate (reversing the order of operators)

q → q ′ = e−χ q e−η.

The norm is clearly left untouched by these six-parameter transformations, thenorm group SO(4). By taking χ = −η, we specialize to the SU (2) subgroup

q → q ′ = eη q e−η, (12.8)

or infinitesimally,

δ q = [ η, q ]. (12.9)

Setting η = em , apply this transformation to the quaternion multiplication law,finding

[ em, ei e j − εi jk ek ] = 0. (12.10)

Page 268: Pierre Ramond Group Theory a Physicists Survey 2010

256 Exceptional structures

The transformations (12.8) leave quaternion multiplication invariant: SU (2) isthe automorphism group of the quaternion algebra. To physicists, quaternionsare simply i times the Pauli matrices, and as such satisfy an associative butnon-commutative algebra.

Octonions

Octonions are numbers with seven imaginary units eα,

ω = a0 +7∑1

aα eα; ω = a0 −7∑1

aα eα, (12.11)

where a0, aα are real numbers. They satisfy the Cayley algebra

eα eβ = −δαβ + ψαβγ eγ , (12.12)

where ψαβγ are totally antisymmetric octonion structure functions, whose onlynon-zero elements are

ψ123 = ψ246 = ψ435 = ψ651 = ψ572 = ψ714 = ψ367 = 1. (12.13)

We define the octonion norm as

N (ω) = √ωω =

√√√√(a0)2 +

7∑α=1

a2α. (12.14)

The ψαβγ are such that the magical norm property

N (ωω′) = N (ω)N (ω′), (12.15)

holds for any two octonions ω, ω′. Like the quaternions, the Cayley algebra isnon-commutative, but a little computation shows that it is not associative as well.

The associator of three octonions ω1,2,3 is defined as

( ω1, ω2, ω3 ) ≡ (ω1 ω2) ω3 − ω1 (ω2 ω3). (12.16)

Using the algebra, we find

( eα, eβ, eγ ) = −2ψαβγ δ eδ, (12.17)

where

ψα1α2α3α4= 1

3!εα1α2α3α4α5α6α7ψα5α6α7

(12.18)

is the dual of the octonion structure functions. The associator is completely anti-symmetric under the interchange of the three octonions: octonions form an alterna-tive algebra. Octonions also satisfy the even less intuitive Moufang identities (withmany more to be found in Gürsey and Tze [8]),

Page 269: Pierre Ramond Group Theory a Physicists Survey 2010

12.2 Matrices over Hurwitz algebras 257

(ω1 ω2) (ω3 ω1) = ω1 (ω2 ω3) ω1,

(ω1 ω2 ω1)ω3 = ω1 (ω2(ω1 ω3)),

ω1(ω2 ω3 ω2) = ((ω1 ω2)ω3) ω2.

Since octonions are described by eight real numbers, the group that leaves thequadratic norm invariant is SO(8). In the quaternion case, the norm group wasSO(4), and we used a transverse quaternion to reduce it to SO(3), which turnedout to be the automorphism group of the quaternion algebra. We follow a similarprocedure for octonions.

Imaginary octonions are given in terms of seven parameters, so we decomposeSO(8) in sevens, and find the sequence

SO(8) ⊃ SO(7) ⊃ G2. (12.19)

Written in terms of the imaginary octonion α, the seven coset generators ofSO(8)/SO(7) which map the real part ω0 into the imaginary units ωα, are simply

SO(8)/SO(7): ω → ω′ = eα ω eα, δ ω = α ω + ωα. (12.20)

Seven of the twenty-one SO(7) transformations that map the seven imaginarycomponents into themselves are given by

SO(7)/G2: ω → ω′ = eβ ω e−β, δ ω = β ω − ωβ, (12.21)

where β is an imaginary octonion. The remaining fourteen transformations gener-ate the exceptional group G2. Its action on octonions is not trivial, reflecting thefact that while group operations are associative, the octonion algebra is not. G2

turns out to be the automorphism group of the octonion algebra. Its infinitesimalaction can be written in terms of two imaginary octonions ζ and η, and involvesthe associator and a double commutator

G2: δ ω = 3( ζ, η, ω ) − [ [ ζ, η ], ω ], (12.22)

and it is not exactly straightforward to prove that this transformation respectsoctonion multiplication.

12.2 Matrices over Hurwitz algebras

In this section, we consider antihermitian matrices whose elements are membersof the Hurwirtz algebras. For the associative Hurwitz algebras, the commutator ofthese matrices are expected to close and form Lie algebras. On the other hand, formatrices whose elements are non-associative octonions, one expects that the Jacobiidentity will fail. Yet in a few special cases, Lie algebras can be generated this way,the four remaining exceptional groups, F4, E6, E7, and E8.

Page 270: Pierre Ramond Group Theory a Physicists Survey 2010

258 Exceptional structures

Antisymmetric matrices over real numbers

The commutator of two antisymmetric matrices is itself antisymmetric. These(2n + 1) × (2n + 1) and (2n × 2n) antisymmetric matrices generate the Lie alge-bras, SO(2n + 1) and SO(2n), respectively. Rotations from the i to the j axis arerepresented by matrices whose elements are zero except for the i j entry equal toone and j i entry equal to minus one.

Antihermitian matrices over complex numbers

The commutator of two such matrices is antihermitian and traceless. The Lie alge-bras An−1, are represented in their simplest form by antihermitian n × n matricesover the complex numbers⎛⎜⎜⎝

ia1 z12 · · · z1 n

−z12 ia2 . . . z2 n

· · · ·−z1 n −z2 n . . . ian

⎞⎟⎟⎠ ,

where zi j are complex numbers and ai are real. The commutator of two such matri-ces yields an antihermitian traceless matrix, which compels us to impose the tracecondition

n∑i

ai = 0.

The n(n − 1)/2 complex off-diagonal elements together with the (n − 1) realdiagonal elements account for the n2 − 1 parameters of SU (n).

Antihermitian matrices over quaternions

What matrices over quaternions? Antihermitian matrices over quaternions are ofthe form

M(Q) =

⎛⎜⎜⎝p1 q12 · · · q1 n

−q12 p2 . . . q2 n

· · · · · ·−q1 n −q2 n . . . pn

⎞⎟⎟⎠ ,

where qi j are quaternions and the pi are imaginary quaternions,

pi = −pi .

In analogy, let us impose the trace condition

n∑i

pi = 0.

Page 271: Pierre Ramond Group Theory a Physicists Survey 2010

12.3 The Magic Square 259

The number of parameters in this matrix is

4 × n(n − 1)

2+ 3 × (n − 1) = 2n(2n + 1)

2− 3,

parameters, since imaginary quaternions have three real parameters. This is threeshort of the number of parameters of the Sp(2n) algebras.

Since quaternions are not commutative, the commutator of two quaternionicmatrices is not automatically traceless, as it is for complex numbers. It is straight-forward to see that the trace of the commutator of two quaternionic matrices is thesum of quaternion commutators which, as per eq. (12.9), can be recast as elementsof the SU (2) automorphism group.

The action of Sp(2n) splits as (n × n) traceless antihermitian quaternionicmatrices, together with the quaternion automorphism group acting diagonally.

12.3 The Magic Square

The natural next step is to consider antihermitian matrices over octonions. Theanalogy with the quaternions fails because their trace is a transverse octonion,which transforms as the 7 of their automorphism group G2. Still, let us focus ontraceless matrices, and consider the automorphism group separately. For (n × n)traceless octonionic matrices, we find

8 × n(n − 1)

2+ 7 × (n − 1) = 4n2 + 3n − 7,

parameters, since an imaginary octonion has seven parameters.But here we encounter an apparently fatal roadblock: octonions are non-

associative, and thus octonionic matrices will not satisfy associativity (Jacobiidentity). This is consistent since matrices over the real, complex and quaternionsaccount for all four infinite families of Lie algebras, leaving out the five exceptionalgroups. The first G2 is accounted for as the octonions’ automorphism group. Whatabout the 52-parameter F4?

We note that the (3 × 3), octonionic traceless antihermitian matrix contains 36parameters, which together with the 14 in G2 yield 52 parameters, the same as F4.This is strong incentive to consider (3 × 3) matrices over octonions!

The story begins, as can be expected, with SO(8) triality. The first point to makeis that the action of the octonion’s norm group SO(8), can be decomposed withrespect to G2 into three different transformations acting on an octonion ω

t = d + lα + rβ, (12.23)

where d is an element of G2,

d: ω → ω + δ ω = ω + 3 ( ζ, η, ω ) − [ [ ζ, η ], ω ], (12.24)

Page 272: Pierre Ramond Group Theory a Physicists Survey 2010

260 Exceptional structures

and

lα: ω → ω + α ω,

rβ : ω → ω + ω β, (12.25)

where ζ , η, α and β are imaginary octonions. This decomposition can be shown tobe unique. To see this, let t = 0, so that

d = −lα − rβ, (12.26)

meaning that lα + rβ is an element of G2. The infinitesimal form of its transforma-tion, eq. (12.22), requires α = −β. Hence

lα − rα: ωω′ → (ωω′) + [α(ωω′) − (ωω′)α],but, it can also be interpreted as an element of G2,

lα − rα: ωω′ → ωω′ + (αω − ωα)ω′ + ω(αω′ − ω′α).

Equating the two forms and the use of alternativity, yields,

(α, ω, ω′) = 0,

which means that α associates with any octonion; it must be zero. We have shownthat t = 0 implies d = α = β = 0. Since dim SO(8)− dim G2 = 28 − 14 = 14,the generality of this decomposition ensues.

Let t1 be an infinitesimal element of SO(8), with action on an octonion ω,

ω → ω + t1(ω).

The principle of triality states that we can associate uniquely to t1 two otherelements of SO(8), t2 and t3, with the property,

t1(ωω′) = t2 (ω)ω′ + ω t3(ω′), (12.27)

for any two octonions ω and ω′. They are uniquely determined according to thefollowing.

t1 t2 t3

d d dlα rα −lα − rαrα −lα − rα lα

The uniqueness can be proven by considering the three cases t1 = d, t1 = l, andt1 = r , separately, and by clever use of the octonion algebra.

Page 273: Pierre Ramond Group Theory a Physicists Survey 2010

12.3 The Magic Square 261

Triality enables us to establish a one-to-one correspondence between any SO(8)element and the set of triples { t1, t2, t3 }. The SO(8) algebra is then realized in theform

[ { t1, t2, t3 }, { u1, u2, u3 } ] = { [ t1, u1 ], [ t2, u2 ], [ t3, u3 ] }. (12.28)

Consider (3 × 3) antihermitian octonionic matrices with only off-diagonalelements; there are 3 × 8 = 24 such matrices. Together with the 28 SO(8)transformations, they assemble into the 52-parameter exceptional group F4.

Write the twenty-four matrices with no diagonal elements

M(ωi ) =⎛⎝ 0 ω1 ω2

−ω1 0 ω3

−ω2 −ω3 0

⎞⎠ , (12.29)

for any three octonions ω1, ω2, ω3.The infinitesimal action on this matrix of the SO(8) triple { t1, t2, t3 }, is defined

to be

δM(ωi ) ≡⎛⎝ 0 t3(ω1) t2(ω2)

−t3(ω1) 0 t1(ω3)

−t2(ω2) −t1(ω3) 0

⎞⎠ , (12.30)

where the ti satisfy eq. (12.27).To satisfy algebraic closure, we need to determine the commutator of two matri-

ces in such a way that it yields off-diagonal matrices, and SO(8) triples. Thecommutator of M(ωi ) with M(ω′

i ) contains a matrix of the same kind, but alsoa diagonal matrix, which needs to be interpreted as an SO(8) transformation.

The naive commutator of two matrices of the same type

Xa =⎛⎝ 0 ωa 0

−ωa 0 00 0 0

⎞⎠ , (12.31)

where ωa are arbitrary octonions, is diagonal. According to our intuition, it mustbe interpreted as an SO(8) triple,

[ X1, X2 ] = {t1 (ω1 ω2), t2 (ω1 ω2)

, t3 (ω1 ω2)}. (12.32)

The form of these transformations is determined from their action on the matricesaccording to eq. (12.27). They are given by

t1 (ω1 ω2): ω → ω + 1

4ω2(ω1ω) − (1 ↔ 2),

t2 (ω1 ω2): ω → ω + 1

4(ωω1)ω2 − (1 ↔ 2),

t3 (ω1 ω2): ω → ω + 1

2(ωω1 − ω1ω)ω2 − (1 ↔ 2). (12.33)

Page 274: Pierre Ramond Group Theory a Physicists Survey 2010

262 Exceptional structures

In deriving these formulæ, much care is to be exercised because of non-associativity. As an application, we have the double “commutator,”

[ [ Xa, Xb ], Xc ] =⎛⎝ 0 t3 (ωa ωb)

(ωc) 0−t3 (ωa ωb)

(ωc) 0 00 0 0

⎞⎠ . (12.34)

We can verify the Jacobi identity by taking the cyclic sum over a, b, c, since∑cycl

t3 (ωa ωb)(ωc) = 0, (12.35)

using eq. (12.33).In general, the “commutator” between two matrices is given by another matrix,

and a set of triples

[ M(ωi ), M(ω′i ) ] = M(ω′′

i )

+ ( t1 (ω3 ω′3), t2 (ω3 ω′

3), t3 (ω3 ω′

3))

+ ( t2 (ω2 ω′2), t3 (ω2 ω′

2), t1 (ω2 ω′

2))

+ ( t3 (ω1 ω′1), t1 (ω1 ω′

1), t2 (ω1 ω′

1)), (12.36)

where

ω′′1 = ω′

2ω1 − ω2ω′1, ω′′

2 = ω′1ω3 − ω1ω

′3, ω′′

1 = ω′3ω2 − ω3ω

′2. (12.37)

Further checks show that the twenty-four M matrices, together with the twenty-eight similarity triples, close on the exceptional group F4. This remarkableconstruction of a composition algebra is due to J. Tits.

The same procedure, applied to real numbers, complex numbers, and quater-nions, yields three classical groups SO(3), SU (3), and Sp(6), respectively. Theirdimensions are given by

3nH + 2(nH − 1) + dH , (12.38)

where nH is the number of parameters in the four Hurwitz algebra (1, 2, 4, 8), anddH is the dimension of the automorphism group of the algebra, 0, 0, 3, 14. Theembeddings

F4 ⊃ SO(3) × G2, F4 ⊃ Sp(6) × SU (2) F4 ⊃ SU (3) × SU (3),

easily follows from these constructions.The F4 construction can be thought of as involving two Hurwitz algebras, the

reals and octonions. In fact, it generalizes to two arbitrary Hurwitz algebras, andproduces the remaining exceptional groups.

Page 275: Pierre Ramond Group Theory a Physicists Survey 2010

12.3 The Magic Square 263

While triality is of course trivial for real and complex numbers, it can still bedefined for quaternions but the decomposition is not unique. With SO(8) replacedby SO(4), it is written in the form

t = d + rq, (12.39)

where d is an element of the automorphism group SU (2), and q = −q is a pureimaginary quaternion. The similarity triples are as follows.

t1 t2 t3

d d dd lq −rqlq 0 −rq0 rq −lq

Starting with two independent Hurwitz algebras, we build antihermitian matriceswith no diagonal elements, whose matrix elements are products of the two algebras,H and H ′:

M(ωi , ρi ) =⎛⎝ 0 ω1ρ1 ω2ρ2

−ω1ρ1 0 ω3ρ3

−ω2ρ2 −ω3ρ3 0

⎞⎠ ,

where ωi (ρi ) belong to H(H ′), respectively. The diagonal part of their commutatoris interpreted as elements of the similarity triples of H and H ′.

These matrices and the two similarity triples generate new composition algebras.If H represents real numbers, this is exactly our construction of F4.

When both H and H ′ are octonions, the composition algebra contains two sim-ilarity triples, which generate the 56 parameters of SO(8) × SO(8). Each matrixelement contains 82 possibilities, so that there are 3(64) = 192 matrices. Alltogether they generate the 248 parameter exceptional group E8.

Since the construction is symmetric, there are ten composition algebras, asshown in the following Magic Square.

R C Q O

R SO(3) SU (3) Sp(6) F4Z SU (3) SU (3) × SU (3) SU (6) E6Q Sp(6) SU (6) SO(12) E7O F4 E6 E7 E8

Page 276: Pierre Ramond Group Theory a Physicists Survey 2010

264 Exceptional structures

We can read off interesting algebraic embeddings along its rows or columns. Forinstance, the algebras in the last column contains those in the third column timesSU (2), and the second column times SU (3), and the first column times G2. Forexample, the last row yields

E8 ⊃ E7 × SU (2), E8 ⊃ E6 × SU (3), E8 ⊃ F4 × G2,

and the same applies to the other three rows.

Page 277: Pierre Ramond Group Theory a Physicists Survey 2010

Appendix 1

Properties of some finite groups

D3 ∼ S3 ∼ Z3 � Z2

Order: 6, permutation/dihedral group

Presentation: <a, b | a3 = e, b2 = e, bab−1 = a−1>

Classes: C [2]2 (b, ab, a2b), C [3]

3 (a, a2)

Permutations: (1 2), (1 2 3)

S3 C1 3C [2]2 2C [3]

3

χ [1] 1 1 1χ [11] 1 −1 1χ [2] 2 0 −1

Irrep:

2: a =(

e2π i/3 00 e−2π i/3

); b =

(0 11 0

)

Kronecker products:

11 × 11 = 1

11 × 2 = 2

2 × 2 = 1 + 11 + 2

265

Page 278: Pierre Ramond Group Theory a Physicists Survey 2010

266 Properties of some finite groups

D4 ∼ Z4 � Z2

Order: 8, dihedral/octic group

Presentation: <a, b | a4 = e, b2 = e, bab−1 = a−1>

Classes: C [2]2 (a2), C [2]

3 (b, a2b), C [2]4 (ab, a3b), C [4]

5 (a, a3)

Permutations: (1 2 3 4), (1 2)

D4 C1 C [2]2 2C [2]

3 2C [2]4 2C [4]

5

χ [1] 1 1 1 1 1

χ [11] 1 1 −1 1 −1

χ [12] 1 1 −1 −1 1

χ [13] 1 1 1 −1 −1

χ [2] 2 −2 0 0 0

Irrep:

2: a =(

i 00 −i

); b =

(0 11 0

)

Kronecker products:

11 × 11 = 12 × 12 = 13 × 13 = 1

1i × 1j = 1k, i �= j �= k,

1i × 2 = 2

2 × 2 = 1 + 11 + 12 + 13

Page 279: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of some finite groups 267

Q ∼ Q4

Order: 8, quaternion, dicyclic

Presentation: <a, b | a4 = e, a2 = b2, bab−1 = a−1>

Classes: C [4]2 (a, a3), C [2]

3 (a2), C [4]4 (b, a2b), C [4]

5 (ab, a3b)

Permutations: (1 2 3 4)(5 6 7 8), (1 6 3 8)(2 5 4 7)

Q C1 2C [4]2 C [2]

3 2C [4]4 2C [4]

5

χ [1] 1 1 1 1 1χ [11] 1 −1 1 1 −1χ [12] 1 −1 1 −1 1χ [13] 1 1 1 −1 −1χ [2] 2 0 −2 0 0

Irrep:

2: a =(

i 00 −i

); b =

(0 1

−1 0

)

Kronecker products:

11 × 11 = 12 × 12 = 13 × 13 = 1

1i × 1j = 1k, i �= j �= k,

1i × 2 = 2

2 × 2 = 1 + 11 + 12 + 13

Page 280: Pierre Ramond Group Theory a Physicists Survey 2010

268 Properties of some finite groups

D5 ∼ Z5 � Z2

Order: 10, dihedral group

Presentation: <a, b | a5 = e, b2 = e, ab = ba−1>

Classes: C [2]2 (b, ab, a2b, a3b, a4b), C [5]

3 (a, a4), C [5]4 (a2, a3)

Permutations: (1 2 3 4 5), (2 5)(3 4)

D5 C1 5C [2]2 2C [5]

3 2C [5]4

χ [1] 1 1 1 1

χ [11] 1 −1 1 1

χ [21] 2 0 2 cos(2π/5) 2 cos(4π/5)

χ [22] 2 0 2 cos(4π/5) 2 cos(2π/5)

Irrep:

2k: a =(

e2kπ i/5 00 e−2kπ i/5

), k = 1, 2; b =

(0 11 0

)

Kronecker products:

11 × 11 = 1

11 × 2k, = 2k, k = 1, 2,

21 × 22 = 21 + 22

21 × 21 = 1 + 11 + 22, 22 × 22 = 1 + 11 + 21

Page 281: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of some finite groups 269

(Q6) ∼ Z3 � Z4

Order: 12, dicyclic group

Presentation: <a, b | a6 = e, b2 = a3, bab−1 = a−1>

Classes: C [6]2 (a, a5), C [3]

3 (a2, a4), C [2]4 (a3), C [4]

5 (b, a2b, a4b), C [4]6 (ab, a3b, a5b),

Permutations: (1 2 3), (1 2)(4 5 6 7)

Q6 C1 2C [6]2 2C [3]

3 C [2]4 3C [4]

5 3C [4]6

χ [1] 1 1 1 1 1 1

χ [11] 1 −1 1 −1 i −i

χ [12] 1 1 1 1 −1 −1

χ [11] 1 −1 1 −1 −i i

χ [21] 2 1 −1 −2 0 0

χ [22] 2 −1 −1 2 0 0

Irrep:

2k: a =(

ekπ i/3 00 e−kπ i/3

); b =

(0 1

(−1)k 0

), k = 1, 2.

Kronecker products:

11 × 11 = 12 × 12 = 1

11 × 12 = 11, 11 × 12 = 11,

11 × 21 = 11 × 21 = 22

11 × 22 = 11 × 22 = 21

12 × 2k = 2k, k = 1, 2,

21 × 21 = 1 + 12 + 22, 22 × 22 = 1 + 12 + 22,

21 × 22 = 11 + 11 + 21

Page 282: Pierre Ramond Group Theory a Physicists Survey 2010

270 Properties of some finite groups

D6

Order: 12, dihedral group

Presentation: <a, b | a6 = e, b2 = e, ab = ba−1>

Classes: C [2]2 (a3), C [2]

3 (b, a2b, a4b), C [4]4 (ab, a3b, a5b), C [3]

5 (a2, a4), C [6]6 (a, a5)

Permutations: (1 2 3)(4 5), (2 3)

D6 C1 C [2]2 3C [2]

3 3C [2]4 2C [3]

5 2C [6]6

χ [1] 1 1 1 1 1 1

χ [11] 1 −1 −1 1 1 −1

χ [12] 1 −1 1 −1 1 −1

χ [13] 1 1 −1 −1 1 1

χ [21] 2 −2 0 0 −1 1

χ [22] 2 2 0 0 −1 −1

Irrep:

2k: a =(

ekπ i/3 00 e−kπ i/3

), k = 1, 2; b =

(0 11 0

)

Kronecker products:

11 × 11 = 12 × 12 = 13 × 13 = 1

1i × 1j = 1k, i �= j �= k,

11 × 21 = 22, 11 × 22 = 21,

12 × 21 = 22, 12 × 22 = 21,

13 × 21 = 21, 13 × 22 = 22,

21 × 21 = 22 × 22 = 1 + 13 + 22

21 × 22 = 11 + 12 + 21

Page 283: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of some finite groups 271

Hol(Z5) ∼ (Z2 × Z2) � Z3

Order: 20, holomorph of Z5

Presentation: <a, b | a5 = e, b4 = e, b−1ab = a3>

Classes: C [2]2 (b2, ab2, a2b2, a3b2, a4b2), C [4]

3 (b, ab, a2b, a3b, a4b),C [4]

4 (b3, ab3, a2b3, a3b3, a4b3), C [5]5 (a, a2, a3, a4)

Permutations: (1 2 3 4 5), (2 4 5 3)

Hol(Z5) C1 5C [2]2 5C [4]

3 5C [4]4 4C [5]

5

χ [1] 1 1 1 1 1

χ [11] 1 −1 i −i 1

χ [12] 1 1 −1 −1 1

χ [13] 1 −1 −i i 1

χ [4] 4 0 0 0 −1

Irrep:

4: a =

⎛⎜⎜⎝e2π i/5 0 0 0

0 e4π i/5 0 00 0 e8π i/5 00 0 0 e6π i/5

⎞⎟⎟⎠ ; b =

⎛⎜⎜⎝0 1 0 00 0 1 00 0 0 11 0 0 0

⎞⎟⎟⎠

Kronecker products:

11 × 11 = 12 × 12 = 1

11 × 12 = 11, 11 × 12 = 11

11 × 4 = 11 × 4 = 12 × 4 = 4,

4 × 4 = 1 + 11 + 11 + 12 + 4 + 4 + 4

Page 284: Pierre Ramond Group Theory a Physicists Survey 2010

272 Properties of some finite groups

A4 ∼ T

Order: 12, alternating group, tetrahedral group

Presentation: <a, b | a2 = b3 = (ab)3 = e>

Classes: C [3]2 (b, ab, aba, ba), C [3]

3 (b2, ab2, b2a, ab2a), C [2]4 (a, b2ab, bab2)

Permutations: a = (1 2)(3 4), b = (1 2 3)

A4 C1 4C [3]2 4C [3]

3 3C [2]4

χ [1] 1 1 1 1

χ [11] 1 e2π i/3 e4π i/3 1

χ [11] 1 e4π i/3 e2π i/3 1

χ [3] 3 0 0 −1

Irrep:

3: a =⎛⎝−1 0 0

0 1 00 0 −1

⎞⎠ ; b =⎛⎝0 1 0

0 0 11 0 0

⎞⎠

Kronecker products:

11 × 11 = 11; 11 × 11 = 11; 11 × 11 = 1

11 × 3 = 11 × 3 = 3,

2 × 31 = 2 × 32 = 31 + 32,

3 × 3 = 1 + 11 + 11 + 3 + 3

Page 285: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of some finite groups 273

S4

Order: 24, permutations group, symmetries of the cube

Presentation: <a, b | a4 = e, b2 = e, (ab)3 = e>

Classes: C [2]2 (b, a2ba2, ba2ba, aba2b, a3ba, aba3), C [2]

3 (a2, ba2b, a2ba2b),C [3]

4 (ab, a3ba2, a3b, aba2, ba3, ba, a2ba3, a2ba), C [4]5 (a, a3, aba, ba2, a2b)

Permutations: (1 2 3 4)

S4 C1 6C [2]2 3C [2]

3 8C [3]4 6C [4]

5

χ [1] 1 1 1 1 1

χ [11] 1 −1 1 1 −1

χ [2] 2 0 2 −1 0

χ [31] 3 1 −1 0 −1

χ [32] 3 −1 −1 0 1

Irreps:

2: a =(

0 e2π i/3

e4π i/3 0

); b =

(0 11 0

)

31: a =⎛⎝ 0 0 −1

0 −1 0−1 0 0

⎞⎠; b =⎛⎝1 0 0

0 0 10 1 0

⎞⎠

32: a =⎛⎝0 0 1

0 1 01 0 0

⎞⎠; b =⎛⎝−1 0 0

0 0 10 1 0

⎞⎠Kronecker products:

11 × 2 = 2; 11 × 31 = 32; 11 × 32 = 31

2 × 2 = 1 + 11 + 2,

2 × 31 = 2 × 32 = 31 + 32,

31 × 31 = 32 × 32 = 1 + 2 + 31 + 32

31 × 32 = 11 + 2 + 31 + 32

Page 286: Pierre Ramond Group Theory a Physicists Survey 2010

274 Properties of some finite groups

Z7 � Z3

Order: 21, Frobenius group

Presentation: <a, b | a7 = e, b3 = e, b−1ab = a4>

Classes: C2(b, ba, ba2, ba3, ba4, ba5, ba6),

C3(b2, b2a, b2a2, b2a3, b2a4, b2a5, b2a6), C4(a, a2, a4), C5(a3, a5, a6)

Permutations: (1 2 3 4 5 6 7), (2 5 3)(4 6 7)

Z7 � Z3 C1 7C [3]2 7C [3]

3 3C [7]4 3C [7]

5

χ [1] 1 1 1 1 1

χ [1′] 1 e2π i/3 e4π i/3 1 1

χ [1′] 1 e4π i/3 e2π i/3 1 1

χ [3] 3 0 0 12 (−1 + i

√7) 1

2 (−1 − i√

7)

χ [3] 3 0 0 12 (−1 − i

√7) 1

2 (−1 + i√

7)

3: a =⎛⎝e2π i/7 0 0

0 e4π i/7 00 0 eπ i/7

⎞⎠; b =⎛⎝0 1 0

0 0 11 0 0

⎞⎠

3: a =⎛⎝e5π i/7 0 0

0 e10π i/7 00 0 e6π i/7

⎞⎠; b =⎛⎝0 1 0

0 0 11 0 0

⎞⎠

Kronecker products:

1′ × 1′ = 1′; 1′ × 1

′ = 1; 3 × 1′ = 3

3 × 1′ = 3; 3 × 3 = (3 + 3)s + 3a

3 × 3 = 1 + 1′ + 1′ + 3 + 3

Page 287: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of some finite groups 275

A5

Order: 60, alternating group, icosahedral group

Presentation: <a, b | a2 = b3 = (ab)5 = e>

Classes: C2(ab), C3((ab)2), C4(a), C5(b)

Permutations: a = (12)(34) , b = (135)

A5 C1 12C [5]2 12C [5]

3 15C [2]4 20C [3]

5

χ [1] 1 1 1 1 1χ [31] 3 ϕ −ϕ −1 0χ [32] 3 −ϕ ϕ −1 0χ [4] 4 −1 −1 0 1

χ [5] 5 0 0 1 −1

ϕ = 12 (1 + √

5), ϕ = 12 (−1 + √

5)

31: a =(−1 0 0

0 −1 0ϕ ϕ 1

); b =

(0 1 00 0 11 0 0

)

32: a =(−1 0 0

0 −1 0−ϕ −ϕ 1

); b =

(0 1 00 0 11 0 0

)

4: a =(1 0 0 0

0 1 0 00 0 −1 00 0 0 −1

); b = 1

4

⎛⎝ −1 −√5 −√

5 −√5√

5 1 1 −3−√

5 −1 3 −1√5 −3 1 1

⎞⎠

5: a =⎛⎜⎝1 0 0 0 0

0 −1 0 0 00 0 −1 0 00 0 0 1 00 0 0 0 1

⎞⎟⎠; b = − 12

⎛⎜⎝1 0 1 −ω2 −ω

0 1 1 −1 −1−1 −1 0 ω ω2

ω 1 ω2 0 −ω

ω2 1 ω −ω2 0

⎞⎟⎠; ω3=1

A5 Kronecker products:

31 × 31 = 31a + (1 + 5)s; 32 × 32 = 32a + (1 + 5)s; 31 × 32 = 4 + 5

31 × 4 = 32 + 4 + 5; 32 × 4 = 31 + 4 + 5

31 × 5 = 31 + 32 + 4 + 5; 32 × 5 = 31 + 32 + 4 + 5

4 × 4 = (31 + 32)a + (1 + 4 + 5)s

5 × 5 = (31 + 32 + 4)a + (1 + 4 + 5 + 5)s; 4 × 5 = 31 + 32 + 4 + 5 + 5

Page 288: Pierre Ramond Group Theory a Physicists Survey 2010

276 Properties of some finite groups

PSL2(7)

Order: 168

Presentation: <a, b | a2 = b3 = (ab)7 = [a, b]4 = e>

Classes: C [1]1 ,C [2]

2 (a),C [3]3 (b),C [4]

4 ([a, b]),C [7]5 (ab),C [7]

6 (ab2)

PSL2(7) C [1]1 21C [2]

2 (a) 56C [3]3 (b) 42C [4]

4 ([a, b]) 24C [7]5 (ab) 24C [7]

6 (ab2)

χ [1] 1 1 1 1 1 1χ [3] 3 −1 0 1 b7 b7

χ [3] 3 −1 0 1 b7 b7χ [6] 6 2 0 0 −1 −1χ [7] 7 −1 1 −1 0 0χ [8] 8 0 −1 0 1 1

b7 = 12 (−1 + i

√7), b7 = 1

2 (−1 − i√

7); η7 = 1

3: a = i√7

(η2 − η5 η − η6 η4 − η3

η − η6 η4 − η3 η2 − η5

η4 − η3 η2 − η5 η − η6

); b = i√

7

(η3 − η6 η3 − η η − 1η2 − 1 η6 − η5 η6 − η2

η5 − η4 η4 − 1 η5 − η3

)

PSL2(7) Kronecker products:

3 × 3 = 3a + 6s; 3 × 3 = 1 + 8;3 × 6 = 3 + 7 + 8; 3 × 6 = 3 + 7 + 8;3 × 7 = 6 + 7 + 8; 3 × 7 = 6 + 7 + 8;3 × 8 = 3 + 6 + 7 + 8; 3 × 8 = 3 + 6 + 7 + 8;6 × 6 = (1 + 6 + 6 + 8)s + (7 + 8)a; 6 × 7 = 3 + 3 + 6 + 7 + 7 + 8 + 8;6 × 8 = 3 + 3 + 6 + 6 + 7 + 7 + 8 + 8;7 × 7 = (1 + 6 + 6 + 7 + 8)s + (3 + 3 + 7 + 8)a; 7 × 8 = 3 + 3 + 6 + 6 + 7 + 7 + 8 + 8 + 8;8 × 8 = (1 + 6 + 6 + 7 + 8 + 8)s + (3 + 3 + 7 + 7 + 8)a

Page 289: Pierre Ramond Group Theory a Physicists Survey 2010

Appendix 2

Properties of selected Lie algebras

We list the basic properties of the Lie algebras that occur frequently in physics.Their Dynkin diagrams are obtained by taking out the node ⊗ with number(0) from the extended Dynkin diagrams. The numbering of the nodes is shown,together with the dimensions of the fundamental representations. Representationsare labeled by their dimensions in boldface type with their Dynkin label as a sub-script whenever needed to avoid confusion. The adjoint representation (Adjt) isidentified with a supersript. Useful properties of the algebra are listed, in particularthe congruence properties of the representations, the marks, co-marks, the Coxeternumber, and the level vector.

Useful embeddings are shown for each algebra; for those which contain a U (1)factor, the U (1) charge appears as a subscript. Much of this information can befound in R. Slansky, Phys. Rep. 79, 1–128 (1981), as well as in the excellent tablesof W. G. McKay and J. Patera, Tables of Dimensions, Indices, and Branching Rulesfor Representations of Simple Lie Algebras, Marcel Dekker, Inc., New York (1981),as well as W. G. McKay, J. Patera, and D. W. Rand, Tables of Representationsof Simple Lie Algebras, Volume I Exceptional Simple Lie Algebras, PublicationsCRM, Montréal (1990). For the reader who finds these appendices insufficient, werecommend Schur, the open source software, from the late B. Wybourne. It can befound at http://sourceforge.net/projects/schur/

277

Page 290: Pierre Ramond Group Theory a Physicists Survey 2010

278 Properties of selected Lie algebras

SU(3)

(1)

3

(2)

3

� �⊗

���(0)

Representations & Dynkin indices

Reps I (2) I (3) I (4)

3(10) 1 1 4

3(01) 1 −1 4

6(20) 5 7 68

6(02) 5 −7 68

8(11) 6 0 72

10(30) 15 27 396

10(03) 15 −27 396

15(21) 20 14 464

15(40) 35 77 1484

27(22) 15 0 1944

Kronecker products

3(10) × 3(10) = 6 s(20) + 3

a(01)

3(10) × 3(01) = 8Adjt(11) + 1(00)

3(10) × 6(20) = 10(30) + 8Adjt(11)

3(01) × 6(20) = 15(21) + 3(10)

6(20) × 6(20) = 15 s(40) + 6

s(02) + 15 a

(21)

6(20) × 6(02) = 27(22) + 8Adjt(11) + 1(00)

3(10) × 8Adjt(11) = 15(21) + 6(02) + 3(10)

8Adjt(11) × 8Adjt

(11) = 27 s(22) + 8Adjt, s

(11) + 1 s(00) + 10 a

(30) + 10a(03) + 8Adjt, a

(11)

Page 291: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 279

Embeddings

SU (3) ⊃ SU (2) × U (1)

3 = 2−1 + 12; 8 = 30 + 10 + 23 + 2−3

SU (3) ⊃ SO(3)

3 = 3; 8 = 3 + 5

Representation congruence: Z3; Coxeter number: 3

Marks: (1, 1, 1); Level vector: (2, 2)

Page 292: Pierre Ramond Group Theory a Physicists Survey 2010

280 Properties of selected Lie algebras

SU(4) ∼ SO(6)

(1)

4

(2)

6

(3)

4

� � �⊗

��

���

���(0)

Representations & Dynkin indices

I (2) I (3) I (4)

4(100) 1 1 9

4(001) 1 −1 9

6(010) 2 0 24

10(200) 6 8 168

15(101) 8 0 192

20(011) 13 −7 381

20(020) 16 0 576

20(003) 21 29 1101

36(201) 33 21 1473

45(210) 48 48 2496

64(111) 64 0 3072

84(202) 112 0 7488

Kronecker products

4(100) × 4(100) = 10 s(200) + 6 a

(010)

4(100) × 4(001) = 15Adjt(101) + 1(000)

4(100) × 6(010) = 20(110) + 4(001)

6(010) × 6(010) = 20 s(020) + 1 s

(000) + 15 a(101)

4(100) × 15Adjt(101) = 4(001) + 20(011) + 36(201)

6(010) × 15Adjt(101) = 4(001) + 10(002) + 10(200) + 64(111)

15Adjt(101) × 15Adjt

(101) = 1 s(000) + 15Adjt, s

(101) + 20 s(020) + 84 s

(202) + 15Adjt, a(101) + 45 a

(210) + 45a(012)

Page 293: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 281

Embeddings

SU (4) ⊃ SU (3) × U (1)

4 = 3−1 + 1 3; 15 = 3−4 + 3 4 + 1 0 + 8 0

SU (4) ⊃ SU (2) × SU (2)

4 = (2 , 2); 15 = (3 , 1) + (1 , 3) + ( 3 , 3 )

SU (4) ⊃ SO(5)

4 = 4; 15 = 5 + 10

Representation congruence: Z4; Coxeter number: 4

Marks: (1, 1, 1, 1); Level vector: (3, 4, 3)

Page 294: Pierre Ramond Group Theory a Physicists Survey 2010

282 Properties of selected Lie algebras

SU(5)

(1)

5

(2)

10

(3)

10

(4)

5

� � � �⊗

�����(0)

Representations & Dynkin indices

I (2) I (3) I (4)

5(1000) 1 1 16

5(0001) 1 −1 16

10(0100) 3 1 72

15(2000) 7 9 328

24(1001) 10 0 400

35(0003) 28 −44 2392

40(0011) 22 −16 1168

45(0101) 24 −6 1224

50(0020) 35 −15 2360

70(2001) 49 29 3544

70(0004) 84 −156 11064

75(0110) 50 0 3200

105(0012) 210 −99 8176

126(2010) 105 75 9000

175(1101) 140 30 11360

200(2002) 200 0 21200

Page 295: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 283

Kronecker products

5(1000) × 5(1000) = 15 s(2000) + 10 a

(0100)

5(1000) × 5(0001) = 24Adjt(1001) + 1(0000)

5(1000) × 10(0100) = 40(1100) + 10(0010)

5(0001) × 10(0100) = 45(0101) + 5(1000)

10(0100) × 10(0100) = 50s(0200) + 5

s(1000) + 45

a(1010)

10(0100) × 10(0010) = 1(0000) + 24Adjt(1001) + 75(0110)

5(1000) × 24Adjt(1001) = 5(1000) + 45(0101) + 70(2001)

10(1000) × 24Adjt(1001) = 10(0100) + 15(2000) + 40(0011) + 175(1101)

24Adjt(1001) × 24Adjt

(1001) = 1 s(0000) + 24Adjt, s

(1001) + 75 s(0110) + 200 s

(2002) + 24Adjt, a(1001) + 126 a

(2010)

+126a(0102)

Embeddings

SU (5) ⊃ SU (4) × U (1)

5 = 4−1 + 1 4; 24 = 4−5 + 4 5 + 1 0 + 15 0

SU (5) ⊃ SU (3) × SU (2) × U (1)

5 = ( 3 , 1 )−2 + ( 1 , 2 )3;24 = ( 8 , 1 ) 0 + ( 1 , 3 ) 0 + ( 1 , 1 ) 0 + ( 3 , 2 )−5 + ( 3 , 2 ) 5

SU (5) ⊃ SO(5)

5 = 5; 24 = 10 + 14

Representation congruence: Z5; Coxeter number: 5;

Marks: (1, 1, 1, 1); Level vector: (4, 6, 6, 4)

Page 296: Pierre Ramond Group Theory a Physicists Survey 2010

284 Properties of selected Lie algebras

SU(6)

(1)

6

(2)

15

(3)

20

(4)

15

(5)

6

� � � � �⊗

�������

�������(0)

Representations & Dynkin indices

I (2) I (3) I (4)

6(10000) 1 1 25

6(00001) 1 −1 25

15(01000) 4 2 160

20(00100) 6 0 270

21(20000) 8 10 560

35(10001) 12 0 720

56(30000) 36 54 4500

70(11000) 33 27 2745

84(01001) 38 −4 2990

105(00101) 52 −22 4480

105(00020) 64 −40 6880

120(20001) 68 38 7220

126(00004) 120 −210 22800

175(00200) 120 0 14400

189(01010) 108 0 10800

210(00110) 131 −37 14315

210(00012) 152 −170 20720

252(00005) 330 −726 87450

280(20010) 192 108 24480

315(30001) 264 258 42960

336(00102) 248 −196 34040

Page 297: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 285

384(11000) 256 80 31360

405(20002) 324 0 49680

420(00021) 358 −308 56470

Kronecker products

6(10000) × 6(10000) = 21 s(20000) + 15 a

(01000)

6(10000) × 6(00001) = 35Adjt(10001) + 1(00000)

6(10000) × 15(01000) = 70(11000) + 20(00100)

6(10000) × 15(01000) = 6(00001) + 84(10010)

6(10000) × 20(00100) = 15(10010) + 105(10100)

15(01000) × 15(01000) = 105s(02000) + 105

a(10100) + 15

s(00010)

15(01000) × 15(00010) = 1(00000) + 189(01010) + 35Adjt(10001)

20(00100) × 20(00100) = 175 s(00200) + 35Adjt, s

(10001) + 1 s(00000) + 189 a

(01010)

15(01000) × 20(00100) = 210s(01100) + 84

s(10010) + 6

s(00001)

35Adjt(10001) × 35Adjt

(10001) = 405 s(20002) + 189 s

(01010) + 35Adjt, s(10001) + 1 s

(00000) + 35Adjt, a(10001)

+280 a(20010) + 280 a

(01002)

Embeddings

SU (6) ⊃ SU (5) × U (1)

6 = 5−1 + 1 5; 35 = 24 0 + 1 0 + 5−6 + 5 6

SU (6) ⊃ SU (4) × SU (2) × U (1)

6 = ( 4 , 1 )−1 + (1 , 2 )2;35 = ( 15 , 1 ) 0 + ( 1 , 3 ) 0 + ( 4 , 2 )−3 + ( 4 , 2 ) 3

SU (6) ⊃ SU (3) × SU (3) × U (1)

6 = ( 1 , 3 ) 1 + ( 3 , 1 )−1

35 = ( 8 , 1 ) 0 + ( 1 , 8 ) 0 + ( 1 , 1 ) 0 + ( 3 , 3 ) 2 + ( 3 , 3 )−2

SU (6) ⊃ SU (3) × SU (2)

6 = ( 3 , 2 ); 35 = ( 8 , 1 ) + ( 1 , 3 ) + ( 8 , 3 )

Page 298: Pierre Ramond Group Theory a Physicists Survey 2010

286 Properties of selected Lie algebras

SU (6) ⊃ SU (3)

6 = 6; 35 = 27 + 8

SU (6) ⊃ SU (4)

6 = 6; 35 = 20(020) + 15

Representation congruence: Z6; Coxeter number: 6

Marks: (1, 1, 1, 1, 1); Level vector: (5, 8, 9, 8, 5)

Page 299: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 287

SO(5) ∼ Sp(4)

(1)

5

(2)

4

� �⊗(0)

Representations & Dynkin indices

I (2) I (4)

4(01) 1 3

5(10) 2 12

10(02) 6 60

14(20) 14 252

16(11) 12 156

20(03) 21 399

30(30) 54 1836

35(04) 56 1680

35(12) 42 924

Kronecker products

5(10) × 5(10) = 1 s(00) + 14 s

(20) + 10Adjt, a(02)

4(01) × 4(01) = 1 a(00) + 5 a

(10) + 10Adjt, s(02)

4(01) × 5(10) = 16(11) + 4(01)

10Adjt(02) × 10Adjt

(02) = 1 s(00) + 5 s

(10) + 14 s(20) + 35 s

(04) + 10Adjt, a(02) + 35 a

(12)

Page 300: Pierre Ramond Group Theory a Physicists Survey 2010

288 Properties of selected Lie algebras

Embeddings

SO(5) ⊃ SU (2) × SU (2)

5 = ( 1 , 1 ) + ( 2 , 2 ); 4 = ( 2 , 1 ) + ( 1 , 2 )

10 = ( 3 , 1 ) + ( 1 , 3 ) + ( 2 , 2 )

SO(5) ⊃ SU (2) × U (1)

5 = 1 2 + 1−2 + 3 0; 4 = 2 1 + 2−1;10 = 1 0 + 3−2 + 3 2 + 3 0

SO(5) ⊃ SU (2)

5 = 5; 4 = 4; 10 = 7 + 3

Representation congruence: Z2; Coxeter number: 4

Marks: (1, 2, 2); Co-marks: (1, 2, 1); Level vector: (4, 6)

Page 301: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 289

SO(7)

� � �(1)

7

(2)

21

(3)

8

⊗ (0)

Representations & Dynkin indices

I (2) I (4)

7(100) 2 24

8(001) 2 18

21(010) 10 216

27(200) 18 600

35(002) 20 528

48(101) 28 780

77(300) 88 5280

105(110) 90 3864

112(003) 108 5148

112(011) 92 3708

168(020) 192 11136

168(201) 170 8826

182(400) 312 28704

189(102) 180 8592

Kronecker products

7(100) × 7(100) = 1 s(000) + 27 s

(200) + 21Adjt, a(010)

7(100) × 21Adjt(010) = 105(110) + 35(002) + 7(100)

8(001) × 8(001) = 1 s(000) + 35 s

(002) + 21Adjt, a(010) + 7 a

(100)

7(100) × 8(001) = 48(101) + 8(001)

8(001) × 21Adjt(010) = 112(011) + 48(101) + 8(001)

21Adjt(010) × 21Adjt

(010) = 1 s(000) + 27 s

(200) + 35 s(002) + 168 s

(020) + 21Adjt, a(010) + 189 a

(102)

Page 302: Pierre Ramond Group Theory a Physicists Survey 2010

290 Properties of selected Lie algebras

Embeddings

SO(7) ⊃ SO(5) × U (1)

7 = 1 2 + 1−2 + 5 0; 8 = 4 1 + 4−1; 21 = 10 0 + 1 0 + 5 2 + 5−2

SO(7) ⊃ G2

7 = 7; 8 = 1 + 7; 21 = 7 + 14

SO(7) ⊃ SU (4)

7 = 1 + 6; 8 = 4 + 4; 21 = 6 + 15

SO(7) ⊃ SU (2 × SU (2) × SU (2))

7 = ( 1 , 1 , 3 ) + ( 2 , 2 , 1 ); 8 = ( 1 , 2 , 2 ) + ( 2 , 1 , 2 )

21 = ( 1 , 1 , 3 ) + ( 1 , 3 , 1 ) + ( 3 , 1 , 1 ) + ( 2 , 2 , 3 )

Representation congruence: Z2; Coxeter number: 6

Marks: (1, 2, 2, 2); Co-marks: (1, 2, 2, 1); Level vector: (6, 10, 6)

Page 303: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 291

SO(8)

(1)

8

(2)

28

(4) 8

(3) 8

� ���

��

��

⊗�

�(0)

Representations & Dynkin indices

I (2) I (4)

8(1000) 2 40

28(0100) 12 480

35(2000) 20 1120

56(1010) 30 1560

112(1000) 108 10800

160(1100) 120 9120

224(2010) 200 18400

294(4000) 420 63840

300(0200) 300 31200

350(1011) 300 26400

Kronecker products

8(1000) × 8(1000) = 1 s(0000) + 35 s

(2000) + 28Adjt, a(0100)

8(1000) × 28Adjt(0100) = 8(1000) + 56(0011) + 160(1100)

8(1000) × 8(0010) = 56(1010) + 8(0001)

8(1000) × 8(0001) = 56(1001) + 8(1000)

28Adjt(0100) × 8(0010) = 8(0010) + 56(1001) + 160(0110)

28Adjt(0100) × 8(0001) = 8(0001) + 56(1010) + 160(0101)

8(0010) × 8(0010) = 1 s(0000) + 35 s

(0020) + 28Adjt, a(0100)

Page 304: Pierre Ramond Group Theory a Physicists Survey 2010

292 Properties of selected Lie algebras

8(0010) × 8(0001) = 8(1000) + 56(0011)

8(0001) × 8(0001) = 1 s(0000) + 35 s

(0002) + 28Adjt, a(0100)

28Adjt(0100) × 28Adjt

(0100) = 1 s(0000) + 300 s

(0200) + 35 s(2000) + 35 s

(0002) + 35 s(0020) + 28Adjt, a

(0100)

+350 a(1011)

Embeddings

SO(8) ⊃ SU (4) × U (1)

8(1000) = 6 0 + 1 1 + 1−1; 8(0001) = 4 1/2 + 4−1/2; 8(0010) = 4−1/2 + 4 1/2

28 = 15 0 + 1 0 + 6 1 + 6−1

SO(8) ⊃ SU (2) × SU (2) × SU (2) × SU (2)

8(1000) = (2, 2, 1 , 1) + (1, 1, 2 , 2); 8(0001) = (1, 2, 2, 1) + (2 , 1, 1, 2);8(0010) = (2 , 1 , 2 , 1) + ( 1 , 2 , 1 , 2)

28 = ( 3 , 1 , 1 , 1 ) + ( 1 , 3 , 1 , 2 ) + ( 1 , 3 , 2 ) + ( 1 , 1 , 1 , 3 )

+( 2 , 2 , 2 , 2 )

SO(8) ⊃ SU (3)

8(1000) = 8; 8(0010) = 8; 8(0001) = 8; 28(0100) = 8 + 10 + 10

SO(8) ⊃ Sp(4) × SU (2)

8(1000) = ( 4 , 2 ); 8(0010) = ( 4 , 2 ); 8(0001) = ( 5 , 1 ) + ( 1 , 3 )

28 = ( 10 , 1 ) + ( 1 , 3 ) + ( 5 , 3 )

SO(8) ⊃ SO(7)

8(1000) = 1 + 7; 8(0010) = 8; 8(0001) = 8; 28 = 21 + 7

Representation congruence: Z2; Coxeter number: 5

Marks: (1, 1, 2, 2, 1); Level vector: (6, 10, 6, 6)

Page 305: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 293

SO(9)

� � � �(1)

9

(2)

36

(3)

84

(4)

16

⊗ (0)

Representations & Dynkin indices

I (2) I (4)

9(1000) 2 40

16(0001) 4 80

36(0100) 14 520

44(2000) 22 1160

84(0010) 42 2040

126(0002) 70 3800

128(1001) 64 3200

156(3000) 130 11960

231(1100) 154 10760

432(0101) 300 21360

450(4000) 550 75800

495(0200) 440 41440

576(2001) 464 40000

594(1010) 462 37560

672(0003) 616 58400

768(0011) 640 55040

910(2100) 910 99320

924(1002) 770 67000

1122(5000) 1870 355640

1650(0110) 1650 174600

1920(3001) 2240 288640

1980(0020) 2310 285960

Page 306: Pierre Ramond Group Theory a Physicists Survey 2010

294 Properties of selected Lie algebras

Kronecker products

9(1000) × 9(1000) = 1 s(0000) + 44 s

(2000) + 36Adjt, a(0100)

9(1000) × 36Adjt(0100) = 9(1000) + 84(0010) + 231(1100)

9(1000) × 84(0010) = 36Adjt(0100) + 126(0002) + 594(1010)

9(1000) × 16(0001) = 16(0001) + 128(1010)

36Adjt(0100)× 36Adjt

(0100) = 1 s(0000)+ 44 s

(2000)+ 126 s(0002)+ 495 s

(0200)+ 594 a(1010)+ 36Adjt, a

(0100)

36Adjt(0100) × 84(0010) = 9(1000) + 84(0010) + 126(0002) + 231(1100) + 924(1002)

+ 1650(0110)

36Adjt(0100) × 16(0001) = 16(0001) + 128(1001) + 432(0101)

84(0010) × 84(0010) = 1 s(0000) + 44 s

(2000) + 126 s(0002) + 495 s

(0200) + 924 s(1002)

+ 1980 s(0020) + 36Adjt, a

(0100) + 84 a(0010) + 594 a

(1010) + 2772 a(0102)

84(0010) × 16(0001) = 16(0001) + 128(1001) + 432(0101) + 768(0011)

16(0001) × 16(0001) = 1 s(0000) + 9 s

(11000) + 126 s(0002) + 36Adjt, a

(0100) + 84 a(0010)

Embeddings

SO(9) ⊃ SU (4) × SU (2)16 = ( 4 , 2 ) + ( 4 , 2 ); 9 = ( 6 , 1 ) + ( 1 , 3 )

36 = ( 15 , 1 ) + ( 1 , 3 ) + ( 6 , 3 )

SO(9) ⊃ SO(8)16 = 8(0010) + 8(0001); 9 = 8(1000) + 1; 36 = 8(1000) + 28

SO(9) ⊃ Sp(4) × SU (2) × SU (2)16 = ( 4 , 1 , 2 ) + ( 4 , 1 , 2 ); 9 = ( 5 , 1 , 1 ) + ( 1 , 2 , 2 )

36 = ( 10 , 1 , 1 ) + ( 1 , 3 , 1 ) + ( 1 , 1 , 3 ) + ( 5 , 2 , 2 )

SO(9) ⊃ SO(7) × U (1)16 = 8 1 + 8−1; 9 = 1 2 + 1−2 + 7 0; 36 = 1 0 + 21 0 + 7 2 + 7−2

SO(9) ⊃ SU (2) × SU (2)16 = ( 4 , 2 ) + ( 2 , 4 ); 9 = ( 3 , 3 )

36 = ( 3 , 1 ) + ( 1 , 3 )) + ( 3 , 5 )) + ( 5 , 3 )

Representation congruence: Z2; Coxeter number: 8

Marks: (1, 1, 2, 2, 2); Co-marks: (1, 1, 2, 2, 1);Level vector: (8, 14, 18, 10)

Page 307: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 295

SO(10)

(1)

10

(2)

45

(3)

120

(5) 16

(4) 16

� � ���

��

��

⊗ (0)

Representations & Dynkin indices

I (2) I (4)

10(10000) 2 60

16(10000) 4 150

45(01000) 16 960

54(20000) 24 1920

120(00100) 56 4560

126(00002) 70 6900

144(10001) 68 5670

210(10000) 112 10560

210(30000) 154 21420

320(11000) 192 21120

560(01001) 364 43170660(40000) 704 145920

672(00003) 616 105180

720(20001) 532 73470

770(02000) 616 92160

945(10100) 672 88320

1050(10002) 840 124800

1200(00101) 940 136050

1386(21000) 1232 208320

1440(00012) 1256 203580

1728(10011) 1344 193920

2640(30001) 2772 559470

2772(00004) 3696 938880

2970(01100) 2706 463260

Page 308: Pierre Ramond Group Theory a Physicists Survey 2010

296 Properties of selected Lie algebras

Kronecker products

10(10000) × 10(10000) = 1 s(00000) + 54 s

(20000) + 45Adjt, a(01000)

10(10000) × 45Adjt(01000) = 10(10000) + 120(00100) + 320(11000)

10(10000) × 120(00100) = 45Adjt(01000) + 210(00011) + 945(10100)

10(10000) × 16(00001) = 16(00010) + 144(10001)

45Adjt(01000) × 45Adjt

(01000) = 1 s(00000) + 54 s

(20000) + 210 s(00011) + 770 s

(02000) + 45Adjt, a(01000)

+945 a(10100)

45Adjt(01000) × 120(00100) = 10(10000) + 120(00100) + 126(00002) + 126(00020)

+320(11000) + 1728(10011) + 2970(01100)

45Adjt(01000) × 16(00001) = 16(00001) + 144(10010) + 560(01001)

120(00100) × 120(00100) = 1 s(00000) + 54 s

(20000) + 210 s(00011) + 770 s

(02000) + 1050 s(10002)

+1050s(10020) + 4125 s

(00200) + 45Adjt, a(01000) + 210 a

(00011)

+945 a(10100) + 5940 a

(01011)

120(00100) × 16(00010) = 16(00010) + 144(10001) + 560(01010) + 1200(00101)

16(00010) × 16(00010) = 10 s(10000) + 126 s

(00002) + 120 a(00100)

16(00010) × 16(00001) = 1(00000) + 45Adjt(01000) + 210(00011)

Embeddings

SO(10) ⊃ SU (5) × U (1)

16 = 5 3 + 10−1 + 1−5; 10 = 5−2 + 5 2

45 = 10−4 + 10 4 + 24 0 + 1 0

SO(10) ⊃ SU (4) × SU (2) × SU (2)

16 = ( 4 , 2 , 1 ) + ( 4 , 1 , 2 ); 10 = ( 6 , 1 , 1 ) + ( 1 , 2 , 2 )

45 = ( 6 , 2 , 2 ) + ( 15 , 1 , 1 ) + ( 1 , 3 , 1 ) + ( 1 , 1 , 3 )

SO(10) ⊃ SO(9)

16 = 16; 10 = 9 + 1; 45 = 9 + 36

SO(10) ⊃ SO(7) × SU (2)

16 = ( 8 , 2 ); 10 = ( 7 , 1 ) + ( 1 , 3 )

45 = ( 7 , 3 ) + ( 21 , 1 ) + ( 1 , 3 )

Page 309: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 297

SO(10) ⊃ Sp(4) × Sp(4)

16 = ( 4 , 4 ); 10 = ( 5 , 1 ) + ( 1 , 5 )

45 = ( 5 , 5 ) + ( 10 , 1 ) + ( 1 , 10 )

SO(10) ⊃ SO(8) × U (1)

16 = 8 1 + 8−1; 10 = 1 2 + 1−2 + 8 0

45 = 8 2 + 8−2 + 28 0 + 1 0

SO(10) ⊃ Sp(4)

16 = 16; 10 = 10; 45 = 10 + 35(12)

Representation congruence: Z2; Coxeter number: 8

Marks: (1, 1, 2, 2, 1, 1); Level vector: (8, 14, 18, 10, 10)

Page 310: Pierre Ramond Group Theory a Physicists Survey 2010

298 Properties of selected Lie algebras

G2

(2)

7

(1)

14

� �⊗(0)

Representations & Dynkin indices

I (2) I (4)

7(01) 2 8

14(10) 8 80

27(02) 18 216

64(11) 64 1216

77(03) 88 1936

77(20) 110 3080

182(04) 312 10608

189(12) 288 8640

Kronecker products

7(01) × 7(01) = 1 s(00) + 27 s

(02) + 7 a(01) + 14Adjt, a

(10)

7(01) × 14Adjt(10) = 7(01) + 27(02) + 64(11)

14Adjt(10) × 14Adjt

(10) = 1 s(00) + 27 s

(02) + 77 s(20) + 14Adjt, a

(10) + 77 a(03)

Embeddings

G2 ⊃ SU (3)7 = 1 + 3 + 3; 14 = 8 + 3 + 3

G2 ⊃ SU (2) × SU (2)7 = ( 2 , 2 ) + (1 , 3 ); 14 = ( 3 , 1 ) + (1 , 3 ) + (2 , 4 )

G2 ⊃ SU (2)7 = 7; 14 = 11 + 3

Representation Chang number: 0,±1 mod(7); Coxeter number: 6

Marks: (1, 2, 3); Co-marks: (1, 2, 1); Level vector: (10, 6)

Page 311: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 299

F4

(1)

52

(2)

1274

(3)

273

(4)

26

� � � �⊗(0)

Representations & Dynkin indices

I (2) I (4)

26(0001) 6 120

52(1000) 18 600

273(0010) 126 5880

324(0002) 162 8280

1053(1001) 648 41760

1274(0100) 882 64680

2652(0003) 2142 185640

4096(0011) 3072 245760

8424(1010) 7452 712080

Kronecker products

26(0001) × 26(0001) = 1 s(0000) + 26 s

(0001) + 324 s(0002) + 52Adjt, a

(1000) + 273 a(0010)

26(0001) × 273(0010) = 4096(0011) + 1274(0100) + 1053(1001) + 324(0002) + 273(0010)

+52Adjt(1000) + 26(0001)

26(0001) × 52Adjt(1000) = 26(0001) + 273(0010) + 1053(2000)

273(0010) × 52Adjt(1000) = 8424(1010) + 4096(0011) + 1053(1001) + 324(0002)

+273(0010) + 26(0001)

52Adjt(1000) × 52Adjt

(1000) = 1 s(0000) + 324 s

(0002) + 1053 s(2000) + 52Adjt, a

(1000) + 1274 a(0100)

Page 312: Pierre Ramond Group Theory a Physicists Survey 2010

300 Properties of selected Lie algebras

Embeddings

F4 ⊃ SO(9)

26 = 1 + 9 + 16; 52 = 36 + 16

F4 ⊃ SU (3) × SU (3)

26 = (3 , 3 ) + (3 , 3 ) + (1 , 8 )

52 = (6 , 3 ) + (6 , 3 ) + (1 , 8 ) + (8 , 1 )

F4 ⊃ Sp(6) × SU (2)

26 = (14 , 1 ) + (6 , 2 )

52 = (21 , 1 ) + (1 , 3 ) + (14 , 2 )

F4 ⊃ G2 × SU (2)

26 = (7 , 2 ) + (1 , 4 )

52 = (7 , 5 ) + (14 , 1 ) + (1 , 3 )

F4 ⊃ SU (2)

26 = 9 + 17

Representation Chang number: 0,±1 mod(13); Coxeter number: 12

Marks: (1, 2, 3, 4, 2); Co-marks: (1, 2, 3, 2, 1);Level vector: (22, 42, 30, 16)

Page 313: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 301

E6

(5) 27

(4) 351

(3) 2925

(2) 351

(1) 27

78

(6)

⊗(0)

Representations & Dynkin indices

I (2) I (4)

27(100000) 6 336

27(000010) 6 336

78(000001) 24 2016

351(000020) 168 23520

351(000100) 150 18480

650(100010) 300 40320

1728(000011) 960 158592

2430(000002) 1620 326592

2925(001000) 1800 332640

3003(000030) 2310 543312

5824(000110) 4032 846720

7371(010010) 5040 1044288

7722(100020) 5076 1271424

Kronecker products

27(100000) × 27(100000) = 27s(000010) + 351

s(200000) + 351

a(01000)

27(100000) × 351(010000) = 5824(110000) + 2925(001000) + 650(100010) + 78Adjt(000001)

Page 314: Pierre Ramond Group Theory a Physicists Survey 2010

302 Properties of selected Lie algebras

27(100000) × 351(000100) = 27(100010) + 351(010000) + 1728(000011) + 7371(100100)

27(100000) × 27(000010) = 1(000000) + 78Adjt(000001) + 650(100010)

27(100000) × 78Adjt(000001) = 27(100000) + 351(000100) + 1728(100001)

78Adjt(000001) × 78Adjt

(000001) = 2430 s(000002) + 650 s

(100010) + 1 s(000000) + 2925 a

(001000)

+78Adjt, a(000001)

Embeddings

E6 ⊃ SO(10) × U (1)

27 = 1 4 + 10−2 + 16 1 ; 78 = 1 0 + 45−2 + 16 3 + 16 3

E6 ⊃ SU (3) × SU (3) × SU (3)

27 = ( 3 , 1 , 3 ) + ( 3 , 3 , 1 ) + ( 1 , 3 , 3 )

78 = ( 8 , 1 , 1 ) + ( 1 , 8 , 1 ) + ( 1 , 1 , 8 ) + ( 3 , 3 , 3 ) + ( 3 , 3 , 3 )

E6 ⊃ SU (6) × SU (2)

27 = ( 15 , 1 ) + ( 6 , 2 )

78 = ( 35 , 1 ) + ( 1 , 3 ) + ( 20 , 2 )

E6 ⊃ G2 × SU (3)

27 = ( 1 , 6 ) + ( 7 , 3 ); 78 = ( 14 , 1 ) + ( 1 , 8 ) + ( 7 , 8 )

E6 ⊃ F4

27(100000) = 1(0000) + 26(0001); 78(000001) = 52(1000) + 26(0001)

E6 ⊃ Sp(8)

27 = 27(0100)

E6 ⊃ G2

27 = 27(02)

E6 ⊃ SU (3)

27 = 27(22)

Representation Chang number: 0, ±1 mod(13); Coxeter number: 12

Marks: (1, 2, 3, 2, 1, 2, 1); Level vector: (16, 30, 42, 30, 16, 22)

Page 315: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 303

E7

(1) 133

(2) 8645

(3) 365750

(4) 27664

(5) 1539

(6) 56

(7)

912

⊗ (0)

Representations & Dynkin indices

I (2) I (4)

56(0000010) 12 1008

27(000010) 6 336

133(1000000) 36 4032

912(0000020) 660 133056

1463(0000100) 150 18480

1539(0000100) 648 120960

6480(1000010) 3240 731808

7371(2000000) 4212 1100736

8645(0100000) 4680 1153152

24320(0000030) 17280 5709312

27664(0001000) 17160 4900896

40755(0000011) 25740 7495488

· · · · · · · · · · · · · · · · · ·365750(0010000) 297000 113097600

Page 316: Pierre Ramond Group Theory a Physicists Survey 2010

304 Properties of selected Lie algebras

Kronecker products

56(0000010) × 56(0000010) = 1 a(0000000) + 1539 a

(0000100) + 1463 s(0000020) + 133Adjt, s

(1000000)

56(0000010) × 133Adjt(1000000) = 56(0000010) + 912(0000001) + 6480(1000010)

133Adjt(1000000) ×133Adjt

(1000000) = 1 s(0000000) +1539 s

(0000100) +7371 s(2000010) +133Adjt, a

(1000000)

+8645 a(0100000)

Embeddings

E7 ⊃ SU (8)

56 = 28 + 28 ; 133 = 70 + 63

E7 ⊃ SO(12) × SU (2)

56 = ( 32(000001) , 1 ) + ( 12(100000) , 2 )

133 = ( 32(000010) , 2 ) + ( 1 , 3 ) + ( 66(0100000) , 1 )

E7 ⊃ SU (6) × SU (3)

56 = ( 20 , 1 ) + ( 6 , 3 ) + ( 6 , 3 )

133 = ( 35 , 1 ) + ( 1 , 8 ) + ( 15 , 3 ) + ( 15 , 3 )

Chang number: 0, ±1 mod(19); Coxeter number: 18

Marks: (1, 2, 3, 4, 3, 2, 1, 2); Level vector: (34, 66, 96, 75, 52, 27, 49)

Page 317: Pierre Ramond Group Theory a Physicists Survey 2010

Properties of selected Lie algebras 305

E8

(7) 3875

(6) 6696000

(5) 6899079264

(4) 146325270

(3) 2450240

(2) 30380

(1) 248

(8)

147250

⊗ (0)

Representations & Dynkin indices

I (2) I (4)

248(10000000) 60 8640

3875(00000010) 1500 371520

27000(20000000) 13500 4432320

30380(01000000) 14700 4656960

147250(00000001) 85500 32996160

779247(10000010) 502740 217183680

1763125(30000000) 1365000 715478400

2450240(00100000) 1778400 870704640

4096000(11000000) 3072000 1157135360

4881384(00000020) 3936600 2154107520

6696000(00000100) 5292000 2834818560

· · · · · · · · · · · · · · · · · ·146325270(00010000) 141605100 93799218240

· · · · · · · · · · · · · · · · · ·6899079264(00001000) 8345660400 6970295566080

Page 318: Pierre Ramond Group Theory a Physicists Survey 2010

306 Properties of selected Lie algebras

Kronecker products

248Adjt(10000000) × 248Adjt

(10000000) = 1 s(00000000) + 3875 s

(00000010) + 27000 s(20000000)

+ 248Adjt, a(10000000) + 30380 a

(01000000)

248Adjt(10000000) × 3875(00000010) = 248Adjt

(10000000) + 3875(00000010) + 30380(01000000)

+147250(00000001) + 779247(10000010)

Embeddings

E8 ⊃ SO(16)248 = 120(01000000) + 128(00000001)

E8 ⊃ SU (9)248 = 80(10000001) + 84(00100000) + 84(00000100)

E8 ⊃ E7 × SU (2)248 = ( 133 , 1 ) + ( 1 , 3 ) + ( 56 , 2 )

E8 ⊃ E6 × SU (3)

248 = ( 78 , 1 ) + ( 1 , 8 ) + ( 27 , 3 ) + ( 27 , 3 )

E8 ⊃ SU (5) × SU (5)

248 = ( 1 , 24 ) + ( 24 , 1 ) + ( 5 , 10 ) + ( 5 , 10 ) + ( 10 , 5 ) + ( 10 , 5 )

E8 ⊃ F4 × G2

248 = ( 52 , 1 ) + ( 1 , 14 ) + ( 26 , 7 )

E8 ⊃ SO(8) × SO(8)

248 = ( 1 , 28 ) + ( 28 , 1 ) + ( 8(1000) , 8(0010) ) + ( 8(0010) , 8(0001) )

+ ( 8(0001) , 8(1000) )

Chang number: 0, ±1 mod(31); Coxeter number: 30

Marks: (1, 2, 3, 4, 5, 6, 4, 2, 3);Level vector: (92, 182, 270, 220, 168, 114, 58, 136)

Page 319: Pierre Ramond Group Theory a Physicists Survey 2010

References

[1] Cahn, Robert N. (1984) Semi-Simple Lie Algebras and Their Representations(Benjamin/Cummings, Menlo Park).

[2] Carmichael, Robert D. (1956) Groups of Finite Order (Dover Publications, NewYork).

[3] Carter, Roger W. Simple Groups of Lie Type (John Wiley & Sons Ltd, London).[4] Dyson, Freeman J. (1966) Symmetry Groups in Nuclear and Particle Physics

(W. A. Benjamin, New York).[5] Fuchs, Jürgen (1992) Affine Lie Algebras and Quantum Groups (Cambridge

University Press, Cambridge).[6] Georgi, Howard (1999) Lie Algebras in Particle Physics (Perseus Books, Boulder).[7] Gilmore, Robert (1974) Lie Groups, Lie Algebras and Some of Their Applications

(John Wiley & Sons, New York).[8] Gürsey, Feza and Tze, Chia-Hsiung (1996) On the Role of Division, Jordan and

Related Algebras in Particle Physics (World Scientific, Singapore).[9] Hall Jr., Marshall (1959) The Theory of Groups (The Macmillan Company, New

York).[10] Hammermesh, Morton (1962) Group Theory and its Applications to Physical

Problems (Addison-Wesley, Reading).[11] Humphreys, James E. (1972) Introduction to Lie Algebras and Representation

Theory (Springer-Verlag, New York).[12] Kac, V. G. (1990) Infinite Dimensional Lie Algebras, 3rd edn (Cambridge University

Press, Cambridge).[13] Kass, S., Moody, R. V., Patera, J., and Slansky, R. (1990) Affine Lie Algebras, Weight

Multiplicities, and Branching Rules (University of California Press, Berkeley).[14] Ledermann, W. (1973) Introduction to Group Theory (Oliver & Boyd, Edinburgh).[15] Lomont, J. S. (1959) Applications of Finite Groups (Academic Press, Toronto).[16] Miller, G. A., Blichfeldt, H. F., and Dickson, L. E. (1916) Theory and Application of

Finite Groups (John Wiley & Sons, New York).[17] Rose, John S. (1978) A Course on Group Theory (Cambridge University Press,

Cambridge).[18] Sternberg, S. (1994) Group Theory and Physics (Cambridge University Press,

Cambridge).[19] Tinkham, M. (1964) Group Theory and Quantum Mechanics (McGraw-Hill, New

York).

307

Page 320: Pierre Ramond Group Theory a Physicists Survey 2010

Index

Abelian group, 13, 23, 193–4, 212, 250adjoint representation, 82, 100, 108, 110, 115, 124,

132, 147, 148, 165–66, 168, 171–72, 174, 217–18,247–48, 277

affine algebras, see Kac–Moodyalternating group, 21, 25, 41, 52, 190, 197, 253, 272,

275annihilation operators, 93, 95, 115, 143–44, 178, 215associativity, 5, 24, 86–87, 193, 259automorphism, 22, 29, 52, 58–59, 89, 203–6, 256–57,

259, 262–63

basic invariants, 66binary dihedral group, 14Bohr atom, 95–97bras, 33, 44, 69, 72–74, 102

Cartan–Killing form, 124, 126Cartan matrix, 107, 109–10, 113, 131, 136, 144–45,

148–89, 175, 177, 208, 209–11, 214, 215Casimir operators, 82, 97, 109, 144, 164, 255center, 4, 5, 11, 12, 16, 119, 197, 211, 253character, 37character table, 41–45, 48, 51–53, 55–58, 64, 66, 201,

251characteristic of a field, 194Chevalley basis, 112–13, 131, 190, 201–2, 204class coefficients, 39, 42classes, 20, 23, 38–44, 46, 49–52, 54, 57, 64Clebsch–Gordan, 45, 65, 86Clifford algebras, 153, 159, 162commutator, 71, 76, 81, 89–90, 94–98, 105, 110,

125–27, 131, 144, 152, 181, 209–11, 218, 257–59,261–63

commutator subgroup 27, 52, 56, 58, 68composition series, 26, 190compound characters, 50–51co-marks, 176–77, 215, 277conjugations, 22coset, 9, 61–63, 157, 199, 257covering group, 68, 89Coxeter groups, 151

Coxeter number, 148, 166, 277creation operators, 94–95, 183–84, 239cycle, 12, 19–23, 31, 41, 135–36, 190–91cyclic group, 4, 6–7, 10, 12, 15, 27, 59

derived subgroup, see commutator subgroupdicyclic group, 17, 67–68, 250, 258, 267, 269dihedral group, 7, 10, 12–13, 15, 29, 41, 58, 199, 250,

265, 266, 268, 270Dirac group, 71Direct product, 8–10, 13, 28–29, 46, 58–61, 83–84,

97, 99, 159, 178, 198, 222, 225dominant weight, 144, 151Dynkin diagram, 107, 109, 134–40, 142, 144, 146–47,

153–54, 156, 158, 164, 174–78, 190, 203–5, 210,252

Dynkin index, 165, 167, 168, 181Dynkin label, 80, 112, 177, 180, 277

eight-fold way, 70, 239Elliott model, 115–7, 170embeddings, 3, 48–51, 56, 64, 168–77, 210, 262, 264,

277embeddings, anomalous, 177embeddings, maximal, 171, 174, 176, 177, 210even (odd) permutation, 21, 30exceptional groups, 3, 166, 180, 203, 254, 257, 259,

262extended Dynkin diagrams, 177, 210, 252, 277

Fermi–Yang model, 99, 235Fermi oscillators, 70–81, 99Feynman diagrams, 166field, 150, 221, 224, 238–43Frobenius groups, 28, 198Frobenius map, 195–96, 203, 205, 274fully reducible, 34fundamental representations, 144, 145, 147, 153,

156–57, 164, 178, 239, 277

Galois field, 150, 193–97, 201–5gauge invariance, 242

308

Page 321: Pierre Ramond Group Theory a Physicists Survey 2010

Index 309

gauge theories, see Yang–MillsGell-Mann–Nishijima relation, 118, 122,

236, 245Gell-Mann matrices, 104, 111, 112, 151,

161, 241general linear group, 196gluons, 244, 245, 247Grand Unified Theories, 247Grassmann numbers, 219–20group generators, 46, 227, 233group lattice, 111, 150, 179group manifold, 87–88group of Lie type, 25, 131, 197

highest weight state, 80, 82–84, 86, 110–12, 116,143–45, 164, 178, 180–81, 183–84, 188, 214

Hilbert space, 2, 35, 46, 61–62, 69–79, 83–84,86–87, 89–90, 93, 102, 110–12, 151, 153, 154, 157,179

holomorph, 29, 58, 59, 271Hurwitz, 13–14, 254–55, 257, 262–63

index, 9, 61, 64, 146–47, 165, 167–68, 181,217, 238

induced representations, 61, 63invariants, 3, 64–67, 103, 164–65, 167, 202, 222,

240–41irreducible polynomials, 194–95irreducible representations, 22, 33–34, 65, 71, 82, 91,

111, 143, 151, 164, 166–67, 178–180, 185, 187,188

isotopic spin, 99

Jacobi identity, 79, 89, 105, 123–127, 216, 218–19,257, 262

Kac–Moody algebras, 3, 176, 210–15kaons, 117kets, 33–34, 44, 69, 72–74, 80, 94, 102, 119,

227, 229Klein’s four-group, 7, 17Kostant counter, 182Kronecker product, 8, 45–46, 48, 58, 65,

85, 169

Lagrange’s theorem, 9–10level vector, 144–45, 148, 277

magnetic moment, 236, 238marks, 166, 176, 177, 214–15, 277McKay correspondence, 251–52Minkowski space, 89, 207Molien, 66–67

non-compact (real) form, 218normal ordering, 215, 230normal subgroup, 30, 71, 150–51, 191, 250

octonions, 193, 254, 256–57, 259–63order, 5–6

P-group, 27partition, 31Pauli spin matrices, 16, 81permutation group, see symmetric grouppions, 99–100, 117Poincaré duality, 146projective linear group, 197projective transformations, 196, 198pseudoreal representations, 47–48

QCD, 239, 244, 246quadrupole, 95, 112, 114, 116quarks, 114, 122, 239, 244–47quaternion group, 2, 13–14, 17, 26–27,

41, 48quaternions, 13–14, 193, 254–56, 258–59, 262–63,

267quotient (factor) group, 24–26, 30, 52, 64, 150–51,

250

rank, 11, 13, 82, 109–10reducible representation, 38, 45, 49, 167regular representation, 22, 33–34, 38–42,

53–54, 57rho-mesons, 235ring, 194, 235root diagram, 106, 132–34, 170, 212roots, 125, 140

Sakata model, 117–18Schur multiplier, 67–68Schur’s lemmas, 35, 37, 39semi-direct product, 28–29, 58–61,

198, 225semi-simple Lie algebra, 129Serre presentation, 113, 208–10, 214–15simple group, 27, 30, 150, 197simple roots, 106, 129singleton representation, 93soluble group, 26special linear group, 196–97spinor representation, 80, 95, 155–59, 162, 178, 219,

248, 250–51sporadic groups, 3, 25, 31, 190, 205–7Standard Model, 1, 3, 77, 101, 221Steiner system, 200, 206structure constants, 123, 128, 131, 179super algebra, 218Sylow theorems, 27–28symmetric group, 19, 273symplectic algebras, 89, 164syzygies, 66

tetrahedral group, 15–17, 64, 66, 250, 272transposition, 11, 21, 30, 41, 47, 112transitivity, 30, 191

Verma basis, 188–89Verma module, 180–85, 188

Page 322: Pierre Ramond Group Theory a Physicists Survey 2010

310 Index

Weyl chamber, 109, 111, 144, 150, 180–81, 184, 186,214

Weyl character formula, 186–87Weyl dimension formula, 187Weyl function, 182Weyl group, 149–51, 185–86, 212Weyl reflections, 128, 132–33

Weyl vector, 141, 148, 150Wigner–Eckhardt theorem, 236

Yang–Mills theories, 244Young Tableaux, 31–32Yukawa couplings, 241, 244