Download PDF
ads:
A monster tale:
a review on Borcherds’ proof of
monstrous moonshine conjecture
by
Alan Gerardo Reyes Figueroa
Instituto de Matem´atica Pura e Aplicada
IMPA
ads:
Livros Grátis
http://www.livrosgratis.com.br
Milhares de livros grátis para download.
ads:
A monster tale:
a review on Borcherds’ proof of
monstrous moonshine conjecture
by
Alan Gerardo Reyes Figueroa
Disserta¸ao
Presented in partial fulfillment of
the requirements for the degree of
Mestre em Matem´atica
Instituto de Matem´atica Pura e Aplicada
IMPA
Maio de 2010
A monster tale:
a review on Borcherds’ proof of
monstrous moonshine conjecture
Alan Gerardo Reyes Figueroa
Instituto de Matem´atica Pura e Aplicada
APPROVED: 26.05.2010
Hossein Movasati, Ph.D.
Henrique Bursztyn, Ph.D.
Am´ılcar Pacheco, Ph.D.
Fr´ed´eric Paugam, Ph.D.
Advisor:
Hossein Movasati, Ph.D.
v
Abstract
The Monster M is the largest of the sporadic simple group s. In 1979 Conway and Norton
published the remarkable paper ‘Monstrous Moonshine’ [38] , proposing a completely unex-
pected relationship between fin i t e simple groups and modular functions, in which related
the Monster to the theory of modular forms. Conway and Norton conjectured in this paper
that there is a close conn e ct io n between the conjugacy classes of the Monster and the action
of certain subgr oups of SL
2
(R) on the upper half plane H. This conjecture im plies that
extensive information on the representations of the Monster is contained in the classical
picture describing the action of SL
2
(R) on the u p per half plane. Monstrous Moonshine
is the collection of questions (and few answers) that these observations had dir ect l y inspired.
In 198 8, the book ‘Vertex Operator Algebras and the Monster’ [67] by Frenkel, Lepowsky
and Meurman appeared. This book gave an explanat io n of why the Monster is related to
the the or y of modular forms, by showing th a t it acts as an automorphism group of a vertex
algebra V
of central charge 24, constr ucted in terms of the Leech lattice. Vertex o perators
also arise in mathematical physics contexts, such as conformal field theory and strin g the-
ory. Although Fren kel et al. did not prove the Conway-Norton conjecture, th ey provided
the raw material with which this conjecture could be approached. The mai n conjecture was
eventually pr oved by Borcherds in 1992 in his paper ‘Monstrous moonshine and monstrous
Lie superalgebras’ [12]. A key role in th e proof was played by a class of i n fi nite dimen si on a l
Lie algebras —called by Borcherds— generalized Kac-Moody algebras. Borcherds proved
properties of the monster Lie algebra which turned out to be sufficient to complete the
Conway-Norton conjecture. A key st ep in the proof was an application of the ‘no-ghost’
theorem from string theory.
Borcherds also obt a ined many other remarkable connections between sporadic finite simple
groups and modular forms in this and other papers. He was awarded a Fiel d s medal in
1998 for his contributions.
The principal interest of Moonshine is that it cons t it utes a new bridge between al geb r a ic
structures and modular apparatus, and a new era of collaboration between mathematics a n d
physics. This work is a modest revi ew of the Moonshine p henomena: the main conjectures,
Borcherds’ proof, open p r o b l em s and actual areas of research.
vii
Resumo
O Monstro M ´e o maior dos grupos simples espor´adicos. Em 1979 Conway e Nor t on pub-
licaram um remarc´avel artigo ‘Monstrous Moonshine’ [38], onde proporam uma rela¸ao
inesperada entre grupos finitos simples e fun¸oes modulares, na qual relacionaram o M on -
stro `a teoria de formas modulares. Conway e Nort on conjetura r am neste artigo que existe
uma forte conex˜ao entre as classes de conjuga¸ao do Monstro e a ao de certos subgru-
pos de SL
2
(R) no semi-plano superi or H. Tal conjetura implica qu e muita informa¸ao
das representa¸oes do M on st r o est´a contida na imagem que descreve a ao de SL
2
(R)
no semi-plano superior. Monstrous Moonshine ´e a cole¸ao de quest˜oes (e umas poucas
respostas) diretamente inspiradas por essas observoes.
Em 198 8, apareceu o livro ‘Vertex Operator Algebras and the Monster’ [67] de Frenkel, Lep-
owsky and Meurman. Este livro deu u m a explica¸ao de por que o Monstro est´a relacionado
com a teoria de formas modulares, most r an d o que M age como um grupo de automorfismos
de uma ´algebra de vˆertices V
de carga central 24, construida em termos do ret´ıculo de
Leech. Os operadores ertice ocorrem tamb´em em contextos da f´ısica matem´atica, tais
como teoria de campos conformes e teoria das cordas. Embora Frenkel et al. ao pr ova r am
a conjetura de Conway-Norton, eles proporcionar a m a ferramenta asica com a qual esta
conjetura pode-se provar. A conjetura principal eventualmente foi provada por Borcherds
em 1992 no seu artigo ‘Monstrous moonshine and monstrous Lie superalgebras’ [12]. Um
papel chave na prova ´e feito por uma classe de ´algebras de Lie infinito d i m en si onales
—chamadas por Borcherds— de ´algebras de Kac-Moody g er al iz ad a s. Borcherds provou
propriedades da ´algebra de Lie monstro que foram suficentes para completar a conjetura
Conway-Norton. Um papel fundamental na prova foi uma apli ca¸ao do teorema ‘no-ghost’
da teor i a das cordas.
Borcherds obteve tamb´em muitas outras conex˜oes entre grupos simples espor´adicos e for-
mas modu l a re s em este e outros artigos. Ele foi galardoado com a medalha Fields em 1998
pelas suas contribu¸oes.
O inter´es princi p a l de Moonshine ´e que ele estabele ce uma nova po nte entre as estruturas
alg´ebricas e as modulares, e come¸ca uma nova era de coopera¸ao entre a matem´atica e
a f´ısica. Esta disserta¸ao ´e uma revis˜ao somera do fen´omeno Moonshine: as conjeturas
principais, a p r ova de Borcherds, problemas abertos e ´areas atuais d e pesquisa.
ix
Preface
Monstrous Moonshine is probably one of the most esoteri c achievements ar i sing in m a t h -
ematics. The fact that th e Monster has connections to other parts of mathematics shows
that there is someth i n g very deep going on here. No one fully understands it, and the links
to the fields of physics and geometry are tantalizing. The Moonshine connections have
spawned a lot of work by a several mathematici ans and mathematical physicists recently,
and h ave opened intriguing conjectures, most of them rem ai ning open until today.
This work grew out in a attempt to explain the mysteries about the Monster and its relati on
with number theory, specially the j modul ar invariant. The idea about work with this topic
was given to me by Prof. Hossein Movasati, with the purpose to have a better understand
on some connections occurring between automorp h i c functions, geometry and physics. The
original i d ea was to write a compl et e and detailed p r oof of the Moonshines conjectu r es, but
by obviously reasons, a complete proof was somewhat voluminous to be given here. I have
decided to restrict attention to the ori gi n a l Conway-Norton conject u r e. Other aspects of
Moonshine are m entioned only as a matter of general cultur e on the subject.
I have written with special attention to the non initiated . In fact, I have includ ed so many
theory on t h e first chapters in order to give sufficient background to u n derstand al l materi al
in the later ones, and to make easier most of the ideas of Borcherds’ proof. For the sake of
background material, the proofs of theorems and propositions are omitted . The idea is to
give only the proof of the Conway-Norton conjecture, making the material accessible to the
level of a second year graduate student. Unfortunately, this work is by no m ean s complete.
There are several references included, with the purpose of encourage readers to specialize
topics of t heir interest. Perhaps, the recent [72] i s the most complete review of the Moon-
shine phenomenon. A good resume of [72] is the paper [71]. For the non-math em at i ci an
reader, [160] is good sou r ce, basically for its d i vulgation style. Main differences between
this work and [72, 71] is the devoted attention to the number th e or y of moonshine, and the
detailed exposition of the proof of the Conway-Norton conjecture.
In Chapter 1 I present an historical introduction, in order to facilitate a quickly access to
the main conjectur e. Several concepts appear, most of them explained more detailed in
subsequent chapters.
Chapters 2 and 3 give background material on algebra, particularly on Lie algebras. Chap-
ter 2 also includes other classic material on algebra, such as representation of finite group s,
modules, tensor products and some con st r u ct i on s . The main topic of Chapter 2 is to ex -
plain affine Li e algebras (Section 2.8), which appear later. Chapter 3 introduce important
concepts in Lie algebra theory, such as the Cartan subalgebras, the root systems and th e
xi
Weyl g r ou p . Main topics of this chapter are the Th eor em of Highest Weights (Section 3.7)
and the Weyl Character formula (Section 3.8).
Chapter 4 introduces the concept of vertex operator algebras, and relate it to th e repre-
sentation theory of Kac-Moody algebras, including a bit of lattice theory. At the fi nal of
Chapter 4, I give some interesting data relating Moonshine to the E
8
classical Lie algebra
and the Leech lattice.
Chapter 5 is devoted exclusively to a partial construction of the Moonshine module V
given
by Frenkel-Lepowsky-Meurman. It also in cl u des some aspects of other ‘unusual’ algebraic
structures, such as the Golay code C
24
, the Leech lattice Λ
24
and the Griess algebra B.
In Ch a p t er s 6 and Chapter 7, I explain in more detail the j-function and the original
Conway-Norton conjecture. This is the only chapter where the proofs are given (mainly
due to the fact that my principal area is number theory, and I have decided give more
attention to this). Chapter 7 includes important useful ma t er i al to understand the ideas of
Borcherds’ p r oof: congruence subgroups of SL
2
(R), replicable functions and the so called
replication formulae.
Finally, Chapter 8 is devoted to the proof of the Conway-Norton conjecture given by
Borcherds (h er e is where a lot of concepts i ntroduced in previous chapters main t h ei r
contribution). I introduce her e the Borcherds Lie algebras (Sec t io n 8.1) and subsequently
construct the Monster Lie algebra M (Section 8.3), and the denom i n at or formulas (Section
8.4). At the end, the proof is concluded by using properti es of the replication formulae.
Chapter 9 presents the actual status of Moonshine. It includes some generalizations of the
Moonshine conjectures, and explores deep connections between Moonshine, number theory,
geometry and physics. I have add ed a lot o f references to facilitate future work. Most of
the mat er i al of this chapter is t aken from [72].
The order of the chapters is by n o means strict. In fact, most of the material can be
omitted in a rapid lecture. For example, the advanced reader may omit Chapters 2 and 3.
Similarly, r ea d er not interested in number theory can omit Chapter 6. A quickly readin g
of Borcherds proof could be:
Chapter 1; Sections 4.1, 4.3, 4.5, 5.3; Sections 7.3, 7.4, 7.5; a nd Chapter 8
Obviously the fin al chapter is interesting by its own.
Acknowledgments
I would l i ke to express my deep-felt gratitude to Hossein Movasati, for his advice, e n cour-
agement, enduring patience and constant support. He was never ceasing in his beli ef in
me, always pr oviding clear explanations , expert in ind i cat e the simplest way to do th e job,
constantly driving me with energy and always, always, giving me his time, in spite of any-
thing else that was g oi n g on. I wish all students the honor and opportunity to experience
xii
his ab i l ity to perform at that job.
I also wish to thank the other members of my committee, Prof. Henrique Bursztyn from
Impa, Prof. Amilca r Pacheco from UFRJ, and Prof. Fed´eric Paugam from Universit´e
Pierre et Marie Curie - Paris VI. Their suggest i on s , comments and additio n al guidance
were invaluable to the completion of this work . Professor Don Zagier, from Max Plank
Institute of Mathematics, give some useful references about equation (7.16) and made im-
portant comments about future work, so I am i n debt with him. Also, Professor Michel
Waldschmidt from Universit´e Pierre et Marie Curie - Paris VI, made additional useful
comments. I also would tank to Pr ofes sor Alfredo Iusem. He was extrem el y helpful and
provided all the addition al support I needed when things were going not so well.
Additionally, I want to thank all my colleagues: Altemar, Antonio, Arturo, Brun o, Cad´u,
Carlos, Carol, Cristiane, Daniel, Diogo, Guillermo, Jos´e V´ıtor, Juan, L eon a r d o, Lucas,
Marco, Mauricio, Phillip, Rafael, Ricardo, S usana and Tiane. They were always expressing
best wishes and advices. In particular , I appreciate the conversations (and e-mails) with
Arturo, Karla and Mabel, who were always giving their moral support. I must also thank
to Juan, Juan Pablo, Cr i st i an, Mauricio, Pedro and Sergio by th ei r constant advices, an d
to Marlon for his motivation, constant interest and help in the progress of my work. Also,
I must thank Angel, by some suggestions on future work. Thanks to Alvaro for his com-
ments and references in the mathematical physics topics. I would like to thank also Mar´ıa
Jose for her encouragement and her valuable indications about how to use the PsTricks
package for the drawings. Mar´ıa Jose and Carol also g ive some helpful tips in how to
use Beamer. As a special note, I would like to express my gratitude to Al t em a r , who’s pa-
tient and h el p fu l indicat io n s conduced me to complete the proof of the replication formulae.
Finally, I must thank Impa professors and staff for all their hard work and dedication,
providing me the means to complete th i s degree and prepare for a career. I exp r ess my
gratitude to CNPq and Impa for their financial support.
A. G. Reyes Figueroa
Rio de Janeiro, maio de 2010.
xiii
Table of Contents
Page
Abstract vii
Resumo ix
Preface xi
List of Tables xix
List of Figures xxi
1 A histo rica l crash course on Monstrous Moonshine 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Finite s im ple groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 The discovery of M . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Modular functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Some on Kleinian groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.6 McKay’s obser vation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.7 The Monst r ou s Moonshine conjecture . . . . . . . . . . . . . . . . . . . . . 14
2 Some bas ics on Lie algebras 17
2.1 Representations of fin i t e groups . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Lie grou ps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3 Lie algeb r as . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.4 Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.5 Tensor pr oducts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.6 Module construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.7 Induced modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.8 Affine Li e algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3 The Root space decomposition 49
3.1 Representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.2 Complete Reducibility and Semi si m p l e algebras . . . . . . . . . . . . . . . 53
3.3 Cartan subalgebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.4 Roots and Roo t Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.5 The Weyl Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.6 Root Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.7 Integral and Dominant integral elements . . . . . . . . . . . . . . . . . . . 63
xv
3.8 The Weyl character formula . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.9 Representations of SU(3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4 Vertex algebras and Kac-Moody algebras 81
4.1 Vertex operator algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.2 The Virasor o algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.3 Kac-Moody algebras and t h ei r representation . . . . . . . . . . . . . . . . 87
4.4 Lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.5 The vertex algebra of an even latt ice . . . . . . . . . . . . . . . . . . . . . 96
4.6 Dynkin Diagrams of classic Lie algebras . . . . . . . . . . . . . . . . . . . . 98
4.7 Lie th eor y and Moonshine . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5 The Moonshine module 105
5.1 The Golay code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.2 The Leech lattice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.3 The Monst er vertex algebra V
and the Griess algebra B . . . . . . . . . . 111
5.4 The const r u ct i on of M . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.5 Improvements o n the Construction . . . . . . . . . . . . . . . . . . . . . . 123
6 Intermezzo: the j function 127
6.1 The modular group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.2 Modular forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
6.3 Lattice functions and modular functi ons . . . . . . . . . . . . . . . . . . . 131
6.4 The space of modular forms . . . . . . . . . . . . . . . . . . . . . . . . . . 135
6.5 The modular invariant j . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
6.6 Expansions at infi nity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
6.7 Theta fun c ti o n s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
7 Conway-Norton fundamental conjecture 149
7.1 Hecke operat o r s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
7.2 Action of T (n) on modul ar functions . . . . . . . . . . . . . . . . . . . . . 152
7.3 Congruence subgroups of SL
2
(R) . . . . . . . . . . . . . . . . . . . . . . . 153
7.4 Replicable functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
7.5 The repl i cat i on formulae . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
8 Borcherds’ proof of Conway-Norton conjecture 165
8.1 Borcherds L i e algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
8.2 The Borcherds character formula . . . . . . . . . . . . . . . . . . . . . . . 168
8.3 The mons t er Lie algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
8.3.1 The no-ghost theorem . . . . . . . . . . . . . . . . . . . . . . . . . 171
8.4 Denominator identities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
8.5 The twisted denominator identity . . . . . . . . . . . . . . . . . . . . . . . 178
xvi
8.6 Replication formulae again, a n d proof’ s end . . . . . . . . . . . . . . . . . 182
9 Concluding Remarks 185
9.1 Orbifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
9.2 Why the Monster? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
9.3 Other fi n i t e groups: Mini-Moonshine . . . . . . . . . . . . . . . . . . . . . 190
9.4 Modular Moonshine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
9.5 The geomet r y of Moonshine . . . . . . . . . . . . . . . . . . . . . . . . . . 193
9.6 Moonshine and physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
9.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
References 199
Appendix
A Proof of the Replication Formulae 207
A.1 Faber polynomial s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
A.2 Replicabl e functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
A.3 Deduction of the replication formulae . . . . . . . . . . . . . . . . . . . . . 210
Index 217
xvii
List of Tables
1.1 All 26 finite sporadic simple groups. . . . . . . . . . . . . . . . . . . . . . . 3
1.2 First d imensions of irreducible characters of M. . . . . . . . . . . . . . . . 13
2.1 Character table for the symmetric group S
3
. . . . . . . . . . . . . . . . . . 22
2.2 Connectedness properties of so m e classical matrix Lie groups. . . . . . . . 26
2.3 Simply connectedness properties of som e classical matrix Lie groups. . . . 27
2.4 Complexifications of some classica l real Lie algebras. . . . . . . . . . . . . 33
3.1 Semisimple properties of some classical complex Lie algebras. . . . . . . . . 56
3.2 Semisimple properties of some classical real Lie algebras. . . . . . . . . . . 57
3.3 Roots for the Lie algebra sl
3
(C). . . . . . . . . . . . . . . . . . . . . . . . . 72
3.4 Roots for sl
3
(C) in terms of the basis = {α
1
, α
2
}. . . . . . . . . . . . . . 72
3.5 Roots for sl
3
(C) in terms of the basis
= {α
1
, α
3
}. . . . . . . . . . . . . . 73
5.1 Third generation of the Happy Family. . . . . . . . . . . . . . . . . . . . . 118
6.1 First Ber n oulli numbers B
k
. . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6.2 Some val u e s ζ( 2 k) for t he Riemann ζ function. . . . . . . . . . . . . . . . . 141
6.3 Expansion of the first Eisenstein series E
k
. . . . . . . . . . . . . . . . . . . 143
8.1 Graded co m ponents of the Monster Lie algebra M. . . . . . . . . . . . . . 174
8.2 Z-graded components of the Monster Lie algebra M. . . . . . . . . . . . . 175
8.3 Cartan m a tr i x for the monster Lie algebra M. . . . . . . . . . . . . . . . . 177
xix
List of Figures
1.1 The grap h G
555
representing the Bim on st er . . . . . . . . . . . . . . . . . . 7
1.2 Fundamental domain D for the action of P SL
2
(Z) on H. . . . . . . . . . . 9
3.1 The root system A
2
for th e Lie algebra sl
3
(C). . . . . . . . . . . . . . . . . 73
3.2 Roots and domi n a nt integral elements for sl
3
(C). . . . . . . . . . . . . . . 74
3.3 Roots and domi n a nt integral elements for sl
3
(C) in the root basis. . . . . . 79
3.4 The Weyl group for sl
3
(C). . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.1 Dynkin diagrams for A
n
, B
n
, C
n
and D
n
. . . . . . . . . . . . . . . . . . . . 99
4.2 The except i on al Dynkin diagrams E
6
, E
7
, E
8
, F
4
and G
2
. . . . . . . . . . . 100
4.3 The affine Dynkin diagrams
ˆ
E
8
,
ˆ
F
4
and
ˆ
G
2
with l abels. . . . . . . . . . . . 103
6.1 Fundamental domain D and some of its images by S a n d T . . . . . . . . . 128
6.2 Integration on the bound a ry of region D. . . . . . . . . . . . . . . . . . . . 136
6.3 Integration on the bound a ry of region D. . . . . . . . . . . . . . . . . . . . 137
9.1 A ‘friend l y ’ process of compactifying and sewing a 4-punctured plane. . . . 188
xxi
Chapter 1
A historical crash course on
Monstrous Moonshine
1.1 Introduction
In 1978, John McKay made the ob ser vation that
196, 884 = 196, 883 + 1. (1.1)
Here, the number 196,884 refers to the first nontrivial coefficient of the automorphic fo r m
or nor m al i zed j-function
J(z) = q
1
+ 196, 884q + 21, 493, 7 60q
2
+ . . .
associated to the modular group SL
2
(Z), that appears in number theory. On the other
hand, the numb er 196,883 refers to the dimension of t h e minimal faithful representation of
the Mon st er group.
The central question of McKay’s equ at i on (1.1) is: What does the j-function (the left
side) have to do with the Monster finite group (the right side)? In general, specialists on
the subject agree that we still do not und er st and completely this phenomenon. Today we
say that there exists a vertex operator algebra, called the Moonshine module V
, which
interpolates between left and right sides of (1.1): its automorphisms group is the Monster
and its graded dimension is the normalized j-function. Since the original conjecture ap-
peared, the Moonshine quest i on triggered new developments in theory and p u sh e d ou t t h e
boundary of mathematical knowledge, bringing with itself m o r e questions than answers.
This thesis work co n si st s main l y in a modest review of the original proble m and the main
proof of Moonshine conjecture, a briefly look of how this theme is related to other areas of
mathematics, and what are the unsolved questions that remains until today.
The purpose of this chapter is to give a quickly un derstanding of the original Moonshine
conjecture. Because Moonshine is a subject that mixes several branches of mathematics
and physics, the reader is warned that this work is not self-contained an d requires extensive
mathematical baggage. Th er e are several references included, some of them by histor i cal
reasons, others only with th e intention of co n d u c e th e interested r ea d er t o specialize par-
ticular topics.
1
1.2 Finite simple groups
The discovery of the Monster was preceded by a long history of development of t h e t h eo r y
of finite groups. It has been of interest to ask for the classification of all fini t e groups,
yielding t he enumeration of all kind of symmetries. Remember that a group G is called
simple if it has no non trivial normal subgroup s. Basically, the importance of finite simple
groups is that by Jordan-H¨older theorem, th ey constitute the ‘primes’ of all fi n i t e groups,
that is, they are the constructive blocks of all finite grou p s. Thus, the core of the probl em
is the classification of finite simple groups.
This classification was announced in 1981 [77]. This result was the culmination of an intense
effort involving several hundred of mathematicians over a period of some 25 years. The
complete classification theorem r e qu i r es between 9,000 and 10,000 pages of mathematical
work, although a more streamlined proof is now on progress. By the end of the nineteenth
century, thanks to works of Jordan, Dickson, Chevalley and others, several families of sim-
ple groups were known. In addition, by 1861, Mathieu discovered five st r an g e finite simple
groups [141]. This groups were first called sporadic in the book of Burnside [23]. Mos t
of the fin i t e simple groups are now called of Lie type or Chevalley groups, and admit a
uniform construct i on in terms of simple Lie algebras, via a systematic treatment discovered
in [27 ].
The modern classification race started with the paper of Feit and Thompson in 1963 [61],
showing that every non-cyclic finite simple group has even order. This gave the great est
impetus to the effort to classify the finite simple gr o u p s . Thompso n followed this up by
another lengthy pa per in which he classified al l the minimal simple groups. This papers
made feasibl e the classificatio n project. In 1972, Gorenstein proposed a strategy for the
classification involving a det ai l ed 16 point programme. Progress was rapid from this stage,
with Gorenstein and Aschbacher playing leading roles in the project. It was finally sh own
in 1981 that every finite simple group is isomorp hic to one of the foll owing:
A cyc li c group Z
p
of pri m e order p.
An alternating group Alt
n
for n 5.
A si m p l e group of Lie type over a finite field, e. g ., P SL
n
(F
q
).
Some one of the 26 sporadic simple groups (see Tabl e 1.1).
There are 26 sporad i c simple groups, n ot definitively organized by a ny simple theme. Fur-
ther d et ai l s about these groups and their classification programme can be fou n d in [77].
The largest of the sporadic simple group s is called the Monster. We shall denote this
group by M. The Monster co ntains among its subquotients twenty of the sporadic simple
2
Group Order Name/Discoverer
M
11
2
4
·3
2
·5·11 Mathieu
M
12
2
6
·3
3
·5·11 Mathieu
M
22
2
7
·3
2
·5·7·11 Mathieu
M
23
2
7
·3
2
·5·7·11·23 Mathieu
M
24
2
10
·3
3
·5·7·11·23 Mathieu
J
1
2
3
·3·5·7·11·19 Janko
J
2
= HJ 2
7
·3
3
·5
2
·7 Hall-Janko
J
3
2
7
·3
5
·5·17·19 Higman-Janko-McKay
HS 2
9
·3
2
·5
3
·7·11 Higman-Sims
McL 2
7
·3
6
·5
3
·7·11 McLaughlin
Suz 2
13
·3
7
·5
2
·7·11·13 Suzuki
He 2
10
·3
3
·5
2
·7
3
·17 Held
Ru 2
14
·3
3
·5
3
·7·13·29 Rudvalis
Co
1
2
21
·3
9
·5
4
·7
2
·11·13·23 Conway
Co
2
2
18
·3
6
·5
3
·7·11·23 Conway
Co
3
2
10
·3
7
·5
3
·7·11·23 Conway
Fi
22
2
17
·3
9
·5
2
·7·11·23 Fisch er
Fi
23
2
18
·3
13
·5
2
·7·11·13·17·23 Fischer
Fi’
24
2
21
·3
16
·5
2
·7
3
·11·13·17·23·29 Fischer
O’N 2
9
·3
4
·5·7
3
·11·19·31 O’Nahn
Ly 2
8
·3
7
·5
6
·7·11·31·37·67 Lyons
J
4
2
21
·3
3
·5·7·11
3
·23·29·31·37·43 Janko
HN 2
14
·3
6
·5
6
·7·11·19 Harada-Norton
Th 2
15
·3
10
·5
3
·7
2
·13·19·31 Thompson
B 2
41
·3
13
·5
6
·7
2
·11·13·17·19·23·31·47 Baby Monst e r
M 2
46
·3
20
·5
9
·7
6
·11
2
·13
3
·17·19·23·29·31·41·47·59·71 The Mons t er
Table 1.1: All 26 finite sporadic simple groups.
3
groups, except for J
3
, Ru, O’N, Ly and J
4
. These twenty constitute the Happy Family,
and they occur naturally in three generations: the family of Mathieu groups, the family
of Conway groups, and the so called Third Generation. Each one of these is related to
an algebraic structure, in particular, the last one is associated to the Monster. The other
five sporadic groups a re called the pariah. Fischer and Griess i n dependently predicted in
1973 the existence and properties of M as t he lar g est of the sporadic groups. It has 194
conjugacy classes and irreducible characters. The character table of M was determined
by F ischer, Livingstone and Thorne in 1978 [62] and can be found in the Atlas of Finite
Groups [ 36]. The degree of the smallest i r r ed u ci ble characters of M are:
d
0
= 1, d
1
= 196883, d
2
= 2129687 6, d
3
= 8426093 26, . . .
In part i cu l a r, the minimal faithful representation of the Monster would have dimension
d
1
= 196, 883. (1.2)
A lot of information about M can be fou nd in [36].
As a remarkable fact, we mention that the greatest of the Mathieu sporadic groups, M
24
was
constructed by Mathieu as the grou p of symm et r i es of th e Golay code C
24
(a 12-dimensional
subspace of a 24-dimensional vector space over F
2
). Similarly, Conway constructed his 3
sporadic groups in 1968, and the greater of them, Co
3
, is realized as the automorphism
group of the Leech lattice Λ
24
(a 24- d i m en s io n al subspace of the 26-dimensional Lorentzian
vector space R
25,1
, wi t h norm ||x||
2
= x
2
1
+. . .+x
2
25
x
2
26
). Norton observed that the minimal
representation of M would have the structure of a rea l commutative non-associative algebra
with an associative form . Finally, the Monster group was first constructed by Griess in
1980 as a group of automorphisms of a commutative non-associative algebra of dim en si on
196,883 over Q with an associative form [80]. Griess’ construction has been simplified i n
works of Tits [174, 175, 176] and Conway [34, 35]; and Tits has in fact proved that M is
the full group of automorphisms of the Griess algebra B. Unfortunately, even this fine
version of Griess algebra does not appear as elegant as Golay code C
24
or Leech lattice Λ
24
,
which have simple characterizations. But, there was some hints that the Monster could be
associated with an elegant canonical str u ct ure. We shall discuss this structure in Chapter
5.
1.3 The discovery of M
We have already mention that Feit and Thompson’s p aper [61] was the starti n g point on
the syst em at i c search of sporadic groups, or sometimes called symmetry atoms, leading
to a whole project known as ‘t h e Classifica ti o n ’ [77]. The idea: to compil ing a list of all
finite symmetry atoms, and showing that the l i st was complete (like a periodic table of
symmetries). Recall that there were already know some of this exceptional the sporadic
4
simple groups. Conway constructed his three sporadic groups Co
1
, Co
2
and Co
3
in 1968,
by loo ki n g various mirror symmetries occurri n g within Leech lattice Λ
24
, and this lattice
had yielded a tot a l of 12 sporadic groups, nine of which had already appeared elsewhere.
These 12 groups, along with Janko’s groups J
1
and J
3
—which had nothing to do with the
Leech lattice— brought the total number of exceptions to 14. By the end of 1972 there
were six more: thr ee found by Fischer; one by Dieter Held; one by Richard Lyons; and one
by Ar u n as Rudvalis. Both Held and Lyons used the cross-secti o n meth od. Rudvalis used
permutations. Th e tot al number o f excep t i on s was now 20. With so much activity and so
much information coming in, it made excellent sense to collect it al l, correct errors, and
present it in a form that was easy to read and readily available. This was ‘the Atlas’ [36]
(exceptions to periodic table).
In fact, Fischer constructed h i s sporadic groups Fi
22
, Fi
23
and Fi’
24
by using the method of
transpositions (adding some new symmetries to the Mathieu groups M
22
, M
23
, M
24
respec-
tively), leading him to the discovery of the Baby Monster B in 1973. Another new group
of symmetries was found by Michael O’Nan the same year. This brought the total to 22,
though not all these groups were yet known to exist. After that, it was expected that no
other new sporadics would appear. Fortunately, Fischer and Griess noticed independently
that this Fischer’s huge new group could appear as a cross-sectio n in a larger group. If a
group emerged from the cross-section method —like B then a great deal of information
needed to be calculated before a construction was possible. Most of this data was encoded
in the form of a square array of numbers called a character table (a character table is a
square array of number s that that express the fund am entally ways the group can operate in
multidimensional space) (see Section 2.1). In fact, a finite group can have a huge character
table, but sporadic groups usual ly have a low number o f rows/columns in their character
tables.
In late 1973, knowing only that the Monster had two cross-section s, using a procedure
called Thompson order formula, the size of the whole thing was within reach. Thompson’s
technique needed detailed computations on how the two cross-sections could intersect.
Fischer used it to show that the size could not be greater than a certain number. Further
calculations made by Conway, led Fischer to four n ew sporadic groups: B, Th, HN, and
M, the last one with size
2
46
· 3
20
· 5
9
· 7
6
· 11
2
· 13
3
· 17 · 19 · 23 · 29 · 31 · 41 · 47 · 59 · 71
Having the size of the Monster was essential before working out it s character table. Nor-
ton and others established that the Monster would have a representation of dimen si on
196, 883 = 47 · 59 · 71. Only with this information, Fischer, Livigstone and Thorne com-
puted th e entire character table of M, in 1974; and then that of the Baby Monster. Final l y,
Sims and Leon co n cl u d e the construct i o n of B in 1977, by creating it on a computer as a
group of permutations on 13,571,955,000 mirrors.
5
After Sims and Leon constr u ct i on , it was natural to ask whether the Monster could be con-
structed in a similar way. Un fo rt unately this seemed to be out of sight, so an alternative
method was needed. Perhaps one could use multidimensional space. Similar methods had
been applied to other exceptional groups, such as Janko’s first group J
1
. But, as J
1
needed
seven dim en s io n s, th e Monster needed 196,883, which means that a single operation in the
Monster would appear as a matrix with nearly 196,883 rows and as many columns.
Suddenly in 1980, Bob Griess announced a cons t ruction. Alread y, Nort on had figure out
that the Monster must preserve an algebra structure in 196,884 dimensions. This structure
would allow any two points to be multiplied together to give a thir d point. Griess’ first
task was to construct a suitable multiplication, solving some problems about signs on th e
algebra structure, by tracking them back into the grou p . This group Gr ie ss i s r efer r i n g to
here is a huge subgroup of the Monster, ext ended by a factor of over 32 million. This group
needs 96,308 dimensions, and is one of the two cross-sections of the Monster (the other
one involves the Baby Monster). Fischer used it earli er in helping to build the character
table of the Monster, and Griess now us ed it to help to const r uct the Monster in 196,884
dimensions. He knew that the action of this huge subgrou p must split the space into three
subspaces of the following dimensions (see (5.19) and (5.20)):
98, 304 + 300 + 98, 280 = 196, 884
The first number is 98, 304 = 2
12
× 24. This is the space needed for the cross-section
mentioned above. The second number is 300 = 24 + 23 + 22 + ... + 3 + 2 + 1. This c om es
from a triangular arrangement of numbers with 24 in the first row, 23 in the second row, 22
in the th ir d row, and so on. The third number i s 98, 280 =
1
2
(196, 560). This comes from
the Leech lat t i ce (see Chapt er 5), where there are 196,560 points closest to a given point,
and they come in 98,280 diametrically opposite pairs. Each pair yields an axis through the
given point, and in the 196,884-dimensional space these axes become i n d ependent of one
another. Griess [80] finally sort out the sign problem for the multiplicat i on , an d proved
that the group of symmetries contains the Monster, by addi n g on e ext r a permutation to
create a larger group.
Two other mathematicians got deeply involved in looking at the Gr i ess construction. One
was Jacques Tits, who found a way of avoidi n g the sign problems [ 17 5 ] . Tits also found a
number of other impr ovements and simp l i fi ed Griess’s construction. The other person who
took a detailed interest in the Griess construction was Conway, giving a construction of his
own, and a very elegant proof of finiteness [34, 35]. In a sense, Tits avoided all calculations
in the Monster, while Conway made easier to per for m such calculations. Conway’s con-
struction is similar to Griess’s in the sense that they both use the same large cross -s ect io n of
the Monster to get started. Thi s splits the 196,884-dimensional sp ace i nto three subspaces.
A more detailed (with a lot of hist o r ica l notes) on the construction of the sporadic grou p s
can be found in [160]. The interested reader may also see [72] to have a complete context
on the Moonshin e phenomenon. The compl et e construction of the Monster group is too
6
voluminous to be explained here . We will present a sketched part i al construction of M in
Section 5.4.
For now, we describe a remarkably simple representation of M. As with any noncy cli c finite
simple group, it is generated by its involutions (i. e., elements of order 2) and so will be
an image of a Coxeter group. Let G
pqr
, p q r 2, be the grap h consisting of three
strands of lengths p + 1, q + 1, r + 1, sharing a common endpoint. Label the p + q + r + 1
nodes as in Figure 1.1. Given any graph G
pqr
, d efine Y
pqr
to be the group consisting of a
generator for each node, obeying the usu al Coxeter group relations (i. e., all generators are
involutions, and the product gg
of two genera t or s has order 2 or 3, depending on whether
or not the two nodes are adjacent), together wi t h one more relation:
(ab
1
b
2
ac
1
c
2
ad
1
d
2
)
10
= 1.
The groups Y
pqr
, for p 5, ar e all i d entified (see [102]). Conway conjectured and , building
on work by Ivanov [103], Norton proved [156] that Y
555
=
Y
444
is the Bimonster, the
wreathed-square M Z
2
(or (M ×M).2 in Conway’s notation), that is, a group with M ×M
as normal sub g ro u p and Z
2
as quotient, with order 2|M|
2
. A closely r e la t ed presentation
of the Bimonster has 26 involuti on s as genera t or s and h as relations given by the incidence
graph of the projective plane of order 3; the Monster itself arises from 21 involutions and
the affine plane of order 3. Likewise, Y
553
=
Y
443
=
M × Z
2
. Other sporadic groups arise
in e. g., Y
533
=
Y
433
=
B is the Baby Monster; Y
552
=
Y
442
is the Fischer grou p Fi’
24
; and
Y
532
=
Y
432
is the Fischer group Fi
23
. The Coxeter group s associated to the other g r ap hs
G
pqr
, p 5 are all finite groups of hyperbolic reflections in R
17,1
and contain copies of the
Weyl group E
8
, giving a rich underlying geometry.
b
1
b
2
b
3
b
4
b
5
c
1
c
2
c
3
c
4
c
5
d
1
d
2
d
3
d
4
d
5
a
Figure 1.1: The graph G
555
representing the Bim o n st er .
7
1.4 Mod ul ar fu ncti on s
Although the Monster group M was discovered within the context of finite simpl e grou ps,
hints later began t o emerge that it might be strongly related to other bran ches of mathe-
matics. One of these is the theory of mod u l ar function s and modular forms.
Consider the action of the group SL
2
(R) on the upper half-plane H. Let
SL
2
(R) =
a b
c d
GL
2
(R) : ad bc = 1
.
SL
2
(R) acts on H = {z C : Im(z) > 0} by
a b
c d
· z =
az + b
cz + d
,
(note that Im(z) > 0 Im
az+b
cz+d
= Im(z) > 0). In particular, the subgroup SL
2
(Z) acts
on H discontinu a ll y.
Since
1 0
0 1
· z =
z
1
= z, we see that P SL
2
(Z) = SL
2
(Z)/I} acts on H. We call
Γ = P SL
2
(Z) the modular group and we denote by H/SL
2
(Z) the set of orbits. Since
T =
1 1
0 1
SL
2
(Z) an d
1 1
0 1
· z = z + 1,
then z and z + 1 are in the same orbit. Thus, each orbi t intersects
{z H : 1/2 Re(z) 1/2}.
Similarly,
S =
0 1
1 0
SL
2
(Z) an d
0 1
1 0
· σ =
1
σ
,
so the elements σ and 1 ar e in the same orbit. Thus, each orbit intersects
{σ H : |σ| 1}.
In particular, if |σ| = 1, then (
0 1
1 0
) ·σ =
1
σ
= ¯σ. In fact, Theorem 6.1.2 shows that we
obtain a fundam ental domain D for the action of SL
2
(Z) on H taking the region
D = {z H : 1/2 Re(z) 1/2, |z| 1},
and identifying z with z + 1, for Re(z) = 1/2; and σ with ¯σ =
1
σ
, for |σ| = 1 (see
Figure 1. 2 ) .
8
Figure 1.2: Fund am ental domain D for the action of P SL
2
(Z) on H.
With these identifications we obtain a set D intersecting each orbit just at one poi nt. In
fact, the canonical map D H/SL
2
(Z) is surjective and its restricti on to the interior of
D is inject i ve. We can prove also that the modular group Γ is generated by S and T (see
Theorem 6.1.3). The set H/SL
2
(Z) of orbi t s has the structure of a Riemann su r fa ce with
one point removed. This is a Riemann su r face of genus 0. When we rem ove one point of it,
we obtain a set which can be identified with C. Thus, we have an isomor p h i sm of Riemann
surfaces
H/SL
2
(Z) C.
This ma p can be extended to an iso m or p hism between compact Riemann surfaces
H/SL
2
(Z) {i∞} C = C {∞}
=
CP
1
,
which maps i → . Such an i som o r p h i sm is not unique. However, when we operate
with Riemann surfaces of genus 0, we can prove that if j is just one of these isomorphisms,
then any other is of the form a(j + b), where a, b are constants and a = 0.
Let j : H/SL
2
(Z)
C be one of such isomorphisms. Then, j defines a map j : H C that
is constant on orbits. Since z and z +1 l ies on the same orbit, then we have j(z) = j(z + 1).
Thus, j is periodic (of period 1). This implies that j has a Fourier expansion
j(z) =
nZ
c
n
e
2πinz
.
9
Writing q = e
2πiz
, then we have
j(z) =
nZ
c
n
q
n
.
Let k be an integer. We say tha t a mer om o r p h i c function f : H C is a modular function
of weight 2k i f
1
f
az + b
cz + d
= (cz + d)
2k
f(z), for all
a b
c d
SL
2
(Z).
If f is holomorp h i c everywhere —including at infinity—, we say t hat f is a modular form.
We give some examples (see Section 6. 3 ) :
Example 1.4.1. Let σ
3
(n) =
d|n
d
3
. We define the Eisenstein series
E
4
(z) = 1 + 240
n1
σ
3
(n)q
n
= 1 + 240q + 2160q
2
+ . . .
Thus, E
4
(z) is a modular form of weight 4.
Example 1.4.2. The Dedekind’s function
η(z) = q
d1
(1 q
n
)
24
= q 24q
2
+ 252q
3
. . .
is a modular form of weight 12.
We now define j : H C by
j(z) =
(E
4
(z))
3
η(z)
.
This is a modular form of weight 0, so it is constant o n orbits of P SL
2
(Z) on H. In fact,
j(z) = q
1
+ 744 + 196, 884q + 21, 493, 760q
2
+ . . . (1.3)
Such map s are usually called fundamental functions or Hauptmodul.
Observe that j(z) i s holomorphic on H and t h a t j(z) has a simple pole at z = i (i . e., at
q = 0). Then, j gives an holomorphic isomorphism between H/SL
2
(Z) and C that can be
extended to a meromorphic isomorphism of compact Rieman n surfaces
H/SL
2
(Z) C,
where
H = H Q {i∞}.
The SL
2
(Z)-orbits of Q {i∞} are called cusps and thei r role is to fill the punctures
1
Some authors say that f is of weight k.
10
of H/SL
2
(Z), compactifying the surface, as there ar e much fewer meromorphic functions
on compact surfaces than on noncompact ones. Since any other isomorphism is of the
form a(j(z) + b), with a, b constants and a = 0, in particular there is just one of such
isomorphisms with leading coefficient c
1
= 1 and constant term c
0
= 0:
J(z) = j(z) 744 = q
1
+ 196, 884q + 21, 493, 760q
2
+ . . . (1.4)
We shall call this function J(z) the canonical isomorphism or the normalized Hauptmodul
of the modular group.
The expansion coefficients of J(z), which are all positive integers (except for the constant
term), mig ht appear unattractive. As we shall see, i t took many years and an accident
before their mea n i n g was finally found.
1.5 Some on Kleinian groups
Even before the discovery of the modular invariant j(z) was made, it was observed th a t
certain characteristics of ell i p t i c functions with per i ods ω
1
, ω
2
were invariant only under a
certain sub gr o u p Γ(2) of Γ = SL
2
(Z). This and other facts led Klein to th e creation of
the theory of congruence subgroups [115]. He introduced a class of principal congruence
subgroups
Γ(n) =
a b
c d
Γ :
a b
c d
1 0
0 1
(mod n)
, for n > 0,
and notions of congruence subgroups Γ
of level n. An example of a congruence subgroup
of level n is the cla ss
Γ
0
(n) =
a b
c d
Γ : c 0 (mod n)
.
At the same time, Poincar´e, influenced by a paper of Fuchs [69], launched a p r ogr a m to
study discrete subgroup of P SL
2
(R) and their corresponding automorphic functions (anal-
ogous t o modul ar functions) , and this laid to the theory of Fuchsian groups.
One of the basics results of this theory is th at for any Kleinian group Γ
, a suitable com-
pactification of H/Γ
has the structure of a compact Riemann surface. In the special case
when the genus of H/Γ
is zero the theory of automorphic functions becomes specially
simple: the field of automorphic functions is generated by only one funct i on J
Γ
(z), called
the Hauptmodul of Γ
. In the par t i cu l ar case of the modular group Γ = P SL
2
(Z), the sur-
face —the Riemann sph e r e— has genus zero, and the Hauptmodul of Γ is just J(z) in (1.4).
Fricke [68] investigated the surfaces associat ed with Γ
0
(n). In particular, congruence sub-
groups Γ
0
(p), for p prime, provide examples of genus zero surfa ces if and only if p 1 | 24.
11
One can obtain more examples adjoining to Γ
0
(n) the Fricke involuti on w
n
(z) = 1/nz,
which of course can be realized as an element of P SL
2
(R). Th a t is
Γ
0
(n)
+
=
Γ
0
(n),
1
n
0 1
n 0
.
The normalizer of Γ
0
(n) in P SL
2
(R) was fully described by Atkin and Lehner [3]. When
n is a prime p, it is just the group Γ
0
(p)
+
generated by Γ
0
(p) and the Fricke involut i on
w
p
. Ogg [158] completed Fricke’s proof [68] that for a prime p, Γ
0
(p)
+
has the genus zero
property if and only if
p = 2, 3, 5, 7, 11, 13, 17, 19, 23, 29, 31, 41, 47, 59, 71 . (1.5)
This strange set of primes could be passed to histor y as another mathematical fact without
any special sign i fi c an ce. It hap pened however that Ogg attended a talk of Tits mentioning a
certain sporadic sim p l e group predicted —bu t not proved— to exi st by Fischer and Griess,
of order
|M| = 2
46
· 3
20
· 5
9
· 7
6
· 11
2
· 13
3
· 17 · 19 · 23 · 29 · 31 · 41 · 47 · 59 · 71,
and noticed that the primes app ea r i n g in l i st (1.5) are exactly the prime divisors of the
order of the Mon st er group M. What a coincidence! It was not realized at th at time, at
1975, t h a t this coincidence was the tip of an iceberg.
1.6 McKay’s observation
Hints that the Mon s te r might in fact be associated with an elegant structure had appeared
before Griess announced h is construction. We had mention already that Ogg, who was
working on the field of modular fun ct i ons, came with some coincidences. From the other
side, McKay, who was working in finite group theory, noticed an o t h er relation between t h e
Monster and modular functions: the near coincidence of the minimal possible representation
of M (1.2) and the first non triv i al coefficient c
1
of the Hauptmodul J(z) in (1.4).
196, 884 = 196, 883 + 1.
Soon, McKay and Thompson found similar relations [172] includ ing anot h er di m en si ons of
irreducible representations of M, for example:
196, 884 = 196, 883 + 1,
21, 493, 760 = 21, 296, 876 + 196, 883 + 1,
864, 299, 970 = 842, 609, 326 + 21, 296, 876 + 2 · 196, 883 + 2 · 1, (1.6)
If we interpret d
0
= 1 as the dimension of the trivi al representation of M, then we have
c
1
= d
0
+ d
1
.
12
where d
1
is the dimension of the next irreducible representation of the Monster. In fact, if
we comput e some more few coefficients of J(z)
J(z) = q
1
+ 196, 884q + 21, 493, 760q
2
+ 864, 299, 970q
3
+
+ 20, 245, 856, 256q
4
+ 333, 202, 640, 600q
5
+ 4, 252, 023, 300, 096q
6
+ . . .
and som e more few dimensions of irreducible characters of M
d
0
= 1 d
1
= 196, 883
d
2
= 21, 296, 876 d
3
= 842, 609, 326
d
4
= 18, 538, 750, 076 d
5
= 19, 360, 062, 527
d
6
= 293, 553, 734, 29 8 . . .
Table 1.2: First dimensions of irreducible characters of M.
then eq u at i ons (1.6) are extended:
c
1
= d
0
+ d
1
,
c
2
= d
0
+ d
1
+ d
2
,
c
3
= 2d
0
+ 2d
1
+ d
2
+ d
3
,
c
4
= 2d
0
+ 3d
1
+ 2d
2
+ d
3
+ d
5
,
c
5
= 4d
0
+ 5d
1
+ 3d
2
+ 2d
3
+ d
4
+ d
5
+ d
6
, (1.7)
. . .
and further relations of this sort. Based on these observat io n s, McKay and Thompso n
conjectured the exist en ce of a natural infinite-dimensional representation of the Monster
V = V
1
V
1
V
2
V
3
. . . ,
such that dim V
n
= c
n
, for n = 1 , 1, 2, 3, . . .. That is, this suggests that there is a graded
vector space acted on by the Monster, and that our Hauptmodul J(z) is in fact what we
call t h e graded dimension of V :
J(z) =
n≥−1
c
n
q
n
= q
1
+
n1
(dim V
n
)q
n
. (1.8)
McKay, also gave a r el at i o n between the j-function and the classical Lie algebra E
8
(see
Section 4.7) . A more elementary observation concerns the Leech lattice Λ
24
. The Leech
lattice is a particularly special one in 24 dimensions. It has some special features. The
book by Conway and Sloane [39] includes a lot o f interesting details relating the Leech
lattice. Another useful reference is [86] . For example, 196,560 is the number of vectors in
13
the Leech lat ti ce with (squa r ed ) norm equal to 4, an d note that this number is also close
to 196,884: In fact, the first coefficients of the Hauptmodul J (1.4) are related to the Leech
lattice
196, 884 = 196, 560 + 324 · 1,
21, 493, 760 = 16, 773, 120 + 24 · 19 6, 560 + 3, 200 · 1,
864, 299, 970 = 398, 034, 000 + 24 · 16, 773, 120 + 324 · 196, 560 + 25, 650 · 1,
where 16,773,120 and 398,034,000 are the numbers of 6-norm and 8-n or m vectors in Λ
24
.
This may not seem as convincing as (1.6), but the same equations hold for any of the
24-dimensional even self-dual lattices, apart from an extra term on the right sides corre-
sponding to 2-norm vectors (there are none of th e se in the Leech).
What conceptually does the Monster, the Leech lat t ice have to do with the j-function?
Is there a common theory explaining this numerology? The answer is yes. In fact, as we
shall see, there is a relation between E
8
to the j-function. In the late 1960’s, Victor Kac
[110] and Robert Moody [1 49 ] independently defined a new class of infinite-dimensional
Lie algebras. A Lie algebra is a vector space with a bilin ear vector-valued product that is
both anti-commutative and anti-associative (Chapter 2). The familiar vector-product u ·v
in three dimensions d efines a Li e algebra, called sl
2
(R), and in fact this algebra generates
all Ka c-M oody algebras. Within a decade it was realised that the graded dim en si o n s of
representations of the affine Kac-Moody algebras are (vector-valued) modular functions for
SL
2
(Z) (se e Theorem 3.2.3 in [72]).
1.7 The Monstrous Moonshine conjecture
Recall the strong connections of previous section, that suggested the existence of the graded
algebra V . Thompson [171] also proposed considering, for any element g M, the modular
properties of t h e series
T
g
(z) =
nZ
tr(g|V
n
)q
n
= q
1
+ tr(g|V
1
)q + tr(g|V
2
)q
2
+ . . . , (1.9)
where q = e
2πiz
and V
n
is t h e n-th graded component of V . This graded trace above is
called the McKay-Thompson series of g, and generalizes our Haup t m odul on (1.4). For
example, if g = 1 (the identity element of M), then T
g
(z) is just J(z).
Working with data fr om the table character of the Monster, remarkable numerology con-
cerning these graded t r aces was collected in the paper Monstrous Moonshine [38 ]. It has
been pointed by Conway that he found some of these interesting series in the classical book
of Jacobi [106].
14
Influenced by Ogg ’s observat i on , Thompson, Conway and Norton realized that all the se-
ries they were discovering (proceedin g exper i m entally by the first few coefficients) were
normalized gener at o rs of genus zero function fields arising from certain discrete subgroups
of P SL
2
(R), th at is the McKay-Thompson series they were discoverin g beh ave as ‘mini
j-functions’, for certain other subgroups of P SL
2
(R). They were led to conjecture that
there exists a graded representation V of t h e Monster —in fact a double- gr aded space—
with all the functions T
g
(z) having this genus zero property. Knowing the function s T
g
(z)
determines the M-m odule V uniquely, and the qu est i on was whether it existed, given the
list of proposed functions T
g
(z), for each of the 194 conjugacy classes of M. Such graded
module was subsequently discovered by Frenkel, Lepowsky and Meurman [66], [67]. It is
called the monster vertex algebra or the moonshine module V
. We shall study it in Chap-
ter 5.
Instead of consider only the modular subgroup P SL
2
(Z) of PSL
2
(R), we shall consid er
other subgroups. A subgroup G of P SL
2
(R) is called commensurable with P SL
2
(Z) if:
[P SL
2
(Z) : P SL
2
(Z) G] is finite;
[G : P SL
2
(Z) G] is finite.
We may consider the action of such a group on the up per half-plane H. We know that if G
is a subgroup of P SL
2
(R) commensu r able with P SL
2
(Z), then the set of orbits H/G has
the structure of a compact Riemann surfa ce, with finitely many points removed. We write
H/G
H/G,
where
H/G is a compact Riemann surface. Let H/G a compact Riemann surface of genus g.
Then,
H/G is homeomorphic to a g-torus. In the case g = 0, H/G is homeomorphic to the
Riemann sph er e CP
1
. A subgroup G is called of genus 0, if the Riemann sur face
H/G has
genus 0. We have already pointed in Section 1.5 that under these particular circumstances,
the field of automorphic functions of
H/G CP
1
is generated by just one element, and
we can take this element as the unique isomorphism of Riemann surfaces
H/G CP
1
with leading coefficient 1 and con s t ant term 0. We denote this by J
G
(z), and is called the
canonical isomorphism or normalized Hauptmodul of G. So J
G
, plays exactly the same role
for G that the J-function plays for SL
2
(Z). For example, i n t h e case of Γ
0
(2), Γ
0
(13) and
Γ
0
(25) —ar e all genus 0 subgroups of P SL
2
(R)—, we obtain Hauptmoduls
J
Γ
0
(2)
(z) = q
1
+ 276q 2048q
2
+ 11202q
3
49152q
4
+ 184024q
5
+ . . . , (1.10)
J
Γ
0
(13)
(z) = q
1
q + 2q
2
+ q
3
+ 2q
4
2q
5
2q
7
2q
8
+ q
9
+ . . . , (1.11)
J
Γ
0
(25)
(z) = q
1
q + q
4
+ q
6
q
11
q
14
+ q
21
+ q
24
q
26
+ . . . (1.12)
We now state formally an in i t ia l form of the Conway-Norton conjecture, or Moonshine
conjecture.
15
Conjecture 1.7.1 ( Moonshine Conjecture). (Conway-Norton, 1979).
For each g M, the McKay-Thompson series T
g
(z) is the normalized Hauptmodul J
G
(z) :
H/G CP
1
for some subgroup G of SL
2
(R) commensurable with P SL
2
(Z).
The first major step in the proof of Monstrous Moonshine was accomplished in the mid
1980’s with the construction by Frenkel-Lepowsky-Meurman [67 ] of th e Moo n sh i n e module
V
, and its interpretation by Richard Borcherds [8] as a vertex operator algebra ( VOA). In
1992, Borcherds [12] com p l et ed the proof of the original Monstrous Moonshine conjectures
by showing that the graded characters T
g
of V
are indeed the Hauptmodu l s identified by
Conway and Norton, and hence that V
is indeed the desired representation V
of M conjec-
tured by McKay and Thompson. The algebraic structure appearing in moonshine typically
arises as the symmetry group of the associated vertex operator algebra, for exam p l e, that
of V
is the Monster M. By Zhu’s Theorem ( Th eor e m 9.2.1), the modular functions appear
as graded dimensions of the (possibly twisted) modules of a vertex operator alg eb r a. In
particular, the answer that this framework provides for what M, E
8
and the Leech latt ic e
have to do with the j-function i s that they each correspond to a vertex operator algebra
with a single simple module; their relation to j is then an immediate corollary to the much
more gen e ra l Zhu’s Theorem.
Moonshine is also profoundly connected with geometry a n d physics (namely conformal field
theory and string theory). Stri n g theory proposes that the el em entary part i cle s (electr ons,
photons, quarks, etc.) are vibrational mod es on a strin g of length about 10
33
cm. These
strings can interact only by splitting apart o r joining together as th e y evolve through time,
these (classical) strings will tr a ce out a surface called the world-sheet. Quantum field theo r y
tells us that the quant u m quantities of interest (amplitudes) can be perturb a t ively com-
puted as weighted averages taken over spaces of these world-sheets. Conformally equivalent
world-sheets should be identified, so we are led t o i nterpret amplitudes as certain integrals
over moduli spaces of surfaces. This approach to stri n g theory leads to a conformally in-
variant quantum field theory on two-dimensional space-time, called conformal field theory
(CFT). The various modular forms and functions arising in Moonshine appear as inte-
grands in some of these genus 1 surfaces appearing i n these conformal theories. We shall
explain a bit of these relation between Moonshine and physics in Ch apter 9. Thus, the
actual importance of the Moonshine pheno m en on: It proposes a conexion between four
branches of math em at i cs, namely algebraic structures, modular structures, geometry and
mathematical physics.
16
Chapter 2
Some basics on Lie algebras
In this chapter we introdu ce Lie algebras termed affine, of which Virasoro algebra plays
a central role throughou t this work. We also present standard constructions of important
classes of modules and algebras and we discuss the concept of graded dimension. We
have introduced a number of elementary concep ts i n first sections to provide sufficient
background for understanding the main concepts. For an extensive exposit i on of b asi c
algebra, the reader may refer to [107, 108] and [127]. A detailed reference on Lie groups
and Li e algebras can be found in [100] an d [179], or the more recent [91 ] .
2.1 Representations of finite groups
Let V be a vector space over the field C of complex numbers and let GL(V ) be the group
of isomorphisms of V onto itself. An element a GL(V ) is, by definition, a linear mapping
of V into V which has an inverse a
1
; this inverse is linear . When V has a finite basis (of
n elements), each linear map a : V V is defined by a square matrix [a
ij
] of order n. Th e
coefficients a
ij
are complex numbers, and we usually write in this case GL(V ) as GL
n
(C).
Suppose now G is a finite group, with i d e ntity element 1 and with composition (s, t) → st.
A linear representation of G in V is a homomorph i sm π from the group G into the grou p
GL(V ). In other words, we associate with each element s G an element π(s) of GL(V )
in such a way that we have the equality
π(st) = π(s) · π(t), for s, t G.
Observe that the preceding formula implies the following: π(1) = 1, and π(s
1
) = π(s)
1
.
When π is given, we say that V i s a representation space of G (or even simply, a repre-
sentation of G). We restrict ourselves to the case where V has finite dimension. When
dim V = n, we say that a representation π : G GL(V ) has degree n.
Let π : G GL(V ) be a linear representation and let W be a vector subspace of V . We
say that W is stable under the action of G (we say also invariant), if π(s)W W , for
all s G. Given a linear representation π : G GL(V ), we say that it is irreducib le or
simple if V is not 0 and if V has no nontrivial stable subspa ces und er G. This condition is
equivalent to saying V is not the direct sum of two representations (except for the trivial
V = 0 V ). We have the following result [166, p . 7]
17
Theorem 2.1.1. Every representation is a direct sum of ir reducible representations.
Let V be a vector space having a basis e
1
, . . . , e
n
, and let a be a linear map of V into itself,
with m at r i x [a
ij
]. By the trace of a we mean the sca la r
tr a =
1in
a
ii
.
It is the sum of the eigenvalues of a (counted with their multiplicities), and does not depend
on the choice of basis {e
i
}. Now let π : G GL(V ) be a linear representation of a finite
group G in the vector space V . For each s G, we put
χ
π
(s) = tr(π(s)).
The complex valued function χ
π
on G thus obtained is called the character of the represen-
tation π; the importance of this functio n comes primarily from the fact that it characterizes
the rep r ese ntation π.
Proposition 2.1.2. If χ i s the character of a representation π of degree n, we have:
1. χ(1) = n.
2. χ(s
1
) = χ(s), for all s G
3. χ(tst
1
) = χ(s), for all s, t G.
Proof. We have π( 1) = I, and tr(I) = n since V has dimension n; hence (1). For (2) we
observe th at π(s) h as finite order; consequently the same is true of its eigenvalues λ
1
, . . . , λ
n
and so these have absolute value equal to 1 (this is also a consequence of the fact that π
can be defin e d by a unitary matrix. Thus
χ(s) = tr(π(s)) =
λ
i
=
λ
1
i
= tr(π(s)
1
) = tr
π
1
(s) = tr
π
(s
1
).
Formula (3) can also be written χ(vu) = χ(uv), putting u = ts, v = t
1
; hence it Follows
from the well known formula tr ( ab) = tr(ba), valid for two arbitrary linear mappings a and
b of V into itself.
A function f on G satisfying identity ( 3 ) above, is called a class function. We shall see later
that each class function is a linear combination of characters. We will use the followi n g
result [166, p.11 ]
Proposition 2.1.3. Let π
1
: G GL(V
1
) and π
2
: G GL(V
2
) be two linear representa-
tions of G, and let χ
1
and χ
2
be their characters. Then:
1. The character χ of the direct sum representation V
1
V
2
is equal to χ
1
+ χ
2
.
2. The character ψ of the tensor product representation V
1
V
2
is equal to χ
1
· χ
2
.
18
If φ and ψ are functions on G, set
φ, ψ =
1
|G|
gG
φ(g
1
)ψ(g) =
1
|G|
gG
φ(g)ψ(g
1
).
We have φ, ψ = ψ, φ. Moreover φ, ψ is lin ear in φ and in ψ. Consider the following
notation. If φ and ψ are two complex-valued functions on G, put
φ|ψ =
1
|G|
gG
φ(g)
ψ(g).
This is a s cal ar product; it is linear in φ, semilinear in ψ; and we have φ|φ > 0, for all
φ = 0. If
˘
ψ is the function d efined by the formula
˘
ψ(g) =
ψ(g
1
), th en we have
φ|ψ =
1
|G|
gG
φ(g)
˘
ψ(g
1
) = φ,
˘
ψ.
In particular, if χ is the character of a representation of G, we h ave ˘χ = χ (Proposit io n
2.1.2), so that φ|χ = φ, χ, for all functions φ on G. So we can use at will φ|χ or φ, χ,
so long as we are concerned with characters. We have an important result on orthogonality
of characters [166, p.15]:
Theorem 2.1.4. 1. If χ is the character of an irreducible represent ation, we have χ|χ =
1 (i. e., χ i s of norm 1).
2. If χ and χ
are the characters of t wo nonisomorphic irreducible re presentations, we
have χ|χ
= 0 (i. e., χ and χ
are orthogonal).
A character of an irreducib l e representation is called an irreducible character. Thus, The-
orem 2. 1. 4 shows that the irreducible characters form an orthonor m al system.
Theorem 2.1.5. Let V be a linear representation of G, with character φ, and suppose V
decomposes into a direct sum of irreducible representat ions:
V = W
1
W
2
··· W
k
.
Then, if W is an irreducible representation with character χ, the number of W
i
’s isomorphic
to W is equal to the scalar product φ|χ.
Proof. L et χ
i
be the charac t er of W
i
. By Proposition 2.1.3, we have φ = χ
1
+ . . . + χ
k
.
Thus, φ|χ = χ
1
|χ+ . . . + χ
k
|χ. But, accord ing to the preceding Theorem 2.1.4, χ
i
|χ
is equal to 1 or 0, depending on whether W
i
is, or is not, isomorph i c to W. The result
follows.
Corollary 2.1.6. The number of W
i
isomorphic to W does not depend on the chosen
decomposition.
19
This number is called the ’number of times that W occurs in V ’ , or the ‘number of times
that W is contained in V ’. It is in this se n se that on e can say that there is uniquen e ss in
the deco m position of a representation into irreducible representations.
Corollary 2.1.7. Two representations with the same character are isomorphic.
The above results reduce the study of repr esentations to that of their characters. If
χ
1
, . . . , χ
k
are the distinct irreducible characters of G, an d if W
1
, . . . , W
k
denote corre-
sponding representations, each r ep r e sentation V of G is isomorphic to a direct sum
V = m
1
W
1
m
2
W
2
. . . m
k
W
k
,
with m
1
, . . . , m
k
0 integers. The character φ of V is equal to m
1
χ
1
+ . . . + m
k
χ
k
and we
have m
i
= φ| χ
i
. We obtain thus a very convenient irreducibi li ty criterion [166, p.17]:
Theorem 2.1.8. If χ is the character of a representation V , φ|φ is a positi ve integer
and we have φ|φ = 1 if, and only if, V is irreducible.
Example 2.1.9 (The regular representation). Let n be the order of G, and let V be a
vector space of dimension n, wi t h a basis {e
g
}
gG
indexed by the elements g of G. For
s G, let π(s) be the linear map of V into V which sends e
g
→ e
sg
; this d efi n e s a linear
representation, which is called th e regular representation of G. Its degree is equal to the
order of G. Note that e
g
= π
g
(e
1
); henc e note that the images of e
1
form a basis of V .
Conversely, let W be a re p r esentation of G containing a vector w such that the π
g
(w), g G,
form a basis of W ; then W is isomorphic to the regular repre sentation (an isomorphism
τ : V W is defined by putt ing τ (e
g
) = π
g
(w).
For the rest of this secti on, the irreduci b l e characters of G are denoted χ
1
, . . . , χ
k
; their
degrees are written n
1
, . . . , n
k
, where n
i
= χ
i
(1) (Proposition 2.1.2). Let R be the regular
representation of G. Recall that it has a basis {e
g
}
gG
such that π
s
(e
g
) = e
sg
. If s = 1,
we have sg = s for all g, which shows that the diagonal terms of the matrix of π
s
are zero;
in particular we have tr ( π
s
) = 0. On the other hand, for s = 1, we have tr(π
s
) = tr(I) =
dim R = n. Whence:
Proposition 2.1.10. The character r
G
of the regular representation is given by the for-
mulas:
r
G
(1) = n = |G|, r
G
(s) = 0, if s = 1.
Corollary 2.1.11. Every irreducible representation W
i
of G is contained in the regular
representation with multiplicity equal to its degree n
i
.
Proof. According to Theorem 2.1.4, this number is equal to r
G
|χ
i
, and we have
r
G
|χ
i
=
1
|n|
gG
r
G
(s
1
)χ
i
(s) =
1
n
r
G
(1)χ
i
(1)
=
1
n
n · χ
i
(1)
= χ
i
(1) = n
i
.
20
Corollary 2.1.12. 1. The degrees n
i
satisfy the relation
i
n
2
i
= n.
2. If s G is different f rom 1, we have n
i
χ
i
(s) = 0.
Proof. By Corollary 2.1.11 , we have r
G
(s) =
i
n
i
χ
i
(s), for all s G. Ta ki n g s = 1 we
obtain (1), and taking s = 1, we obtain (2).
Note that the above result can be used in determining th e irreducible representations of
a gro u p G: suppose we have constructed some mutually non isomorphic irreducible repre-
sentations of degrees n
1
, . . . , n
k
; in order that they be all the irreducible representations of
G (u p to isomorphism ) , it is necessary and sufficient that n
2
1
+ . . . + n
2
k
= n. Also, we will
see below that the degrees n
i
divide the order n = |G|.
Recall that a function f on G is called a class function if f(tst
1
) = f (s) for all s, t G.
Denote by H the space of class funct i on s on G. In particular, the irreduci b l e characters
χ
1
, . . . , χ
k
belong to H. In fa ct , [166, p.19]
Theorem 2.1.13. The irreducible characters χ
1
, . . . , χ
k
form an orthonormal basis of H.
Recall that two el em ents g and g
of G are said to be conjugate if there exists s G su ch
that g
= sgs
1
; this is an equivalence re la t io n , which partitions G into classes (al so called
conjugacy classes). We have the following (see [166, p.19–20])
Theorem 2.1.14. The number of irreducible representations of G (up to isomorphism) is
equal to the number of conjugacy classes of G.
Proposition 2.1.15. Let s G, and let c(s) be the number of elements in the conju-
gacy class of s. Then,
i
χ
i
(s)χ
i
(s) =
n
c(s)
, and for g G not conjugate to s, we have
i
χ
i
(s)χ
i
(g) = 0.
We now give an exam p l e of a character table of a finite group.
Example 2.1.16 (Table of characters of S
3
). Take for G the group of symmetric grou p
of three elements S
3
. We have n = 6, and there are three conjugacy classes: the element
1 = (1); the three transpositions (1 2), (2 3), ( 3 1); and the two cyclic permutations (1 2 3),
(1 3 2). Let t be a transposition and c a cyclic permutation. We have t
2
= 1, c
3
= 1, and
tc = c
2
t (equivalently, ctct = 1). Whence th er e are just two characters of degree 1: the
unit character χ
1
and the character χ
2
giving the sig n at u r e of a pe rmutation (that is 1 for
the even per mutations and 1 for the odd ones). Theorem 2.1.14 above shows that there
exists one other irredu ci b l e character χ
3
; if n is its degree we must have 1 + 1 + n
2
= 6,
hence n = 2.
The values of χ
3
can be deduced from the fact that χ
1
+ χ
2
+ 2χ
3
is the character of t h e
regular representation r
S
3
of S
3
(Proposition 2.1.10 ) . We thus get the character table of
S
3
:
21
character 1 t c
χ
1
1 1 1
χ
2
1 -1 1
χ
3
2 0 -1
Table 2.1: Charac te r table for the symmetric group S
3
.
2.2 Lie groups
Consider the general linear group over the real numbers, denoted GL
n
(R), i. e., the group of
all n ×n invertible matrices with real entries. Similarly, we consider the the general linear
group over the complex numbers of all n × n invertible matrices with complex entries,
denoted GL
n
(C). The gen er al linear groups are indeed groups under the operation of
matrix multiplication: the product of two invertible matrices is invertible, the identity
matrix is an identity for the grou p , an invertible matrix has (by d efinition) an inverse, and
matrix multiplication is associative.
Let C
n×n
denote the space of all n × n matrices with com plex entries, and let {A
k
}
k
be a
sequence of complex matrices in C
n×n
. We say that {A
k
}
k
converges to a matrix A if each
entry of (A
k
)
ij
of A
k
converges to the corresponding entry of A
ij
of A.
Definition 2.2.1. A matrix Lie group is any subgroup G of GL
n
(C) with the following
property: If {A
k
}
k
is any sequence of mat r i ces in G, and {A
k
} converges to some m at r i x
A then either A G, or A is not invertible.
The condition on G amounts to saying that G is a closed subgroup of GL
n
(C) (although
it does not necessarily mean that G is closed as a subset of C
n×n
. Most of the ‘interesting’
subgroups of GL
n
(C) have this property, we give some classical examples:
Example 2.2.2. The general linear gr ou ps GL
n
(R) and GL
n
(C) are th emselves matrix
Lie gro u p s.
Example 2.2.3. (The special linear gr oups SL
n
(R) and SL
n
(C))
The special linear group (over R or C) is the group of n × n invert ible m at r i ces (wit h rea l
or complex entries) having determin ant one. That is SL
2
(C) = {A GL
2
(C) : det A = 1}.
Both of these are subgroups of GL
n
(C). Furthermore, if {A
k
}
k
is a sequence of matrices
with determinant one an d {A
k
} converges to A, then continuity of determinant fun ct io n
implies that A also has determinant one.
Example 2.2.4. (The or t hogonal and special orthogonal groups O
n
(R) an d SO
n
(R))
An n × n real matrix A is said to be orthogonal if the col u m n vectors that make up A are
22
orthonormal, that is, if
n
=1
A
ℓj
A
ℓk
= δ
jk
, for all 1 < j, k < n.
Equivalently, A is orthogonal if it preserves the inner product, namely if x, y = Ax, Ay,
for all vectors x, y R
n
. Since det A
tr
= d et A, we see that if A is orthogonal, then
(det A)
2
= d et( A
tr
A) = det I = 1. Hence, det A = ±1, for all orthogonal matrices A.
In particu l ar , every orthogonal matrix must be invertible. However, if A is an ort hogonal
matrix, then
A
1
x, A
1
y = A(A
1
)x, A(A
1
)y = x, y.
Thus, the inverse of an orthogon a l matrix is orthogonal. Furthermore, the product of two
orthogonal matrices is o r t h ogo n al , since if A and B both preserve inner products, then so
does AB. Thus, the set of orthogonal matrices forms a group.
The set of all n × n real orthogonal matrices is the orthogonal group O
n
(R), and i t is a
subgroup of GL
n
(C). The limit of a sequence of orthogonal matrices is orthogonal, because
the rela t io n A
tr
A = I is preserved under taking l i m i t s. Thus, O
n
(R) is a matrix Lie group.
The set of n ×n ort hogonal matrices with determinant one is the special orthogonal group
SO
n
(R). This is also a matrix Lie group. Sometimes O
n
(R) and SO
n
(R) are si m p l y de-
noted by O(n) and SO( n) , respect i vely.
Example 2.2.5. (The complex orthog on a l groups O
n
(C) and SO
n
(C))
Similarly to the example above, we can define orthogonal groups of complex matrices.
Consider the bilinear form ·, · on C
n
given by x, y =
k
x
k
y
k
. We define the complex
orthonormal group as
O
n
(C) = {A GL
n
(C) : Ax, Ay = x, y}.
The complex special orthogonal group SO
n
(C) is defined to be the set of all matrices
A O
n
(C) with det A = 1. These groups are also matrix Lie gro u p s.
Example 2.2.6. (The u nitary group s U(n) and SU(n))
An n×n complex matrix A is said to be unitary if the column vectors of A are orthonormal,
that i s, if
n
=1
A
ℓj
A
ℓk
= δ
jk
, for all 1 < j, k < n.
Equivalently, A is unit a r y if it preserves the inner pro d uct x, y
C
=
k
x
k
¯y
k
, namely if
Ax, Ay
C
= x, y
C
for all vectors x, y C
n
. Still another equivalent definiti on i s tha t A is
unitary if A
A = I, where, A
=
A
tr
is the adjoint (herm i t i an ) of A. Since det A
= det A,
we see that if A is unitary, then det(A
A) = |det A|
2
= 1. Hence, |det A| = 1, for all
23
unitary matrices A. In particu la r , this shows that every unitary matrix is invertible. The
same argument as for the orthogonal group shows tha t the set of unitary matrices forms a
group.
The set of all n × n unitary matrices is the unitary group U(n), and it is a subgroup o f
GL
n
(C). The limit of unitary matrices i s unitary, so U(n) is a matrix Lie gr ou p. The set
of u n i t ar y matrices with determin a nt one is the special unitary group SU(n), a n d it can
be p r oved that SU(n) is a m at r i x Lie group.
Example 2.2.7. (The sy m p l ect i c groups Sp
n
(R), Sp
n
(C), an d Sp(n))
Consider the skew-symmetric bilinear form [·, ·] on R
2n
defined as follows:
[x, y] =
n
k=1
(x
k
y
n+k
x
n+k
y
k
). (2.1)
The set of all 2n × 2n matrices A which preserve [ · , ·] i s called the real symplectic group
Sp
n
(R), and it is a subgroup of GL
2n
(C). In fact, Sp
n
(R) is a matrix Lie group. This
group arises natu r al l y in the study of classical m e chanics: If J is the 2n × 2n matrix
J =
0 I
I 0
,
then [x, y] = x, Jy, and it is possib l e t o check that a 2n ×2n real m at r ix A is i n Sp
n
(R) if
and on ly if A
tr
JA = J. Taking the determinant of this identity gives (det A)
2
det J = det J,
or (det A)
2
= 1. This shows that d et A = ±1, for all A Sp
n
(R). In fact, det A = 1 for all
A Sp
n
(R).
One can define a bilinear form on C
2n
by the same formula (2.1), since this form involves
no complex conjugates. The set of 2n × 2n complex matrices which preserve this form i s
called the complex symplectic group Sp
n
(C). Finally, we have the compact symplectic group
Sp(n) defined as
Sp(n) = Sp
n
(C) U(2n).
Example 2.2.8. (The Hei senberg group H)
The set of all 3 × 3 real matrices A of the form
1 a b
0 1 c
0 0 1
, wh er e a, b, c R,
is called the Heisenberg group H. In fact, H is a matrix Lie group. The reason for the
name Heisenberg group is that the Lie algebra of H gives a realization of the Heisenberg
commutation relations of quantum mechanics.
24
Example 2. 2.9 . The groups R
×
, C
×
, S
1
, R, and R
n
can be thought as subgroups of matri-
ces. For example, the group R
×
of non-zero real numbers under multiplication is isom o rphic
to GL
1
(R). Thus, we will regard R
×
as a matr i x Lie group . Similarly, the group C
×
. The
group S
1
of compl ex numbers module one is isomorphic to U(1).
The additive grou p R is isomorp h i c to GL
1
(R)
+
(1 ×1 real matrices with positive determi-
nant) vi a the map x → [e
x
]. In a similar way, the additive group R
n
is isom or phic to the
group of diagonal real matrices with positive diagonal entries, via the map
(x
1
, . . . , x
n
) →
e
x
1
0
.
.
.
0 e
x
n
.
Now we discuss some notions of compact n ess and connectedness for matrix Lie groups.
Definition 2.2.10. A matrix Lie group G is said to be compact if the following two
conditions are sat is fi ed :
1. [closeness] If {A
k
}
k
is any sequence of matrices i n G and {A
k
}
k
converges to a matrix
A, then A is in G.
2. [boundedness] There exists a constant C such that for all A G, |A
ij
| < C for all
1 < i, j < n.
This is not the usual t opological definition of compactness. Thinking the set of all n × n
complex matrices as C
n
2
, the above definition says that G is compact if it is a closed,
bounded subset of C
n
2
. It is a stand ar d theo r em from eleme ntary analy si s th a t a subset
of C
n
2
is compact if and only if it is closed and bounded. All of our examples of matrix
Lie groups except GL
n
(R) and GL
n
(C) have property (1). Thus, it is the boun d edness
condition (2) t h at is most important.
The groups O
n
(C) and SO
n
(C) are compact. Property (1 ) is satisfied because the limit of
orthogonal matrices is orthogonal and the limit of matrices with determinant one has de-
terminant one. Pr operty (2) is satisfied because if A is orthogonal, then the column vectors
of A have norm one, and hence |A
ij
| < 1, for all 1 < i, j < n. A similar argument shows
that U(n), SU(n), and Sp(n) are compact , this includes th e unit circle, S
1
=
U(1). On
the oth er hand, the groups GL
n
(R) and GL
n
(C) are noncompact, si n c e a limit of inver t ible
matrices could be noninvertible. The grou p s SL
n
(R) and SL
n
(C) violate boundedness,
except in the trivial case n = 1. The following groups also violate (2), and hence are
noncompact: O
n
(C) and SO
n
(C); the Heisenberg group H; Sp
n
(R) and Sp
n
(C); R, C, R
×
and C
×
.
Definition 2.2.11. A matr i x Lie group G is said to be connected if given any two matrices
A and B in G, there exi st s a continuous path γ : [0, 1] G, with γ(0) = A and γ(1) = B.
This property is what is called path-connected in topology, which is not (in gener a l) t he
same as connected. However, it is a fact that a mat r i x Lie grou p is connect ed if and
25
only if it is pat h - con nected. A matrix Lie group G which is n ot connected ca n be decom -
posed (uniquely) as a un i on of several pieces, called components, such that two el em ents
of the same component can be joined by a continuous path, but two elements of di ffe re nt
components cannot.
It is a known result that the general linear g ro u p GL
n
(C) is connected for all n 1, but
the group GL
n
(R) is not. In Table 2.2 we list the examples of matrix Lie grou p s a bove,
indicating their connectedness properties.
Group Connected? Number of Components
GL
n
(C) yes 1
SL
n
(C) yes 1
GL
n
(R) no 2
SL
n
(R) yes 1
O
n
(C) yes 1
SO
n
(C) yes 1
O
n
(R) no 2
SO
n
(R) yes 1
U(n) yes 1
SU(n) yes 1
Heisenberg yes 1
Table 2.2: Conne ct ed n es s propert i es of some classical matrix Lie gr oups.
In a similar manner, we can st u dy the notion of simply connected for mat r i x Lie groups:
Definition 2.2.12. A matrix Lie group G is sai d t o be simply connected if it is connected
and, in addition, every loop in G can be shrunk continuously to a point in G. More precisely,
if gi ven any continuous path γ : [0, 1] G, with A(0) = A( 1) , there exists a continuous
function H(s, t) : [0, 1] × [0, 1] G having the foll owing properties:
(a) H(s, 0) = H(s, 1) for all s [0, 1].
(b) H(0, t) = γ( t), for all t [0, 1].
(c) H(1, t) = H(1 , 0), for all t [0, 1].
Table 2.3 resumes the examples o m at r i x Lie groups above, indicat i n g their simply con-
nectedness properties.
We conclude this section with the definition of a Lie group:
Definition 2.2.13. A Lie group is a differentiable manifold G which is also a group and
such tha t the group product
G × G G
26
Group Simply connected? Fundamental group
SO
2
(R) no Z
SO
n
(R), n 3 no 2Z
U(n) no Z
SU(n) yes {1}
Sp(n) yes {1}
GL
n
(C) no Z
SL
n
(C) yes {1}
GL
n
(R)
+
, n 2 no same as SO
n
(R)
SL
n
(R) no same as SO
n
(R)
SO
n
(C) no same as SO
n
(R)
Sp
n
(C) yes 1
Sp
n
(R) no Z
Table 2.3: Sim p l y connectedne ss propert i es of some classical matrix Lie gr oups.
and the inverse map g → g
1
are differ entiable.
A manifold is an object that looks locally like a p i ece of R
n
, that is, a topological space that
is Hausdorff, second cou ntable, and locally Eucl i d ean. An example would be a torus. For
a precise definition, see [128] or [129]. We conclude with a result that establishes whether
every matr i x Lie group is a Lie group [91, p.52].
Theorem 2.2.14. Every matrix Lie group is a smooth embedded submanifold of C
n×n
and
is thus a Lie group.
2.3 Lie algebras
A nonassociative algebra is a vector space A (over a field F of characteristic 0) equipped
with a bilinear map, called product from A × A to A. The algebr a A is called associative
if it contains an identity element 1 for multiplication, so that
1a = a = a1, a A,
and i f the associative law a(bc) = (ab)c holds, for all a, b, c A. The algebra A is said to
be commutative if the commutative law ab = ba holds, a, b A.
For subspaces B, C of an a lg eb r a A, we write BC for the subspace of A spann ed by the
products bc, for b B, c C. Similarly, we denote by [B, C] the subspace BC CB =
27
{[b, c] : b B, c C} spanned by al l th e commutators [b, c]. A subalgebra of A is a subspace
B of A such that B
2
B, and 1 B in th e associative case. Equivalently, a subalgebra of
A is a subset B of A which is an algebra under the linear and product structures induced
from A.
For algebras A and B, a lin ear map f : A B is an homomorphism if f(ab) = f( a) f (b),
for a, b A, and if i n addition f(1) = 1 in the as sociative case. The homomorph i sm f is
an isomorphism if it is a li near isomorphism . In case A = B, an homomorphism is called
an endomorphism, and an isom o r p h i sm is called an automorphism. Two algebras A and B
are said isomorphic if there is an isomorphism f : A B. In thi s case, we write A
=
B.
A linear endom or phism d : A A of an algebra A is called a derivation if for all x, y A
we have d(xy) = d(x)y + xd(y).
Definition 2.3.1. A Lie algebra is a nonassociat ive algebra g such that:
(i) ( al t er n at e axiom) [x, x] = 0, x g.
(ii) (Jacobi’s id entity) [x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0, x, y, z g.
The alternate pr operty is equivalent to the skew-symmetry condition [x, y] + [y, x] = 0,
x, y g. The Jacobi identity is equivalent to the condition that the adjoint map
ad x : g g, given by ad x(y) = [x, y]
be a derivation, for all x g.
A Lie algebra g is called commutative or abelian if [g, g] = 0. Two elements x, y of a L ie
algebra g are said to commute if [x, y] = 0. In particular, all pairs of elements of an abelian
Lie algebra commute. For Lie algebras g and k, a linear map f : g k is an homomorphism
if f is an algebra homomorphism and it sat i sfi es
f([x, y]) = [f(x), f(y)], x, y g.
We define in obvious way t h e term isomorphism, endomorph i sm and automorphism.
Example 2.3.2. An associative algebra is a Lie algeb r a wit h t h e br acket [x, y] = xy yx.
Example 2.3. 3. Let B be a nonassociative algeb r a , and set A = End B. Then, the space
of deri va ti o n s of B forms a Lie subal gebra of A.
Example 2.3.4. Every one- d i m ensional Lie algebra is abelian.
Lie algeb r as are simpler than matrix Lie groups, because (as we have seen) the Lie algebra
is a linear space. Thus, we can understand much about Lie algebras just by doing linea r
algebra. In fa ct , to each matrix Lie grou p G we can associat e a L i e algebra, just by
28
considering the expon en ci a l of a matrix. Let X be an n × n real or complex m at r i x. We
define t h e exponenti al matrix of X, d en ot e d e
X
or exp X, by the usual power series
e
X
=
m0
X
m
m!
. (2.2)
We know from the differential equation th e or y that th e exponencial series of any ma tr i x
A GL
n
(C) conver ges uniforml y in every compact s u b se t of C, since it i s the solution of
the homogeneous linear equation given by X
= AX, X(0) = I (see [168] for a proof).
Also, t h e exponen cia l satisfies the usual properties:
1.
d
dt
e
tA
= Ae
tA
;
2. e
(s+t)A
= e
sA
e
tA
, for all s, t R;
3. e
tA
=
m0
t
m
A
m
m!
;
4. BC = CA impli es e
tB
C = Ce
tA
;
5. e
t(A+B)
= e
tA
e
tB
if and only if AB = BA. In particular , (e
tA
)
1
= e
tA
;
for all A, B, C GL
n
(C) and all s, t C (see [168] or [26]).
Definition 2.3.5. Let G be a matr i x Lie group in C
n×n
. The Lie algebra of G, denoted
by g, is the set of all matrices X C
n×n
such tha t e
tX
is in G for all real numbers t.
This means that X is in g if and only if the one-parameter subgroup generated by X lies
in G. Note that even though G i s a subgroup of GL
n
(C) (and not necessarily of GL
n
(R)),
we do not require that e
tX
be in G for all complex numbers t, but only for all real numbers
t. Also, it is not enou gh to have just e
X
G. It can be given an exam p l e of an X an d a
G such that e
X
G, but such that e
tX
/ G for some real va l u es of t. Such an X is no t in
the Li e algebra of G.
On the other han d , since all matrix Lie group G is itself a Lie group, and thus a smooth
manifold (Theorem 2.2 .1 4), we can obtain the tan gent space of G on the point I, the iden-
tity matrix of G, namely T
I
G. It is a well kn own result that, if X is on T
I
G, then e
tX
G,
for all t R (see [134] for a proof).
For A, B T
I
G, we define the bracket [ · , ·] by [A, B] = AB BA. Since we ca n write
e
tA
= I +
tA +
1
2
tA
2
+ o(t),
e
tB
= I +
tB +
1
2
tB
2
+ o(t),
e
tA
= I
tA +
1
2
tA
2
+ o(t),
e
tB
= I
tB +
1
2
tB
2
+ o(t),
29
and set λ(t) = e
tA
e
tB
e
tA
e
tB
, then we have
λ(t) = e
tA
e
tB
e
tA
e
tB
= I +
t
(A + B) ( A + B)
+ t
A
2
+ 2AB + B
2
(A + B)
2
+ o(t)
= I + t[A, B] + o(t),
with lim
t0
o(t)
t
= 0. Th en , λ
(0) = [A, B]. Since λ(t) G, for all t R, we see that
[A, B] T
I
G. We have also proved that T
I
G admits a Lie algebra structure. In fact, th e
Lie alg eb r a g of a matrix L ie group G is given by
g = T
I
G,
and it directly follows that T
X
G = T
I
G · X, for a ny X G (see [129 ] or [134]).
Note that we can a lso see the Lie algeb r a g of a Lie group G as the set of all left-invariant
vector fields X : G T G (or linear derivations X : C
(G) R), wi t h some bracket
[X, Y ] (see [91 ]) . Because g is a real sub al gebra of th e space GL
n
(C) we h ave the following
results, [91, p. 55] and [179, p.237]
Proposition 2.3.6. The Lie algebra g associated to the Lie group G is a real Lie algebra.
Theorem 2.3.7 (Ad o ) . Every finite-dimensional real L ie algebra is isomorphic to a subal-
gebra of gl
n
(R). Every finite-dimensional complex Lie algebra is isomorphic to a subalgebra
of gl
n
(C).
We present some examples of Lie algebras associated to some Lie groups st u died in previous
section:
Example 2.3.8 (The g en er al linear groups). If X is any n × n complex matrix, then e
tX
is invertible. Thus, the Lie al geb r a of GL
n
(C) is the space of all n × n compl ex matrices.
This Lie algebra is denoted gl
n
(C). If X is any n×n real mat r i x, th en e
tX
will be invertible
and real. On the other hand, if e
tX
is r eal for all real numbers t, then X =
d
dt
e
tX
t=0
will
also be real. Thus, the Lie algebra of GL
n
(R) is the space of all n × n real matrices, and
is den ot ed gl
n
(R).
Example 2.3.9 (The special linear groups). It is a well known restult that det e
X
= e
tr X
.
Thus, if tr X = 0, then det e
tX
= 1 for all real numbers t. On the other hand, if X is any
n × n matrix such that det e
tX
= 1 for all t, then e
t tr X
= 1 for all t. This means that
t tr X is an integer multiple of 2πi for all t, which is only possi b l e if tr X = 0. Thus, the
Lie algebra of SL
n
(C) is the space of all n × n complex matrices with trace zero, denoted
sl
n
(C).
Similarly, the Lie algebra of SL
n
(R) is the space of all n ×n real matrices with trace zero,
denoted sl
n
(R).
30
Example 2.3.10 (The unitary groups). Recall that a matr i x U is unitary if and only if
U
= U
1
. Thus, e
tX
is un it a r y if and only if
(e
tX
)
= (e
tX
)
1
= e
tX
. (2.3)
Since (e
tX
)
= e
tX
, th en (2.3) becomes e
tX
= e
tX
. Clearly, a su fficient condition for th i s
last identity to hold is that X
= X. On the o ther hand, if e
tX
= e
tX
holds for all t,
then by differentiating at t = 0, we see that X
= X is necessary. Thus, the Lie al gebra
of U(n) is the space of all n × n complex matrices X such that X
= X, and is denoted
by u(n).
Combining the two previous computations, we see that the Lie algebra of SU(n) is the
space of all n × n complex matrices X such that X
= X and tr X = 0, denoted su(n).
Example 2.3.11 (The orthogonal groups). The identity component of O(n) is just SO(n).
Since the exponential of a matrix in the Lie algebra is automatically in the i d entity com-
ponent, the Lie algebra of O(n) is the same as the Lie algebra of SO(n). Now, an n × n
real mat r ix R is orthogonal if and only if R
tr
= R
1
. So, given an n × n real matrix X,
e
tX
is orth og on al if and only if (e
tX
)
tr
= (e
tX
)
1
, or
e
tX
tr
= e
tX
. (2.4)
Clearly, a sufficient condition for th i s to hold is that X
tr
= X. If (2.4) holds for all
t, then by differentiating at t = 0, we must have X
tr
= X. Thus, the Lie alg eb r a of
O(n), as well as the Lie algebra of SO(n), is the space of all n × n real matrices X with
X
tr
= X, denoted by so
n
(R) (or simply so(n)). Note that t he condition X
tr
= X forces
the di ag on al entries of X to be zero, and so, n ecess ar i ly the trace of X is zero.
The same argument shows that the Li e algebra of SO
n
(C) is the space of n × n complex
matrices satisfying X
tr
= X, denoted by so
n
(C). Th i s is not the same as su(n).
Example 2.3.12 (The symplectic groups). These are d en o te d sp
n
(R), sp
n
(C), and sp(n) .
The calculation of t hese Lie algebras is sim i l ar to that of the generalized orthogonal group s,
and we will just record the result h er e. Let J be the matrix in the definition of the symplectic
groups. Then, sp
n
(R) is the space of 2n × 2n real matrices X such th at JX
tr
J = X,
sp
n
(C) is the space of 2n ×2n complex matrices satisfying t he same condition, and sp(n) =
sp
n
(C) u( 2n) .
Example 2.3.1 3 (The Heisenb er g group). Recall that the Heisenberg group H is the group
of all 3 × 3 real matrices A of the form
1 a b
0 1 c
0 0 1
, wh er e a, b, c R.
Computing the exponential of any matrix of the form
X =
0 α β
0 0 γ
0 0 0
, wh er e α, β, γ R, (2.5)
31
we can see that e
tX
is i n H. On the other hand, if X is any matrix such that e
tx
is i n H
for all t, th en all of the entries of X =
d
dt
e
tX
t=0
which are on or below the diagonal must
be zero, so th at X is o f form (2.5) . Thus, the Lie algebra of the Heisenberg gr ou p is the
space of all 3 × 3 real matrices that are strictly upper triangular.
Definition 2.3.14. If V is a fini t e- d i m en si onal real vector space, then t h e complexification
of V , den ot ed V
C
, is the space of formal linear combinations
v
1
+ iv
2
, wit h v
1
, v
2
V.
This becomes a real vector space in the obvious way and becomes a complex vector space if
we defi ne i(v
1
+ iv
2
) = v
2
+ iv
1
. We will regard V as a real subspace of V
C
in the obvious
way. Now let g be a finite-dimension al real Lie algebra and g
C
its com p l ex ification (a s a
real vector space). Then, the bracket operation on g has a unique extension to g
C
which
makes g
C
into a complex Lie algebra. In fact, the uniqueness of the extens io n is obvious,
since if the bracket operation on g
C
is to be bilinear, then it must be gi ven by
[x
1
+ ix
2
, y
1
+ iy
2
] =
[x
1
, y
1
] [x
2
, y
2
]
+ i
[x
1
, y
2
] + [x
2
, y
1
]
. (2.6)
This bracket defined above is really bilinear and skew symmetric and it s at i sfi es the Jacob i
identity. It is clear from (2.6 ) is re al b i l inear, and skew-symmetric. The skew symmetry
means that if (2.6) is complex linear in the first factor, it is also complex linear in th e
second fact or . Thus, we need only show that
[i(x
1
+ ix
2
), y
1
+ iy
2
] = i[x
1
+ ix
2
, y
1
+ iy
2
]. (2.7)
The left-hand side of (2.7) is [x
2
+ix
1
, y
1
+iy
2
] = ([x
2
, y
1
][x
1
, y
2
])+i([x
1
, y
1
]+[x
2
, y
2
]),
whereas t h e right-hand side of (2.7) is just
i
[x
1
, y
1
] [x
2
, y
2
]
+ i
[x
1
, y
2
] + [x
2
, y
1
]
= ([x
2
, y
1
] [x
1
, y
2
]) + i( [x
1
, y
1
] + [x
2
, y
2
]),
and, in d ee d , these are equal. To check the Jacobi identity, note that the Jacobi identity
holds if x, y, and z are in g. However, observe that t h e expression on the left-hand side of
the Jacobi identity in Definition 2.3.1 is (complex!) linear in x for fixed y an d z. It follows
that the Jacobi identity holds if x is in g
C
, and y and z are in g. The same argum e nt shows
that we can extend to y i n g
C
, a n d then to z in g
C
. Thus, the Jacobi identity holds i n g
C
.
Thus, we have
Proposition 2.3.15. Let g be a finite-dimensional real Lie algebra and g
C
its complexifi-
cation (as a real vector space). Then, the bracket operation on g has a unique extension to
g
C
which makes g
C
into a complex Lie algebra.
Definition 2.3.16. The complex Lie algebra g
C
is called the complexification of the real
Lie alg eb r a g.
We give the complexifications of some real Li e algebras in Table 2.4
32
Lie algeb r a Complexification
gl
n
(R) gl
n
(R)
C
= gl
n
(C)
u(n) u(n)
C
= gl
n
(C)
su(n) su(n)
C
= sl
n
(C)
sl
n
(R) sl
n
(R)
C
= sl
n
(C)
so(n) so(n)
C
= so
n
(C)
sp
n
(R) sp
n
(R)
C
= sp
n
(C)
sp(n) sp(n)
C
= sp
n
(C)
Table 2.4: Compl ex ifications of some classical real Lie algebras.
2.4 Mod ul es
Definition 2.4.1. Let A be an associative algebra and let V be a vector sp a ce. We s ay
that V is an A-module if there is a bilinear map A×V V (denoted by a dot (a, v) → a·v)
such tha t
(i) [ identity] 1 · v = v, for all v V .
(ii) [associativity] (ab) · v = a · (b · v), for all a, b A, and all v V .
For a A, let π
a
the correspond i n g linear endomorphism of V , such that π
a
(v) = a · v.
Then, t h e map
π : A End V, such that π(a) = π
a
is an homomorp h i sm of associative algebras. Such an homomorphism is called a represen-
tation of A on V . Sometim es, V is called a representation of A.
Example 2.4.2. Note that the associative algebra A ha s a natural representation on itself,
given by the left multiplication action: a · b = ab, a, b A.
Definition 2.4.3. Analogously, let g be a Lie algebra and let V be a vector space. Then V
is called a g-module if there is a b i l i n ear map g ×V V (denoted by a dot (x, v) → x ·v)
such tha t
(i) ( b r acket preserving) [x, y] · v = x · (y · v) y · (x · v), for all x, y g, an d all v V .
If for x g, we denote π
x
the corr esponding linear endomorphism of V , such that
π
x
(v) = x · v.
Then, t h e map
π : g End V, such that π(x) = π
x
is a Lie algebra homomorp h i sm . Such an h o m om o rphism is called a representation of g on V
(sometimes V is called a representation of g). The notio n s of g- m odule and representation
33
of g are equivalent.
The Lie algebra g has a natural representation on itself, called the adjoint representation,
by the map
ad : g End g, such that x → ad x.
In p ar t i cu l a r, taking g = End V , every Lie subalgebra of End V has a natural representa-
tion on V .
Let g be an associative or Lie algebr a , and let V be a g-module. For subspaces h of g and
W of V , we denote h · W the linear span of all x · w, for x h, w W . A submodule of V
is a sub space W of V such that g · W W , or equivalently, a subset W of V which is a
g-module under the linear stru ct ure a n d g-module action induced from V . A subsp ac e W
of V i s invariant (under g) if it is a su b m odule. The module V is said irreducible or simple
if V = 0 and if V has no proper nonzero invariant subspaces. The module V is called
indecomposable if it cannot be decomposed as a direct sum of two nonzero submo d u l es.
Clearly, an irreduci b l e module is indecomposable. Let V and W be g-modules. A linear
map f : V W is called a g-module homomorphism or g-module map if
f(x · v) = x · f(v), x g, v V.
Such a map f is called a g-module isomorphism or g-module equivalence if it is a linear iso-
morphism. Two modul es V and W are isomorphic or equivalent if there is an isomorphism
f : V W . In that case we write V
=
W .
Definition 2.4.4. A subspace a of a Lie algebra g is called an ideal of g if [g, a] a.
Equivalently, an i d eal of g is a subm odule under the adjoint representation.
An ideal is a su b al ge b r a. Given an ideal a of g, the quotient vector space g/a becomes a
Lie algebra, called the quotient Lie algebra, by means of the well defined nonassocia t ive
product
[x + a, y + a] = [x, y] + a, x, y g.
The canonical map π : g g/a is an homomorphism, and we have an exact sequence of
Lie alg eb r as
0 a g
π
g/a 0.
A Lie algebra g is said to be simple if g is non zer o and has no proper nonzero ideals (equiv-
alently, the adjoint representation is simple), and if dim g > 1 (i. e., g is not abelian) .
Remember that a sub sp a ce I of an associative algebra A is called a left ideal (respectively
right ideal) of A if AI I (respectively IA A), and an ideal if is both left and right ideal.
Since an i d eal I need not to contain 1, it need not be a subalgebra. Given and ideal I of
A, the qu ot i ent vector space A/I becomes an associative algebra —the quotient algebra
in an obvious way.
34
Example 2.4.5. The kernel of any homomorp hism of a Lie (respectively associative) al-
gebra g is an ideal.
In particular, the kernel of the adjoint representation of a Lie algebra g is an important
ideal c al le d the center of g and de n ot ed by Cent g:
Cent g = {x g : [x, g] = 0}.
Any subs p ace of Cent g is an ideal in g and is said to be a central ideal.
Definition 2.4.6. Given the Lie algebras a and b, an extension of a by b is a Lie algebra
g together with an exact sequence
0 b g a 0,
(note that b is an ideal and g/b
=
a). This extension is said to be central if b is a central
ideal o f g.
Two extensions g
1
and g
2
of a by b are equivalent if there is an i so m or p h i sm g
1
=
g
2
,
making the following diagram commute
g
1
=
?
?
?
?
?
?
?
?
0
//
b
??
?
?
?
?
?
?
?
?
a
//
0.
g
2
??
If a and b are ideals of a Lie algebra g, then a + b, a b and [a, b] are ideals also. In
particular, [g, g] is an ideal of g, called the commutator ideal, and is denoted by
g
= [g, g].
Given two Lie algebras a and b, their direct product is the Lie alge b r a a ×b, which is a b
as a vector space, with a and b retaining their original bracket structures and commuting
with one another. In particular, a and b are ideals of a × b. We define the di r ect product
of fini t el y many Lie algebras analogously.
More generally, suppose that we have a representation π : a End b of a Lie algebra a on
a Lie algebra b by derivations, i. e., π(x) = π
x
is a derivation of b, for all x a. Then, the
vector sp a ce a b carries a unique Lie algebra structure such that a and b are subalgebras
and [x, y] = π
x
(y), for all x a, y b (observe that b is an ideal, but n o t necessarily a).
This Lie algebra is called the semidirect product of a and b, and we denote this by a b. A
Lie algebra is a semidirect product whenever it i s a vector space direct sum of a subalgebra
and a ideal. In fact, a b = a × b if and only i f π 0. A semidirect pr oduct a b is an
extension of a by b. An extension of a Lie algebra a by a Lie algebra b is said trivi al if it
is equi valent to a semidirect product a b.
35
Example 2.4.7. We mention a particular kind of semidirect product. Given a L i e algebra
g and a derivation d of g, we can form Fd g. This procedure is called adjoining the
derivation d to g.
Definition 2.4.8. Let S be a set. A vector space V is sai d to be S-graded if it is the direct
sum
V =
αS
V
α
,
of subsp aces V
α
. In this case, the elements o f V
α
are said to be homogeneous of degree α,
and V
α
is called the homogeneous subspace of degree α or the α-graded component of V .
For v V
α
(including v = 0) we write
deg v = α.
Given another S-graded vector space W , a linear map f : V W is called grading
preserving if
f : V
α
W
α
, for all α S.
If such a map i s a linear isomorph is m , V and W are said to be graded-isomorphic. If S is
an abelian group, a linear map f : V W is said to be homogeneous of degree β if
f : V
α
W
α+β
, for all α S. (2.8)
In that case, we write deg f = β. Not e that f is grading-pr es er vi n g if and only if deg f = 0.
If S is an abel ia n group which is a subgroup of the add i t i ve group of F, we can define the
degree operator d : V W by th e condition
d(v) = αv, for v V
α
, α S. (2.9)
Note that a linear map f : V W is grad i n g- p r es er vi n g if and on l y if [d, f ] = 0, and is
homogeneous of degr ee β if and only if [d, f] = βf.
A subspace W of a S-graded vector space V is graded if W =
αS
W
α
, where W
α
=
W V
α
, for α S. In this case, the quotient V /W is graded in a natural way:
V/W =
αS
(V/W )
α
=
αS
V
α
/W
α
.
Given a family of S-graded vector spaces {V
i
}
i
, the direct sum X =
i
V
i
is natural ly
S-graded if we t a ke
X
α
=
i
V
i
α
, for α S.
36
2.5 Tensor products
Definition 2.5.1. Let V be a vector space over a field F wit h a basis {e
i
}
i
and let W be
a vector space over F with a basis {f
j
}
j
. The tensor product V W is a vector space over
F, with a basis {e
i
f
j
}
i,j
.
If x =
i
x
i
e
i
V and y =
j
y
j
f
j
W , we have a tensor multiplication defined by
x y =
i, j
x
i
y
j
(e
i
f
j
) V W.
The tensor map τ : V × W V W d efined by (x, y) → x y satisfies bilinearity:
(i) ( x + y) z = (x z) + (y z);
(ii) x (y + z) = (x y) + (x z);
(iii) (λx) y = λ(x y) = x (λy).
Conversely, if β is a bil i n ea r mapping of V ×W into another vector space X, then β(x, y) =
β
i
x
i
e
i
,
j
y
j
f
j
=
i,j
x
i
y
j
β(e
i
, f
j
) by bilinearity, and the linear transformation T :
V W X t h at sends e
i
f
j
to β(e
i
, f
j
) also sends x y to β(x, y), for all x V, y W .
In fact, T is the unique linear tr an sfo rmation such that β = T τ. In thi s sense, every
bilinear mapping of V × W factors uniquely through β (universal property):
V × W
T
//
τ
X
V W
β
;;
v
v
v
v
v
v
v
v
v
Tensor products of modules over a commutative rin g R can be defined by the same universal
property (which does not require bases). More explicitly, we have [88, p.434]
Proposition 2.5.2 (Universal property for tensor produ ct s) . Let R be a commutative ring
and let A, B, C be R-modules. For a mapping β : A × B C the following conditions are
equivalent:
1. β is bilinear;
2. a → β(a, ·) is a module homomorphism of A into Hom
R
(B, C);
3. b → β(·, b) is a module homomorphism of B into Hom
R
(A, C).
Bilinear mappings can be defined in th e same way for left R-modules over an arbitrary r i n g
R, but then l ose properties (2) and (3) above, if only because Hom
R
(A, C) and Hom
R
(B, C)
are only abelian g r ou p s . It is more fruitful to keep properties (2) and (3), and to forgot
bilinearity unless R is commutative. In the simplest form of ( 2 ) and (3), C is an abelian
group; i f B is a left R-module, then Hom
Z
(B, C) is a right R-module and A needs to be a
right R-module; so B and Hom
Z
(A, C) are left R-modul es. Also, [88, p.435]
37
Proposition 2.5.3. Let R be a ring, let A be a right R-module, let B be a left R-module,
and let C be an abelian group. For a mapping β : A ×B C the following conditions are
equivalent:
1. For all a, a
A, b, b
B and r R
(i) [biadditive] β(a + a
, b) = β(a, b) + β(a
, b), β(a, b + b
) = β(a, b) + β(a, b
),
(ii) [balanced] β(ar, b) = β(a, rb);
2. a → β(a, ·) is a module homomorphism of A into Hom
Z
(B, C);
3. b → β(·, b) is a module homomorphism of B into Hom
Z
(A, C).
We say that a bihomomorphism of modules is a mapping that satisfies the equivalent
conditions in Proposition 2.5.3. For example, the left a ct i on (r, x) → rx of R on any left
R-module M is a bihomomorphism of R × M into the underlying abelian group M. If
β : A × B C is a bihomomorphism and ϕ : C D is an homomorp h i sm of abelian
groups, then ϕ β : A × B D is a bihomomorph is m . The tensor product of A and
B is an abelian group A
R
B with a bihomomorphism τ of A × B, fr om which every
bihomomorphism of A B can be recovered uniquely in this fashion:
A × B
β
$$
H
H
H
H
H
H
H
H
H
τ
A
R
B
¯
β
//
C.
Definition 2.5.4. Let A be a right R-module and let B be a left R-module. A tensor
product of A and B is an abelian group A
R
B together with a bihomomorphism τ :
A × B A
R
B given by (a, b) → a b, the tensor map, such t hat, for every abelian
group C and bihomomorph is m β : A × B C there exists a unique homomorph is m
¯
β : A
R
B of abeli an groups such that β =
¯
β τ.
If S i s an abelian group and if V and W are S-graded vector spaces, then V W acquires
a uni que S-grading by t h e condition
V
α
W
β
(V W )
α+β
, for α, β S.
Using t h e symbol d
U
for th e degree operator in the space U, we have
d
V W
= d
V
1 + 1 d
W
.
This tensor product grading extends to an arbitrary finite number of tensor factors.
Now, let G be an abelian group and let A be a nonassociative algebra. Then A is a G-graded
algebra if it is G-graded as a vector space, so that
A =
αG
A
α
,
and if A
α
A
β
A
α+β
, for α, β U.
38
2.6 Mod ul e constr ucti o n
Fix an associative Lie algebra g. Let V a g-module and U V a submodule. Then the
quotient vector space V /U becomes a g-module, ca l led the quotient module, by means of
the well defined action
x · (v + U) = x · v + U, for x g, v V.
We have an exact sequence of g- m odules
0 U V V/U 0.
Given two g-modules V
1
and V
2
, their direct sum V
1
V
2
is the g-module which is V
1
V
2
as a vector space, with V
1
and V
2
retaining their original modu l e structures. In particular,
V
1
and V
2
are submodules of V
1
V
2
. The direct sum of any collection {V
i
}
i
of g-m odules
is defi n ed analogously and is denoted by
i
V
i
.
A g-module is called completely reducible or semisimple if it is a direct sum of irreducible
submodules (here the null sum is allowed, so that the zero-dim ensional module is considered
completely reducible). Let G an abelian group and suppose that g is G-graded. A g-module
V is G-graded if it is G-graded as a vector space, so that V =
αG
V
α
, and if
g
α
· V
β
V
α+β
, for α, β G,
i. e., g
α
acts as operators of d egr ee α (s ee (2.8)). Quotients and direct sums of G-g ra d ed
modules are graded (as modules). In case g i s a G-graded Lie algebra, with G a subgroup
of the additive group of F, let d be the degree derivation of g. Then a G-graded g-module
V becomes an Fd g-module when d is required to act as t he degree operator (2.9) on V
(observe t h a t d plays two diffe re nt compatible roles).
The grading of a graded module can be shifted in the following sense. Suppose that G is a
subgroup of an abelian group B a n d that V is a G-graded A- m odule, A a G-graded nonas-
sociative alg eb r a. Let β B. Then for each α G, V
α
can be renamed V
α+β
, gi vi ng V the
structure of a B-graded m odule with A
γ
= 0 for γ B G and V
γ
= 0 for γ B (G + β).
Now let g be a Lie algebra. In p r eparation for constructing the tensor produ ct of g-mod u l es,
we first note that if π
1
and π
2
are two representations of g on V which commute in the
sense th at
[π
1
(x), π
2
(y)] = 0, for all x, y g,
then π
1
+ π
2
is a representation of g on V .
Definition 2.6.1. Given two g-modules V and W , we define the tensor product module
V W to be the vector space V W wi t h the acti on of x g determin e d by the condition
x · (v w) = ( x · v) w + v (x · w), for v V, w W.
39
This is a g-module action because the equations
x · (v × w) = ( x · v) w and x · (v × w) = v (x · w)
clearly define two commuting g-module structures on the vector space V W . The tensor
product of finitely many g-modules is defined a n al og ou sl y. If th e tensor factors are G-graded
modules (G an abelian group), then so is the te n sor prod u ct .
2.7 Induced modules
Let B a subalgebra of an associative algebra A and let V be a B-mod u l e. We denote by
A
B
V the quotient of th e vector space A
F
V by the subspace sp an ned by the el em ents
ab v a b · v, for a A, b B, v V , and we again write a b for the image of
a v A
F
V in A
B
V . The space A
B
V car ri es a natu r al A-m odule structure
determined by the condition
c · (a v) = ca v for a, c A, v V,
and A
B
V is called the A-module induced by the B-mod ule V . It is sometimes denoted
as Ind
A
B
V .
There is a canonical B-module map i : V A
B
V , given by v → 1 v, and Ind
A
B
V has
the foll owing universal property: Given any A-module W a n d B-m odule map j : V W ,
there i s a unique A-module map f : Ind
A
B
V W making the foll owing diagram commute
Ind
A
B
V
f
//
W
V
i
OO
j
;;
w
w
w
w
w
w
w
w
w
w
This property characterizes the A- m odule Ind
A
B
V and the map i up to canonical isomor-
phism. In fact, if I
is another A-module with a B-module map i
: V I satisfyin g the
same condition, the we obtain A-module maps f : Ind
A
B
V I, g : I
Ind
A
B
V . But g f
and the identity map both make the diagram
Ind
A
B
V
//
Ind
A
B
V
V
i
OO
i
99
s
s
s
s
s
s
s
s
s
s
s
commute, so that g f is th e identity map by u niqueness. Similarly, f g is the identity on
I
. If the algebra A and subalgeb r a B are G- gr a d ed and if V is a G-graded B-module (G
an abelian g r ou p ) , t h en the induced module Ind
A
B
V is a G-graded A-module in a natural
40
way.
Given a group G we defi ne its group algebra to be the associative algebra F[G], which is
formally the set of finite linear combinations of elements o f G. That is, F[G] has the set
G as a linear basis, and multiplication in F[G] is simply defined by line ar extension of
multiplication in G. The identity element of F[G] is just the identity element of G.
Definition 2.7.1. A representation of the group G on a vector space V is a group homo-
morphism π : G Aut V . Th e space V is called a G-module or representation of G, and
just as for associative and Lie algebras, we often use the dot notation
g · v = π(g)v for g G, v V.
We have 1 ·v = v and (gh) ·v = g ·(h·v), for all g, h G, and all v V . We have the usual
module-theoretic concepts such as irredu ci b i l ity and equivalence. If π(G) = 1, π i s called a
trivial representation. Given G-m odules V
1
, . . . , V
n
, their tensor product is the vector space
V
1
. . . V
n
, with G-action determined by
g · (v
1
. . . v
n
) = (g · v
1
) . . . (g · v
n
) for g G, v
i
V
i
.
The group G has a natural representation on its own group algeb r a , given by the left mul-
tiplication action. This is called the left regular representati on of G.
Any G-module V becomes a F[G]-module in a canonical way —by extending the map
π : G Aut V by linearity to an algebra homomorphism of F[G] to End V —. In fact,
the G-modules are essentially the same as the F[G]-modules. For exampl e, the left regula r
representation of G corresponds to the left multiplication representation of F[G]. If the
group G is an abelian group written additively, such as the group Z, there can be confusion
as to whether t he symbol a + b means the sum in G or t h e sum in F[G], for a, b G. For
this reason we use t he exponential notation for the elements of G viewed as elements of
F[G] when G is such a group: we write e
a
for the element of F[G] corresponding to a G.
In part i cu l a r, e
0
= 1 and e
a
e
b
= e
a+b
, for a, b G.
Given a subgroup H of a group G and a H-module V , we d efi n e th e G-module induced by
V to be th e G-modul e associated with the induced F[G]-module F[G]
F[H]
V . We write
Ind
G
H
V = F[G]
F[H]
V.
There is a canonical H-module map i : V F[G]
F[H]
V given by v → 1 v, and the
induced module is characterized by the following universal property: Given any G- m odule
W and H-module map j : V W , there is a unique G-module map f : Ind
G
H
V W such
that t h e diagram
Ind
G
H
V
f
//
W
V
i
OO
j
;;
w
w
w
w
w
w
w
w
w
w
41
commutes. It is clear that if X G contains exactly one element from each of the left
cosets gH of H in G, the we have a linear isomorphism
Ind
G
H
V
=
F[X]
F
V.
Here we denote by F[X] the linear span of X i n F[G] (even i f X is not a subgroup) . We shall
construct the analogue for a L i e algebra of the grou p algebra of a group —the universal
enveloping algebra—. First, we construct the tensor algebra T (V ) of a vector space V . For
n 0 define T
n
(V ) to be the n-th tensor power of V , i. e., the vector space
T
n
(V ) = V . . . V

n times
.
Here it is understood that T
0
(V ) = F and T
1
(V ) = V .
Definition 2.7.2. For a vector space V , we defin e the tensor algebra T (V ) by
T (V ) =
n0
T
n
(V ), (2.10)
with t h e associative algebra structure given by t he condition
(v
1
. . . v
m
)(w
1
. . . w
n
) = v
1
. . . v
m
w
1
. . . w
n
T
m+n
(V ).
Then T (V ) becomes a Z-graded associative algebra with T (V )
n
= T
n
(V ), for n 0, and
T (V )
n
= 0, for n < 0. Such algebra is characterized by the following u n i versal property:
given any associative a l geb r a A and linear map j : V A, there exists a unique algebra
map f : T (V ) A for wh i ch the diagram
T (V )
f
//
A
V
i
OO
j
<<
z
z
z
z
z
z
z
z
z
commutes, where i is the inclusion map of V into T (V ). In the same sense that the tensor
algebra is the ‘universal associative algebra over V ’, the symm etr ic algebra S(V ) is the
universal commutative associative algebra over V . To construct it, let I be the ideal of
T (V ) generated by all the elements v w w v, for v, w V , so that I is the line ar
span of the products a(v w w v)b, for a, b T (V ), v, w V . Form the algebra
S(V ) = T (V )/I. Since I is spanned by homogeneous elements, then it is clear t hat S(V )
is a Z-graded commutative algebra of the form
S(V ) =
n0
S
n
(V ), (2.11)
42
where S
n
(V ) = S(V )
n
, called the n-th symmetric power of V , is the image in S(V ) of
T
n
(V ). We have S
0
(V ) = F an d S
1
(V ) = V . The algebra S(V ) is characterized by a
universal property analogous to th e one above, but for li n e ar maps of V int o commutative
associative algebras. Given a basis {v
j
}
jJ
(J a totally ord er ed index set) of V , S(V ) has
a basis con si st i n g of the products V
j
1
···V
j
n
, for n 0, j
J, j
1
. . . j
n
. The space
S
n
(V ) has an obvious basis. If V is G-graded (G an abelian group) then T (V ) and S(V )
acquire unique algebra G-gradings (different from (2.10) and (2.11)) extending the grading
of V . We now turn to universal envelop i n g algebras.
Definition 2.7.3. Given a Lie algebra g, the universal enveloping algebra U(g) is con-
structed as th e quotient associative algebra of T (g) by the ideal generated by the elements
x y y x [x, y], for x, y g. That is
U(g) = T (g)/I.
Clearly, F embeds in U(g). There is a canonical linear map i : g U(g) which is an
homomorphism of Lie algebras, and U(g) is characterized by the following universal prop-
erty: Given any associative algebra A and Lie algebra map j : g A, there is a unique
associative algebra m ap f : U(g) A maki n g the diagram
U(g)
f
//
A
g
i
OO
j
==
z
z
z
z
z
z
z
z
z
commute. In particular, every g-modul e is a U(g) - m odule in a natural way and conversely.
If the Lie algebra g is G-graded ( G an abelian group), th en U(g) becomes a G-graded
algebra in a canonical way via the G-grading of T (g). If the Lie algebra g is abelian, then
U(g) is just the symmetric algebra S(g), and in particular, the map i : g U(g) is an
inclusion and we know a basis of U(g). For a general Lie algebra g, the correspond i ng
result is not trivial, and we need th e following result [116, p.168]:
Theorem 2.7.4 (Poincar´e-Birkhoff- Wi t t) . The canoni cal map i : g U(g) is injective.
Furthermore, let {x
j
}
jJ
(J a totally ordered index set) be a basis of g. Then the universal
enveloping algebra U(g) has a basis consisting of the ordered products x
j
1
···x
j
n
, for n 0,
j
J, j
1
. . . j
n
.
Now we turn to induced Lie algebr a modul es.
Definition 2 .7 .5. Given a subalgebra h of a Lie algebra g, and a h-module V , the g-module
induced by V i s by definition the g-module corresponding to the U(g)-mod u l e
Ind
g
h
V = U(g)
U(h)
V.
43
There is a canonical h-module map i : V Ind
g
h
V , given by v → 1 v, and the induced
module is characterized by the following universal property: For any g-modu le W and
h-module map j : V W , t here is a unique g-module map f : Ind
g
h
V W making the
diagram
Ind
g
h
V
f
//
W
V
i
OO
j
;;
x
x
x
x
x
x
x
x
x
x
commute. If g, h and V are G-graded (G an abelian group), t h en so is Ind
g
h
V , in a canonical
way.
Suppose t hat k and h are subalgebras of g, such that g = k h as vector sp a ces. Then the
Poincar´e-Birkhoff-Witt theorem (Theorem 2.7.4) implies that the linear map
U(k)
F
U(h) U(g)
defined by x y → xy is a linear isomorphism (using a basis of g made u p by bases of k
and h). It follows that the linear m a p
U(k)
F
V U(g)
U(h)
V
defined by x v x v is a linear isomorphism . The acti on of k on Ind
g
h
V carries over
to th e left multiplication action of k on U(k) V .
We mention an important special construction . Sup pose that V is a finite-dim e n si on a l
vector space, with a non -si n g u l ar symmetric bilinear form ·, ·. L et {v
1
, . . . , v
n
} be a basi s
of V and let {v
1
, . . . , v
n
} be the corresponding dual basis of V
, defi n ed by
v
i
, v
j
= δ
ij
, for i, j = 1, . . . , n.
Thus, th e element
ω
0
=
n
j=1
v
j
v
j
V V,
is independent of the choice of basis. In fact, consider the linear isomorph i sm i : V
V
from the dual V
to V determi n e d by ·, ·, and the canonical linear isomorphism j :
End V V
V given by
n
i=1
a
i
v
i
, v →
n
i=1
a
i
(v
i
v) .
Then, ω
0
= ((i 1) j)(1
V
), where 1
V
is denoting the identity in End V . The canonical
image
ω
1
=
n
j=1
v
j
v
j
S
2
(V )
44
of ω
0
in the symmetric square of V is also indepen d e nt of the bas is . In particular, if V
admits an orthonormal basis {e
1
, . . . , e
n
} (for instance, if F is algebraically closed) then
ω
0
=
i
e
i
e
i
and ω
1
=
i
e
2
i
.
2.8 Affine Lie algebras
Let g be a Lie algebra and ·, · a bilinear form on g —a bilinear map from g × g t o F—.
Then · , · is said to be invariant o r g-invariant i f
(i) ( associativity) [x, y], z = x, [y, z], for x, y, z g.
Suppose that ·, · is an invariant symmetric bilin ea r form on g. To the pair (g, ·, ·  )
we shall associate two infinite-dimension al graded Lie algebras
ˆ
g and
˜
g, called the ‘affine
Lie alg eb r as’ .
Let F[t, t
1
] be the commutative associative algebra of Laurent polynomials in an in deter-
minate t. For a Laurent polynomial
f =
nZ
a
n
t
n
, a
n
F
the sum being finite, set f
0
= a
0
. Let d be the derivation
d = t∂
t
(2.12)
on F[t, t
1
]. Note that (df )
0
= 0. Consider the vector space
ˆ
g = g
F
F[t, t
1
] Fc,
where Fc is a one-dimensional sp ace. Ther e is an (alternat i n g ) bilinear map [·, ·] :
ˆ
g×
ˆ
g
ˆ
g
determined by the condition s
[c,
ˆ
g] = [
ˆ
g, c] = 0;
[x f, y g] = [x, y] fg + x, y(df · g)
0
c; (2.13)
for x, y
ˆ
g, f, g F[t, t
1
], or equivalently,
[c,
ˆ
g] = [
ˆ
g, c] = 0;
[x t
m
, y t
n
] = [x, y] t
m+n
+ x, ym δ
m+n,0
c; (2.14)
for all x, y
ˆ
g, m, n Z. It follows directly from th e symme t ry and invariance of ·, · that
ˆ
g is a Lie algebra.
45
Definition 2.8.1. We will call this Lie algeb r a
ˆ
g the affine algebra or the untwisted affine
algebra associated with g and ·, ·.
Give the space g F[t, t
1
] the Lie algebra structure by:
[x t
m
, y t
n
] = [x, y] t
m+n
, for x, y g, m, n Z.
Then, t h er e is an exact sequence of Lie algebras via the canonical maps
0 Fc
ˆ
g g F[t, t
1
] 0,
so that
ˆ
g is a central extension of g F[t, t
1
].
For x g, we shall sometimes write x for the element x t
0
of g F[t, t
1
].
Suppose that g is not assumed to be a Lie algebra, but only a nonassociative algebra under
[·, ·]. Suppose also that the for m ·, · on g is not assumed symmetric or invariant, but only
bilinear. We can repeat th e construction of th e vector space
ˆ
g and of t h e nonas sociative
algebra structure [·, ·] given by (2.1 4 ) . In this case,
ˆ
g is a Lie algebra if and only if g is a
Lie alg eb r a and the form ·, · on g is symmetric and g-i nvariant.
Led d also denote the derivation of
ˆ
g deter m i n ed by
d(c) = 0,
d(x f ) = x df, (2.15)
for x g, f F[t, t
1
]. Form the semidirect product Lie algebra
˜
g =
ˆ
g Fd,
called the extended affine algebra associated wi t h g and ·, ·, or just the affine algebra, i f
no con fu si on is possible. We obtain a natural gradation
˜
g =
nZ
˜
g
n
by considering the eigenspaces
˜
g
n
= {x
˜
g : [d, x] = nx}, n Z,
of ad d. Then, d is th e degree derivation with respect to this grading, and
˜
g =
g t
n
for n = 0
g Fc Fd for n = 0
where we write g t
0
as g. We also have a gradatio n of
ˆ
g,
ˆ
g =
nZ
ˆ
g
n
46
via
ˆ
g
n
=
˜
g
n
ˆ
g.
When h is a subalgebra of g, we shall consider
ˆ
h and
˜
h as suba lg eb r as of
ˆ
g and
˜
g in
the obvious way. We shall also consider the analogue of affinization by ‘twisting’ by an
involution of g.
Definition 2.8.2. An automorphism θ of a Lie algebra (or another algebraic stru ct ure) is
called an involution if
θ
2
= 1.
Let θ be an involution of a L i e algebra g which is also an isometry with respect to the form
·, ·, i. e.,
θx, θy = x, y for x, y g.
For i Z
2
, set
g
(i)
= {x g : θx = (1)
i
x}.
Then,
g = g
(0)
g
(1)
;
[g
(0)
, g
(0)
] g
(0)
, [g
(0)
, g
(1)
] g
(1)
, [g
(1)
, g
(1)
] g
(0)
;
g
(0)
, g
(1)
= 0.
Consider th e algebra F[t
1/2
, t
1/2
] of Laurent polynomials in an indeterminate t
1/2
(whose
square is t), and extend d to a derivation of F[t
1/2
, t
1/2
] via
d : t
n/2
→
n
2
t
n/2
, n Z, (2.16)
and form
l = g
F
F[t
1/2
, t
1/2
] Fc.
The formulas (2.13) and (2.14) again make l into a Lie algeb r a. Let ν the automorphism
of F[t
1/2
, t
1/2
] such that
ν : t
1/2
→ t
1/2
,
and let θ be the automorphism of l determined by
θ :
c → c
x f → θx νf
,
for x g, f F[t
1/2
, t
1/2
]. The formula θ(xf) = θxf will define another automorph is m
of l.
Definition 2.8.3. The twisted affine algebra
ˆ
g[θ] is then the subalgebra
ˆ
g[θ] = {x l : θx = x}
of fixed points of θ in l.
47
We have
ˆ
g[θ] = g
(0)
F[t, t
1
] g
(1)
t
1/2
F[t, t
1
] Fc.
We can again adjoin the derivation d determined by (2.16 ) as in (2.15) and set
˜
g[θ] =
ˆ
g[θ] Fd,
the extended twisted affine algebra associated with g, ·, · and θ. Th e eigenspaces of ad d
make
ˆ
g[θ] and
˜
g[θ] into
1
2
Z-graded Lie algeb r as. Not e that if θ = 1, then
˜
g[θ] degenerates
to th e untwisted affine algebra
˜
g.
Finally, we remark that the process of twisted affinization can be extended to any auto-
morphism of fini t e order of g, which is an isometry with respect to ·, ·.
48
Chapter 3
The Root space decomposition
Now we will consi d e r a class of Lie algebras (t h e complex semisimple ones), that their
representations can be described, similarly to sl
3
(C), by a ‘theorem of the highest weight’.
We develop the structures needed to state the theorem of the highest weight. Although
this chapter could be understood simply as a description of the structure of semisimple
Lie algebras, without any mention of representation theory, it is helpful t o have the repre-
sentations in mind. The represe ntation theory, especially in light of sl
n
(C), motiva t es the
notions of Cartan subalgebras, roots, and the Weyl group. See [179] or [91] a more detailed
treatment of these topics.
3.1 Representations
Definition 3.1.1. Let G be a matrix Lie group. Then, a (finite-dimension al ) complex
representation of G i s a Lie group h om om o r p h i sm Π : G GL
n
(C) (with n 1 ) or, in
other words, a Lie group homomorphi sm Π : G GL(V ), where V is a finite-dimensi on a l
complex vector space (wit h dim V 1). A finite-dimensional real representation of G
is a Lie group homomorphism Π of G into GL
n
(R) or into GL(V ), where V is a finite-
dimensional real vector space.
Definition 3.1.2. If g is a real or complex Lie algebra, th e n a fin i t e-dimensional complex
representation of g is a Lie al gebra homom or p h i sm π : g gl
n
(C) (o r into gl(V )), wh er e
V i s a finite-dimensional complex vector space. We can define a real representation of g in
a similar way. If Π or π is a on e- t o- one homomorphism, then the representation is called
faithful.
One should think of a representation as a linear action of a group or Lie algebra on a vector
space (since to every g G, there is associated an operator Π(g), which acts on the vec-
tor space V ). If g is a real Lie algebra, we will consi d er mainly complex representations of g.
Definition 3.1.3. Let Π be a finite-dimensional real or com p l ex representation of a m a tr i x
Lie group G, acting on a space V . A subspace W of V is called invariant if Π(A)w W ,
for all w W and all A G. An invariant subspace W is call ed nontrivial if W = {0}
and W = V . A r ep r esentation with no nontrivial invariant subspaces is called irreducible.
The terms i nvariant, nontrivial, and irreducible are defined analogously for representations
of Lie algebras.
49
Definition 3.1.4. Let G be a matrix Lie group, let Π be a representation of G acting on
the space V , and let Σ be a representation of G acting on the sp ace W . A linear map
ϕ : V W is called an intertwining map of representations if
ϕ
Π(A)v
= Σ(A)ϕ( v),
for all A G and all v V . The analogous proper ty defines intertwining maps of repre-
sentations of a Lie algebra. If ϕ is an intertwining map of representations and, in addition,
ϕ is invertible, then ϕ is said to be an equivalence of representati on s . If there exists an
isomorphism between V and W , then the representations are said to be equivalent.
If G is a matrix Lie group with Lie algebra g and Π is a (finite-dimension a l real or complex)
representation of G, acting on the space V , then there is a un i qu e representation π of g
acting on the same space, such that
Π(e
x
) = e
π(x)
, for all x g.
This representation π can be comp u t ed as
π(x) =
d
dt
Π(e
tx
)
t=0
,
and satisfies π( AxA
1
) = Π(A)Π(x)Π(A)
1
, for all x g and all A G. We state the
following two propo si ti o n s. For a proof of them, see [91, p.93].
Proposition 3.1.5. 1. Let G be a connected matrix Lie group with Lie algebra g. Let
Π be a representation of G and π the associated representation of g. Then, Π is
irreducible if an d only if π is irreducible.
2. Let G be a connected matrix Lie group, let Π
1
and Π
2
be representations of G, and
let π
1
and π
2
be the associated Lie algebra representations. Then, π
1
and π
2
are
equivalent if and only Π
1
and Π
2
are equivalent.
Proposition 3 .1 .6. Let g be a real Lie algebra and g
C
its complexification. Then, every
finite-dimensional complex representation π of g has a unique extension to a complex-linear
representation of g
C
, also denoted π and given by
π(x + iy) = π(x) + (y), for all x, y g.
Furthermore, π is irreducible as a representation of g
C
if and only if it π irreducible as a
representation of g.
We give some examples of representations:
Example 3.1.7. (The st andard representation)
A matrix Lie group G is, by defini ti o n , a su bset of some GL
n
(C). The inclusion map of G
into GL
n
(C) (i. e., Π(A) = A) is a r ep r es entation of G, called the standard representation
50
of G. If G happens to be contained in GL
n
(R) or GL
n
(C), then we can think of the stan-
dard representation as a real representation if we prefer. Thus, for example, the standard
representation o f SO
3
(C) is the o n e in which SO
3
(C) acts in the usual way on R
3
and the
standard representation of SU(2) is the one in which SU(2) acts on C
2
in the usual way.
If G is a subgroup of GL
n
(R) or GL
n
(C), then its Lie algebra g will be a subalgebra of
gl
n
(R) or gl
n
(C). The inclusion of g into gl
n
(R) or gl
n
(C) is a representation of g, called
the standard representation.
Example 3.1.8. (The t r i vi al representation)
Consider the one-d im e n si on a l complex vector space C. Given any matrix Lie group G, we
can define the trivial representation of G, Π : G GL
1
(C), by the formula
Π(A) = I, for all A G.
Of course, this is an irreducib l e representatio n , since C has no nontrivial subspaces, and
thus no nontrivial invariant subspaces. If g is a Lie algebra, we can also define the trivial
representation of g, π : g gl
1
(C), by
π(x) = 0, for all x g.
This is an irreducible representation.
Example 3.1.9 (The adjoint representation). Let G be a ma tr i x Lie group with Lie algebra
g. We define the adjoint mapping Ad : G GL(g) by the formula
Ad
A
(X) = AXA
1
.
Since Ad is a Lie group homomorphism into a group of invertible operators, we see that,
in fact, Ad is a repr ese ntation of G, actin g on the space g. We call Ad the adjoint repre-
sentation of G. The adjoint representation is a real representation of G (if g is a complex
subspace of C
n×n
, th en we can think of t h e adjoint representation as a complex r ep r es en -
tation). Similarly, i f g is a Lie algebra, we have the adjoint map ad : g gl(g) defined by
the formula (see Chapter 2)
ad
x
(y) = [x, y].
We know that ad is a Lie algebra homomorphism and is, therefore, a representation of g,
called the adjoint representation. In the case that g is the Lie algebra of some matrix Lie
group G, Ad and ad are related by e
ad
x
= Ad
e
x
.
Now we will discuss a cla ssi cal example, n am el y the irreducible complex representations of
the Lie algebr a su(2). This computation is im portant for several reasons. In the first place,
su(2)
=
so(3) and the representations of so(3) are of physical significance, particularly
51
in quantum mechanics [91]. In the second place, the representation theory of su(2) is an
illuminating example of how one uses relations to determine the representations of a Lie
algebra. Also, in determining the representations of semisimple Lie a lg eb r as (Section 3.2),
it usually uses the representation of su(2).
Example 3.1.10. (The Ir r educible Repr esentations of sl
2
(C))
Every finite-dimensi onal complex r epresentation π of su(2) extends to a comp le x- l inear
representation (also called π) of the com p l exi fi ca t io n of su(2), namely sl
2
(C). The extensio n
of π to sl
2
(C) is ir r ed u c ible if and only if the original representation is irreducible. We will
use the following basis for sl
2
(C):
h =
1 0
0 1
, x =
0 1
0 0
, y =
0 0
1 0
;
which have the commutation relations
[h, x] = 2 x,
[h, y] = 2y,
[x, y] = h.
If V is a (finite-dimensional complex) vector space and A, B and C are operators on V
satisfying
[A, B] = 2B,
[A, C] = 2C,
[B, C] = A,
then because of the skew symmetry and bilineari ty of brackets , the linear map π : sl
2
(C)
gl(V ) satisfying
π(h) = A, π(x) = B, π(y) = C,
will be a representation of sl
2
(C).
We state an important result we will use later. For a proof see [179 , p.268].
Theorem 3.1.11. Suppose π is any finite-dimensional, complex-linear representation of
sl
2
(C) acting on a space V . Then, we have the following results:
1. Every eigenvalue of π(h) must be an integer.
2. If v is a nonzero element of V such that π(x)v = 0 and π(h)v = λv, then λ is
a non-negative integer. Furthermore, the vectors v, π(y)v, . . . , π(y)
λ
v are linearly
independent and their span is an irreducible invariant subspace of dimension λ + 1.
52
Analogously to the previous example, we can think of study the representations of su(3) by
studying the representations of its complexification sl
3
(C). We will discuss this particu l ar
case in Section 3.9. In general, studying the irreducible representations of su(n) is equiva-
lent to st udying the ir r educible (complex-linear) representations o f sl
n
(C). Passing to the
complexified Lie algebra m akes our comput a t io n s easier, and we can find a nice basis for
sl
n
(C) tha t has no counterpart among the b ase s of sl(n).
3.2 Complete Reducibility and Semisi mpl e algeb ra s
Definition 3.2.1. A finite-dimensional representation of a group or Lie algebr a is said to
be completely reducible if it is isomorphic to a direct sum of a finite number of irreducible
representations. A group or Lie algebra is said to have the complete reducibility property if
every fin it e- d i m ensional repres entation of it is completely reducible.
The complete reducibility property is a very special on e that most groups and Lie algebras
do n ot have. If a group or Lie algebra does have the complete reducibili ty property, then
the study of its representations redu ces to the study of its irreducible representations, which
simplifies the an al ys is considera b l y. The following results are useful [91, p.1 19]
Proposition 3.2.2. If V is a completely reducible representation of a group or Lie algebra,
then the following properties hold:
1. Every invariant subspace of V is completely reducible.
2. Given any invariant subspace U of V , there is another invariant subspace
˜
U such that
V = U
˜
U.
Proposition 3.2.3. Let G be a matrix Lie group. Let Π be a finite-dimensional unitary
representation of G, act ing on a finite-dimensional real or complex Hilbert space V . Then,
Π is completely reducible.
If Π is a repr esentation of a finite gr ou p G, acting on a space V , we can choose an ar b i t r ar y
inner product ·, · on V . Then, we can defin e a new inner product ·, ·
G
on V by
v
1
, v
2
G
=
gG
Π(g)v
1
, Π(g)v
2
. (3.1)
Furthermore, if h G, then
Π(h)v
1
, Π(h)v
2
G
=
gG
Π(g)Π(h)v
1
, Π(g)Π(h)v
2
=
gG
Π(gh)v
1
, Π(gh)v
2
However, as g rang es over G, so d oes gh. Thus, in fact, Π(h)v
1
, Π(h)v
2
G
= v
1
, v
2
G
;
that is , Π is a uni t ar y representation with respect to the inner product ·, ·. Thus , Π is
isomorphic to a direct sum of irreducibles, by Proposition 3.2.3 and we have
53
Proposition 3.2.4. Every finite group has the complete reducibility property.
There is a variant of the above argument which can be used to prove the following result
[116]:
Proposition 3.2.5. If G is a compact matrix Lie group, G has the complete reducibility
property.
The argument below is sometimes called ‘Weyl’s Unitarian trick’. Its proof requires the
notion of Haar measure (see, for example, [116]). A left H aar measure on a matrix Lie group
G is a nonzero measure µ on the Borel σ-alg eb r a in G with the fol lowing two properties:
It is locall y finite (i. e., every point i n G has a neighborhood with finite measure),
It is left-transl at i on invariant: µ(gE) = µ(E), for al l g G and all Bor el sets E G.
It is a fact , wh i ch we cannot prove here, that every matrix Lie group has a left Haar measure
and t h at this measure is unique up to multiplication by a constant. One can analogously
define right Haar measure, and a similar theorem holds for it. Left Haar measure and right
Haar measure may or may not coincide (a group for which they do is cal l ed unimodular).
Now, the key fact for our purpos e is t h a t le ft Ha ar measure is finite if and only if the group
G is compact. Suppose, then, that Π is a finite-dimensional representation of a compact
group G acting on a space V . Let  · , · be an arbitrary inner product on V and d efine a
new in n er produ ct ·, ·
G
on V (analogous to (3.1)) by
v
1
, v
2
G
=
G
Π(g)v
1
, Π(g)v
2
(g), (3.2)
where µ is a left Haar measure. Again, it is possible to check that ·, ·
G
is an inn er product.
Furthermore, if h G, then by the left-invariance of µ,
Π(h)v
1
, Π(h)v
2
G
=
G
Π(g)Π(h)v
1
, Π(g)Π(h)v
2
(g)
=
G
Π(gh)v
1
, Π(gh)v
2
(g) = v
1
, v
2
G
.
So, Π is a unitary representation with respect to  · , ·
G
and t hus completely reducible (note
that t h e integral in (3.2) is convergent beca u se µ is finite).
Recall from Chapter 2 t h a t if g is a complex L i e algebra, then a n ideal in g is a co m p l ex
subalgebra h of g with the property that for all x g and all h h, we have [x, h] h.
Recall also that a comp l ex Lie algebra g is called indecomposable if the only ideal s in g
are g and (0) . A complex Lie algebra g is call ed simple if g is indecomposable and dim g 2.
There is an analogy between finite-dimensional Lie algebras an d fin i t e group s. Subalgebras
in the Lie algebra setting are the analogs of subgroups in the finite group setting , and ideals
54
in the Lie algebra setting are the analogs of normal subgroups in the fin i t e group set t i n g. In
this ana lo gy, th e one-dimension a l Lie algebras (which are precisely the Lie algebras having
no nontrivial subal g eb r as) are the analogs of the c ycl i c groups of prime order (which are
precisely the gr ou ps having no nontrivial subgroups). However, there is a discrepancy in
terminology: cyclic groups of prime order a r e called simple but one-dimensional Lie algebras
are not called simple. This terminological convention is importa nt to bear in mind in the
following definition.
Definition 3.2.6. A complex Lie algebra is calle d reductive if i t is isomorphic to a direct
sum of indecomposable Lie algeb r as . A compl ex Lie algebra is called semisimple if it
isomorphic to a direct sum of simple Lie algeb r a s.
Note that a reductive Lie algebra is a direct sum of indecomposable algebra s, which are
either si m p l e or one-dimensional commutative. Thus, a reductive Li e algebra is one that
decomposes as a direct sum of a semisimple algebra (coming from the sim p l e terms in the
direct sum) and a commutative al ge b r a (coming from the one-dimensional terms in the
direct sum). We will assume that the complex semi si m p l e Lie algebras we study are given
to us as subalgebras of some gl
n
(C), since by Ado’s Theorem 2.3.7 every finite-dimensional
Lie algebra has a faithful finite-dimensional representation. In fact, for semisimple Lie
algebras, the adjoi nt representation is always faithful, as is shown in [91, p.158]
Proposition 3.2. 7. A complex Lie algebra g is reductive precisely if the adjoint represen-
tation is completely reducible.
In fact, the complexification of the Lie algebra of a connected compact matrix Lie gr ou p
is reductive. This follows from the above proposition and the p r operty that connected
compact groups have the complet e redu cibility property (P r oposition 3.2.5). Note that
the Lie algebra of a compact Lie group may be only reductive and not semisimple. For
example, the Lie algeb r a of S
1
is on e dimensional and, thus, not semisimple. We have the
following characterization result of semisimp l e Lie algebras (see [179, p.348])
Theorem 3.2.8. A complex Lie algebra is semisimple if, and only if, it is isomorphic to
the complexification of the Lie algebra of a simply-connected compact matrix Lie group.
We have already seen that if g is the complexification of the Lie algebra of a compact
simply-connected group K, then g is reductive, even if K is not simply connected. Thus
g = g
1
g
2
, wit h g
1
semisimple and g
2
commutative. It can be shown that the Lie algebra k
of K decomposes as k = k
1
k
2
, where g
1
= k
1
+ ik
1
and g
2
= k
2
+ ik
2
. Then, K dec om poses
as K
1
×K
2
, where K
1
and K
2
are simply connected and where K
2
is commutative. However,
a simply-connected commutative Lie g r ou p is isomorphic to R
n
, which is noncompact for
n 1. Thus, the compactness of K means that k
2
= {0}, in which case g
2
= {0} and
g = g
1
is semisi m ple.
For the other direction, given a complex sem i si m p l e Lie algebra, we must find the correct
real form whose correspondi n g simply-con n ect ed group is compact (c. f. [179]).
55
Definition 3.2.9. If g is a complex semisimple Lie algebra, then a compact real form of
g is a real subalgeb r a k of g with the property that every x g can be written uni quely
as x = x
1
+ ix
2
, with x
1
and x
2
in k and such that there is a compact simply-connected
matrix Lie group K
such tha t the Lie algebra k
of K
is isomo r p h i c to k.
We have the following important fact [91, p . 159 ].
Proposition 3.2.10. Let g be a complex semisimple Lie algebra. If g is a subalgebra of
gl
n
(C) and k is a compact real form of g, then the connected Lie subgroup K of GL
n
(C)
whose Lie algebra is k is compact.
In particular, every complex semisimple Lie algebra has the complete reducibility property.
This last statement holds because the representations of g are in one-to-one correspondence
with the representations of K, and compact groups h ave the complete reducib il i ty prop-
erty (Proposi t i on 3.2.5). Actually, only the semisimple ones have the complete reducibility
property, and thus, complete redu c ibility is sometimes taken a s the definition of sem i si m -
plicity for Lie algebras. For an algebraic proof of complete red u ci b i l i ty of semisimp l e Lie
algebras, see [100]). Up to now, we have considered only complex semisimple Lie algebras,
since these are the ones whose representations we will co n si d er . Nevertheless, we can d efi ne
the terms ideal, indecom posable, simple, r ed uctive, and semisimple for real Lie algebras in
precisely the same way as for the comp l ex case.
Let us consider some examples of Lie algebr as that are red u c t ive or semisimple. The
following table lists some of the co m p l ex Lie algebras that we have encountered alr eady that
are either reductive or semisimple (see [22]). Her e, ‘reductive’ means actually ‘reductive
but n o t semisimple’ .
Group Reductive/Semisimple
sl
n
(C) (n 2) semisimple
so
n
(C) (n 3) semisimple
so
2
(C) reductive
gl
n
(C) (n 1) reductive
sp
n
(C) (n 1) semisimple
Table 3.1: Semi si m p l e proper t i es of some classical complex Lie al geb r a s.
The other Lie algebras we have examined, such as the Lie algebras of th e Heisenberg group,
are nei t h er reductive nor semisimple.
56
Group Reductive/Semisimple
su(n) (n 2) semisimple
so(n) ( n 3) semisimple
so(2) reductive
sp(n) ( n 1) semisimple
sp
n
(R) (n 1) semisimple
sl
n
(R) (n 2) semisimple
gl
n
(R) (n 1) reductive
Table 3.2: Sem i si m p l e propert i es of some classical real Lie algebr a s.
3.3 Cartan subalgebras
Definition 3.3.1. If g is a comp l ex semisimple Lie alge b r a, then a Cartan subalgebra of g
is a complex subspace h of g with the following properties:
(i) For all h
1
and h
2
in h, [h
1
, h
2
] = 0.
(ii) For all x g, if [h, x] = 0 for all h h, then x is in h.
(iii) For all h h, ad h is diagonalizable.
Condition (i) says that h is a commutative subalgebra of g. Condition ( ii ) says that h is
a maximal commutative subalgebra, i. e., not contained in any larger commutative sub-
algebra. Condition (iii) says that each ad h (h h) is diagonalizable . Since the h’s in h
commute, the ad h’s also commute, and thus they are simultaneously diagonalizable. (It is
a standard result in linear algebra that any commuting fami l y of diagonalizable matrices is
simultaneously di ago n al i zable; see [97]) . Of cour se , the definition of a Cartan subalgebra
makes sense in any Lie algebra, semisimple or not. However, if g is not semisimple, then g
may not have any Cartan su b a l geb r a s. Even i n the semisimple case we must pr ove tha t a
Cartan subalgebra exists (see [91, p.163]).
Proposition 3.3.2. Let g be a complex semisimple Lie algebra, let k be a compact real form
of g, and let t be any maximal commutative subalgebra of k. Define h g to be h = t + it.
Then, h is a Cartan subalgebra of g.
Note that k (or any other Lie algebra) contains a maximal commutative subalgebra. After
all, let t
1
be any one-dimensional subspace of k. Then, t
1
is a commutative subal geb r a of k.
If t
1
is maximal, then we are done; if not, then we choose some commutative subalgebra t
2
properly containing t
1
. Th en , i f t
2
is maxim al , we are done, and if not, we choose a com-
mutative subalg eb r a t
3
properly containing t
2
. Since k is fini t e dimensional, this process
cannot go on forever and we will eventually get a maximal commutative subalgebra.
57
It is possible to prove that every Cartan subalgebra of g arises as in Proposition 3.3.2 (for
some compact real form k and some maximal co m mutative subalgebra t of k) and also th at
Cartan subalgebras are unique up to conjugation. In particular, all Cartan subalgebra s of
a given complex semisimple Lie al gebra have the same di m en si o n . In light of this result,
the foll owing definiti on makes sense.
Definition 3.3.3. If g is a complex semisimple Lie algebra, then the rank of g is t he
dimension of any Cartan subalgebra.
3.4 Roots and Root Spaces
From now on we assume that we have chosen a compact real form k of g and a maximal
commutative subalgebra t of k, and we consider the Cartan su b a lg eb r a h = t + it. We
assume also that we have chosen an inner product on g t h at is invariant under the adjoint
action of K and that takes real values on k.
Definition 3.4.1. A root of g (relative to the Cartan subalgebra h) is a nonzero linear
functional α on h such that there exists a nonzero element x of g with
[h, x] = α(h)x,
for all h h.
The set of all roots is denoted by Φ. The cond it i o n on x says that x is an eigenvector for
each ad h, with eigenvalue α(h). Note that i f x is actually an eigenvector for each ad h with
h h, then the eigenvalues must depend lin ea rl y on h. That is why we insist that α be a
linear functional on h. So, a root is just a (nonzero) collection of simultaneous eigenvalues
for the ad h’s. Note that any elem ent of h is a simultaneous eigenvector for all the a d h’s,
with all eigenvalues equal to zero, but we only call α a r oot if α i s nonzero. Of cours e, for
any root α, some of the α(h)’s may be equal to zero; we just require that not all of them
be zero. Note tha t th e set of li n ear functionals on h that are imaginary on t forms a real
vector space whose real dimension equals the complex dimension of h. If t
denotes the
space of real-valued linear functionals on t, then the r oots ar e contained in it
h
.
Definition 3.4.2. If α is a root of the Lie algebra g (relati vely to the subalgebra h), then
the root space g
α
is the space of all x g for which [h, x] = α(h)x, for all h h. An
element of g
α
is called a root vector (for the root α).
More generally, i f α is any element of h
, we define g
α
to be the space of all x g for which
[h, x] = α(h)x, for all h h (but we do not call g
α
a root space unless α is actually a
root). Taking α = 0, we see that g
0
is the set of all elements of g th at commute with every
element of h. Since h is a maximal commutative subalgebra, we conclude t h a t g
0
= h. If
α is not zero and not a root, then g
α
= {0}. Now, since h is commutative, the operators
58
ad h, h h, all commute. Furthermore, by th e definition of Cartan subalgebra, each ad h,
h h, is diagonalizable. It follows that the ad h’s, are simultaneously diagonalizable. As a
result, g can be decomposed as the direct sum of h and the root spaces g
α
:
g = h
αΦ
g
α
.
This means that every el em ent of g can be written uniquely as a sum of an element o f h
and on e element from each root space g
α
.
We resume some elementary properti es of roots. You usually can find a proof of theses
properties in almost books abo u t r epresentation theory. See for example [91] or [59] for the
proofs.
Proposition 3.4.3. (i) For any α and β in h
, [g
α
, g
β
] g
α+β
.
(ii) If α h
is a root, then so i s α.
(iii) If α is a root, then the only multiples of α that are roots are α and α.
(iv) The roots span h
.
(v) If α is a root, then the root space g
α
is one dimensional.
(vi) For each root α, we can find nonzero elements x
α
g
α
, y
α
g
α
, and h
α
h such
that
[h
α
, x
α
] = 2x
α
, [h
α
, y
α
] = 2y
α
, [x
α
, y
α
] = h
α
.
The element h
α
is unique, i. e., in dependent of the choice of x
α
and y
α
.
Last point of the p r o position above tells us that x
α
, y
α
, and h
α
span a subalgebra of g
isomorphic to sl
2
(C). The elements h
α
of h are called the co-roots. Their properti es are
closely related to the properties of the roots themselves.
Given any linear functional α h
(not n eces sar i l y a root), there exists a unique element
h
α
h such that
α(h) = h
α
, h,
for all h h, where we take the inner product to be linear in the second factor. The map
α → h
α
is a one-to- o n e and onto correspondence between h
and h. However, this correspon-
dence is not linear but ra ther conjugate-linear, since the inner product is conjugate-linear
in the first factor (where h
α
is). It is convenient to permanently identify each root α h
with the corresponding el em ent h
α
h. Having done t his, we then omit the h
α
notation
and d en ot e that element of h simp ly as α.
The reader can find some more proper t i es and relations of roots and co-roots in [91]. We
only mention that if α h is a root in the sense o f last paragrap h and h
α
is the corresponding
co-root, then α and h
α
are rela t ed by the formulas
h
α
=
2α
α, α
, α =
2h
α
h
α
, h
α
. (3.3)
59
The real content of this proposition is that once we use the inner product to identify h
with h (so that the root s and co-roots now live in t h e same space), α and h
α
are multiples
of one another. Once this is known, the normalizat i on is determined by the conditio n
that α, h
α
= 2, which reflects that [h
α
, x
α
] = 2x
α
. Observe t h a t both formulas ( 3 .3 ) are
consistent with the relation α, h
α
= 2. We conclude with the following [91, p.173]
Theorem 3.4.4. For all roots α, β h (in the notation above), the quantities
2
α, β
α, α
and 2
h
α
, h
β
h
α
, h
α
are integers and, furthermore,
2
α, β
α, α
= 2
h
α
, h
β
h
α
, h
α
.
3.5 The Weyl Group
We use here the comp act - gr oup approach to defining the Weyl group, as opposed to th e
Lie algebra approach. The compact-group approach makes certain aspects of the Weyl
group more transparent. Never t h el ess , th e two approaches are equivalent. We continue
with the setting o f the previous section. Thus, g is a complex semisimple Lie algebra given
to us as a subalgebra of some gl
n
(C). We have chosen a compact real form k of g and
we let K be the compact subgroup of GL
n
(C) whose Lie algebra is k. We have chosen a
maximal commutative subalgebra t of k, and we work with the associated Cartan subal gebra
h = t + it. We have chosen an inner product on g that is invariant un der the adjoint action
of K and that takes real values on k. Consid er the following two subgroups of K:
Z(t) = {A K : ad A( h) = h, h t},
N(t) = {A K : ad A(h) t, h t}.
Clearly, Z(t) is a normal su bgroup of N(t). If T is the con n ect ed Lie subgroup of K with
Lie algebra t, then T Z( t) , since T is generated by elements of the form e
h
with h t.
It tur ns out that, in fact, Z(t) = T . See [22].
Definition 3.5.1. The Weyl group for g is t h e quotient group W = N(t)/Z(t).
We can define an action of W o n t as follows. For each element w W , choose an element
A of the corresponding equivalence class in N(t). Then for h t we define the act io n of w
on h by
w · h = ad A(h).
In fact, this action is well defined (i. e., independent of the choice of A in a given equivalence
class). Since h = t+it, each linear transformation of t extends uniquely to a complex-linear
transformation of h. Thus, we also think o f W as acting on h. If w is an element of the
60
Weyl group, then we write w · h for t he action of w on an element h of h. It can be seen
that W is isomorphic to the group of linear transformations of h that can be expr ess ed as
ad A for some A N( t). The following states basic properties of t h e Weyl group [91, p.174]
Proposition 3.5.2. 1. The inn er product ·, · on h is invariant under the action of W .
2. The set Φ h of roots is invariant under the action of W .
3. The set of co-roots is invariant under the action of W , and w · h
α
= h
w·α
, for all
w W , α Φ.
4. The W eyl group is a finite group.
We state some an important property, leading to a ‘dual ’ nature of roots [91, p.178]:
Proposition 3.5.3. For each root α, there exists an element w
α
of W such that
w
α
· α = α
and such that
w
α
· h = h,
for all h h with α, h = 0.
Note that since h
α
is a multiple of α, saying w
α
· α = α is equivalent to saying that
w
α
·h
α
= h
α
. The linear operator corresponding to the action of w
α
on h is ‘the reflection
about the hyperplane perpendicular to α’. This means that w
α
acts as the identity on the
hyperplane (of codimen si on one) perpendicular to α and as minus the identity on the span
of α. We can work out a formula for w
α
as follows. Any vector β can be decomposed
uniquely as a multiple of α plus a vector orthogonal to α. This decomposition is given
explicitly by
β =
α, β
α, α
α +
β
α, β
α, α
α
, (3.4)
where the second term is indeed orthogona l to α. Now, t o o b t ai n w
α
·β, we should change
the sign of the part of β parallel to α and leave alone the part of β that is orthogonal to α.
This means that we change the sign of the first ter m on the right-hand side of (3.4), giving
w
α
· β = β 2
α, β
α, α
α. (3.5)
We now have another way of thinking about the quantity 2
α,β
α,α
in Theorem 3.4.4: it is the
coefficient of α in the expression for w
α
·β. So, we can r e- ex p r ess Theorem 3.4.4 as follows.
Corollary 3.5.4. If α and β are roots, then β w
α
· β is an integer multiple of α.
Finally, we state a useful characterization of the Weyl group [116, p.208]:
61
Theorem 3.5.5. The W eyl group W is gen erated by the elements w
α
as α ranges over all
roots.
That is to say, th e smallest subgroup of W that contains all of the w
α
’s is W itself. This is
somewhat involved to prove and we will not do so here; see [22] or [116]. In the Lie algebra
approach to the Weyl group, the Weyl group is defined as the set of line ar transform at i ons
of h generated by the reflections w
α
. Theorem 3.5.5 shows that the Lie algebra definition
of the Weyl group gives the same group as the compact-group approach.
3.6 Root Systems
In the previous section we have esta b l is h ed several properties of roots. For example, we
know that the roots are imaginary on t, which, after transferring the roots from h
to h,
means that t h e roots live in it h. The inner product ·, · was constructed to t ake real
values on k, and hence on t. The inner product th en also takes real val u e s on it, since
ix, iy = (i)iw, y = x, y. So, the roots live in the real inner-product space E = it.
From Proposition 3.4.3 we know that the roots span it and that if α is a root, then α
is the only other multiple of α also root. Furthermore, Theorem 3. 4. 4 tell us that for any
roots α and β, the number 2
α,β
α,α
is an integer. Finally, we have established that the roots
are invariant under th e action of the Weyl group, and Theorem 3. 5. 5 tells us that the
Weyl group contains the refl ect i on about the hyperp l an e orthogonal to each root α. We
summarize these results in the following theorem.
Theorem 3.6.1. The roots of g form a finite set of nonzero elements of a real inner-product
space E and have the following properties:
1. The roots span E.
2. If α is a root, then α is a root and the only multiples of α that are roots are α and
α.
3. If α is a root, let w
α
denote the linear transformation of E given by w
α
·β = β2
α,β
α,α
α.
Then, for all roots α and β, w
α
· β is also a root.
4. If α and β are roots, then the quantity 2
α,β
α,α
is an integer.
Definition 3.6.2. Any collection R of vectors in a finite-dim e n si on a l real inner-product
space h aving these properties of Theorem 3.6.1 is call ed a root system.
The Weyl grou p for a root system R is the group of linear transformations of E generated
by the w
α
’s. Note that it em 4 is equiva le nt to saying that β w
α
· β must be an integer
multiple of α for all roots α and β. We have also established certain important properties
of the root spaces that are not properti es of the roots th em se lves, namely that each root
space g
α
is on e dimensi on a l and that out of g
α
, g
α
, and [g
α
, g
α
], we can form a subal geb r a
62
isomorphic to sl
2
(C). Finally, we claim that the co-roots h
α
themselves form a root system.
Theorem 3.4.4 tells us that the co-roots satisfy property 4 and Proposition 3.5.2 tells us
that the set of co-roots is invari ant under the Weyl group a n d hence, in particular, under
the refle ct io n s w
α
. However, note that since h
α
is a multiple of α, the refl ect i o n generated
by h
α
is the same a s the reflecti on gen er at ed by α. Thus, the set of co-roots satisfies prop-
erty 3. Properties 1 and 2 for the co-roots follow from the corresponding properties for th e
roots, since ea ch h
α
is a multiple of α. The set of co-roots is called the dual root system
to t h e set of roots. See Section 3.9 for more information on root systems, including some
pictures.
In the next section, we will present the irreducible representations of g in terms of a ‘highest
weight’. What we need is simply some consistent not i on o f higher and lower that will allow
us to divide the root vectors x
α
into ‘raising operato rs ’ and ‘lowering oper a t or s’ . This
should be done in such a way that the commut at o r of two raising oper at o rs is, again, a
raising operator and not a lowering operator. This means that we want to divid e the roots
into two groups, one of wh i ch will be called ‘positive’ and the other ‘negative’. This should
be done is s u ch a way that if th e sum of positive root s is again a root, that root shoul d
be positive. There is no un i que way to make the division into positive and negative; any
consistent div i si on will do. The uni q u en e ss theorems of th e next section show that it does
not reall y matter which choice we make. The following definition and theorem shows that
it is possibl e to make a good choice.
Definition 3.6.3. Suppose that E is a finite-dimensional real inner-product space and that
R E is a root system. Then, a base for R is a subset = {α
1
, . . . , α
r
} of R such that
forms a basis for E as a vector space and such that for each α R, we have
α = n
1
α
1
+ . . . + n
r
α
r
,
where n
j
Z and either all n
j
0 or all n
j
0.
Once a base has been chosen, the α’s for which n
j
0, j, are called t h e positive roots
(with respect to the given choice of ∆) and t h e α’s with n
j
0, j, are cal l ed the negative
roots. The elements of are called the positive simple roots. We will d en o t e R
+
the set
of all positive root s, and R
the set of all negative roots, so then R is the disjoi nt union
R = R
+
R
. To be a ba se (in the sense of root systems), R must in p ar t i cu l a r be
a basis for E in the vector sp ace sense. In addition, the exp a n si on of any α R in terms
of the elements of must have integer coefficients and all of the nonzero coefficients must
be of the same sign.
3.7 Integral and Dominant integral elements
Definition 3.7.1. An element ω of h is called an in tegral ele m ent i f ω, h
α
is an integer,
for each root α.
63
As explained in next, the integral elements are preci sel y the elements of h that arise as
weights of finite-dimensional rep r ese ntations of g. In fact, we can prove that the set of
integral elements is invariant under the action of the Weyl group. Checking that (ω, h
α
)
is an i nteger for every root α is a rather tiresome process. Fortunately, it suffices to check
just for the positive simple roots: if ω is an element of h for which ω, h
α
is an integer
for all positive simple roots α, then ω, h
α
is an integer for al l roots α, and thus ω is an
integral element.
Recalling the expression ( 3. 3) for h
α
in terms of α, we may restate a characterizat i on of
integral elements as follows: an element ω h is integral if and on l y if
2
ω, α
α, α
is an integer for each positive simple root α. In particular, every root is an integral element.
Recall now from elementary linear algebra that if ω and α are any two elements of an inner-
product space, then th e orthogonal projection of ω onto α is given by
ω
α,α
α. Thus, we
may reformulate the notion of an integral element yet again as: ω is integral if and only
if the orthogonal projection of ω onto each positive simple root α is an integer or half-
integer multiple of α. Th i s characterization of the integral elements will help us visualize
graphically what the set of integral elements looks like in example (see Section 3.9).
Definition 3.7.2. An element ω of h is called a dominant integral element if ω, h
α
is a
non-negative integer for each positive simple root α.
Equivalently, ω is a dom i n ant integral element if 2
ω
α,α
is a non-negative integer for each
positive simple root α. If ω is dominant integral, then ω, h
α
will au t om a t ica l ly be a
non-negative integer for each positive root α, not just th e positive simple ones.
Definition 3.7.3. Suppose π is a fi nite-dimensional representation of g on a vector space
V . Then, ω h is called a weight for π if there exists a nonzero vector v V such that
π(h)v = ω, hv,
for all h h. A nonzero vector v satisfying condition above is called a weight vector for the
weight ω, and the set of all vectors (zero or nonzero) satisfying this conditio n is called the
weight space with weight ω. The di m en s io n of the weight space is call ed the multiplicity of
the weight.
To understan d this definition, suppose that v V is a simultaneous eigenvector for each
π(h), h h. This means that for each h h, there is a number λ
h
such that π(h)v = λ
h
v.
Since the representation π(h) is li n ea r in h, the λ
h
’s must depend linearly on h as well;
that is, the map h → λ
h
is a linear fun ct i onal on h. Then, there is a u n i qu e element ω of
h su ch that λ
h
= ω, h. Thus, a weight vector is nothing but a simultaneous eigenvector
64
for all the λ
h
’s and the vector ω is simp l y a conven ient way of encod ing the eigenva l u es.
Note that the roots (in the dual notation) are pre cis el y the nonzero weights of the adjoint
representation of g. It can be shown that two equivalent representations h ave th e same
weights and multiplici t i es. It is tru e, although by no means obvious, that ever y integral
element actually arises as a weight of some finite-dimension al representation of g, see [91].
Then, for any finite-dimensional representation π of g, the weights of π and t heir multiplicity
are inva r ia nt under the action of the Weyl group.
Definition 3.7.4. Let ω
1
and ω
2
be two elements of h. We say that ω
1
is higher than ω
2
(or, equival ently, ω
2
is lower than ω
1
) if there exist non-negative real numbers a
1
, . . . , a
r
such tha t
ω
1
ω
2
= a
1
α
1
+ . . . + a
r
α
r
,
where = {α
1
, . . . , α
r
} is the set of positive simple root s . This relationship is oft en written
as ω
1
ω
2
or ω
2
ω
1
.
If π is a representation of g, th en a weight ω
0
for π is said to be a highest weight if for all
weights ω of π, ω ω
0
. We now state an important result of this chapter [91, p.197]
Theorem 3.7.5 (Th eo re m of the Highest Weight). We have:
1. Every irreducible represen tation has a highest weight.
2. Two irreducible representati ons with the same highest weight are equivalent.
3. The highest weight of every irreducible representation is a dominant integral element.
4. Every dominant integral element occurs as the highest weight of an irreducible represen-
tation.
3.8 The Weyl character formula
Let Σ be a finite-dimensional irreducib l e rep r esentation of K acting on a vector space V ,
then we co n si d er the space of matrix entries of Σ. Suppose we choose a basis {u
k
} for V .
Then, for each x K, the linear operator Σ(x) can be expressed as a mat r i x with respect
to this basis; we d en ot e the entries of this matrix as Σ
kℓ
. Then, a mat ri x entry for Σ is a
function on K that can be expressed in the form
f(x) =
dim V
k,ℓ=1
α
kℓ
Σ(x)
kℓ
, (3.6)
for some set of c on st a nts a
kℓ
. We can describe the space of matrix entries in a basis-
independent way as the space of fu nctions that can be expressed in the for m
f(x) = tr(Σ(x)A) = t r ace( Σ( x) A) , (3.7)
65
for some linear operator A on V . To see the equivalence of these two forms, let A
kℓ
be the
matrix for t he operator A in the basis {u
k
}. Then, the matrix for Σ(x)A is given by the
matrix product (Σ(x)A)
kℓ
=
m
Σ(x)
km
A
mℓ
, so
tr(Σ(x)A) =
dim V
k,m=1
Σ(x)
km
A
mℓ
.
Thus, every function of the form (3.7) can be expressed in form (3.6) with α
kℓ
= A
ℓk
, and
conversely.
Definition 3.8.1. Let K be a simply-connected compact Lie group and let Σ be a finite-
dimensional irreducible representation of K. Then, the character of Σ is the function on
K given by
char Σ(x) = tr(Σ(x)) = tr ac e( Σ(x)).
This function is a matrix entry, ob t ai n ed by taking A = I in (3.7) or tak ing α
kℓ
= δ
kℓ
in
(3.6). The character is special because it satisfies
tr(Σ(xyx
1
)) = tr(Σ(x)Σ(y)Σ(x)
1
) = tr(Σ(y)),
for all x, y K. Rec al l from Section 2.1 that any function f satisfying f(xyx
1
) = f(y),
x, y K, is called a class function (constant on each conjugacy class of K). Onl y as
a note, the Peter-Weyl Theorem states that th e family of class functions tr Σ forms an
orthonormal basis for L
2
(K, µ), wher e Σ ranges over the equivalence classes of irreducible
finite-dimensional representations of K.
We now assume that K is a simply -co n n ec te d compact Lie group (there is also a version
of the result for connected com p a ct Lie groups that are not simply connected). We choose,
as usual, a maximal commutative subalgebra t of k an d we let T be the connected Lie
subgroup of K whose Lie algebra is t. It can be shown that T is a closed subgroup of K
(called a ‘maximal t or u s’ ) . It can further be shown that every el em ent of K is conju ga t e
to an element of T . This means that the values of a class function on K are, in principle,
determined by its values on T . The Weyl character formula is a formula for the restriction
to T of the character of an irred ucible representation of K.
We let g denot e the complexification of the Lie algebra t of K, so that g is a complex
semisimple Lie algebra. Then , h = t + it is a Cartan subalgebra in g. If we follow dual
notation and regard the roots as elements of h (not in h
), we know that if α h is a root,
then α, h is imaginary for all h t, which means that α itself is in it. It is then convenient
to introduce the real roots, which are simply
1
i
times the ordinary roots. This means that
a real root is a nonzero element α o f t with the property that there exist s a nonzero x g
with
[h, x] = iα, hx,
66
for all h t (or, equival ently, for all h h). We can also introduce the real co-roots as the
elements of t of the form h
α
=
2α
α,α
, where α is a real root.
In the same way, we will consider the real weights, which we t h i n k of as elements of t in
the same way as for the roots. So, if , V ) is an irreducible representation of g, then an
element ω t is call ed a real weight for Σ if there exists a nonzero vector v V such that
σ(h)v = iω, hv,
for all h t. Here, σ is the Lie algebra representation associated to the group represen-
tation Σ. An element ω of t is said to be integral if ω, h
α
is an integer for each real
co-root h
α
. (All of the ‘real’ objects are simply
1
i
times the corresponding objects without
the qualifier ‘real’). The real weights of any finite-dimensional representation of g must be
integral. For the rest of this section, all of roots and weight s will be assumed real, even if
this is not explicitly stated.
If α is an integral element, then it can be shown that there is a function f on T satisfyin g
f(e
h
) = e
iα,h
, (3.8)
for all h h. Note that because T is connected and commutative, every element t T can
be exp r essed as t = e
h
. However, a given t can be expressed as t = e
h
in many different
ways; the content of the above assertion is t hat the right-hand side of (3.8) is independent
of the choice of h for a given t. This means that we want to say that the right-hand side
of (3.8 ) defines a function on T , not just on t.
Next, we introduce the element δ of t defined to be half the sum of the positive roots:
δ =
1
2
αΦ
+
α.
It can be shown that δ is an integral element. (Clearly, ρ = 2δ is integral, but it is not
obvious that δ itself is integral). Finally, if w is any element of the Weyl group W , we think
of w as an orthogonal linear transformation of t in which case, ǫ(w) = det(w) = ±1. We
are now ready to state the Weyl character formula [91, p.213].
Theorem 3.8.2 (Weyl Character Formula). If Σ i s an irreducible representation of K with
highest real weight ω, then we have
char Σ(e
h
) =
wW
ǫ(w)e
iw·(ω+ρ),h
wW
ǫ(w)e
iw·ρ,h
,
for all h t, for which the denominator of right-hand side above is nonzero. Or equivalently,
wW
ǫ(w)e
iw·ρ,h
char Σ(e
h
) =
wW
ǫ(w)e
iw·(ω+ρ),h
,
for all h t. Here, ρ denotes the sum of the positive real roots.
67
The set of points h for which the denominato r of the Weyl character formula, the so-called
Weyl denominator, is nonzero is dense in t. At points where the denominator is zero, there
is an apparent singularity in th e formula for char Σ. However, actually at su ch points the
numerator is also zero and the character itself rema i n s finite (as must be the case since,
from the definition of the character, it is well defined and finite at every point). Note that
the character formula gives a formula for the restriction of char Σ to T . Since char Σ is a
class function and since (as we have asserted but not p r oved) every element of K is conju-
gate t o an element of T , knowing char Σ on T determin es, char Σ on all of K. A sketch of
the pr oof of the Weyl character formula can be found in [22] or [91].
In fact, the expression
w
ǫ(w)e
iw·ω,h
appearing on denominator of the Weyl character
formula can be written in an alternate way. This is established in the next r esult [116,
p.264]
Theorem 3.8.3 (Weyl Denominator Identity). On the same assumption of Weyl character
formula, we have
wW
ǫ(w)e
iw·ρ,h
= e
ρ
αΦ
+
(1 e
α
).
3.9 Representations of SU(3)
As an illustration of the con cepts introduced in this chapter, we will discuss the special
case of the representation theory of SU(3). The main result we have done in this chapter
is Theorem 3.7.5, which states th at an irr educible finite-dimensional representation of a
semisimple Lie al geb r a can be classified in terms of its highest weight.
The group SU(3) is simply connected , and so the finite-dim ensional representations of
SU(3) are in one-to-one cor r espondence with the finite-di m en s io n al representations of the
Lie algebra su(3). Meanwhile, the complex representations of su(3) are in one-to-one
correspondence with the complex-linear representations of th e complexified Lie algebra
(su(3))
C
= sl
3
(C) (Tab l e 2.4). Moreover, a repr esentation of SU(3) is irreducible if and
only if the associated representation of su(3) is irreducib le, and this holds if and on l y if the
associated complex-linear representation of sl
2
(C) is irreducible. (This follows from Pr opo-
sition 3.1.5, Proposition 3.1.6, and th e connectedness of SU(3)). This correspondence is
determined by the proper ty that
Π(e
x
) = e
π(x)
, for all x su(3) sl(3).
Since SU(3) is compact, Proposition 3.2.5 tells us that all of the finite-dimensional repre-
sentations of SU(3) are direct sums of ir r ed u c ible representations. The above paragraph
then implies that the same holds for sl
3
(C), that is, sl
3
(C) has the complete reducibility
property. We can apply th e same reaso n i n g to th e simply-connected group SU(2), its Lie
68
algebra su(2), and it s complexified Lie algebra sl
2
(C). Thus, every finite-dimensional rep-
resentation of sl
2
(C) or sl
3
(C) decomposes as a direct sum of irreducible invariant subspa ces.
We will use the following basis for sl
3
(C):
h
1
=
1 0 0
0 1 0
0 0 0
, h
2
=
0 0 0
0 1 0
0 0 1
;
x
1
=
0 1 0
0 0 0
0 0 0
, x
2
=
0 0 0
0 0 1
0 0 0
, x
3
=
0 0 1
0 0 0
0 0 0
;
y
1
=
0 0 0
1 0 0
0 0 0
, y
2
=
0 0 0
0 0 0
0 1 0
, y
3
=
0 0 0
0 0 0
1 0 0
.
Note that the span of {h
1
, x
1
, y
1
} is a subalgebra of sl
3
(C) which is isomorphic to sl
2
(C)
(Example 3.1.10) by ignoring the third row and column in each matrix. Similarly, the
span of {h
2
, x
2
, y
2
} is a subal gebra isom or p hic to sl
2
(C). Thus, we h ave the following
commutation relations:
[h
1
, x
1
] = 2x
1
, [h
2
, x
2
] = 2x
2
,
[h
1
, y
1
] = 2y
1
, [h
2
, y
2
] = 2y
2
, (3.9)
[x
1
, y
1
] = h
1
, [x
2
, y
2
] = h
2
.
We now list all of the commutation relations among th e basis elements which involve at
least on e of h
1
and h
2
(this includes some repetition s of the above commutation relat i on s ) .
[h
1
, h
2
] = 0;
[h
1
, x
1
] = 2x
1
, [ h
1
, y
1
] = 2y
1
,
[h
2
, x
1
] = x
1
, [ h
2
, y
1
] = y
1
;
[h
1
, x
2
] = x
2
, [ h
1
, y
2
] = y
2
,
[h
2
, x
2
] = 2x
2
, [ h
2
, y
2
] = 2y
2
;
[h
1
, x
3
] = x
3
, [ h
1
, y
3
] = y
3
,
[h
2
, x
3
] = x
3
, [ h
2
, y
3
] = y
3
.
69
Finally, we list all of the remaining commutation relations.
[x
1
, y
1
] = h
1
, [ x
2
, y
2
] = h
2
[x
3
, y
3
] = h
1
+ h
2
;
[x
1
, x
2
] = x
3
, [y
1
, y
2
] = y
3
,
[x
1
, y
2
] = 0, [x
2
, y
1
] = 0;
[x
1
, x
3
] = 0, [y
1
, y
3
] = 0,
[x
2
, x
3
] = 0, [y
2
, y
3
] = 0;
[x
2
, y
3
] = y
1
, [ x
3
, y
2
] = x
1
,
[x
1
, y
3
] = y
2
, [ x
3
, y
1
] = x
2
.
All of the an a ly si s we will do for the representations of sl
3
(C) will be in terms of t h e above
basis. From now on, all representations of sl
3
(C) will be assumed to be finite dimensional
and com p l ex linear.
Now, den ot e by h the com p l ex subal gebra generated by the element s {h
1
, h
2
}. Observe
from the relations listed above that:
(i) [ h
1
, h
2
] = 0;
(ii) [x
i
, h
j
] = 0 and [y
i
, h
j
] = 0, for i = 1, 2, 3 and j = 1, 2;
(iii) The oper a t or s ad h
1
and ad h
2
are di ago n al i zable.
Thus, the algebra h satisfies the conditions (i ) , (ii) and (iii) of Defini t io n 3.3.1, so h is a
Cartan subalgebra of sl
3
(C).
Recall from Section 3.7 that a weight for a representation π of the algebra sl
3
(C) (respect
to the Cartan subalgebra h) is an element ω sl
3
(C) such that there exists a vector v
satisfying
π(h)v = ω, hv, for all h h.
If we denote m
1
= ω, h
1
and m
2
= ω, h
2
, then we can see ω as the ordered pair
ω = (m
1
, m
2
), and the above condition says that ω is a weight if
π(h
1
)v = m
1
v, π(h
2
)v = m
2
v. (3.10)
A nonzero vector v satisfying relations (3.10) is called a weight vector corresponding t o the
weight ω = (m
1
, m
2
). Recall also that the multiplicity of ω a weight is th e di m en si on of
the cor r esponding weight sp ac e, i. e. the space of all vectors v satisfying (3.10). Thus, a
weight is simply a pair of simultaneous eige nvalues for π(h
1
) and π(h
2
). It can be shown
that equivalent repr es entations have the same weights and multiplicities.
70
Here is the advantage of work with the c om p l ex ification of Lie algebras: since we are
working over the complex numbers, π( h
1
) has at lea st one eigenval u e m
1
C. If W V
is the eigenspace for π(h
1
) with eigenval u e m
1
, since [h
1
, h
2
] = 0, π(h
2
) commutes with
π(h
1
), and, so, π(h
2
) must map W into itself. Thus, π(h
2
) can be viewed as an operator
on W, and its restriction of to W must have at least one eigenvector w with eigenvalue
m
2
C, giving w a si multaneous eigenvector for π(h
1
) and π(h
2
) with eigenvalues m
1
and
m
2
, respectively. He n ce, we have
Proposition 3.9.1. Every representation π of sl
3
(C) has at least one weight.
Now, every representation π of sl
3
(C) can be viewed, by restriction, as a representation of
the subalgebra {h
1
, x
1
, y
1
}
=
sl
2
(C). Note that even if π is irreducible as a repre sentation
of sl
3
(C), there is no reason to expect that it will still be irreducible as a representation of
the subalgebra {h
1
, x
1
, y
1
}. Nevertheless, π restricted to {h
1
, x
1
, y
1
} must be some finite-
dimensional representation of sl
2
(C). The same reasoning app l i es to the rest r ic ti o n of π to
the su b al gebra {h
2
, x
2
, y
2
}, which is also isomorphic to sl
2
(C). Now, recall Theorem 3.1.11,
which tells us that in any finite-dimensional representation of sl
2
(C), irr educible or not, all
of the eigenvalues of π(h) must be integers. Applying this result to the restriction of π to
{h
1
, x
1
, y
1
} and to the restriction of π to {h
2
, x
2
, y
2
}, we can state the foll owing corollary
Corollary 3.9.2. If π is a representation of sl
3
(C), then all of the weights of π are of the
form ω = (m
1
, m
2
), with m
1
and m
2
being integers.
Recall now that a root of sl
3
(C) (relative to the Cartan subalgebra h) is a nonzero linear
functional α on h such that there exists a nonzero element z of sl
3
(C) with
[h, z] = α(h)z, for all h h.
We can see α as an ordered pair α = (a
1
, a
2
) C
2
, and the above condit i on says that a
nonzero α is a root of sl
3
(C) if
[h
1
, z] = a
1
z, [h
2
, z] = a
2
z. (3.11)
The element z satisfying rel at i on s (3.1 1 ) is called a root vector corresponding to the roo t
α = (a
1
, a
2
). Thus, z is a simultaneous eigenvec to r for ad h
1
and ad h
2
. This means that z
is a weight vector for the adjoint rep r esentation with weight (a
1
, a
2
). Taking into account
the nonzero condition for (a
1
, a
2
), we may say that the roots are prec is ely the nonzero
weights of the adjoint representation. Corollary 3.9.2 then tells us that for any root, both
a
1
and a
2
must be integers, which we can also see dir ect l y in Table 3.3. The commutation
relations (3.9) t el l us what the roots for sl
3
(C) are. There are six roots:
Note that h
1
and h
2
are also simultaneous eigenvectors for ad h
1
and ad h
2
, but th ey are not
root vectors because the simultaneous eigenvalues are both zero. Si n ce the vectors in Table
3.3 together with h
1
and h
2
form a basis for sl
3
(C), it is not hard to show that the roots
listed in Table 3.3 are the only roots . These six roots form a root syst e m , conventionally
71
root α root vector z
(2, 1) x
1
(1, 2) x
2
(1, 1) x
3
(2, 1) y
1
(1, 2) y
2
(1, 1) y
3
Table 3.3: Roots for the Lie algebra sl
3
(C).
called A
2
. It is convenient to single out a basis consis t ing on the two roots correspon ding
to x
1
and x
2
and give them special names:
α
1
= (2, 1),
α
2
= (1, 2). (3.12)
The roots α
1
and α
2
are called the posit ive simple roots, because they have the property
that all of the roots can be expressed as linear combinations of α
1
and α
2
with integer
coefficients, and t h e se coefficients are (for e ach root) either all nonnegative or nonposi t i ve.
This is verified by direct computation:
root α linear combination z
(2, 1) α
1
(1, 2) α
2
(1, 1) α
1
+ α
2
(2, 1) α
1
(1, 2) α
2
(1, 1) α
1
α
2
Table 3.4: Roo t s for sl
3
(C) in terms of the basis = {α
1
, α
2
}.
The decision to designate α
1
and α
2
as the positive simple roots is arbitrary; any other pair
of roots with similar properties would do ju st as well. For examp l e, if we set α
3
= α
1
+ α
2
,
then we can define
= {α
1
, α
3
} as another basis. Figure 3.1 shows the root system for
sl
3
(C).
The significance of the roots for the representation theory of sl
3
(C) is contained in the
following lemma.
72
root α linear combination z
(2, 1) α
1
(1, 2) α
1
α
3
(1, 1) α
3
(2, 1) α
1
(1, 2) α
1
+ α
3
(1, 1) α
3
Table 3.5: Roo t s for sl
3
(C) in terms of the basis
= {α
1
, α
3
}.
m
1
m
1
m
2
m
2
α
1
α
1
α
2
α
3
Figure 3. 1: The root system A
2
for th e Lie algebra sl
3
(C).
Lemma 3.9.3. Let α = (a
1
, a
2
) be a root and z
α
a corresponding root vector in sl
3
(C). Let
π be a representation of sl
3
(C), ω = (m
1
, m
2
) a weight for π, and v = 0 a corresponding
weight vector. Then,
π(h
1
)π(z
α
)v = (m
1
+ a
1
)π(z
α
)v,
π(h
2
)π(z
α
)v = (m
2
+ a
2
)π(z
α
)v.
Thus, either π(z
α
)v = 0 or π(z
α
)v is a new weight vector with weight ω + α = (m
1
+
a
1
, m
2
+ a
2
).
Proof. In fact, the definition of a root tells us that we have the commutation relation
[h
1
, z
α
] = a
1
z
α
. Thus,
π(h
1
)π(z
α
)v =
π(z
α
)π(h
1
) + a
1
π(z
α
)
v = π(z
α
)(m
1
v) + a
1
π(z
α
)v
= (m
1
+ a
1
)π(z
α
)v,
and a similar argument allows us to comput e π(h
2
)π(z
α
)v.
73
We see then that if we have a representation with a weight ω = (m
1
, m
2
), then by applying
the root vectors x
1
, x
2
, x
3
, y
1
, y
2
and y
3
, we can get some new weights of th e form ω + α,
where α is the root. O f course, some of the time π(z
α
)v will be zero, in which case ω + α is
not necessarily a weight. In fact, since our repres entation is finite di m ensional, there can
be only finitely many weights, so we must get zero quite often. Now we would like to single
out in each representation a highest weight. Recall from a pr evi ous section that if ω
1
and
ω
2
are two weights, then ω
1
is higher than ω
2
(ω
1
ω
2
) if ω
1
ω
2
can be written in the
form
ω
1
ω
2
= a
1
α
1
+ a
2
α
2
, wit h a
1
, a
2
0,
and recall that maximal elements of this pa r t ia l relation are called highest weights. Note
that the relation of ‘hi gher’ is only a partial ordering; th at is, one can easil y have ω
1
and ω
2
such that ω
1
≻ ω
2
neither ω
1
≺ ω
2
. For example, α
1
α
2
is neither higher nor lower than
0. This, in particular, means that a finit e set of weights need not have a hi ghest element
(e. g., t he set {0, α
1
α
2
} h as no highest element). Note also that the coefficients a
1
and a
2
do not have to be integers, even if both ω
1
and ω
2
have integer entries. For example, (1, 0)
is higher than (0, 0) since (1, 0) (0, 0) = (1, 0) =
2
3
α
1
+
1
3
α
2
. Recall also that an ordered
pair (m
1
, m
2
) with m
1
and m
2
being non-negative integers i s called a dominant integral
element.
m
1
m
2
α
1
α
2
Figure 3.2: Root s and dominant integral elements for sl
3
(C).
Theorem 3.7.5 tell us t hat the highest weight of each irre d u ci b l e representation of sl
3
(C)
is a dominant integral element and, conversely, that every dominant integral el em ent oc-
curs as the highest weight of some irreducible representation. Since (1, 0) =
2
3
α
1
+
1
3
α
2
and (0, 1) =
1
3
α
1
+
2
3
α
2
, we see that every dominant integral element is high er than zero.
However, if ω has integer coefficients and is higher than zero, this does not necessarily
mean t h at ω is dominant integral (for example, α
2
= (2, 1) is higher than zero, but is
not dominant integral). Figure 3.2 shows the roots and dominant integral elements for
sl
3
(C). This picture is made using the obvious basis fo r the space of weights; that is, the
74
x-coordinate is the eigenval u e of h
1
and the y-coordinate is the e ig envalue of h
2
.
It is possible to obtain much more informat i on about the irreducible repr esentations besides
the high e st weight. For example, we have the foll owing formula for the di m en si on of the
representation with highest weight (m
1
, m
2
). It is a consequence of the Weyl character
formula (see [100]):
Theorem 3.9.4. The irreducible representation with highest weight (m
1
, m
2
) has dimension
1
2
(m
1
+ l)(m
2
+ 1)(m
1
+ m
2
+ 2).
There is an import ant symmetry to the representations of sl
3
(C) involving the Weyl group.
To understand the idea behi n d the Weyl group symmet r y, let us observe th at the repr e-
sentations of sl
3
(C) are, in a certain sense, invari ant under th e adjoint action of SU(3).
This mea n s the following: let π be a finite-dimensional representation of sl
3
(C) actin g on
a vector space V and let Π be the associated representation of SU(3) acting on the same
space. For any A SU(3), we can define a new representation π
A
of sl
3
(C), acti n g on the
same vector space V , by setting
π
A
(x) = π(AxA
1
).
Since the adjoint action of A on sl
3
(C) is a Lie algebr a automorph i sm , this is, again, a
representation of sl
3
(C). Thi s new representation is to be equival ent to the original repre-
sentation; and direct calculation shows that Π(A) is an intertwining map between (π, V )
and (π
A
, V ) . We may say, th en , tha t the adjoint action of SU(3) is a symmetry of the set
of equivalence classes of representations of sl
3
(C).
Now, we have analyzed the representations of sl
3
(C) by simultaneously diagonalizing the
operators π(h
1
) and π(h
2
). Of course, this means that any lin ear combination of π(h
1
) and
π(h
2
) is also simultaneously diagonalized. So, what really count s is the two-dimensional
subspace h of sl
3
(C) spann ed by h
1
and h
2
, the Car t an subalgebra of sl
3
(C). In general,
the adjoi nt action of A SU(3) does not preserve the space h and so the equival ence of
π and π
A
does not (in general) tell us anything about the weights of π. However, there
are elements A SU(3) for which ad A does preser ve h. We have already seen that these
elements make up the Weyl group for SU(3) and give rise to a symmetry of the set of
weights of any representation π. So, we may say that the Weyl group is the ‘ r esi d u e ’ of the
adjoint symmetry of the representations (discussed in the previous para gr ap h) that is l eft
after we focus our attention on the Cartan subalgebra h of sl
3
(C).
In the case of the two-dimension a l subspace of sl
3
(C) spanned by h
1
and h
2
, let Z be the
subgroup of SU(3) consisting of those A SU(3) such that Ad A(h) = h, for all h h.
Let N be the subgroup of SU(3) consisting of those A SU(3) such that Ad A(h) is an
element of h, for all h h. In fact , Z and N ar e actually subgroups of SU(3) and Z is
75
a normal subgroup of N. This leads us to the Weyl group W = N/Z of SU(3). We can
define an action of W on h as follows. For each element w of W , choose an element A of
the corr esponding equivalence class in N. Then for h h we define the action w · h by
w · h = Ad A(h).
To see that this action is well defined, suppose B is another element of the s am e equivalence
class as A. Then B = AC, with C Z and thus, Ad B(h) = Ad AC(h) = Ad A Ad C(h) =
Ad A(h), by the definition of Z. It can be proved that W is isomorphic to the group of
linear transformations of h that can be expressed as Ad
A
for so m e A N. In fact, the
group Z consists pr eci sel y of the diagonal matrices insid e SU(3), namely the ma t r ic es of
the form
A =
e
0 0
0 e
0
0 0 e
i(θ+ φ)
, for θ, φ R.
The group N consists of precisely those mat r i ces A SU(3) such that for each k = 1, 2, 3
there exist {1, 2, 3} and φ R such that Ae
k
= e
e
. Here, {e
1
, e
2
, e
3
} is the standard
basis for C
3
. Hence, th e Weyl group W = N/Z is iso m or p hic to the permutation group on
three el em ents (see [91] for a proof).
In the case of SU(3), it is pos si b l e to identify the Weyl group with a cert ai n subgroup
of N , instead of as the quotient group N/Z. We want to show that the Weyl group is a
symmetry of the weights of any finite-dimensional representation of sl
3
(C). To understand
this, we adopt a less basis-dependent view of the weights. We have defi n e d a weight as a
pair (m
1
, m
2
) of simultaneous eigenvalues for π(h
1
) an d π(h
2
). However, if a vector v is an
eigenvector for π(h
1
) and π(h
2
) then it is also an eigenvector for π(h) for any element h of
the space h span n ed by h
1
and h
2
. Furthermore, the ei genvalues must depend lin ear l y on
h since if h and j are any two elements of h and π(h)v = λ
1
v and π(j) = λ
2
v, then
π(ah + bj)v = ((h) + (j))v = (
1
+
2
)v.
So, we may make the following basis-independent notion of a weight.
Definition 3.9.5. Let h be the subspace of sl
3
(C) spann ed by h
1
and h
2
and let π be a
finite-dimensional representation of sl
3
(C) acting on a vector space V . A linear functional
µ h is called a weight for π if there exists a nonzero vector v V such that π(h)v = µ(h)v,
for all h h. S u ch a vector v is called a weight vector with weight µ.
So, a weight is just a col le ct io n of simultaneous eigenvalues of all the elements h of h, which
must depend linearly on h and, therefore, defi n e a linear functional on h. Since h
1
and h
2
span h, the linear functional µ i s determined by the value of µ(h
1
) and µ(h
2
), and thus our
new notion of weight is equivalent to our o ld notion of a weight as just a pair of simultaneous
eigenvalues of π(h
1
) and π(h
2
). The reason for adopting this basis-in d ependent approach
is that the action of the Weyl group does not preserve the basis {h
1
, h
2
} for h. The Weyl
76
group is (or m ay be thought of as) a group of linear transformations of h. This m ea n s that
W acts linearly on h, and we denote this action as w ·h. We can define an ass ociated acti on
on the dual space h
in the following way: For µ h
and w W , we define w · µ to be
the elem ent of h
given by
(w · µ)(h) = µ(w
1
· h).
The main point of the Weyl grou p from the poi nt of view of representation theory, name ly
that the weights of any representation are invariant under th e a ct i on of the Weyl group.
More expl i cit l y, suppose that π is any finite-dimensional representation of sl
3
(C) and that
µ h
is a weight for π. The n , for any w W , w · µ is also a weight of h
, and the
multiplicity of w ·µ is the same as the multiplicity of µ. In other words, si n ce the roots are
nothing but the nonzero weights of the adjoint representation, this re su l t tells us that the
roots are invariant under the action of the Weyl group. In order to visualize the action of
the Weyl group, it is convenient to identify h
with h by means of an inner product on h
that is invariant under the action of the Weyl group. Recall that h is a su b space of the space
of diagonal matrices, and we can u se the Hilbert- S chmidt inner prod uct A, B = tr(A
B).
Since the Weyl group acts by permuting the diagonal entries, this inner product (restricted
to th e subspace h) is preserved by the action of W .
We now use this inner prod u c t on h
to identify h. Given any el em ent α of h, the map
h → α, h is a linear fun ct i onal on h (i. e., an element of h
). Every linear fun ct i on a l o n
h can be represented in this way for a unique α in h. Identifying each linear functional
with the corresponding element of h, we will now regard a weight for (π, V ) as a no n zer o
element of h with the property that th er e exists a nonzero v V such t hat
π(h)v = α, hv,
for all h h. This is the sa m e as Definition 3.9.5 except that now, α lives in h an d we write
α, h instead of α(h) on the right. The roots, being weights for th e adjoint representation,
are viewed in a simila r way. Now that the roots and weights live in h instead of h
, we can
use the above inner prod u ct o n h, and with this new point of v i ew the roots α
1
and α
2
are
identified with th e following elements of h:
α
1
=
1
1
0
, α
2
=
0
1
1
.
To check this, we note th at these matrices are indeed in h since the diagonal entries sum
to zero. Then, direct calculation shows tha t α
1
, h
1
= 2, α
1
, h
2
= 1, α
2
, h
1
= 1
and α
2
, h
2
= 2, in agreement with our earlie r definit i on of α
1
and α
2
in (3 .1 2 ) . So, then,
we can compute the lengths and angles as | |α
1
||
2
= α
1
, α
1
= 2, ||α
2
||
2
= α
2
, α
2
= 2,
and α
1
, α
2
= 1. This means that (with respect to this inner product) α
1
and α
2
both
have length
2 and the angle θ between them satisfies cos θ =
1
2
, so that θ = 120
. We
now consid er the dominant integral elements, which are th e possible highest weights of
77
irreducible representations of sl
3
(C). With our new point of view, these are the elements
µ of h such that µ, h
1
an d µ, h
2
ar e non-n egat i ve integers. We begin by considering th e
fundamental weights µ
1
and µ
2
defined by
µ
1
, h
1
= 1, µ
1
, h
2
= 0,
µ
2
, h
1
= 0, µ
2
, h
2
= 1.
These can be expressed in terms of α
1
and α
2
as follows:
µ
1
=
2
3
α
1
+
1
3
α
2
,
µ
2
=
1
3
α
1
+
2
3
α
2
,
obtaining
µ
1
=
2
3
1
3
1
3
, µ
2
=
1
3
1
3
2
3
.
A calculation then shows that µ
1
and µ
2
each have length
6
3
and that the angle between
them is 60
. The set of dominant integral elements is then precisely the set of linear
combinations of µ
1
and µ
2
with non-negat i ve integer coefficients. Note that µ
1
+ µ
2
=
α
1
+ α
2
, an observation that helps i n drawing Figure 3.3 below. Figure 3.3 shows the same
information as Figure 3.2, namely, the roots and t h e dominant integral elements, but now
drawn relative to a Weyl-invariant inner product. We draw only the two-dimen si onal real
subspace of h consisting of those elements µ such that µ, h
1
and µ, h
2
are real, since all
the roots and weights have this property. In this fi gure, the arrows indicate the roots, the
black dots indicate dominant integral elements (i. e., points µ such that µ, h
1
and µ, h
2
are non-negative integers), an d the triangular grid indicates integral elements (i. e., points
µ such that µ, h
1
and µ, h
2
are integers).
Let us see how the Weyl group acts on Figure 3.3 . Let (1 2 3) denote the cyclic permutation
of 1,2 and 3, an d let w
(1 2 3)
denote the corresponding Weyl group element. Then, w
(1 2 3)
acts by cyclically permuting th e diagonal entries of each element of h. Thus, w
(1 2 3)
takes
α
1
to α
2
and takes α
2
to (α
1
+ α
2
). This action is a 120
rotation, counterclockwise in
Figure 3.3. Next, let (1 2) be the permutation that interchanges 1 and 2 and let w
(1 2)
be the corresponding Weyl grou p element. Then, w
(1 2)
acts by interchanging the first two
diagonal entries of each element of h and thus takes α
1
to α
1
and takes α
2
to α
1
+ α
2
.
This correspon ds to a reflection about the line perpendicula r to α
1
. The read er is invited
to calculate the action of the remaining Weyl group elements, and observe that the Weyl
group consists of six elements: the symmetry group of an equilateral triangle centered at
the ori gi n , as indicated in Figure 3.4:
78
α
1
α
2
Figure 3.3: Root s and dominant integral elements for sl
3
(C) in the root basis.
α
1
α
2
Figure 3. 4: The Weyl group for sl
3
(C).
79
Chapter 4
Vertex algebras and Kac-Moody
algebras
The notion of a vertex algebras was introduced by Borcherds in [ 8 ] . This is a r i go r ou s
mathematical definition of the chiral part of a 2-d i m en s io n al quantum field t h eor y studied
by physicist since the landmark paper of Belavin, Polyakov and Zamolodchikov [4]. Basi-
cally, vertex algebras are the rigorous formalization of the bosonic theory in mathematical
physics. The main objective of this chapter is to give a quickly understand i n g of what
vertex algebras and vertex operat or algebras are. For a more detailed study, the reader
may r efer to [113], [67] and [7]. In Chapter 9 we shall discuss some other interesting topics
—for our purposes— relat i n g conformal field theory and the Moonshine phenomenon.
Subsequently, we will describe h ow vertex algebras arise on the mat h em at i ca l scene, partic-
ularly in the developing of the re p r esentation th eor y of affine Kac-Moody algeb r as. In this
part it will be useful to have in mind some of the results obt a i n ed in the previou s chapter.
The read er may refer to [112], [ 182 ] or [75] for a complete pr esentation of these topics.
4.1 Vertex operator algeb ra s
Definition 4.1.1. L et F be a field. A vertex algebra over F is a vector space V over F with
a collect io n of bilinear maps V × V V
(u, v) → u
n
v,
for all n Z, an d satisfying the following axioms
(i) u
n
v = 0, for n sufficiently large, i. e., there exists n
0
N (depending on u and v) such
that u
n
v = 0, for all n n
0
;
(ii) There exists an element 1 V such that
1
1
1
v = v, 1
n
v = 0 for all n = 1,
v
1
1 = v, v
n
1 = 0 for all n 0;
(iii) (Borcherds’ identity) for all u, v, w V and m, n, k Z
i0
m
i
(u
n+i
v)
m+ki
w =
i0
(1)
i
n
i
u
m+ni
(v
k+i
w) (1)
n
v
k+ni
(u
m+i
w)
.
1
This disti ngu is he d vector 1 usually appears as |0 in physicist notation. For ex ample see [113].
81
We will work only with the case F = R. Relation (iii) is called Borcherds’ identity. Since
it is som ewh at reminiscent of the Jacobi identity for Lie algebras (see Definition 2.3.1), it
is called sometimes the Jacobi identity for vertex algebras.
Denote by End V [[z, z
1
]] th e set of formal series ( i n the indeterminat e z) of the for m
nZ
ϕ
n
z
n1
,
where ϕ
n
End V , for all n Z. For ea ch u V , we can define the vertex operator
Y (u, z) : V V by
Y (u, z) =
nZ
u
n
z
n1
, (4.1)
where u
n
End V is given by v → u
n
v. We denote by V [[z]] the set of all formal series
i0
v
i
z
i
, for v
i
V . Also, consider the formal expression
δ(z w) = z
1
nZ
w
z
n
F[[z, z
1
, w, w
1
]].
Then, the axioms for a vertex algebra can be written in terms of such vertex operators as
follows:
(i) Y (u, z)v has coefficient of z
n
equal to 0 for all n sufficiently small (i. e., there exists
n
0
N dependi n g on u and v such that u
n
v = 0, n n
0
);
(ii) There exists an element 1 V such that
Y (1, z) 1
V
is the identity on End V ,
Y (v, z)1 V [[z]], for all v V ,
lim
z0
Y (v, z)1 = v.
(iii) δ(z
1
z
2
)Y (u, z
1
)Y (v, z
2
) δ(z
2
z
1
)Y (v, z
2
)Y (u, z
1
) = δ(z
1
z
0
)Y
Y (u, z
0
)v, z
2
, for
all u, v V .
For a series a(z) =
n
a
n
z
n
V [[z, z
1
]], we will denote by
a(z) =
n
na
n
z
n1
the form al derivat i ve of a(z). Let T : V V the linear map defined by
T (v) = v
2
1.
82
Denote by [·, ·] the usu a l bracket [a, b] = ab ba defined in End V [[z, z
1
]]. No t e that if
[T,
n
a
n
z
n
] =
n
[T, a
n
]z
n
, for an element
n
a
n
z
n
End V [[z, z
1
]], t h e n
[T, Y (u, z)] =
T,
n
u
n
z
n1
=
n
[T, u
n
]z
n1
=
n
(T u
n
u
n
T )z
n1
. (4.2)
Also,
Y (u, z) =
n
u
n
z
n1
=
n
(n 1) u
n
z
n2
=
n
(n)u
n1
z
n1
. (4.3)
On the othe r hand, i f we take w = 1, and m = 0, k = 2 on Borcherds’ identity, we obtain
i0
0
i
(u
n+i
v)
2i
1 =
i0
(1)
i
n
i
u
ni
(v
2+i
1) ( 1 )
n
v
2+ni
(u
i
1)
=
i0
(1)
i
n
i
u
ni
(v
2+i
1).
Note that t h e sum on the left-hand sid e above involves only one term (when i = 0), and the
right-hand side actual l y has only two terms (when i = 0, 1), since v
n
1 = 0, for all n 0.
In fact, we have nu
n1
v = (u
n
v)
2
1 u
n
(v
2
1) and it follows that
nu
n1
v = (u
n
v)
2
1 u
n
(v
2
1) = (T u
n
u
n
T )v, for all n Z. (4.4)
From equation ( 4. 2 ) , (4.3) and (4.4), then we deduce that nu
n1
= [T, u
n
], and hence
[T, Y (u, z)] = Y (u, z). We summarize th i s and other similar results in the next s t at em ent.
See [113 , p.117–118] for a complete proof of this fact.
Theorem 4.1.2. We have the following equivalent axioms for a vertex algebra:
(i’) (translation covariance) [T, Y (u, v)] = Y (u, z), for all u V ;
(ii’) (vacuum) Y (1, z) = 1
V
and Y (u, z)1|
z=0
= u, for all u V ;
(iii’) (locality) (z w)
n
Y (u, z)Y (v, w) = (z w)
n
Y (v, w)Y (u, z) for n sufficiently large
(depending on u and v).
Now, we apply the operator T rep e at ed l y to the equation T v = v
2
1. Since T v
n
=
[T, v
n
] + v
n
T , we have that
T (v
n
1) = (n)v
n1
1 + (v
n
T )1 = (n)v
n1
1 + v
n
(1
2
1) = (n)v
n1
1, (4.5)
and so
T
2
v = T (v
2
1) = 2v
3
1 so that v
3
1 =
1
2
T
2
v,
T
3
v = T (2v
3
1) = (2 · 3)v
4
1 so that v
4
1 =
1
3!
T
3
v.
83
Continuing this process on identity (4. 5 ) , an induction on n guar antees t hat v
n
1 =
1
(n1)!
T
n1
v, for all n 1. Hence
Y (u, z)1 =
n
(u
n
1)z
n1
=
n0
(u
n
1)z
n1
+
n<0
(u
n
1)z
n1
=
n0
(u
(n+1)
1)z
n
=
n0
1
n!
T
n
u z
n
=
e
zT
u,
for all u V .
Remark 4.1.3. The bracket operation defined by [u, v] = u
0
v makes V/T V into a Lie
algebra (see [12] for a proof of this). Borcherds’ identity is described in [67] as being ‘very
concentrated’. It can be shown that it is equivalent to three simpler identities [113 ]. For
θ, ϕ End V , we define [θ, ϕ] = θϕ ϕθ. Also, if a(z) =
n
a
n
z
n1
End V [[z, z
1
]], we
define
a(z)
+
=
n<0
a
n
z
n1
= a
1
+ a
2
z + a
3
z
2
+ . . .
a(z)
=
n0
a
n
z
n1
= a
0
z
1
+ a
1
z
2
+ a
2
z
3
+ . . .
Definition 4.1.4. Given a(z), b(z) End V [[z, z
1
]], we also define t h ei r normal ordered
product as
: a(z)b(z) : = a(z)
+
b(z) + b(z)a(z)
.
In te r m s of the notation just introduced, Borcherds’ identity can be shown t o be equivalent
to th e following three simpler identities:
(a) [u
m
, Y (v, z)] =
i0
m
i
Y (u
i
v, z)z
mi
, for all u, v V , m Z;
(b) : Y (u, z)Y (v, z) : = Y (u
1
v, z), for all u, v V ;
(c) Y (T u, z) = Y (u, z), for all u V .
Definition 4.1.5. An element ω of a vertex algebra V is a conformal vector of central
charge c, if is an even vector satisfying:
ω
0
v = T v, for all v V ;
ω
1
ω = 2ω;
84
ω
2
ω = 0;
ω
3
ω =
c
2
1;
ω
i
ω = 0, for all i 4;
V =
nZ
V
n
, where each V
n
is the set of eigenvectors of the line ar operat or ω
1
V
n
= {v V : ω
1
v = nv}
corresponding to the eigenvalue n Z.
In oth er words, ω is a conformal vector if the corresponding vertex operator Y (ω, z) is a
Virasoro field with central charge c, i. e., a formal series L(z) satisfying
L(z)L(w) =
c/2
(z w)
4
+
2L(w)
(z w)
2
+
L(w)
z w
.
In particular, for a conformal vector ω we have, 1 V
0
and ω V
2
. Note th a t wh en
a vertex algebra V has a conformal vector, it admits an action of the Lie algebr a called
Virasoro algebra ( see next section).
Definition 4.1.6. A vertex algebra en d owed with a conform al vector ω as in Definition
4.1.5 i s called a vertex operator algebra (or a conformal vertex algebra) of rank c.
4.2 The Virasoro algebra
Let p(t) F[t, t
1
] and consider the derivation
T
p(t)
= p(t) (4.6)
of F[t, t
1
]. The linear space of all derivations of F[ t, t
1
] of type (4.6) has th e structure of
a Lie algebra with respect to the natural Lie bracket
[T
p(t)
, T
q(t)
] = T
p(t)q
(t)p
(t)q(t)
for p(t), q(t) F[t, t
1
]. We denote this algebra by d and we choose th e following basis of
d:
{d
n
= t
n+1
= t
n
d : n Z},
where d is t h e derivation in (2.12). Then, the commutators have the form [d
m
, d
n
] =
(m n)d
m+n
, for m, n Z.
In fact, if T End F[t, t
1
] is a derivation, and we set
p(t) = T (t), (4.7)
85
then we have T (1) = T (1 · 1) = T ( 1) + T (1), so that
T (1) = 0, (4.8)
and 0 = T (tt
1
) = T (t)t
1
+ tT (t
1
) impl i es that
T (t
1
) = t
2
T (t). (4.9)
Since formulas (4.7), (4.8) and (4.9) also hold for T
p(t)
, we see that T
p(t)
and T agree on all
powers of t. Thus, we have the following statement explaining the importance of d.
Proposition 4.2.1. The derivations of F[t, t
1
] form precisely the Lie algebra d.
Any th r ee generators of d of t h e form d
n
, d
0
, d
n
, n Z
+
span a subalgeb r a of d isomorphic
to the Lie algebra sl
2
(F) of 2 × 2 matrices over F of trace 0. We shall single out the
subalgebra
p = Fd
1
+ Fd
0
+ Fd
1
.
As in the case of affine Lie algebras, we consider central extensi ons. We denote by v the
following one-dimensional central extension of d with basis con si st i n g of a central element
c and elements L
n
, n Z, corresponding to the basi s elements d
n
, of d. For m, n Z
[L
m
, L
n
] = (m n)L
m+n
+
1
12
(m
3
m)δ
m+n,0
c (4.10)
= (m n)L
m+n
+
1
2
m + 1
3
δ
m+n,0
c; (4.11)
and [L
m
, c] = 0, fo r all m Z.
Definition 4.2.2. The Lie algebra d above is called the Virasoro algebra.
Note th at the central extension (4.10) is trivial wh en restricted to the subalgebra p of d.
We can form equivalent extensions of d by setting
L
n
= L
n
+ β
n
c, fo r β
n
F, n Z. (4.12)
Then, the exten si on (4.10) is modified by the subtraction of the term (m n)β
m+n
c.
Although, the significance of the extension (4.10) is given by the next result (see [67, p.33]
for a proof).
Proposition 4.2.3. The extension (4.10) of the Lie algebra d is the unique nontrivial
1-dimensional extension up to isomorphism.
Now we return to the situation in which V is a vertex operator algebra with conformal
vector ω. Define the l i n ea r map L
i
: V V by
L
n
= ω
n+1
, for n Z.
86
The pr operties of ω imply t h a t
[L
m
, L
n
] = (m n)L
m+n
+
1
2
m + 1
3
δ
m+n,0
c 1
V
, (4.13)
in analog y to (4.11) (see [113] for details). Thus, V is a module for the Virasoro algebra in
which th e central element c i s represented by c 1
V
, where c is the central charge of ω.
4.3 Kac-Moody algebras and their representa t io n
We shall now describe how vertex operators arose in the representation theory of affine
Kac-Moody algeb r as . In order to explain th is we first recall some ideas from the th eo r y of
Lie alg eb r as (see Chapter 3).
We su p pose first that g is a fini t e dimen s io n al sim p l e Lie algebra over C. Then, g has a
decomposition
g = n
+
h n
, (4.14)
where
n
+
=
αΦ
+
g
α
, n
=
αΦ
g
α
,
and all g
α
satisfy
dim g
α
= 1 and [h, g
α
] = g
α
.
Here, h is a Cartan subalgebra of g, and the 1-dimensional spaces g
α
are the root spaces of
g with respect to h, each of which gives rise to a 1-dimension a l representation α of h gi ven
by
[x, x
α
] = α(x) · x
α
, for all x
α
g
α
, x h.
Recall that the set Φ = Φ
+
Φ
is the set of roots of g. Φ contains a subset = {α
1
, . . . , α
r
}
of simple roots, where r = di m h. The roots in Φ
+
are linear combinations of elements o f
with non-negative integer coefficients and roots in Φ
are linear combinations of eleme nts
of with non-positive integer coefficients. The free abeli an group
Q = Zα
1
+ Zα
2
+ . . . + Zα
r
is called the root lattice. The real vector space Q
Z
R can be given the str u ct u r e of a
Euclidean space in a natural way. Let w
i
be t h e reflection in the wall orthogonal to α
i
.
Then, we h ave seen that the group of isometri es of Q R generated by w
1
, . . . , w
r
is the
Weyl gr oup W of g. Reme mber that this is a finite group which permutes the elements of
Φ (Proposition 3.5.2). Each root is the image of some si m p l e root under an element of W .
We have
w
i
(α
j
) = α
j
A
ij
α
i
, for A
ij
Z. (4.15)
87
Definition 4.3.1. The matrix A = [A
ij
] is called the Cartan matrix of g (relatively to the
Cartan subalgebra h).
Any Cartan matrix [A
ij
] satisfies the following conditions:
A
ii
= 2, for each i;
A
ij
{0, 1, 2, 3} if i = j;
A
ij
= 0 if and only if A
ji
= 0;
A
ij
{−2, 3} A
ji
= 1.
The Lie algebra g can be defin e d by generators and rel at i ons depen d i n g only on the Cartan
matrix A. The Cartan subalgebra h has a basis h
1
, . . . , h
r
of elements satisfying
α
j
(h
i
) = A
ij
.
We also recall h ow to describe the fi n i t e dimensional irreducible g-modules. These are
parametrized by dominant integral weights. Remember th at a weight is an element of
h
. A weight ω is said dominant if ω(h
i
) is a non-negative real number for all i. Finally,
ω is dom i n ant integral if each ω(h
i
) is a non - n eg at i ve integer. The fund a m ental weights
ω
1
, . . . , ω
r
are defi n ed by
ω
i
(h
j
) = δ
ij
, for i, j)1, 2, . . . , r.
Thus, each dominant integral weight λ is a non-negati ve integral combination of the fun -
damental weights.
For each dominant integral weight λ there i s a corresponding fi n i t e dimensional irreducible
g-module M
λ
. Th i s module M
λ
is a d i r ect sum of 1-dimen s io n al h-m odules which, when
collected together, give rise to a weight space decompos it i o n
M
λ
=
µ
M
µ
λ
,
where M
µ
λ
= {v M
λ
: xv = µ(x)v, x h} (here the sum is over all µ h
). In a simi la r
form to Definitio n 3.8.1, we define a not i on of character for this m odule M
λ
Definition 4.3.2. The character of the h-module M
λ
is the function given by
char M
λ
=
µ
(dim M
µ
λ
)e
µ
.
Note that the character defined above is an element of the group ri n g F[G] of the fr ee
abelian group G gen e ra t ed by the fundamental weights. As usual, we write this gr ou p
multiplicatively, replacing a weight µ by e
µ
and su ch that
e
µ
e
ν
= e
µ+ν
.
88
The Weyl character formula (Theorem 3.8.2) applied to this new definition of character
asserts t h a t
wW
ǫ(w)w(e
ρ
)
char M
λ
=
wW
ǫ(w)w(e
λ+ρ
), (4.16)
where ǫ = det 1}, is the homomorphism given by ǫ(w
i
) = 1, for all i, and ρ =
r
i=1
ω
i
. In an analo gous way to Theorem 3.8.3, the expression
w
ǫ(w)w(e
ρ
) appearing
in the denominator or char M
λ
can be written in an alternative form. In this case, Weyl’s
denominator identity a sser t s that
wW
ǫ(w)w(e
ρ
) = e
ρ
αΦ
+
(1 e
α
). (4.17)
The well known theory of finite dimensional simple Lie algebras over C which we have just
outlined was generalized by Kac [110] a n d Moody [149] to give the t h eor y of Kac-Moody
algebras. In order to obtain such Lie algebras, we begin with the notion of generalized
Cartan matrix.
Definition 4.3.3. A generalized Car tan matrix (sometimes abbreviated GCM) is any ma-
trix A = [A
ij
] satisfy ing the conditions
A
ij
Z;
A
ii
= 2, for all i;
A
ij
0, if i = j;
A
ij
= 0 if and only if A
ji
= 0.
A Lie algebra is then defined by generators and relations depending on the generalized
Cartan matrix A, just as in the case of finite dimensional simple Lie algebras. This Lie
algebra is then extended by ou t er derivations to ensure that the simple roots are lin ea rl y
independent, even th ough the matrix A is singular (may occu r ) .
Definition 4.3.4. The resulting Lie algebra above is called the Kac-Moody algebra given
by the generalized Cartan matrix A.
The mai n differences from the finite dimensional case are as follows:
1. The Lie algeb r a g can have infinite dimension.
2. The root spaces g
α
can have dimension greater than 1.
3. The Weyl group W can be infinite.
4. There can be both real roots and imaginary roots.
89
In thi s case, a root α is called real if α, α > 0 and imaginary if α, α 0. All simple
roots of a Kac- Moody algebra are real, and any real root can be obtained from a simple
root just tran sfor m i n g by some appropriate element of the Weyl group.
A general i zed Cartan matrix A is said symmetrisable if A = DB, where B is symmetric and
D is a non-singular diagonal matrix. We shall restrict attention to Kac-Moody algebras
with symmetrisable GCM. For each dominant integral weight λ, t h er e is a cor r esponding
irreducible module M
λ
for such a Kac-Moody algebr a. The character of this module M
λ
is
given by Kac’s character formula
wW
ǫ(w)w(e
ρ
)
char M
λ
=
wW
ǫ(w)w(e
λ+ρ
), (4.18)
Of course th i s look very much like Weyl’s character formula (4.16) in the fi n i t e dimensional
case, but here the sums over W are infi n i t e. Th e denominator of Kac’s character formula
can be written in an alternati ve way, giving Kac’s denominator identity (analogous to
(4.17))
wW
ǫ(w)w(e
ρ
) = e
ρ
αΦ
+
(1 e
α
)
mult α
. (4.19)
The left-hand side is an infinite sum an d the right-hand side is an infin i t e p roduct. The
multiplicity mult α is given by
mult α = dim g
α
.
In the finite dimensional cas e, all the roots have multiplicity 1 and so equation (4.19 ) re-
duces to Weyl’s denominator identity (4.17).
There is a remarka b l e trichotomy in the theory of Kac-Mood y algebras. Let A be an n ×n
generalized Cartan matrix. Given a vector u R
r
we write u 0 if all coordinates u
i
of u are positive; and u 0 if all u
i
are negative. It turn s out that if the GCM A is
indecomposable, exactly one of the following conditions holds:
There exist u 0 with Au 0;
There exist u 0 with Au = 0;
There exist u 0 with Au 0.
We say that A has finite type if the first condition holds; affine type if the second condition
holds, and indefinite type if we have t h e third condi t i on . The indecomposable Kac-Moody
algebras of finite type corr espond to the finite dime n si on a l simple Lie algebras. All Kac-
Moody algebras of fini te or affine type are s ym m et r i sable.
Examples of affine Kac-Moody algeb r as are the non-twisted affine a l geb r a s (Section 2.8).
These may be constructed as follows. We start with a fi nite d i m ensional simple Lie algebra
g over C. We then consider the algeb r a g C[t, t
1
]. Thi s algebra has a C-basis
h
i
t
m
, x
α
t
m
, for i = 1, . . . , r, m Z,
90
where x
α
is a non-zero vector or g
α
. Also, this algebra h a s a non-trivial 1-dimension al
central ext ension. This is the Lie algebra
g
C
C[t, t
1
] Ck,
in which multiplication is given by
[x t
m
+ λk, y t
n
+ λ
k] = [ x, y] t
m+n
+ x, y
m+n,0
k,
for x, y g. Here  ·, · is a non-degenerat e invar i ant symmet r i c bi l i n ear for m o n g (r ecal l
equation (2.14)). We have already called this Lie algebra
ˆ
g the untwisted affine algebr a
associated wi t h g. If we extend this Lie algebra by adjoining a 1-dimensional space of outer
derivations, thus we obtain the Lie algebra
˜
g =
ˆ
g Cd = g
C
C[t, t
1
] Ck Cd,
with multiplication
[xt
m
+λk +µd, y t
n
+λ
k +µ
d] = [x, y]t
m+n
xµ
mt
m
+y µnt
n
+x, y
m+n,0
k.
We have called this the extended affin e algebra associated with g (Section 2.8). This Lie
algebra
˜
g is an affine Kac-Moody algebra whose generalized Cart an matrix has degree
n + 1. Its GCM is obtained from the Cartan matrix of g by adjoini n g an additional row
and column. Let α the highest root of the Lie algebr a g and consider α
0
= α. The Cartan
matrix of g is the n × n matrix A = [A
ij
], wher e
A
ij
= 2
α
i
, α
j
α
i
, α
i
, for i, j = 1, 2, . . . , r;
and the generalized Cartan matrix of t h e affi n e L i e a lg eb r a
˜
g is the (n + 1) ×(n + 1) matrix
A = [A
ij
], for i, j = 0, 1, 2, . . . , r,
given by the same formula. Thus, the GCM of
˜
g gives rise to a corresponding irreducible
˜
g-module M
λ
, as described above. In part i cu l ar , we obtain such modul es associated with
the fun d a m ental weights ω
0
, ω
1
, . . . , ω
r
of
˜
g. The fundamental representation of
˜
g associ -
ated to the weight ω
0
is called the basic representation of
˜
g. The basic representation plays
a crucial role in the many application s of the repr esentation theory of affine Kac-Moody
algebras. We sh a ll describe a modul e giving the basic repr ese ntation of
˜
g in the case when
g is simply laced, i. e., when A
ij
= 0 or 1 for all i, j {1, 2, . . . , r} with i = j.
Suppose this is so and let Q be the root lattic e of g. There i s a symmetric Z-bilinear form
Q × Q Z given by (α, β) → α, β uniquely determined by the con ditions
α
i
, α
j
= A
ij
, for α
i
, α
j
.
91
This lat t i ce Q has a unique ce ntral extension
ˆ
Q with
1 C
2
ˆ
Q Q 1
in which
aba
1
b
1
= (1)
¯a,
¯
b
, for a, b
ˆ
Q,
where a → ¯a is the map
ˆ
Q Q. For each γ Q, we choose an element e
γ
ˆ
Q such that
e
γ
= γ and e
0
= 1. Then we have
e
α
e
β
= e
α+β
ǫ(α, β), for all α, β Q.
The map ǫ : Q×Q 1} is a 2-cocycle. It is known fr om the theory of finite dimensional
Lie alg eb r as that the root vectors x
α
of g can be chosen so tha t
[x
α
, x
β
] = ǫ(α, β)x
α+β
,
whenever α, β, α + β Φ.
We are now in position to describe a
˜
g-module giving the basic r ep r es entation. Let
˜
h
the
subalgebra of
˜
g given by
˜
h
=
n<0
(h t
n
),
and let S(
˜
h
) be the symmetric algebra on
˜
h
(recall Section 2.7). Let C[Q] the group
algebra of Q with basis e
γ
for γ Q. We write
V
Q
= S(
˜
h
)
C
C[Q]. (4.20)
It t u r n s out that V
Q
can be made into a
˜
g-module affording the basic representation. The
central element K of
˜
g acts as the identity ma p on V
Q
. Th e elements h t
n
for h h act
as follows:
h t
n
acts as
h(n) 1 if n = 0,
1 h(0 ) if n = 0.
Here, h(0) End C[Q] is given by h(0)e
γ
= γ(h)e
γ
, and h(n) End S(
˜
h
) is given by the
following rule
If n < 0, then h(n) i s multiplication by h t
n
.
If n > 0, then h(n) i s the derivation of S(
˜
h
) deter m i n ed by
h(n)(x t
n
) = n(x, h), for x h,
h(x)(x t
m
) = 0, if m = n.
92
We now consider the action of the elements x
α
t
n
˜
g on V
Q
. Let x
α
(n) End V
Q
be the
endomorphism induced by x
α
t
n
. The en domorphisms x
α
(n) turn out to be complicated
expressions not appealing to the intuition. However, it is possible to make sense of them
by combining then into a vertex operator
Y (α, z) =
nZ
x
α
(n)z
n1
,
which is given by an explicit for mula. In order to explain th i s, we introduce the following
notation.
There is an isomorphism h
h given by λ → h
λ
, where µ(h
λ
) = λ, µ. In this way
h
may be identified with elements of h ( t he corresponding co-roots). Thus, the elements
α(n) End S(
˜
h
) are defined as above for n = 0. The vertex operator Y (α, z) is then
given by
Y (α, z) = exp
n<0
α(n)
n
z
n
exp
n>0
α(n)
n
z
n
e
α
z
α
.
Here, z
α
End V
Q
is given by 1 z
α
for z
α
End C[Q]; and z
α
e
γ
= z
α,γ
e
γ
. Also,
e
α
End V
Q
is given by 1 e
α
for e
α
End C[Q]; and e
α
e
γ
= ǫ(α, γ)e
α+γ
(see for example
[112]).
We conclude this section wi t h the definit i on of automorp h i sm of a vertex algebra. We will
see in Chapter 5 that the Mons t er group M occur as a group of au t o m or p hisms of a vertex
algebra.
Definition 4.3.5. Let V be a vertex algebra with conformal vector ω. An automorphism
of V is an invertible linear map g : V V satisfying
gY (v, z)g
1
= Y (gv, z), for all v V ;
g(ω) = ω.
This then, is how vertex oper a t or s first appeared in the representation theory of affine Kac-
Moody algebras. In fact, as we shall see, the vector space V
Q
can be made into a vertex
algebra.
4.4 Lattices
Definition 4.4.1. By a lattice of rank n N we shall mean a rank n free abelian group L
provided with a rational-val u e d symmetric Z-bilin ea r form
·, · : L × L Q
93
A lattice isomorphism is sometimes called an isometry. A lattice L is said non-degenerate
if its form ·, · is non -degenerate in the sense that for α L
α, L implies α = 0.
Given a lattice L, we see by choosing a Z-base of L that L, L
1
r
Z for some r Z
+
.
That is
·, · : L × L
1
r
Z.
We canoni cal l y embed L in the Q-vector space
L
Q
= L
Z
Q, (4.21)
which is n-dimensional since a Z-b as is of L is a Q-basis of L
Q
, and we extend ·, · t o a
symmetric Q-bilinear form
·, · : L
Q
× L
Q
Q. (4.22)
Note that every element of L
Q
is of the form α/N for some α L and N Z
×
. The lattice
L i s non- d eg en er at e if and only if the form (4.22) is non-degenerate, and this amounts to
the con d i t io n
det[α
i
, α
j
]
i,j
= 0, (4.23)
for a Z-base {α
1
, . . . , α
n
} of L. A lattice may be equivalently defined as the Z-span of a
basis of a finite-dimension al rational vector space equipped with a symmetric bilinear form.
Let L be a lattice. For m Q, we se t
L
m
= {α L : α, α = m}. (4.24)
The latti ce L is s ai d to be even if α, α 2Z, for all α L. It is said integral if α, β Z,
for all α, β L; and is said positive definite if α, α > 0, for all α L{0}, or equivalently,
for α L
Q
{0}. The polarization formula
α, β =
1
2
α + β, α + β α, α β, β
, (4.25)
shows that an even lattice is integral. The dual of L is the set
L
= {α L
Q
: α, L Z}. (4.26)
This set is again a lattic e if an d only if L is non-degenerat e, and in this case, L
has as a
base th e dual base {α
1
, . . . , α
n
} of a given base {α
1
, . . . , α
n
} of L, defined by
α
i
, α
j
= δ
ij
for i, j = 1, . . . , n.
Note that L is integral if and only if L L
. The lattice L is said to be self-dual if L = L
.
This is equivalent to L being integral and unimodular, which means th a t
det[α
i
, α
j
]
i,j
= 1.
94
In fact, if L is integral and non-deg en er a te , the [α
i
, α
j
]
i,j
is the matrix of the embedding
map L L
with respect to the given base and its du a l base, and the unimodula r ity
amounts to the condition that this embedding be an isomorphism of abelian groups.
Generalizing (4.21) and (4. 22), we embed L i n the E-vector space
L
E
= L
Z
E,
for any field of characteristic zero, and we extend ·, · t o the symmetric E-bilinear form
·, · : L
E
× L
E
E.
Then, L is positive definit e if and only if the real vector space L
R
is an Euclidean space.
In thi s case,
|L
m
| < , for m Q, (4.27)
since L
m
is the intersection of the discrete set L with a compact set (a sphere) in L
R
.
Using the Schwarz inequality, we observe that if the lattice L is integral and positive definite,
and if α, β L
2
, then
α, β {0, ±1, ±2}
and
α, β = 2 if and only if α + β = 0,
α, β = 1 if and only if α + β L
2
,
α, β 0 if and only if α + β / L
2
{0}.
Let L be an even lattice. Set
˘
L = L/2L,
and view the elementary abelian 2 -g r ou p
˘
L as a vector space over the field F
2
. Denot e by
L
˘
L the canonical map α → ˘α = α + 2L. Since a Z-base of L reduces to an F
2
-basis of
˘
L, the n
dim
˘
L = rank L.
There is a canonical Z-bilinear form c
0
: L × L Z/2Z g iven by
(α, β) → α, β + 2Z
on L, an d c
0
is alternating because L is even. the form c
0
induces a (well-defined) altern a t ing
F
2
-bilinear map c
1
:
˘
L ×
˘
L Z/2Z given by
(˘α,
˘
β) → α, β + 2Z,
for α, β L. There i s a canonical quadratic form q
1
on
˘
L with associated bilinear form c
1
:
q
1
:
˘
L Z/2Z, with ˘α →
1
2
α, α + 2Z, (4.28)
95
for α L. this form is well-defined: if ˘α =
˘
β, then β α = γ 2L, and
1
2
β, β =
1
2
α, α + α, γ +
1
2
γ, γ
1
2
α, α (mod 2).
It is clear that c
1
is the associated form. Denote by q
0
: L Z/2Z, where
α →
1
2
α, α + 2Z,
the pu l l b a ck of q
1
to L. There exist Z-bilinear forms ǫ
0
: L × L Z/2Z su ch that
ǫ
0
(α, α) = q
0
(α) =
1
2
α, α + 2Z, forα L,
and con seq u ently
ǫ
0
(α, β) ǫ
0
(β, α) = c
0
(α, β) = α, β + 2Z,
for α L. Note that q
1
(or equivalently, c
1
) is nonsingular if, and only if, the determinant
(4.23) is odd —in p a r t icular, if L is unimodu l ar — . These considerations are important in
the con st r u ct i on of central extensions.
Let L be a positive definite lattice. The theta function θ
L
associated to L is defined to be
the form al series in the variable q = e
2πiz
(see Sect i on 6.7) given by
θ
L
(q) =
αL
q
α,α/2
=
mQ
|L
m
|q
m/2
, (4.29)
(recall (4.24) and (4.27)). If L is even and uni m odular, the theta function θ
L
has important
modular transformation properties under the modular group SL
2
(Z), when z is a comp l ex
variable in th e upper half plane H (see Ch apter 6).
4.5 The vertex algebra of an even lattice
Let L be an even lattice, i. e., a free abelian group of finite ra n k with a symmetric non-
degenerate bilinear form L × L Z
(α, β) → α, β
such tha t α, α 2Z for all α L.
Example 4.5.1. The root lattice Q of a simply laced simple Lie algebra considered in
Section 4.3 is even.
Let h = L
Z
C and consider
˜
h
=
n<0
(h t
n
).
96
Let C[L] be the group algebra of L, with basis e
γ
for γ L. Define V
L
by equati on (4.20)
V
L
= S(
˜
h
)
C
C[L].
then, it is possib l e t o make V
L
into a vertex algebra with the property that for all α L
we have
Y (1 e
α
, z) = exp
n<0
α(n)
n
z
n
exp
n>0
α(n)
n
z
n
e
α
z
α
,
where the elements α(n), e
α
and z
α
of End V
L
are defin e d as in Section 4.3 (c. f. [67] or
[113]).
More gen er al l y, for
v = (h
1
t
n
1
) ···(h
k
t
n
k
) e
α
V
L
we have
Y (v, z) = :
1
(n
1
1)!
(
d
dz
)
n
1
1
h
1
(z) ···
1
(n
k
1)!
(
d
dz
)
n
k
1
h
k
(z) Y (1 e
α
, z) :
where the normal ordered product of more th an two factors is defin ed inductively from the
right in terms of Definition 4.1.5, that is, for examp l e
: a
1
(z)a
2
(z)a
3
(z) : = : a
1
(z)
: a
2
(z)a
3
(z) :
: .
Thus, for h h and m > 0 we have
Y
(h t
m
) 1, z
=
1
(m 1)!
d
dz
m1
h(z),
where h(z) =
nZ
h(n)z
n1
and h(n) End V
L
is defined as in Section 4.3. The element
1 End V
L
is 1 1. Also, V
L
has a conformal vector ω given by
ω =
1
2
r
i=1
(h
i
t
1
)(h
i
t
1
)
1 V
L
,
where h
1
, . . . , h
r
is a basis for h, and h
1
, . . . , h
r
is the dual basis with respect to ·, ·. The
element ω is independent of this choice of basis ( c. f. [67]).
The map s L
i
: V
L
V
L
defined by L
i
= ω
i+1
satisfy
[L
i
, L
j
] = (i j)L
i+j
+
1
2
i + 1
3
dim h δ
i+j,0
1
V
L
.
Thus, V
L
is a mod ule for the Virasoro algebra under which the central element c acts as
multiplication by dim h = rank L (see [67]). Since the subspace of the Virasoro algebra
spanned by L
1
, L
0
and L
1
is a 3-dimensional subalgeb r a isomorphic to sl
2
(C), we may
regard V
L
as a sl
2
(C)-module. Then, elements of the 1-dimensional subspace spanned by
1 V
L
are ann i h i l at ed by sl
2
(C) and are, in fact, the only elements of V
L
with th i s property.
97
4.6 Dynkin Diagrams of classic Lie algebras
In this sect i on , we discuss informally (withou t proof) the cl ass ification, up to equ ivalence,
of root systems. This lead s to a classification, up to equivalence, o f semisim p l e Lie algebras.
The classification of r oot systems is given in terms of an object called the Dynkin diagram.
Suppose A = {α
1
, . . . , α
r
} is a base for a root system R. Then, the Dynkin diagram for
R (relative to the base A) is a graph having vertices v
1
, . . . , v
r
. Between any two vertices,
we place either no edge, one edge, two edges, or three edges as follows. Consider distinct
indices i and j. If the correspon d i n g roots α
i
and α
j
are orthogonal, th en we put no edge
between v
i
and v
j
. In the cases where α
i
and α
j
are not orthogona l, we put
one edge between v
i
and v
j
if α
i
and α
j
have the same length,
two edges if the longer of α
i
and α
j
is
2 longer than the shorter, and
three edges if the longer of α
i
and α
j
is
3 longer than the shorter.
In addition, if α
i
and α
j
are not orthogonal and not of the same length, then we decorate
the edges between v
i
and v
j
with an arrow pointing from the vertex associated to the longer
root toward the vertex assoc ia t ed to the shorter root ( thinking of the arrow as a ‘greater
than’ sign). Th ese are all the possible values, since we have [91, p.246–249]
Proposition 4.6.1. Suppose that α, β are roots (of a root system R), α i s n ot multiple of
β, and α, α β, β. Then, one of the following holds:
1. α, β = 0;
2. α, α = β, β, and the angle between α and β is 60
o
or 120
o
;
3. α, α = 2β, β, and the angle between α and β is 45
o
or 135
o
;
4. α, α = 3β, β, and the angle between α and β is 30
o
or 150
o
.
Proposition 4.6.2. If α and β are distinct elements of a base for R, then α, β 0.
Thus, Propositi on 4.6.1 tells us that if α
i
and α
j
are not orthogonal, then the only possible
length ratios are 1,
2 and
3. Furthermore, Proposition 4.6.2 says that these three cas es
correspond to angles of 120
, 135
, and 150
, respectively.
Two Dynkin diagram s are said to be equivalent if there is a one-to-on e, onto map of the
vertices of one to the vertices of the other that preserves the number of bonds and t he
direction of the arrows. Recall from S e ct io n 3.5 th a t any two bases for the same root
system can be mapped into one another by the action of the Weyl group. This implies
that the equivalence cl ass of the Dynkin diagra m is independent of the choice of base. As
we wil l see, only graph s of certain very special forms arise s as Dynkin di ag r am of a root
system. The following characterizes root systems via their Dynkin diagram [91, p. 27 0]
98
Theorem 4.6.3. A root system is irreducible if and only if its Dynkin diagram is connected.
Two root systems with equivalent Dynkin diagrams are equivalent. If R
is the dual root
system to R, then the Dynkin diagram of R
is the same as that of R (with direction of
each arrow reversed).
So, t h e classification of irreducible root systems amounts to classifying all the connected
diagrams that can arise as Dynkin diagrams of root systems. It is well known the classi-
fication of the D yn k i n diag r am s for the classical Lie algebras, sl
n
(C), so
n
(C), and sp
n
(C)
(see Figu r e 4.1):
A
n
: The root system A
n
is the root system of sl
n+1
(C), whi ch has rank n.
B
n
: The root system B
n
is the root system of so
2n+1
(C), wh i ch has rank n.
C
n
: The root system C
n
is the root system of sp
n
(C), wh i ch has rank n.
D
n
: The root system D
n
is the root system of so
2n
(C), whi ch has rank n.
. . .
. . .
. . .
. . .
A
n
B
n
C
n
D
n
Figure 4. 1: Dynki n diagram s for A
n
, B
n
, C
n
and D
n
.
Certain special things happen in low rank. In r a n k one, t h er e is only one possible Dynkin
diagram, reflecting that there is only one isomorphism class of complex semisimple Lie
algebras in rank one. The Lie algebra so
2
(C) is not semisimple and th e remaining three,
sl
2
(C), so
3
(C), and sp
1
(C), are isomorphic. In r a n k two, the Dynkin diagram D
2
is discon-
nected, reflecting that so
4
(C)
=
sl
2
(C) sl
2
(C). Also, the Dynkin diagr am s B
2
and C
2
are
isomorphic, reflecting that so
5
(C)
=
sp
2
(C). In ra n k three, the Dynkin diagrams A
3
and
D
3
are isomorp h i c, thus sl
4
(C)
=
so
6
(C). We may obser ve certain things about the short
and long roots in root systems where more than one length of root occurs. The lon g roots
in B
n
form a root system by t hemselves, namely D
n
. The short roots in B
n
form a root
system by themselves, na m el y A
1
× ··· × A
1
. In C
n
, it is the reverse: the long roots for m
A
1
× ··· × A
1
and the short roots form D
n
.
In ad d i t io n to the root systems associated to the classical Lie algebras, there are five
‘exceptional’ irreducible root syst em s, denoted G
2
, F
4
, E
6
, E
7
, and E
8
, whose Dynkin
99
diagrams are shown in Figure 4.2. For t h e con st r u ct i on of th es e except i onal root systems,
see Humphreys [100]. Thus we have the following class ification theorem for irreducible root
systems [ 91 , p.272]:
E
6
E
7
E
8
F
4
G
2
Figure 4.2: The exceptional Dynkin diagrams E
6
, E
7
, E
8
, F
4
and G
2
.
Theorem 4.6.4 (Classification of root systems). Every irreducible root system is isomor-
phic to precisely one root system from the following :
1. The classical root systems A
n
, n 1;
2. The classical root systems B
n
, n 2;
3. The classical root systems C
n
, n 3;
4. The classical root systems D
n
, n 4;
5. The exceptional root systems G
2
, F
4
, E
6
, E
7
, and E
8
.
In the language of semisimple Lie theory (recall Chapter 3), suppose that h is a Cartan
subalgebra of a r ed u ct i ve semisimple Lie algebra g. Suppose also that is the root system
of g with respect to h. We know that is identified with the dual h
by means of the
bilinear form ·, ·, and the x
α
(for α ∆) are the correspondi n g root vectors. Together
with a basis of the root system ∆ , the x
α
’s form a Chevalley basis of g. The sublattice
Q = Z =
n
i
α
i
: n
i
Z, α
i
(4.30)
of some lattice L, generated by is the root lattice of g, and its dual
Q
= {α h : α, Q Z} (4.31)
100
is the so called weight lattice. We now list examples of positive definite even l a tt i ces L
such that δ = L
2
spans h and is indecomposable. Thus, the corresponding Lie algebra g is
simple. In each case, L is generated by ∆, that is, L = Q = Z∆. The not at i ons A
n
, D
n
and E
n
correspond to t he standard designation s of the simple Lie algebras given above. In
each case, n = rank Q = dim h, and recall that dim g = n + ||.
For 1, denote by V
an -dimensional rational vector space equipped with a positive
definite symmetric form ·, · and an orthonormal basis {v
1
, . . . , v
}.
Type A
n
, n 1: In V
n+1
take
Q
A
n
=
n+1
i=1
m
i
v
i
: m
i
Z,
m
i
= 0
.
Then, = (v
i
v
j
) : 1 i < j n + 1}, || = n(n + 1) and dim g = (n + 1)
2
1. As
we have mentioned above, the case A
1
is the case g = sl
2
(C).
Type D
n
, n 3: In V
n
take
Q
D
n
=
n
i=1
m
i
v
i
: m
i
Z,
m
i
2Z
.
Then, = v
i
± v
j
: 1 i < j n}, || = 2n(n 1) and dim g = n(2n 1). As we
have mentioned above, the case D
3
is t h e same as the case A
3
. The sam e construct i on for
n = 2 gives A
1
× A
1
.
Type E
8
: In V
8
take
Q
E
8
= Q
D
8
+
1
2
Z
8
i=1
v
i
=
n
i=1
m
i
v
i
: either m
1
, . . . , m
8
Z, or m
1
, . . . , m
8
Z +
1
2
;
m
i
2Z
.
Then, = v
i
± v
j
: 1 i < j 8}
8
i=1
m
i
v
i
: m
i
= ±
1
2
,
m
i
2Z
, || = 240
and d im g = 248.
Type E
7
: In V
8
take
Q
E
7
= Q
E
8
8
i=1
m
i
v
i
: m
i
1
2
Z,
m
i
2Z
.
101
Then, = v
i
±v
j
: 1 i < j 6}(v
7
v
8
)}
6
i=1
m
i
v
i
: m
i
= ±(v
7
v
8
), m
i
=
±
1
2
,
m
i
2Z
, || = 126 and dim g = 133.
Type E
6
: In V
8
take
Q
E
6
= Q
E
7
8
i=1
m
i
v
i
: m
i
1
2
Z, m
6
= m
7
.
Then, = v
i
± v
j
: 1 i < j 5}
±
5
i=1
m
i
v
i
+
1
2
(v
6
+ v
7
v
8
)
: m
i
=
±
1
2
,
m
i
2Z
1
2
, || = 72 and dim g = 78.
4.7 Lie theory and Moonshine
McKay not only noticed (1.1), but also observed t hat
j(z)
1/3
= q
1/3
(1 + 248q + 4, 124q
2
+ 34, 752q
3
+ . . .). (4.32)
The point is that 248 is the dimension of the defining rep r esentation of the E
8
simple Lie
algebra, while 4, 124 = 3, 875 + 248 + 1 and 34, 752 = 30, 380 + 3, 875 + 2 · 248 + 1. Inci-
dentally, j
1/3
is the Hauptmodul for the genus 0 congruence group Γ(3). Thus Moonshi n e
is relat ed somehow to Lie theory.
McKay later foun d independent r el at i o n sh i p s with Lie theory [142], [17], [74], reminiscent
of his famous A-D-E correspondence with finite subgroups of SU
2
(C). As mentioned in
Chapter 1, M has two conjugacy classes of involutions. Let K be the small er one, called
‘2A’ in [36] (the alternati ve, class ‘2B’, has almost 100 mill i on times more elements). The
product of any two elements of K will lie in one of nine conjugacy cla sses : n a m el y, 1A, 2A,
2B, 3A, 3C, 4A, 4B, 5A, 6A, cor r esponding respectively to elements of orders 1, 2, 2, 3, 3,
4, 4, 5 , 6. It is surpr is ing that, for such a complicated group as M, that list stops at only 6
—we ca l l M a 6-transposi t i on grou p for this reason—. The punchline: McKay noticed that
those ni n e numbers ar e precisely t h e labels of the affine
ˆ
E
8
Dynkin diagram (see Figure
4.3). Thus we can attach a conjugacy class of M to each vertex of the
ˆ
E
8
diagram. An
interpretation of t he edges in the
ˆ
E
8
diagram, in ter m s of M, is unfortunately not known.
We can’t get the affine
ˆ
E
7
labels in a similar way, but McKay noticed that an order two
folding of affi n e
ˆ
E
7
gives the affine
ˆ
F
4
diagram, and we can ob t ai n its label s using the
Baby Monster B (the second largest sporad i c) . In parti cu l ar , let K now be the smallest
conjugacy class of involutions in B (also labeled ‘2A’ in [36]); the conjugacy classes in K
have or d er s 1, 2, 2, 3, 4 (B is a 4-transposition group), a n d these are the labels of the
ˆ
F
4
affine Dynkin diagram. O f course we prefer
ˆ
E
7
to
ˆ
F
4
, but perhaps that two-folding has
something to do with the fact that an or d er - two central extensi on of B is the centraliser of
an elem ent g M of order two.
102
3 2 1
2 4 3 2 1
2 4
3
6 5 4 3 2 1
ˆ
E
8
ˆ
F
4
ˆ
G
2
Figure 4.3: The affine Dynkin diagrams
ˆ
E
8
,
ˆ
F
4
and
ˆ
G
2
with l abels.
Now, the triple-folding of the affine Lie algebra
ˆ
E
6
is th e affine
ˆ
G
2
. The Monster has three
conjugacy classes of order three. The smallest of these (‘3A’ in Atlas notation) has a cen-
traliser which is a triple cover of the Fischer group Fi’
24
. Taking the smalle st conjugacy
class of involutions in Fi’
24
, and multiplying it by it s elf, gives conjugacy classes with orders
1, 2, 3 (hence Fi’
24
is a 3-tran sposition group) and those not surprisingly are the labels
of
ˆ
G
2
. Although we now understand (4.32) (see Chapter 9) and have proved the basic
Conway-Norton conjecture (see Chapter 8), McKay’s observation about
ˆ
E
8
,
ˆ
F
4
and
ˆ
G
2
diagrams still have no explanati on . In [74] these patterns are extended, by relating variou s
simple groups to the
ˆ
E
8
diagram with de let ed nodes.
Shortly after McKay’s E
8
observation, Kac [111] and James Lepowsky [132] independently
remarked that the unique level-1 highest - weight representation L(ω
0
) of the affine Kac-
Moody algebra E
(1)
8
has graded dimension j(z)
1/3
. Since each homogeneous piece of any
representation L(λ) of the affine Kac-Moody algebra X
(1)
(in Kac’s notat i o n ) must carry a
representation of the associated finit e- d i m en si onal Lie group X
(C), and the graded dime n -
sions (multiplied by an appropriate power of q) of an affine algebra are modular functions
for some G SL
2
(Z), this explained McKay’s E
8
observation. His observation (1.1) took
longer to clarify because so much of the mat h em a ti cs needed was still to be developed.
103
Chapter 5
The Moo nshi ne module
Following [67], we shall construct in this chapter a module with an associated vertex oper-
ator algebra V
on which th e Monster group M acts as group of automorphis m s. We shall
call this the Monster vertex algebra or the Moonshine module. We cannot describe this
vertex algebra in detail here —in d e ed a lengthy book i s needed to do this—. However, we
shall mention some of its most basic properties.
The vertex operator algebra associated with V
crowns the sequ ence of exceptional struc-
tures starting with the Golay error-correcting code and continuing with the Leech lattice.
The corresponding sequence of their automorphisms consist of the Mathi eu sporadic group
M
24
, the Conway sporadic group Co
0
and the Monster M.
We begi n this chapter with introductions to the two exceptional structures mentioned
above. For more det ai l s, we refer to the papers [1 30 ] - [1 31 ] , [29]-[33], to the books [86],
and to the extensive collection [39]. The history of the discoveries related to to Monstrous
Moonshine is rev i ewed in Chapter 1.
5.1 The Golay code
Let a finite set with n elements. The power set (Ω) = {S : S } can be viewed as an
F
2
-vector space under the operation + of symmetric difference. By a (binary linear) code
we shall understand an F
2
-subspace of (Ω). An isomorphis m of codes is defined in the
obvious way. The ca r d i n al i ty |C| of an element C of a code is called the weight of C. A
code C is said to be of type I if
n 2Z, |C| 2 Z for all C C and C ,
and C is said to be of type II if
n 4Z, |C| 4 Z for all C C and C .
The typ e II codes will be seen as analogues of even lattices, and will help us construct them .
For a code C , the dual code C
is given by
C
= {S : |S C| 2Z, for all C C }.
105
Thus, C
is the annihi l at or of C in (Ω) with respect to the natural n on s ingular symmetric
bilinear form
(S
1
, S
2
) → |S
1
S
2
| + 2Z (5.1)
on (Ω) . Hence
dim
F
2
C
= n dim
F
2
C .
We call C self-dual if C
= C , in which case n is even and dim
F
2
C =
n
2
.
Consider the subs p ace
E(Ω) = {S : |S| 2Z}.
The map q : E(Ω) Z/2Z = F
2
given by S →
|S|
2
+ 2Z is a quadratic form on E(Ω), with
associated bilinear form given by (5. 1 ) . In case n 2Z, F
2
is the radical form of q. A
subspace of a space with a quadratic form is called totally singular if the form vanishes on it.
Remark 5.1.1. In case n 4Z, the self-dual codes of type II correspon d to the (maximal )
totally singular subspaces of E(Ω)/F
2
of dimension
n
2
1. Equivalently, t h e type II self-
dual codes ar e the (maximal) totally singular subspaces of E(Ω) of dimensi on
n
2
.
For a code C , set
w(C ) =
CC
q
|C|
Z[q]
the weight distribution of C . A code with the following properties is called a Hamming
code (see [67, p.300]):
Theorem 5.1.2. There is a self-dual code of type II on an 8-element set .
It can be proved t h a t the weight distribution of th e constru ct ed Hammin g cod e is 1+14q
4
+
q
8
. The Hamming code is unique up to isomor p h i sm (see, e. g. [ 13 6 ] ).
Definition 5.1.3. A code with th e following properties is called a (binary) Golay code [67,
p.301]:
Theorem 5.1.4. There is a self-dual code C of type II on a 24-element set, such t hat C
has no elements of weight 4.
It is not hard to see that the weight distribution of the constructed Golay code is 1 +
759q
8
+ 2, 576q
12
+ 759q
16
+ q
24
(see [46]). Note that 759 =
24
5
/
8
5
. Th e Golay code is
unique up to isomorphism (see, e. g. [136] ) . The 759 elements of the Golay code of weight
8 are called octads. We m ention some properties of the Golay code [67 , p.302].
Proposition 5.1.5. Let be a 24-element set and let C be a Golay code in E(Ω).
(a) Every 5-element subset of is included in a unique octad in C .
(b) Let T
0
be a 4-element subset of . Then T
0
lies in exactly 5 octads. These are of the
form T
0
T
i
, for i = 1, . . . , 5, where =
5
i=0
T
i
is a disjoint union of 4-element sets (called
a sextet in [32]). The union of any pair of T
i
is an octad.
106
From thi s, it follows directly that the octads generate the Golay code. The group of
automorphisms of the Golay code C is called the Mathieu group M
24
:
M
24
= Aut C .
This is a nonabelian simple group (see Table 1.1). In fact, in the natural rep r ese ntation of
M
24
on (Ω) , the complete list of submodules is
0 C E(Ω) (Ω).
In part i cu l a r, C / and E(Ω)/C are faithful irreducible modules for M
24
.
5.2 The Leech lattice
A self-dual code C of type II based on a set gives rise to an even un im odular lattice as
follows. L et
h =
k
Fα
k
.
be a vector space with basis {α
k
: k } and provide h t h e symmetric bilinear form ·, ·
such tha t
α
k
, α
= δ
kℓ
, for k, .
For S Ω, set α
S
=
kS
α
k
. Defin e
Q =
k
Zα
k
,
and for a code C based on define the positive de fi n i t e lattice
L
0
=
k
m
k
α
k
: m
k
1
2
Z, {k|m
k
Z +
1
2
} C
. (5.2)
(there should be no confusion with notation in (4.24)). Then L
0
is even if and only if |C|
4Z, for all C C . Moreover, the dual lattice L
0
of L
0
(recall (4.26)) i s the correspondin g
lattice based on the dual code C
:
L
0
=
k
m
k
α
k
: m
k
1
2
Z, {k|m
k
Z +
1
2
} C
.
Thus we have
Proposition 5.2.1. A code C is self-dual of type II if, and only if, the corresponding lattice
L
0
is even self-dual, or equivalently, even unimodular.
107
Now consider the following modification of the lattice L
0
associated with a code C , which
we now assume contains Ω:
L
0
=
CC
Z
1
2
α
C
+
k
Z(
1
4
α
α
k
)
=
CC
Z
1
2
α
C
+
k,ℓ
Z(α
k
α
) + Z(
1
4
α
α
k
0
), (5.3)
where k
0
is a fixed element of Ω. Since the lattice
L
0
L
0
=
CC
Z
1
2
α
C
+
k,ℓ
Z(α
k
α
) (5.4)
has index 2 in both L
0
and L
0
, then L
0
is unim odular i f and only if L
0
is. A necessary and
sufficient condition for L
0
to be even is that
n = || 8(2Z + 1),
since
1
4
α
α
k
,
1
4
α
α
k
=
n
8
+ 1, (5.5)
for k Ω, and we find
Proposition 5.2.2. If n 8(2Z + 1) and the code C is self-dual of type II, then the
corresponding lattice L
0
is even unimodular.
Definition 5.2.3. The Leech lattice is the even unim odular lattice
Λ = L
0
for th e case n = 24 and C th e Golay code (recall Theorem 5.1.4).
The lattice Λ has no ‘short’ elements, i. e., Λ
2
= Ø (using notation (4.24)). This can be
checked from equation (5.3) and the correspon d i n g property of the Golay code, but we
shall prove Λ
2
= Ø and recon st r u ct Λ inst ead by using another principle, which will be an
analogue for latt i ces of Remark 5.1.1 for codes.
Let L be an even uni m odular latti ce of rank n and with form · , ·. For our speci al purpose,
now we provide L with the following rescaled form:
α, β
1/2
=
1
2
α, β for α, β L (5.6)
and by abuse of notation we drop the subscript 1/2. With respect to the new form ·, ·, L
has th e following properties:
α, α Z, for all α L.
det[α
i
, α
j
]
i,j
=
1
2
n
, for a base {α
1
, . . . , α
n
} of L.
108
As we did in Section 4.4 (bu t keeping in mind the rescaled no rm ), set
˘
L = L/2L,
an n-dimensional vector space over F
2
, and write α → ˘α for the canonical m ap. Since the
original lattice L is even, the map q
1
:
˘
L Z/2Z = F
2
given by
˘α → α, α + 2Z
defines a quadratic form on
˘
L with associated biline ar form c
1
:
˘
L ×
˘
L F
2
determined by
(˘α,
˘
β) → 2α, β + 2Z.
These forms are nonsingul ar since the original l at t ic e L is unimodular. From these defini-
tions we have
Proposition 5.2.4. Let M be a lattice such that 2L M L. Then, M is even unimod-
ular with respect to the new form ·, · if, an d only if,
˘
M = M/2L is a (maximal) totally
singular subspace of
˘
L of dimension
n
2
.
We shall apply this last principle to the direct sum of three copies of the r oot lattice Q
E
8
,
(recall Section 4. 6 ) which is an even unimodular lattice of rank 8. For brevity, set
Γ = Q
E
8
and provide Γ with the rescaled form (5.6) as above. Using the Hamming code (Theorem
5.1.2), we shall first show that
˘
Γ contains complement ar y 4-dimensional totally singular
subspaces.
We can describe the lattice Γ as follows:
Γ =
k,ℓ
Z(
1
2
α
k
±
1
2
α
) + Z
1
4
α
=
k
m
k
α
k
: either m
1
, . . . , m
8
1
2
Z, or m
1
, . . . , m
8
1
2
Z +
1
4
;
m
k
Z
.
Now we identify with the projective line over the 7-element field: = P
1
(F
7
) = F
7
{∞}.
Consider the sets of squares and non-squares
Q = {x
2
: x F
7
} = {0, 1, 2, 4},
N = Q = {3, 5, 6, ∞};
and d efi n e subspaces
C
1
= N + i|i F
7
,
C
2
= −N i|i F
7
,
109
of E(Ω). Then it follows that C
1
and C
2
satisfy the prop er t i es of Theorem 5.1.2. Now,
consider the latti ces
Φ = L
0
for C = C
1
,
Ψ = L
0
for C = C
2
.
Then, Φ and Ψ are even unimodular lattices by Propositions (5.2.1) and (5.2.2). Moreover,
Φ, Ψ Γ, and since C
1
+ C
2
= E(Ω), we see that
Φ + Ψ = Γ.
Proposition (5.2.4) thus gives
Proposition 5.2.5. We have a decomposition
˘
Γ =
˘
Φ
˘
Ψ (5.7)
into 4-dimensional totally singular subspaces. In particular, Φ + Ψ = Γ and Φ Ψ = .
Using the decomposition (5.7) we shall now reconstruct the L e ech lat t i ce by analogy with
the con st r u ct i on of the Golay code in Theorem 5.1.4. Set
Γ
3
= Γ Γ Γ
the orthogonal direct sum of three copies of Γ, equipped with the modified form (5.6). In
Γ
3
, set
Λ = {(φ, φ, 0) : φ Φ} {(φ, 0, φ) : φ Φ} {( ψ, ψ, ψ) : ψ Ψ}. (5.8)
Then we have [67, p.306]
Theorem 5.2.6. The lattice Λ in (5.8) is an even unimodular latti ce such that Λ
2
= Ø.
Moreover, it coincides with the Leech lattice in Definition 5.2.3.
The Leech lattice is the uniqu e positive definite even unimod u l ar lattice Λ of rank 24 with
Λ
2
= Ø, up to isometry (see [153], [30]), and the E
8
-root lattice is the unique positive
definite even unimodula r lattice of rank 8 up to isometry. Thus the lattices Γ, Φ and Ψ
with its or i g inal bil inear form, are all isometric. In fact, the construction (5.8) expresses
the Leech lattice as the non-orthogonal direct sum or t h r ee rescaled copies of the E
8
-root
lattice. Sometimes the L eech lattice Λ is denoted as Λ
24
.
The grou p of isometries of the Leech latti ce is called the Conway group Co
0
(or ·
0
)
Co
0
= Aut(Λ, ·, ·)
= {g Aut Λ : gα, gβ = α, β for all α, β Λ}.
Its quot i ent by the central subgroup ±1 is called the Conway group Co
1
Co
1
= Co
0
/±1.
We cite some basic facts about this gr oups [31, 32]:
110
Co
0
equals i t s commutator subgroup: Co
0
= [Co
0
, Co
0
];
Cent Co
0
= ±1;
Co
1
is a nonabeli an simple group (see Table 1.1);
Co
1
acts fait h fu l l y and irreducibly on λ/2Λ.
Only as a comment, we have constructed t h e L ee ch lattice as the lattice L
0
based on the
Golay code. The lattice L
0
based on the Golay code is also even unimodular, but (L
0
)
2
= Ø.
In fact, (L
0
)
2
= α
i
: i }. This lattice L
0
is called the Niemeier lattice of type A
24
1
and is somet im es written as L
0
= N(A
24
1
). We have encountered a t h i r d even unimod u l ar
lattice of r an k 24, namely, Γ
3
. Altogether, there are 24 even unimodular lattices of rank
24, u p to isometry, called the Niemeier latti ces (see [153]).
Next, we shall describe and co u nt th e shortest nonzero elem ents of the Leech lattice. In
the notation of the beginning of this section, for S let ǫ
S
be the involution of h given
by
ǫ
S
: α
k
→
α
k
if k S
α
k
if k / S
,
for k Ω. It follows from relations (5.2), (5.3), (5.4) and (5.5) that Λ
4
is composed of three
types of elements:
Λ
4
= Λ
1
4
˙
Λ
2
4
˙
Λ
3
4
,
where
Λ
1
4
= {
1
2
ǫ
S
α
C
: C C , |C| = 8, S C, |S| 2Z};
Λ
2
4
= α
k
± α
: k, , k = };
Λ
3
4
= {ǫ
C
(
1
4
α
α
k
) : C C , k }.
Counting, we find that
|Λ
4
| = |Λ
1
4
| + |Λ
2
4
| + |Λ
3
4
| (5.9)
= 759 · 2
7
+
24
2
· 2
2
+ 24 · 2
12
(5.10)
= 196, 560. (5.11)
5.3 The Monster vertex algebra V
and the Griess al-
gebra B
Now that we have the Leech lattice available, we can construct one of our main objects of
study. We have alread y s een t hat t he Leech lattice i s t h e unique even unimodular lattice
111
of rank 24, such that has no elements of norm 2.
First, using the Leech lattice we form the untwisted space
V
Λ
= S(
˜
h
)
Z
C[Λ] (5.12)
as in (4.20). Let
ˆ
L be a central extension of a l at t ic e L by a finite cyclic group κ =
{k|k
s
= 1} of order s, and den ot e by
c
0
: L × L Z/sZ
the associated commutator map, so that
aba
1
b
1
= κ
c
0
a,
¯
b)
, for a, b
ˆ
L.
We make the following special choices: we fix the central extension
1 κ
ˆ
Λ Λ 1, (5.13)
where κ
2
= 1 , κ = 1 (i. e., s = 2), and where the com mutator map is the alternating
Z-bilinear map
c
0
(α, β) = α, β + 2Z, for α, β Λ. (5.14)
Then, t aki ng a 2-th primitive root of u nity ξ = 1, we have
C[Λ] = C[
ˆ
Λ]/(κ + 1)C[
ˆ
Λ]
and
c(α, β) = ( 1)
α,β
, for α, β Λ,
where c : Λ × Λ C
×
is the map (α, β) → ξ
c
0
(α,β)
. So that, as operators on V
Λ
, κ = 1
and
ab = (1)
¯a,
¯
b
ba, for a, b
ˆ
Λ.
The automorphisms of
ˆ
Λ which induce the involution 1 on Λ are automatically involutions
and are parametrized by t h e quadratic forms on Λ/ with associated form induced by
(5.14). Among these, we fix the distinguished involution θ
0
determined by the canonical
quadratic form q
1
given in (4.28)
q
1
: Λ/ Z/2Z, with α + →
1
2
α, α + 2Z.
Then, we have an involution θ
0
:
ˆ
Λ
ˆ
Λ given by
a → a
1
κ
¯a,¯a/2
.
Besides the general properties θ
2
0
= 1 and θ
0
(a
2
) = a
2
, for all a
ˆ
Λ, we observe that
θ
0
(a) = a
1
if ¯a Λ
4
.
112
We also form the twisted space
V
T
Λ
= S(
˜
h
)
Z+
1
2
T,
where T is any Λ-module such that κ · v = ξv, for all v T . Keepin g in mind the choices
above, we set
K = {θ
0
(a)a
1
: a
ˆ
Λ} = {a
2
κ
¯a,¯a/2
: a
ˆ
Λ},
which is a central subgroup of
ˆ
Λ such that K = 2Λ. Then,
ˆ
Λ/K is a finite group which is
a central extension
1 κ
ˆ
Λ/K Λ/ 1,
with commutator map indu ced by (5.14) and with squaring m ap the quadratic form q
1
.
Since Λ is unimodular, q
1
is non - si n gular, and
ˆ
Λ/K i s an extraspecial 2-group with
|
ˆ
Λ/K| = 2
25
. (5.15)
In fact, in the twisted space V
T
Λ
we take T to be the can o n i cal
ˆ
Λ-module described in (5.12).
Of cour se for a
ˆ
Λ, we have θ
0
(a) = a as operators on T .
Now we can define the Moonsh ine module —the space on which the Monster group will
act—. Recal l that θ
0
acts in a natural way on V
Λ
given by
θ
0
: x i(a) → θ
0
(x) i( θ
0
(a)), for x S(
˜
h
) and a
ˆ
Λ; (5.16)
and on V
T
Λ
, by
θ
0
: x τ → θ
0
(x) ( τ ) = θ
0
(x) τ, for x S(
˜
h
) and τ T. (5.17)
Let V
θ
0
Λ
and (V
T
Λ
)
θ
0
be the subspaces of V
Λ
and V
T
Λ
of θ
0
-invariant elements. We know
that for v V
θ
0
Λ
the component operators of both the untwisted and the twisted vertex
operators Y (v, z) preserve the respective fixed spaces V
θ
0
Λ
and (V
T
Λ
)
θ
0
.
Definition 5.3.1. We define the Moonshine module to be the space
V
= V
θ
0
Λ
(V
T
Λ
)
θ
0
.
The symb o l is for ‘natural’. For v V
Λ
, we form the vertex operator
Y (v, z) = Y
Z
(v, z) Y
Z+
1
2
(v, z)
acting on the larger space
W
Λ
= V
Λ
V
T
Λ
. (5.18)
Similarly, for the component operators of Y (v, z) we write v
n
= v
n
v
n
and x
v
(n) =
x
v
(n) x
v
(n) on W
Λ
, for v V
Λ
, n Q. Then
v
n
· V
V
, x
v
(n) · V
V
,
113
if v V
θ
0
Λ
. In fact, V
can be given the structure of a vertex operator algebra, the Monster
vertex algebra. V
has a conformal vector ω of central charge 24. Thu s, the linear maps
L
i
: V
V
given by L
i
= ω
i+1
satisfy
[L
i
, L
j
] = (i j)L
i+j
+ 12
i + 1
3
δ
i+j,0
1
V
,
and thu s give a representation of the Virasoro algebra in which the central element c is
represented by 24 · 1
V
(recall (4.13)).
Because dim h = 24 and Λ
2
= Ø, the space V
has some special structu ra l features. First,
V
is integrally graded, with degrees bounded below by 1:
V
=
nZ
V
n
,
with V
n
= 0, for all n < 1. In particular,
V
1
= F
(1),
V
0
= 0,
V
1
= f p,
where, i n more detail
f = S
2
(h)
α
ˆ
Λ
4
Fx
+
α
,
p = h T.
Definition 5.3.2. We define the Griess module to be the space
B = V
1
= f p.
We have counted the elements of Λ
4
in (5. 11 ) , and we find that
dim V
1
= 1,
dim V
0
= 0,
dim B = dim V
1
= 196, 884;
since
dim f = 300 +
1
2
(196, 560) = 300 + 98, 280 = 98, 580 (5.19)
and
dim p = 24 · 2
12
= 98, 304. (5.20)
114
In fact we have (recall Chapter 1)
nZ
(dim V
n
)q
n
= J(z) = q
1
+ 0 + 196, 884q + 21, 493, 760q
2
+ . . . ,
where q = e
2πiz
, z H.
It is shown in the last four chapters of [67] that the Monster group M acts as a group
of automorphisms of V
. The subgroup of M preserving the subspaces V
θ
0
Λ
and (V
T
Λ
)
θ
0
is
the centralizer of an involution in M. This is an exten s io n of th e extraspecial group
ˆ
Λ/K
of order 2
25
in (5.15) by Conway’s sporadic group Co
1
, which we have seen is re la t ed to
the Leech lattice Λ. A crucial part of the Frenkel-Lepowsky-Meurman constructi o n is to
find an involution in the Monster M which acts on V
, but which does not preserve the
subspaces V
θ
0
Λ
and (V
T
Λ
)
θ
0
.
The confor m al vector ω V
lies in a 1-dimensional subspace Cω V
1
, invariant under
the action of M. The co m p l em entary submodule of Cω in V
1
gives the smallest non-
trivial representation of M of degree 196,883. This then, is the explanation of McKay’s
observation (1.1) con s idered Chapter 1. The Monster vertex algebra V
is a graded module
whose graded components have dimension given by th e coefficients of the J(z) function,
and the Monster M acts on each graded component. On V
, the Jacobi identity takes the
following simple form for vertex operators parametrized by V
θ
0
Λ
[67, p .3 16] :
Theorem 5.3.3. For v V
θ
0
Λ
, we have
Y (v, z) =
nZ
v
n
z
n1
on V
,
that is, Y (v, z) involves only integral powers of z. For u, v V
θ
0
Λ
, we have
[Y (u, z
1
) ×
z
0
Y (v, z
2
)] = z
1
0
δ
z
1
z
2
z
0
Y (u, z
1
)Y (v, z
2
) z
1
0
δ
z
2
z
1
z
0
Y (v, z
2
)Y (u, z
1
)
= z
1
2
δ
z
1
z
0
z
2
Y
Y (u, z
0
)v, z
2
on V
. In particular, V
θ
0
Λ
is a vertex operator algebra of central charge 24 and (V
T
Λ
)
θ
0
is a
V
θ
0
Λ
-module.
Moreover, we know the following [ 67 , p.317]:
Theorem 5.3.4. The space f is a commutative nonassoc iati ve algebra with identity under
the product
u × v = u
1
· v,
and the bilinear form
u, v = u
3
· v
115
is nonsingular, symmetric and associative. The space V
is a graded module for the com-
mutative affinization
ˆ
f of f under the action π :
ˆ
f End V
defined by
π :
u t
n
→ x
u
(n) for u f, n Z
e → 1.
(5.21)
One of the main results obtained in the work of Frenkel, L epowsky and Meurman [67] is
an extension of this action to a rep r esentation
ˆ
B End V
(5.22)
of a la r ger commutative affi n i zat i on by cross-bracket on V
, where the space B is given the
structure of a commutative nonassoci a ti ve algebra with identity, and with a nonsingular
symmetric associative form in the fol l owing way: the product × an d the form ·, · on B
extend those on f. For u f and v p, we use the product in Theorem 5.3.4 and the
commutativity of this product on f as mot i vation to define
u × v = v × u = u
1
· v.
Similarly, we use the bili n e ar form in Theorem 5.3.4 and the symmetry of this form on f as
motivation to d efine
u, v = v, u = u
3
· v = 0
(the fact that u
3
·v = 0 is obtained by consideration of the gradation of (V
T
Λ
)
θ
0
). Now, the
identity element
1
2
ω on f is also an identity element on B. We defi n e next a nonsingular
symmetric bilinear form  ·, · on p = h T by the formula
h
1
τ
1
, h
2
τ
2
=
1
2
h
1
, h
2
τ
1
, τ
2
( 5. 23)
for h
i
h, τ
j
T . Finall y, we define a product × on p so that p × p f, and uniquely
determined by the nonsingul a r ity of the form on f and the associativity condition
u, v × w = u × v, w, for u, v p, w f. (5.24)
The commutativity of this product on p follows from the explicit formula for it given below.
Definition 5.3.5. The resulting nonassociative alg eb r a B equipped with this form ·, · is
called the Griess algebra.
Here, the Griess algebra B = f p is actually a slight modification, with a natural identity
element, of the algebra defined in [79], [80]. Summarizing,
Proposition 5.3.6. The Griess algebra B is a commutative nonassociative algebra with
identity element
1
2
ω f, and the form ·, · on B is nonsingular, symmetric and associative.
We have
f × f f,
f × p p,
p × p f,
116
with explicit formulas given by:
g
2
× h
2
= 4g, hgh,
g
2
× x
+
a
= g, ¯a
2
x
+
a
,
x
+
a
× x
+
b
=
0 if ¯a,
¯
b = 0, ±1
x
+
ab
if ¯a,
¯
b = 2
¯a
2
if ab = 1
,
for g, h h, a, b
ˆ
L
4
and x
+
a
= i(a) + i(θa) = i(a) + θi(a);
(x
+
a
)
1
· (h τ) =
1
8
(h 2¯a, h¯a) a · τ,
(g
2
)
1
· (h τ) = (g, h +
1
8
g, gh) τ,
for g, h h, a
ˆ
L
4
, τ T ;
(h
1
τ
1
) × (h
2
τ
2
) =
1
8
2h
1
h
2
+ h
1
, h
2
1
2
ω
τ
1
, τ
2
a · τ +
+
1
128
a
ˆ
Λ
4
h
1
, h
2
2¯a, h
1
¯a, h
2
τ
1
, a · τ
2
x
+
a
,
for h
i
h, τ
j
T . We also have
f, p = 0,
and explicit formulas for the form on f and p given by:
g
2
, h
2
= 2g, h
2
,
g
2
, x
+
a
= 0,
x
+
a
, x
+
b
=
0 if ¯a = ±
¯
b
2 if ab = 1
,
for g, h, a, b as above and (5.23)
h
1
τ
1
, h
2
τ
2
=
1
2
h
1
, h
2
τ
1
, τ
2
.
The identity element satisfies
1
2
ω,
1
2
ω = 3.
The significance of B is that Griess, who introduced thi s al gebra, constructed a group
of automorphisms of it, preserving the form ·, ·, and showed this group to be a finite
simple group: the Monster M (see [80]). In fact, Tits showed that the Monster is the full
automorphism group of B (see [174], [175]). Having reconstr u ct ed the Griess algebra using
properties of vertex operators, we shall also reconstruct the Mons t er using properties of
vertex operators, and exh i b i t a natural action of it on V
.
117
5.4 The construction of M
We now sketch the discoveries and con st r u c ti o n s of some of the generation t h r ee of the
Happy Family sporadic simple groups and the pariahs, and make some comments on their
properties. The first generation is related to the Golay code C and the Mathieu five spo-
radic groups. The second one, is related to the Leech lattice and the Conway sporadic
groups. For a detailed discussion on thi s two other generations of the Happy Family see
the book of Griess [86].
The third generation of the Happy Family [80] consist of eight simple groups (see Tabl e 5 .1 )
which are involved in the largest sporadic simple group, the Monster M. The constru c ti o n
of M depends on constructio n and analysis of a certain commutative nonassociative algebra
B, of dimension 196,884, over the rational field Q. This is the Griess algebra we alread y
have st u d i ed in Section 5.3 . This algeb r a B plays the role of the Golay code in the First
generation, and the Leech lattice in the second.
Group Discoverer (d a t e) First con st r u c t io n (date)
He D. Held (1968) G. High m an, J. McKay (1968?)
Fi
22
B. Fischer (1968) B. Fischer (1969)
Fi
23
B. Fischer (1968) B. Fischer (1969)
Fi
24
B. Fischer (1968) B. Fischer (1969)
HN K. Harad a (1973) S. Nor to n (1974)
Th J. Thompson (1973) P. Smith (1974)
B B. Fischer (1973) J. Leon , C. Sims (1977)
M B. Fischer, R. Griess (1973) R. Gri ess (January, 1980)
Table 5.1: Third generation of the Happy Family.
In a sequence of steps, we now proceed to define and establi sh the basic properties of a
group C which will act naturally on V
, B and
ˆ
B, and which moreover will act compatibly
with the appropriate vertex operators. This group will be the centralizer of an involution
in th e Monster.
Starting with t h e central extension
ˆ
Λ (5.13 ) we first set
C
0
= {g Aut
ˆ
Λ : ¯g Co
0
}, (5.25)
where ¯g is the automorphi sm of the Leech lattice Λ induced by g and Co
0
is the isometry
group of Λ (5.9). We know that gκ = κ automatically. We have also mentioned that the
sequence
1 Hom(Λ , Z/2Z)
C
0
Co
0
1 (5.26)
118
is exact, where the pullback λ
:
ˆ
Λ
ˆ
Λ is given by
a →
λa
,
for λ Hom(Λ, Z/2Z). Moreover, we have natural identifications
Hom(Λ, Z/2Z) = Λ/ = Inn
ˆ
Λ,
so that the exact sequence (5.26) can be writt en
1 Inn(
ˆ
Λ) ֒ C
0
Co
0
1. (5.27)
Now, C
0
induces a group of automorphisms of the extraspecial g r ou p
ˆ
Λ/K since C
0
preserves
K, and we have a natural isomorphism ϕ : C
0
Aut(
ˆ
Λ/K). We claim that
Ker ϕ = θ
0
.
In fact, it is clear that θ
0
Ker ϕ. On the other hand, Ker ϕ Inn
ˆ
Λ = 1, since K κ = 1.
Hence, Ker ϕ is isomorph i c to it s image in Co
0
by (5.27). But,
Ker ϕ acts triviall y on λ/2Λ,
so that by the faithfulness of the action of Co
1
on Λ/2Λ, we have
Ker ϕ ±1, p r oving
the clai m .
Set C
1
= ϕ(C
0
) Aut(
ˆ
(Λ)/K). Then we have an exact sequence
1 Inn(
ˆ
Λ) C
1
Co
1
1. (5.28)
Here Co
1
acts in the natural way on Inn
ˆ
Λ = Λ/2Λ, and it follows from the properties of
the Conway group Co
0
that C
1
equals i t s commutator subgroup and has trivial center:
C
1
= [C
1
, C
1
];
Cent C
1
= 1.
We recall (see [67]) that the extraspecial group
ˆ
Λ/K, gives the exact sequence
1 F
×
N
Aut T
(π(
ˆ
Λ/K))
int
Aut(
ˆ
Λ/K) 1, (5.29)
where π denotes the faithful representation of
ˆ
Λ/K on T , N
Aut T
(π(
ˆ
Λ/K)) is the normalizer
of π(
ˆ
Λ/K) on Aut T , and
int(g)(x) = gxg
1
,
for g Aut T , x
ˆ
Λ/K = π(
ˆ
Λ/K) (T the Λ-module used in the twisted space V
T
Λ
). Set
C
= {g N
Aut T
(π(
ˆ
Λ/K)) : int(g) C
1
},
119
so that we have the commutative d i agr a m with exact rows
1
//
F
×
//
C
//
C
1
//
1
1
//
F
×
//
N
Aut T
(π(
ˆ
Λ/K))
//
Aut(
ˆ
Λ/K)
//
1.
Also set C
T
= [C
, C
]. We shall now show that C
T
contains 1 and in fact all of π(
ˆ
Λ/K).
Since
int(π(
ˆ
Λ/K)) = Λ/ = Inn
ˆ
Λ, (5.30)
we see that π(
ˆ
Λ/K) C
, ans so 1 = π(κK) [π(
ˆ
Λ/K), π(
ˆ
Λ/K))] C
T
. But since Co
1
acts irr e d u ci b l y on Λ/2Λ, we have Inn
ˆ
Λ = [Inn
ˆ
Λ, C
1
], and it follows that
π(
ˆ
Λ/K)) = [π(
ˆ
Λ/K)), C
] C
T
. (5.31)
We claim that the sequence
1 ±1 ֒ C
T
int
C
1
1 (5.32)
is exact. By the proper t ie s of C
1
, all we need to show it that C
T
F
×
= ±1. To see this,
we use the fact that the
ˆ
Λ/K-module T has a Q-for m ,
T
=
Ind
ˆ
Λ/K
ˆ
Φ/K
Q
ψ
0
Q
F,
where ψ
0
is any rational-valued character of
ˆ
Φ/K such that ψ
0
(κK) = 1, and this gives
us a
ˆ
Λ-invariant Q-subspace T
Q
of T such that the canonical map
T
Q
Q
F T
is an isomorphism . Let C
,Q
= C
Aut T
Q
. We have an exact sequence
1 Q
×
C
,Q
C
1
1,
and so C
= C
,Q
F
×
, and C
T
C
,Q
, implying that C
T
F
×
C
,Q
F
×
= Q
×
. But since
det C
T
= 1, we also have C
T
F
×
{µ F
×
: µ
2
12
= 1}, proving the claim.
Using ( 5. 28 ) we have a ma p
int =
int : C
T
Co
1
,
and by (5.30) and (5.31), th en π(
ˆ
Λ/K) Ker int. Consideration on the order of C
T
shows
that t h e sequence
1
ˆ
Λ/K
π
C
T
int
Co
1
1 (5.33)
120
is exact. Summari zing, we have an extension C
0
of Co
0
by Λ/ (5.27), an extension C
1
of Co
1
by Λ/ (5.28), and an extension C
T
of Co
1
by the extraspecial group
ˆ
Λ/K (5.33).
Now, form the pullback
ˆ
C = {(g, g
T
) C
0
× C
T
: ϕ(g) = int(g
T
)},
so that we have the commutative d i agr a m of surjections
ˆ
C
π
1
//
π
2
C
0
ϕ
C
T
int
//
C
1
.
Set
ˆ
θ
0
= (θ
0
, 1),
ˆ
θ = (1, 1)
ˆ
C. Then
Ker π
1
=
ˆ
θ,
Ker π
2
=
ˆ
θ
0
,
and
Cent
ˆ
C =
ˆ
θ
0
×
ˆ
θ = Ker(ϕ π
1
),
since Cent C
1
= 1 and θ
0
Cent C
0
from the definition.
We are finally ready to define the gro u p C: Set
C =
ˆ
C/
ˆ
θ
0
ˆ
θ. (5.34)
Then, t h e diagram above enlarges to the commutative diagram of surjection s
ˆ
C
π
1
//
π
2
π
0
A
A
A
A
A
A
A
A
C
0
ϕ
C
σ
B
B
B
B
B
B
B
B
C
T
int
//
C
1
.
Also, Ker π
0
=
ˆ
θ
0
ˆ
θ an d
Cent C = z|z
2
= 1 = Ker σ,
where
z = π
0
(
ˆ
θ
0
) = π
0
(
ˆ
θ) (5.35)
(no confusion should arise between this notation and our formal variable notation). We
have the exact sequence
1 z ֒ C
σ
C
1
1.
121
Proceeding as in (5.33), we have a map ¯σ =
σ : C Co
1
from (5. 28 ) . Moreover, there
is a canonical embedding ν :
ˆ
Λ/K
ˆ
C given by
gK → (int(g), π(g)),
since ϕ( i nt(g)) = int(π(g) ) . The result is an exact sequen ce
1
ˆ
Λ/K
π
0
ν
C
¯σ
Co
1
1,
and we have proved
Proposition 5.4.1. The group C is an extension of Co
1
by the extraspecial group
ˆ
Λ/K.
The nontrivial central element of
ˆ
Λ/K identifies with the nontrivial central element of C.
Now that the grou p C is constructed we shall set up its canonical acti on on the Moonsh i n e
module V
. Fi r st , we shall define an action of th e larger group
ˆ
C on t h e larger space
W
Λ
(see (5.18)). For g Co
0
and Z = Z or Z +
1
2
, let g also denote the unique algebra
automorphism
g : S(
˜
h
Z
) S(
˜
h
Z
),
(where S(
˜
h
Z
) is an abbreviation for S(
˜
h
)
Z
T , for the appropr i at e m odule T ), such that
g agrees with its na t u r al action on
˜
h
. For g C
0
, let g a lso denote the operator
g : F[Λ] F[Λ] given by i(a) → i(ga),
for a
ˆ
Λ; note that thi s is well defi n e d sin ce gκ = κ. For k = (g, g
T
)
ˆ
C, let k also denote
the operator
k = ¯g g ¯g g
T
on W
Λ
= S(
˜
h
)
Z
F[Λ] S(
˜
h
)
Z+
1
2
T . This clearly gives a faithful representation of
ˆ
C
on W
Λ
. In fact, this representation is even faithful on a small subspace of W
Λ
, for i n st ance
T p.
The action of
ˆ
C on W
Λ
extends the action of θ
0
similar to that d efined in (5.16) and (5.17)
for the case W
Λ
, in such a way t hat this operator corresponds to the elem ent
ˆ
θ
0
ˆ
θ = (θ
0
, 1)
of
ˆ
C. We have
ˆ
C · V
V
.
From the definitions of C and V
and the last paragraph, we see that C acts i n a natur a l
way on V
: for k = (g, g
T
)
ˆ
C, π
0
(k) acts as the operator
π
0
(k) = ¯g g ¯g g
T
. (5.36)
This action of C on V
is faithful, even on p = hT . The decomposition V
= V
θ
0
Λ
( V
T
Λ
)
θ
0
in Definition 5.3.1 is t h e eigenspace decomposition with respect to the central involuti on z
in C (5.35), and we introduce corresponding notation
V
θ
0
Λ
= V
z
= {v V
: z · v = v},
(V
T
Λ
)
θ
0
= V
z
= {v V
: z · v = v}.
122
Note that V
= V
z
V
z
and C · V
z
V
z
, C · V
z
V
z
.The actions of C on V
and of
ˆ
C on W
Λ
preserve t h e homogeneou s subspaces with respect to the gradings. The group C
acts on the algebra B = V
1
(Definition 5.3.2) and in fact preserves the summands f and p.
From the definiti on of the product × and the form ·, · on B (see Proposition 5.3.6), we
find [6 7 , p.328]:
Proposition 5.4.2. The group C ac ts faithfully as automorphisms of the Griess algebra
B and as isometries of ·, ·.
In fact, in checking that the form (5.23) on p is preser ved by C, we use the fact that the
form  ·, · on T is C
T
-invariant:
gτ
1
, gτ
2
= τ
1
, τ
2
, for all g C
T
, τ
i
T.
Recall from Theorem 5.3.4 that V
is a gra d ed module for the commutative affinization
ˆ
f
of f. We shall relate this structure to the action of C. Given a commutative nonassociati ve
algebra b with a symmet r ic form and given a group G of linear automorphisms of b, we let
G act as linear automorphisms of
ˆ
b by
g · e = e,
g · (u t
n
) = (g · u) t
n
,
for g G, u b and n Z. If G acts as algebra automorphisms and isometries of b, then
G acts as algebra automorphism s of
ˆ
b. Suppose now that V is a graded
ˆ
b-module and that
G acts as linear automorphisms of V , preserving each homogeneous subspace V
n
. Then,
we call V a graded (G,
ˆ
b)-module if
gxg
1
= g · x as operators on V,
for g G, x
ˆ
b. By Propositi on 5.4.2 we have
Proposition 5.4.3. The Moonshine module V
is a graded (C,
ˆ
f)-module, and C acts as
automorphisms of
ˆ
f, and in fact of
ˆ
B.
5.5 Improvements on the Construction
We have already seen in (5.36) and Proposition 5.4.2 that the group C acts as automor-
phisms on these main structur es: the Griess al ge b r a B and the Moonshine module V
. In
fact, th e idea for complete the const r u c ti o n of the Monster group M lies in enlarge C to a
group M (the Monster) of aut o m or p hisms and isometries of B (and hence automorphisms
of
ˆ
B), and make V
into a ( M,
ˆ
B)-module. Furt h er m or e , we should also define vertex op-
erators Y (u, z) on V
, for all v V
in order to extend an action of C on V
(consequence
of the action of
ˆ
C on W
Λ
) of type
kY (v, z)k
1
= Y (kv, z), for all k C, v V
θ
0
Λ
,
123
to all M and V
. We will not describe the complete construction of M (this requi r es in fact
a complete book). For details on the complete construction, the reader may consult the
paper of Griess [80], or the last three chapters of [67]. Although, we mention some notably
properties.
Under t h e action of C, B breaks into the following invariant subspaces:
Fω, {u S
2
(h) : u, ω = 0},
a
ˆ
Λ
4
Fx
+
a
, h T (5.37)
of dim en si on s 1, 299, 98280 and 98304, r espectively. Of cou r se, C in fact fixes ω:
C · ω = ω.
It can be shown that each of the invariants subspaces is absolutely irreducible under C; for
instance, h T is irreducible since T is irreducible under t h e extraspeci al group
ˆ
Λ/K and
h is irreducible under Co
0
. Before the Monster was proved to exist, it was pos t u l at ed to be
a fini te simple group containing the group C as the centralizer of t h e involutio n z C and
it was believed to have a 196,883-dimensional irreducible modul e con si st i n g of the direct
sum of the las three C- m odules listed in (5.37), or rather, abstract C-modules isomorphic
to them. B y 1976, Norton proved the exist ence of an i nvariant commutative nonassociative
algebra and nonsingular associative symmetric bilinear form on this module if it and the
Monster existed (c. f. [ 80 ] ) , though his met h od gave no descrip t i on (he also made assump-
tions on the conjugacy classes). By constraining th e possibilities for such an algebr a and
form on the direct sum of the C-modules, Griess was able to determine an algebra and a
form admitting an automorphism outside th e group C. The group generated by C and this
automorphism had the requ i re d proper t i es.
The origi n al definiti on of the algebra structure in [80] was complicated, due mainly to sign
problems. The existence o f the mentioned i r r ed ucible representation of the Monster of
degree 196,883 was predicted in 1974 by Griess [78]. With the Norton’s improvement of
the existence for such a nontrivial commutative non associative algebra struct ure o n this
module, an d the st u d y of the automorphisms outside of C, by surveying the actions of
many subgroups of C on the space B, the result was the system of structure con st ants in
[80, Table 6.1], and the formula for an extra automorphism in [80, Table 10.2]. Prove that
the linear automorphism so defined preserves th e algebra structure was the hardest part o f
the con st r u ct i on in [80].
Many improvements on [80] were made by Tits [174, 175, 176], who showed that some
definitions of [80] based on guesswork may be based on a more thorough analysis. A new
style construction was made by Conway in [34], [39], using a Moufang loop to finesse the
sign prob l em s so prominent in th e original version. This loop (a nonassociati ve group)
has order 2
13
and is a kind of 2-cover of the binary Golay code; its creation was an idea
124
of Richard Parker (see [81]), which defines and gives the foundat i on of th e class of loops
called code loops. This theory is usually developed in the theory of p-locals in sporadic
groups and Lie groups [81, 83, 84, 85].
The algebra B is not a classic nonassociative algebra. An algebra of dimension n satisfies
a nontrivial polynomial identity of degree at most n + 1; B satisfies no nontrivial identity
in co m muting varia b l es of degree less than 6 [82]. In [80], the subgroup of Aut B generated
by C and the particular extra aut om or phism was identified as a simp le group of the right
order, thus proving t h e existence of a simple group of the right order and local properties.
The full automorph is m group of a fini t e dimensional algebra is an algebraic group. In [175],
Tits showed th a t Aut B was exact l y the Monster M. In [34], Conway gave a short argument
with idempotents in B that Aut B i s finit e and in [175], Tits i d entified the centralizer of
an involution in Aut B as C (not a larger group). The proof that the group order is right
involves quoting harder theorems (see chapter 13 of [80]). In 1988, a uniqueness proof for
the Mon st er was given by Griess, Meierfrankenfeld and Segev [87].
We close our discussion with this uniqueness result. First, we define a grou p of Monster-type
to be a finite group G containing a pai r of involutions z, t s u ch that
C(z)
=
2
1+24
+
Co
1
;
C(t) is a double cover of Fischer’s {3, 4}-transposition group (discovered by Fischer
in 1973, and later called the Baby Monster B).
Is follows that such a grou p is simple. See [87] for a fuller discussion of the hypotheses.
Theorem 5.5.1 (Uniqueness of the Monster). A group of Monster-type i s unique up to
isomorphism.
125
Chapter 6
Intermezzo: the j functio n
The j function plays an im portant role in modern number theor y. It classifi es the family
of ellipti c curves over the complex field C. In fact, the j functions serves as the moduli
space of this family of cur ves. In this and next chapters, our purpose is to detail in a better
form some aspects of the ori gi n a l Conway-Norton conjecture. First, we will give here some
background theory of modular forms and Hecke operators, and subsequently we ou t li n e
some featu r es of replicable functions introduced in [38].
6.1 The mod ul ar gro up
Let H denote the upper half-plane of C we have introduced in Chapter 1. That is H =
{z C : Im(z) > 0}. Let SL
2
(R) be the group of matrices (
a b
c d
) with real entries, such
that ad bc = 1. We made SL
2
(R) act on C by: if G = (
a b
c d
), th en
Gz =
a b
c d
· z =
az + b
cz + d
.
One can easily checks the formula
Im(Gz) =
Im(z)
|cz + d|
2
,
from which can be showed that H is stable under the action of SL
2
(R). We have also
seen in Chapter 1 that the element I =
1 0
0 1
of SL
2
(R) acts trivially on H. Thus
we can consider the grou p PSL
2
(R) = SL
2
(R)/I} which operates —and in fact, acts
faithfully—, and one can even show that it is the group of all analytic automo r p h i sm s of
H.
Let SL
2
(Z) be the subgroup of SL
2
(R) consisting of the matrices with integer entries. It
is a discrete subgroup of SL
2
(R), t hus it acts discontinually on H.
Definition 6.1.1. The group Γ = P SL
2
(Z) = SL
2
(Z)/I} is c al le d the modular group.
It is the image of SL
2
(Z) on P SL
2
(R). We denote by H/Γ the set of action orbits of Γ on
H.
For simplicity, if G is an element of SL
2
(Z), we will use the same symbol to denote it s
image in the modular gr oup Γ. Let S = (
0 1
1 0
) and T = (
1 1
0 1
) denote the el em ents of Γ
127
introduced in Chapter 1. We have seen that
Sz =
1
z
, T z = z + 1,
S
2
= 1, (ST )
3
= 1.
On the other hand, let D the subset of H formed of all points z such that |z| 1 and
|Im(z)| 1 /2. The Figure 6.1 below represents the transforms of D by some of the
elements of the group Γ.
1
1/2
0
1/2
1
τ
τ + 1
σ
1
D
T
1
T
T
1
S
S T S
ST S ST
ST
1
T ST
H
Figure 6. 1: Fun d am ental domain D and some of its images by S and T .
We will show that D is a fundamental domain for the action of Γ on the half-plane H.
More pr eci sel y :
Theorem 6.1.2. 1. For every z H there exists G Γ such that Gz D.
2. Suppose that t wo distinct points z, z
of D are congruent modulo Γ. Then, Re(z) = ±
1
2
and z
= z ± 1, or |z| = 1 and z
=
1
z
.
3. Let z D and let I(z) = {G Γ : Gz = z} be the stabilizer of z in Γ. One has
I(z) = {I}, except in the following three cases:
z = i, in which case I(z) is the group of order 2 generated by S,
z = ω = e
2πi/3
, in which case I(z) is the group of order 3 generated by ST ,
z = ω
2
= e
πi/3
, in which case I(z) is the group of order 3 generated by T S.
128
Theorem 6.1.3. The modular group Γ is generated by S and T .
Proof. To prove Theorems 6.1.2 and 6.1.3, consider the subgroup Γ
of Γ, generated by S
and T , and let z H. If G = (
a b
c d
) is an element of Γ
, then Im(Gz) = Im(z)/|cz + d|
2
.
Since c and d are integers, the number of pairs (c, d) such that |cz + d| is less that a
given number is fin i t e. This shows tha t there exist G Γ
such that Im(Gz) is maximum.
Choose n ow an integer n such that T
n
Gz has real part between
1
2
and
1
2
. The element
z
= T
n
Gz belongs to D; ind eed , it suffices to show that |z
| 1, but if |z
| < 1 then the
element
1
z
would h ave an imaginary part strictly greater than Im(z
), which is im possi-
ble. Thu s, the element T
n
G of Γ has th e desired property. This proves (1) in Theorem 6.1.2.
We now prove assertions (2) and (3). Let z D and let G = (
a b
c d
) Γ such that Gz D.
Without lose of generality (replacing ( z, G) with (Gz, G
1
) if necessary) we may suppose
that Im(Gz) Im(z), i. e., that |cz + d| 1. This is clearly impossible if |c| 2, leaving
the cases c = 0, 1, 1. If c = 0 , then d = ±1 and G is the tran sl at i on by ±b. Since Re(z)
and Re(Gz) are both between
1
2
and
1
2
, this implies that either b = 0 and G = I, or
b = ±1 in which case o n e of the numbers Re(z) or Re( Gz) must be equal to
1
2
and the
other to
1
2
. If c = 1, the fact that |z + d| 1 implies d = 0, exc ep t if z = ω (respectively if
z = ω
2
), in wh i ch case we can have d = 0, 1 (respectively d = 0, 1). The case d = 0 gives
|z| 1, hence |z| = 1; on the other hand, ad bc = 1 implies b = 1, hence Gz = a
1
z
and the first part of discussion proves that a = 0, except in the cases Re( z) = ±
1
2
, i. e., if
z = ω, ω
2
, in which cases we have a = 0, 1 or a = 0, 1. The case z = ω, d = 1 gives
a b = 1 and = a
1
ω+1
= a + ω, hence a = 0, 1. We argue similar when z = ω
2
,
d = 1. Finally, the case c = 1 leads to the case c = 1 by changing the signs of a, b, c, d
(which does not change G seen as an element of Γ). This complet es th e verification of (2)
and (3 ) .
To complete the proof of Theorem 6.1.3, it remains to prove that Γ
= Γ. Choose an
element z
0
int D (for exam p l e z
0
= 2i), and let z = Gz
0
. We have seen above that
there exists an element G
Γ
such that G
z D. The points z
0
, G
z = G
Gz
0
of D are
congruent modulo Γ (they li e on the same orbit in H/Γ), and one of them is interior to D.
By (2) and ( 3) of Theorem 6.1.2, it follows that these points coincide, so that G
G = I.
Thus, G = (G
)
1
Γ
, which completes the proof.
Corollary 6.1.4. The canonical map D H/Γ is surjective, and its restriction to int D
is injective. In particular, D is a fundamental domain for the action of Γ on H.
Thus, D is a set intersecting each orbit of H/Γ just at one point. As a remark, one can show
that S, T | S
2
, (ST )
3
is a presentation of Γ, or , equivalently, that Γ is the free pr oduct
of the cyclic group of order 2 genera t ed by S and th e cyclic group of order 3 generated by
ST .
129
6.2 Mod ul ar fo rm s
Definition 6.2.1. Let k be an integer. We say that a function f : H C is weakly
modular of weight 2k if f is meromorphic on the upper half-plan e H and if it verifies the
relation
f
az + b
cz + d
= (cz + d)
2k
f(z), for all
a b
c d
SL
2
(Z). (6.1)
Let G be the image in Γ of (
a b
c d
). We have
d(Gz)
dz
= (cz + d)
2
. The relation (6.1 ) can then
be wr it t en as
f(Gz)
f(z)
=
d(Gz)
dz
k
or
f(Gz)d(Gz)
k
= f (z)dz
k
, (6.2)
where dz
k
= dz . . . dz (k times). It means that the tensor product of differential 1-
forms f(z)dz
k
is invariant under Γ. Since Γ is g en er at ed by the elements S and T (Theorem
6.1.3), it suffic es to check the invariance by S and by T . This gives:
Proposition 6.2.2. Let f be a meromorphic function on H. Then, f is weakly modular
of weight 2k if, and only if, for all z H it satisfies the two relations:
f(z + 1) = f(z), (6.3)
f(1/z) = z
2k
f(z). (6.4)
Suppose the relation (6.3) is verified. We have already noticed that f can be expressed as
a function of q = e
2πiz
, function wh i ch we will denot e by
˘
f; i t is meromorphic in the di sk
|q| < 1 wit h the o r ig in removed. We can think the valu e q = 0 making z = i
H, and,
if
˘
f extends to a meromorphic (r espectively holomorphic) function at the origin, we say,
by abuse of language, that f is meromorphic (respectively holomorphic) at infinity. This
means t h a t
˘
f admits a Laurent expansion in a neighb o r h ood of the origin
˘
f(q) =
−∞
a
n
q
n
,
where th e a
n
’s are zero for n small en ou g h (respect i vely for n < 0).
Definition 6.2. 3. If a weakly modular function f is meromorphic at infinity, we say that
f is a modular function. When f is holomorphic at infinity, we set f() =
˘
f(0). This is
the val u e of f at infinity.
Definition 6.2.4. A modular function which is holomorphic everywhere —including at
infinity— is called a modular form; if such a function is zero at infinity, it is called a cusp
form (or Spitzenform, or forme parabolique).
130
Thus, a modular form of weight 2k is given by a series f(z) =
n0
a
n
q
n
=
n0
a
n
e
2πinz
,
which converges for |q| < 1 (i. e., for Im(z) > 0) and that verifies the identity f(1/z) =
z
2k
f(z). It is a cu sp form if a
0
= 0.
Example 6. 2.5 . If f and g are modular forms of weight 2k, then any C-linear combination
αf + βg is a modular form of weight 2k. Thus, the modular forms of weight 2k forms a
C-vector sp ace .
If f and f
are modular forms of weight 2k and 2k
, then the product fg is a modular form
of weight 2k + 2k
.
Example 6.2.6. We will see later that the function
q
n1
(1 q
n
)
24
= q 24q
2
+ 252q
3
1472q
4
+ . . . , q = e
2πiz
,
is a cusp form of weight 12.
6.3 Lattice functions and modular functions
Recall th at a lattice in a r ea l vector space V of finite dimension is a subgroup Λ of V
satisfying one of the following equivalent conditions:
Λ is discrete and V/Λ is compact;
Λ is discrete and generates the R-vector space V ;
There exists an R-basis e
1
, . . . , e
n
of V which is a Z-basis for Λ, tha t is Λ is the n-rank
free abelian group Λ = Ze
1
. . . Ze
n
.
Let L be th e set of lattices of C considered as a r e al vector space. Let M be the set of
pairs (ω
1
, ω
2
) of elements of C
×
such that Im
ω
1
ω
2
> 0. To such a pair we associate the
lattice
Λ(ω
1
, ω
2
) = Zω
1
Zω
2
with basis {ω
1
, ω
2
}. We thus obtain a map M L which is clearly surjective. Let
G = (
a b
c d
) SL
2
(Z) and let (ω
1
, ω
2
) M. We put
ω
1
ω
2
=
a b
c d
ω
1
ω
2
=
1
+
2
1
+
2
.
It is clear that {ω
1
, ω
2
} is a basis for Λ(ω
1
, ω
2
). Mo r eover, if we set z =
ω
1
ω
2
and z
=
ω
1
ω
2
we
have
z
=
az + b
cz + d
= Gz.
This shows that Im(z
) > 0, hence that (ω
1
, ω
2
) belongs to M. In fact, we have the following
result
131
Proposition 6.3.1. For two elements of M to define the same lattice it is necessary and
sufficient that they are congruent modulo SL
2
(Z).
Proof. We just saw tha t the condition is sufficient. Conversel y, if (ω
1
, ω
2
) and (ω
1
, ω
2
) are
two elements of M which define the same lattice, there exists an integer matrix G = (
a b
c d
)
of determin a nt ±1 which tran sfo r m s the first basis into the second. If det(G) < 0, the sign
of Im(z
) would be the opposite of Im(z
) as one sees by an immediate computation. The
two signs being the same, we have necessarily det(G) = 1.
Hence we can identify th e set L of lattices of C with the quotient of M/SL
2
(Z). Make
now C
×
act on L (resp ect i vely on M) by:
Λ → λΛ (respectively (ω
1
, ω
2
) → (λω
1
, λω
2
)), for λ C
×
.
The quot ie nt M/C
×
is identified with H by ( ω
1
, ω
2
) → z =
ω
1
ω
2
and th i s identification
transforms the action of SL
2
(Z) on M into that of the modular group Γ = P SL
2
(Z) on H.
Thus:
Proposition 6.3.2. The map (ω
1
, ω
2
) →
ω
1
ω
2
gives by passing t o the quotient, a bijection of
L /C
×
onto H/Γ. (Thus, an element of H/Γ can be identified with a lattice of C defined
up to a homothety.)
In numb er theory it is frequent to associate to a lattice Λ of C the elliptic curve E
Λ
= C/Λ.
It is easy to see that two lattices Λ and Λ
define isomorphic elliptic curves if and only
if they are homothetic. This gives a third descrip t i on of H/Γ = L /C
×
: it is the set of
isomorphism classes of ellipti c curves.
Let us now pass to modular functions. Let F be a function defined on L , with complex
values, and let k Z. We say that F is a modular lattice function of weight 2k if
F (λΛ) = λ
2k
F (Λ), for all lattices Λ and all λ C
×
. (6.5)
Let F be such a function. If (ω
1
, ω
2
) M, we denote by F (ω
1
, ω
2
) the value of F on the
lattice Λ(ω
1
, ω
2
). The formula (6.5) translates to:
F (λω
1
, λω
2
) = λ
2k
F (ω
1
, ω
2
). ( 6 .6 )
Moreover, F (ω
1
, ω
2
) is invariant by the action of SL
2
(Z) on M. Formula (6.6) shows that
the product ω
2k
2
F (ω
1
, ω
2
) depends only on z =
ω
1
ω
2
. There exists then a functio n f on H
such tha t
F (ω
1
, ω
2
) = ω
2k
2
f(ω
1
2
). (6.7)
Writing that F is invariant by SL
2
(Z), we see that f satisfies the identity (6.1). Conversely,
if f verifies (6.1), formula (6.7) associates to it a functio n F on L which is of weight 2k.
We can thus identify modul ar functions of weight 2k with som e lattice fun ct io n s of weight
2k.
132
Example 6.3.3 ( Ex am ple of modular functions: Eisenstein series).
Lemma 6.3.4. Let Λ be a lattice in C. The series
γΛ
1
|γ|
σ
is convergent for σ > 2. (The symbol Σ
signifies that summation runs over all nonzero
elements of Λ).
Proof. We can proceed as with the seri es
1
n
σ
, i. e., majorize the series unde r consideration
by a multiple of the double integral
dxdy
(x
2
+y
2
)
σ/2
extended over the plane without a disk
with center at 0. The double integral is co m p u t ed using the classical technique with polar
coordinates. Another equivalent method consists in remarking that the number of elements
of Λ such that |γ| is between two consecutive integers n and n + 1 is O(n); the convergence
of the series is then reduced to that of the series
1
n
σ1
.
Now let k be an integer > 1. If Λ is a lattice of C, put
G
k
(Λ) =
γΛ
1
γ
2k
. (6.8)
This series is absolut e ly convergent, thanks to Lemma 6.3.4. Observe that G
k
is a modular
lattice function of weight 2k. It is call ed the Eisenstein series of index k. As i n the
preceding section, we can view G
k
as a function on M, given by
G
k
(ω
1
, ω
2
) =
m,n
1
(
1
+
2
)
2k
. (6.9)
Here again t h e symbol Σ
means that the sum m a ti o n runs over all pairs of integers (m, n)
distinct from (0, 0). The funct io n on H corresponding to G
k
(by th e procedure given in
previous section) is denoted also by G
k
. By formulas (6.7) and (6.9), we have
G
k
(z) =
m,n
1
(mz + n)
2k
. (6.10)
Proposition 6.3.5. Let k an integer > 1. The Eisenste in series G
k
(z) is a modular form
of weight 2k. We have G
k
() = 2ζ(2k), where ζ denotes the Riemann zeta function.
Proof. The ab ove arguments show that G
k
(z) is weakly modular of weight 2k. We have
to show that G
k
is holomorphic everywhere. First suppose that z is contained in the
fundamental domain D. Then
|mz + n|
2
= (mz + n)(
mz + n) = (mz + n)(m¯z + n) = m
2
z¯z + 2mn Re(z) + n
2
m
2
mn + n
2
= | n|
2
.
133
By Lemma 6.3.4, the series
1
|n|
2k
is conver gent. This shows that the series G
k
(z)
converges in D, thus also (a p p l yi ng the result to G
k
(G
1
z) with G Γ) in each of the
transforms GD of D by Γ. Since these sets cover H, we see that G
k
is holomorphic on H.
It remains to verify that G
k
is holomorphic at infin i ty (and to find the value at this po i nt).
This reduces to prove that G
k
has a limit for Im(z) . But one may su p pose that z
remains i n the fundamental domain D; in view of the unifor m convergence in D, we can
make the passage to t he limit term by term. The terms
1
(mz+n)
2k
relative to m = 0 give 0;
the oth er s give
1
n
2k
. Thus
lim
z→∞
G
k
(z) =
n
1
n
2k
= 2
n1
1
n
2k
= 2ζ(2k).
The Eisenstein series of lowest weights are G
2
and G
3
, which are of weight 4 and 6. It is
conveniently (because of the theory of elliptic curves) to replace these by some multiples:
g
2
= 60G
2
, g
3
= 140G
3
.
We have g
2
() = 120ζ(4) and g
3
() = 280ζ(6). Using the known val u es of ζ(4) and ζ(6)
(see for example Table 6.2), one finds that
g
2
() =
4
3
π
4
, g
3
() =
8
27
π
6
.
If we put
= g
3
2
27g
2
3
, (6.11)
then we have ∆() = 0 ; that is to say, is a cusp form of weight 12.
In fact, all these stuff is related to t h e theory of elliptic curves. Let Λ be a la t ti ce of C and
let
Λ
(u) =
1
u
2
+
γΛ
1
(u γ)
2
+
1
γ
2
, (6.12)
be the correspondin g Weierstrass function. Then, G
k
(Λ) occur into the Laurent expansion
of
Λ
:
Λ
(u) =
1
u
2
+
k2
(2k 1)G
k
(Λ)u
2k2
.
If we put x =
Λ
(u) an d y =
Λ
(u), we have
y
2
= 4x
3
g
2
x g
3
, (6.13)
with g
2
= 60G
2
, g
3
= 140G
3
as above. Up to a numerical factor, = g
3
2
27g
2
3
is equal
to the discriminant of the polynomi al 4x
3
g
2
x g
3
. Usually one proves that the cubic
defined by equat io n (6.13) in the projective plane is isomorphic to the ellip t i c curve C/Λ.
In part i cu l a r, it is a nonsingular curve, and th i s shows that = 0.
134
6.4 The space of modular forms
Let f be a m e ro m or p hic fu n c ti o n on H, not identically zero, and let p be a po i nt of H.
The integer n such that
f
(zp)
n
is holomorphic and non-zero at p is called the order of f at
p and is denoted by ν
p
(f). When f is a modu la r function of weight 2k, th e identity (6.1)
shows that ν
p
(f) = ν
Gp
(f) if G is in the modular group Γ. In other terms, ν
p
(f) depends
only on the image of p in H/Γ. Moreover one can define ν
(f) as the order for q = 0 of
the function
˘
f(q) associ at ed to f. Finally, we will denote by e
p
the order of the stabilizer
of the point p; we have that e
p
= 2 (respectively e
p
= 3) if p is congruent modulo Γ to i
(respectively to ω =
1+
3i
2
), an d e
p
= 1 otherwise. Also, we have
Theorem 6.4.1. Let f be a modular function of weight 2k, not identically zero. Then
ν
(f) +
pH/Γ
1
e
p
ν
p
(f) =
k
6
. (6.14)
(We can also write this formula in the form
ν
(f) +
1
2
ν
p
(i) +
1
3
ν
p
(ω) +
pH/Γ
1
e
p
ν
p
(f) =
k
6
, (6.15)
where the symbol Σ
means a summation over the points of H/Γ distinct from the classes
of i and ω.)
Proof. Observe first that the sum written in Theorem 6.4.1 makes sense, i. e., that f has
only a finite number of zeros and poles modulo Γ. Indeed, since
˘
f is meromorphic, there
exists r > 0 such that
˘
f has no zero nor pole for 0 < |q| < r; and this means that f has no
zero nor pole for Im(z) >
1
2π
log
1
r
. Now, the part D
r
of the fun d a m ental domain D defined
by the inequality Im(z) <
1
2π
log
1
r
is comp a ct ; since f is meromorp hic in H, it has only a
finite number of zeros and of poles in D
r
, hen ce our assertion.
To prove Theorem 6.4.1, we will i ntegrate
1
2πi
df
f
on the bound ar y of D. More precisely:
1. Suppose that f has no zero nor pole on the boundary of D except possibly i, ω, and
ω
2
. There exists a contour γ as represented in Figure 6.2, whose interior contains
a representative of each z er o or pole of f not congruent to i or ω. By the residue
theorem we have
1
2πi
γ
df
f
=
pH/Γ
ν
p
(f).
On the other hand
the change of variables q = e
2πiz
transforms the arc EA into a circle η centered
at q = 0, with negative orientation, and not enclosing any zero or pole of
˘
f
except possibly 0. Hence
1
2πi
A
E
df
f
=
1
2πi
η
df
f
= ν
(f).
135
Figure 6.2: Integration on the boundary of region D.
The integral of
1
2πi
df
f
on the circle which contains the arc BB
, oriented nega-
tively, has the value ν
ω
(f). Wh e n the radiu s of this circle tends to 0, the angle
B
ω
B
tends t o
2π
6
. Hence
1
2πi
B
B
df
f
1
6
ν
ω
(f).
Similarly, when the radii of the arcs CC
and DD
tend t o 0:
1
2πi
C
C
df
f
1
2
ν
i
(f), and
1
2πi
D
D
df
f
1
6
ν
ω
2
(f).
T transforms the arc AB into the arc ED
; since f(T z) = f(z), we get:
1
2πi
B
A
df
f
+
1
2πi
E
D
df
f
= 0.
S transforms the arc B
C onto the arc DC
; since f(Sz) = z
2k
f(z), we get:
df(Sz)
f(Sz)
= 2k
dz
z
+
df(z)
f(z)
,
136
hence
1
2πi
C
B
df
f
+
1
2πi
D
C
df
f
=
1
2πi
C
B
df(z)
f(z)
df(Sz)
f(Sz)
=
1
2πi
C
B
2k
dz
z
2k
1
12
=
k
6
,
when the radii of the arcs BB
, CC
, DD
tend to 0. Writing now that the
two expressi ons we get for
1
2πi
γ
df
f
are equal, and passing to the limit, we find
formula (6.15).
2. Suppose that f has a zero or a pole λ on the half line {z H : Re(z) =
1
2
, Im(z) >
3
2
}.
We repeat the above proof with a contour modified in a neighborhood of λ and of T λ
as in Figure 6.3 (the arc circling a r ou nd T λ is the transfor m by T of the arc cir cl ing
around λ)
Figure 6.3: Integration on the boundary of region D.
We proceed in an analogous way if f has several zeros or poles o n the boundary of
D, concluding the proof.
As a remark, we on l y mention that in fact this somewhat laborious proof could have been
avoided by introducing a com plex analytic structure on the compactifi c at i on
H/Γ of H/Γ,
as we mention in Chapter 1 (see for example [20]).
137
If k is an integer, we denote by M
k
(respectively M
0
k
) th e C-vector space of modular forms
of weight 2k (respectively of cusp forms of weight 2k). By Definition 6.2.4, M
0
k
is the kernel
of the linear form f → f() on M
k
. Thus we h ave dim M
k
/M
0
k
1. Moreover, for k 2,
the Eisenstein series G
k
is an elem ent of M
k
such that G
k
() = 0, by Proposition 6.3. 5 .
Hence we have
M
k
= M
0
k
C G
k
, for k 2.
Finally, recall th at one denotes by the element g
3
2
27g
2
3
of M
0
6
, where g
2
= 60G
2
and
g
3
= 140G
3
.
Theorem 6.4.2. 1. We have M
k
= 0 for k < 0 and k = 1.
2. For k = 0, 2, 3, 4, 5, M
k
is a vector space of dimension 1 with basis 1, G
2
, G
3
, G
4
, G
5
;
we have also M
0
k
= 0.
3. Multiplication by defines an isomorphism of M
k6
onto M
0
k
.
Proof. Let f be a n on ze r o element of M
k
. All the term s on the left side of the formula
(6.15)
ν
(f) +
1
2
ν
p
(i) +
1
3
ν
p
(ω) +
pH/Γ
1
e
p
ν
p
(f) =
k
6
are 0. Thus, we have k 0 and also k = 1, since
1
6
cannot be wri t te n in the form
n +
n
2
+
n
′′
3
, with n, n
, n
′′
0. This proves (1).
Now apply equation (6.15) to f = G
2
. We can write
2
6
in th e form n +
n
2
+
n
′′
3
, n, n
, n
′′
0,
only n = 0, n
= 0, n
′′
= 1. This shows that ν
p
(G
2
) = 0, for p = ω (modulo Γ). The same
argument applies to G
3
and proves that ν
i
(G
3
) = 1 and that all the others ν
p
(G
3
) are zero.
This already shows that is not zero at i, h en ce is not identically zero. Since the weight
of is 12 and ν
(∆) 1, formula (6.1 5 ) im p l i es th at ν
p
(∆) = 0 for all p = and th at
ν
(∆) = 1. In other words, does not vanish on H and has a simple zero at infinity. If f
is an element of M
0
k
and if we set g = f/∆, it is clear that g is of weight 2k 12. M or eover,
the formula
ν
p
(g) = ν
p
(f) ν
p
(∆) =
ν
p
(f), if p = ;
ν
p
(f) 1, if p = ;
shows that ν
p
(g) p for all p, thus that g belongs to M
k6
, which proves (3). Finally,
if k < 5, we have k 6 < 0 and M
0
k
= 0 by (1) and (3); this shows th a t dim M
k
1.
Since 1, G
2
, G
3
, G
4
and G
5
are nonzero elements of M
0
, M
2
, M
3
, M
4
, M
5
, respectively, we
have dim M
k
= 1 for k = 0, 2, 3, 4, 5, which proves (2).
Corollary 6.4.3. We have for any k 0
dim M
k
=
k/6, if k 1 (mod 6);
k/6 + 1, if k ≡ 1 (mod 6).
(6.16)
138
Proof. Formula (6.16) is tru e fo r 0 k < 6 by (1) and (2). Moreover, the two expressions
increase by one unit when we r ep l a ce k by k + 6 by (3). The formula is thus true for all
k 0.
Corollary 6.4.4. The space M
k
has for basis the family of monomials G
α
2
G
β
3
, with α, β 0
integers such that 2α + 3β = k.
Proof. We show first th a t these monomials generate M
k
. This is clear for k 3 by (1) and
(2). For k 4, we argue by induction on k. Choose a pair (γ, δ) of integers 0 such that
2γ + 3δ = k (this is possible for all k 2). The modular form g = G
γ
2
G
δ
3
is not zero at
infinity. If f M
k
, there exists λ C such that f λg is a cusp form, hence equal to h,
with h M
k6
. One then applies the inductive hypothesis to h. It remains to see that
the above monomials are lin e ar l y independent; if they were not, the function G
3
2
/G
2
3
would
verify a nontrivial algeb r ai c equati on with coefficients in C, thus would be constant, which
is absu r d because G
2
is zero at ω but not G
3
.
As a remark, Let M =
M
k
be the graded al gebra which is the direct sum of th e M
k
and
let ε : C[x, y] M be the homomorphism which maps x → G
2
and y → G
3
. Corollary
6.4.4 is equivalent to saying that ε is an isomorphism. Hence, one can identify M with the
polynomial algebra C[G
2
, G
3
].
6.5 The mod ul ar inva ri a nt j
We define the j function as
j(z) = 1728
g
3
2
(z)
∆(z)
=
1728g
3
2
(z)
g
3
2
(z) 27g
2
3
(z)
. (6.17)
Proposition 6.5.1. The function j is a modular function of weight 0. Moreover, it is
holomorphic on H and has a simple pole at infinity. It defines by passage to quotient a
bijection of H/Γ onto C.
Proof. First assertion comes from the fact that g
3
2
and ar e bo t h of weight 12. The
holomorphicity of j follows from the fact that = 0 on H and has a simple pole at infinity,
while g
2
is nonzero at infinity. To pr ove the last assertion, on e has to show that if λ C,
the m odular form f
λ
= 172 8g
3
2
λ has a unique zero modulo Γ. To see this, one applies
formula (6.15) to f
λ
and k = 6 . The only decompositions of
k
6
= 1 in the form n +
n
2
+
n
′′
3
,
with n, n
, n
′′
0 correspond s to
(n, n
, n
′′
) = (1, 0, 0) or (0, 2, 0) or (0, 0, 3) .
This sh ows that f is zer o at one and only one point of H/Γ.
Theorem 6.5.2. Let f be a meromorphic function on H. The following are equivalent:
139
1. f is a modular function of weight 0;
2. f is a quotient of two modular forms of the same weight;
3. f is a rational function of j.
Proof. The implications (3) (2) (1) are immediat e. We show that (1) (3). Let f be
a modular funct i on. Being free to multiply by a su i t ab l e polynomial in j, we ca n suppose
that f is holomorph ic on H. Since is zero at infinity, there exists an integer n 0
such that g =
n
f is h ol om o r p h i c at infin i ty. The function g is then a modular form of
weight 12n. By Corollary 6.4.4, we can writ e g as a linear combination of the G
α
2
G
β
3
, with
2α + 3β = 6n. By linear i ty, we are red uced to the case g = G
α
2
G
β
3
, i. e., f = G
α
2
G
β
3
/
n
.
But the rela t io n 2α + 3β = 6n shows that p =
α
3
and q =
β
2
are integers and one has
f = G
3p
2
G
2q
3
/
p+q
. Thus, we only need to see that G
3
2
/ and G
2
3
/ are rational functions
of j, which is obvious from the definitions of g
2
, g
3
and ∆.
As we stated in Chapter 1, it is possible to define in a natural way a structure of com p l ex
analytic manifold on th e compactification
H/Γ of H/Γ. Proposition 6.5.1 thus means that
j defines an isomorph is m of
H/Γ onto the Riemann sphere C = C {∞}. As for Theorem
6.5.2, it amou nts to the well known fact that the only meromorphic functions on
C are the
rational functions.
The coefficient 1728 = 2
6
· 3
3
has been introduced in or d e r that j has a resi d u e equal to 1
at infinity. More precisely, the series expansion of n ex t section shows t h a t the q-expansion
of j (recall equation (1.3)) is
j(z) = q
1
+ 744 +
n1
c
n
q
n
, (6.18)
where q = e
2πiz
, z H. One has c
1
= 2
2
·3
3
·1823 = 196884, c
2
= 2
11
·5·2099 = 21493760.
All c
n
are integers (this follows from the definition of j and the q-expansion formulas of
g
2
, g
3
), an d they enjoy remarkable divisibility properties, see for example [2 ] or [38].
6.6 Expansions at infinity
Eisenstein series and the j function are closely related to the Riemann zeta function. Only
to give a basic idea, we will present some of this results, omitting most of the proofs.
Interested reader can consult [117], [167].
Consider the Bernoulli numbers B
k
, defin ed by the series
x
e
x
1
= 1
x
2
+
k1
(1)
k+1
B
k
x
2k
(2k)!
. (6.19)
140
B
1
=
1
6
B
2
=
1
30
B
3
=
1
42
B
4
=
1
30
B
5
=
5
66
B
6
=
691
2730
B
7
=
7
6
B
8
=
3617
510
B
9
=
43867
798
B
10
=
283·617
330
Table 6.1: First Bernoulli numbers B
k
.
There are a lot of properties rel at ed to Bernoulli numbers, (see [101 ] ) . For example, the
B
k
give the values of the Riemann zeta function for th e positive even integers ( an d also for
the neg at i ve odd integers):
Proposition 6.6.1. If k 1 is an integer, then
ζ(2k) =
2
2k1
(2k)!
π
2k
B
k
. (6.20)
Proof. The identity
z cot z = 1
k1
B
k
2
2k
z
2k
(2k)!
(6.21)
follows from the definition of the B
k
by putting x = 2iz. Moreover, ta ki n g the logarithmic
derivative o f
sin z = z
n1
1
z
2
n
2
π
2
, (6.22)
we get:
z cot z = 1 + 2
n1
z
2
z
2
n
2
π
2
= 1 2
n1
k1
z
2k
n
2k
π
2k
. (6.23)
Comparing (6.21) and (6.23 ) , we get (6.20).
Table 6.2 shows some values of the Riemann ζ function
ζ(2) =
π
2
2·3
ζ(4) =
π
4
2·3
2
·5
ζ(6) =
π
6
3
3
·5·7
ζ(8) =
π
8
2·3
3
·5
2
·7
ζ(10) =
π
10
3
5
·5·7·11
ζ(12) =
691π
12
3
6
·5
3
·7
2
·11·13
ζ(14) =
2π
14
3
6
·5
2
·7·11·13
. . .
Table 6.2: Some values ζ(2k) for the Riemann ζ function.
We now give the Taylor expansion of t h e Eisenstein series G
k
(z) with respect to q = e
2πiz
.
Let us start with the well known formula
π cot πz =
1
z
+
m1
1
z + m
+
1
z m
.
We have on the other hand
π cot πz = π
cos πz
sin πz
= πi
q + 1
q 1
= πi
2πi
1 q
= πi 2πi
n0
q
n
,
141
and com p ar i n g , we get:
1
z
+
m1
1
z + m
+
1
z m
= πi 2πi
n0
q
n
. (6.24)
By successi ve differentiations of (6.24), we obtain t h e following formula (valid for k 2 )
mZ
1
(z + m)
k
=
1
(k 1)!
(2πi)
k
n1
n
k1
q
n
. (6.25)
Denote now by σ
k
(n) =
d|n
d
k
the sum of k-th powers of positive divisors of n.
Proposition 6.6.2. For every integer k 2, one has:
G
k
(z) = 2ζ(2k) + 2
(2πi)
2k
(2k 1)!
n1
σ
2k1
(n)q
n
. (6.26)
Proof. To prove this, we expand G
k
(z) =
(n,m)
1
(nz+m)
2k
= 2ζ(2k) + 2
n1
mZ
1
(nz+m)
2k
.
Applying (6.25) with z replaced by nz, we get
G
k
(z) = 2ζ(2k) + 2
2(2πi)
2k
(2k 1)!
d1
a1
d
2k1
q
ad
= 2ζ(2k) + 2
(2πi)
2k
(2k 1)!
n1
σ
2k1
(n)q
n
.
Definition 6.6.3. For k 1, we define the Eisenstein series E
k
(z) by
E
k
(z) = G
k
(z)/2ζ(2k), (6.27)
where G
k
(z) is the Eisenstein series of index k defined in (6.8) .
Corollary 6.6.4. For k 2, one has
E
k
(z) = 1 + γ
k
n1
σ
2k1
(n)q
n
, (6.28)
where γ
k
= (1)
k
4k
B
k
.
Proof. When taki n g the quotient of G
k
(z) by 2ζ(2k) in equation (6.26), it is clear that
E
k
(z) is given by (6.28). Th e coefficient γ
k
is compu t e d using Proposition 6.6.1
γ
k
=
(2πi)
2k
(2k 1)!
1
2ζ(2k)
=
(2π)
2k
(1)
k
(2k 1)!
(2k)!
2
2k1
π
2k
B
k
= (1)
k
2k
B
k
.
142
E
2
= 1 + 240
n1
σ
3
(n)q
n
E
3
= 1 504
n1
σ
5
(n)q
n
E
4
= 1 + 480
n1
σ
7
(n)q
n
E
5
= 1 264
n1
σ
9
(n)q
n
E
6
= 1 +
65520
691
n1
σ
11
(n)q
n
E
7
= 1 24
n1
σ
13
(n)q
n
Table 6.3: Expan si o n of the first Eisenstein series E
k
.
Table 6.3 gives the q-expansion of some Eisenstein series E
k
:
Since g
2
= 60G
2
, g
3
= 140G
3
, we have
g
2
=
60(2π
4
)
2 · 3
2
· 5
E
2
=
(2π)
4
2
2
· 3
E
2
, (6.29)
g
3
=
140(2π
6
)
3
3
· 5 · 7
E
3
=
(2π)
6
2
3
· 3
3
E
3
. ( 6. 30)
Recall now that
= g
3
2
27g
2
3
=
(2π)
12
2
6
· 3
3
(E
3
2
E
2
3
) = (2π)
12
q 24q
2
+ 252q
2
1472q
3
. . .
.
In fact, we can write a co m p ac t form for the last expansion , due to Jacobi. We will not
prove thi s here. The reader can look [117] or [167].
Theorem 6.6.5 (Jaco b i ) .
= (2π)
12
q
n1
(1 q
n
)
24
. (6.31)
Note that the cusp form η(z) = (2π)
12
∆(z) = q
n1
(1 q
n
)
24
, is the Dedekind fun ct i on
mentioned in Example 1 .4 .2. Usually τ(n) denote the coefficient of q
n
in the expansion of
η(z), thus
η(z) = q
n1
(1 q
n
)
24
=
n1
τ( n) q
n
.
The function n → τ (n) is cal l ed the Ramanujan’s function. Observe also that from the
definition of j (6.17) and equations (6.29), (6.31) we can rewrite
j = 1728
g
3
2
= 1728
(2π)
4
12
3
E
3
2
(2π)
12
η
= 1728
(2π)
12
1728
E
3
2
(2π)
12
η
=
E
3
2
η
,
143
so we obtain the expression (1.3) for j given in Chapter 1.
A useful estimate for the coeffi ci ents of modular functions, which s h ows that the quotient
|a
n
|/n
k
remains bounded when n , is given in the following result due to Hecke. For
a proof see [165, p.94]
Theorem 6.6.6. If f =
a
n
q
n
is a cusp form of weight 2k, then a
n
= O(n
k
).
Remark 6.6.7. The theo r y of quasi-modular forms extends the cl ass ica l theory of modular
forms, when it is equipped with the differencial /∂z. See for example [152].
6.7 Theta functions
Let V be a real vector space of finite dimension n endowed with an invariant measure µ.
Let V
be the dual of V . Let f be a rapidly decreasing smooth function on V (see[163]).
The Four ier transform
ˆ
f of f is defined by the formula
ˆ
f(y) =
V
e
2πix,y
f(x) (x). (6.32)
This is a rapidly decreasing smooth function on V
. Let now Λ be a lattice in V
. We
denote by Λ
the lattice in V
dual to Λ; that is the set of y V
such that x, y Z, for
all x Λ. Observe that Λ
may be identified with t h e Z-dual of Λ (hence the terminology).
Proposition 6.7.1 (Poisson formula). Let ν = µ(V/Λ). One has:
xΛ
f(x) =
1
ν
yΛ
ˆ
f(y).
After replacing µ by ν
1
µ, we can assume that µ(V/Λ) = 1. By taking a basis e
1
, . . . , e
n
of
Λ, we identify V with R
n
, Λ with Z
n
, and µ with the p r oduct measure dx
1
···dx
n
. Thus
we have V
= R
n
, Λ
= Z
n
and we are reduced to the classical Poisson formula (see [163])
We suppose now that V is endowed with a symmetric bilinear form x, y which is positi ve
and non degenerate (i. e., x, x > 0 if x = 0). We identify V with V
by means of this
bilinear form. The lattice Λ becomes thus a lattice in V ; and one has y Λ if and only if
x, y Z, for all x Λ. To a lattice Λ, we associate the following functio n defined on R
+
Θ
Λ
(t) =
xV
e
πtx,x
. (6.33)
We choose th e invari ant measure µ on V such that, if e
1
, . . . , e
n
is an or t h on o r m al basis
of V , the unit cube defined by the e
i
has volume 1. The volume ν of th e lattice Λ is th e n
defined by ν = µ( V /Λ).
144
Proposition 6.7.2. We have the identity
Θ
Λ
(t) = t
n/2
ν
1
Θ
Λ
(t
1
).
Proof. Let f = e
πx,x
. It is a rapidly d ecr ea sing smooth function on V . The Fourier
transform
ˆ
f of f is equal to f. Ind ee d , choose an orthon o r m al basis of V and use this
basis to identify V with R
n
; the measure µ becomes the measure dx = dx
1
···dx
n
and the
function f is
f = e
π(x
2
1
+...+x
2
n
)
.
We are thus reduced to showing that the Fourier tr an sfo rm of e
πx
2
is e
πx
2
, which is well
known. We now apply Proposit i on 6.7.1 to the function f and to the lattice t
1/2
Λ; th e
volume of this lattice is t
n/2
v and its dual is t
1/2
Λ
; hen ce we get the formula.
We can give a matrix interpretation. L et e
1
, . . . , e
n
be a basis of Λ. Put a
ij
= e
i
, e
j
. The
matrix A = [a
ij
] is positive, non degenerate and symmetric. If x =
x
i
e
i
is an element of
V , then
x, x =
i,j
a
ij
x
i
x
j
.
The fun ct i on Θ
Λ
can be writt en as
Θ
Λ
(t) =
x
i
Z
e
πt
a
ij
x
i
x
j
. (6.34)
The volume ν of Λ is given by ν = (det A)
1/2
. This can be seen as follows: Let ε
1
, . . . , ε
n
be an orthon or m al basis of V and put
ε = ε
1
. . . ε
n
, e = e
1
. . . e
n
.
We have e = λε, with |λ| = ν. Moreover, e, e = det Aε, ε, and by comparing, we obtain
ν
2
= det A. Let B = [b
ij
] be the matrix inverse to A. One checks immediately that the
dual b a si s {e
i
} to {e
i
} is given by the formulas
e
i
=
j
b
ij
e
j
, for i = 1, . . . , n.
The {e
i
} form a basis of Λ
. The matrix [e
i
, e
j
] is equal to B. Th i s shows in particular
that i f ν
= µ(V /Λ
), then we have νν
= 1.
We will be interested in pairs (V, Λ) wh ich have the following two properties:
1. The dual Λ
of Λ is equal to Λ.
2. We have x, x 0 (mod 2), for all x Λ.
145
Condition (1) says that one has x, y Z, for all x, y Λ and that the form x, y
defines an isomorphism of Λ onto its dual. In m a t r ix terms, it means that the mat r i x
A = [e
i
, e
j
] has integer coefficients and that its determinant equals 1, or equivalently, to
ν = 1. Condition (2) means that the diagonal terms of A are even. We have cal le d these
lattices even unimodular (see Section 4.4). Suppos e that the pair (V, Λ) satisfies condi t i on s
(1) and (2) above, that i s, Λ is even unimodular. Let m 0 be an integer, and denote by
Λ
m
the set of elements x Λ such that x, x = 2m. It can be seen that |Λ
m
| is bounded
by a polyn om i al in m (a crude volume argument gives for instance |Λ
m
| = O(m
n/2
)). This
shows that the series with integer coefficients
θ
Λ
(q) =
xΛ
q
x,x/2
=
mZ
|L
m
|q
m
, (6.35)
defined in equation (4.29), converges for |q| < 1. Thus one can define a function θ
Λ
(z) on
the half-plane H by the formula ( 6. 35), with q = e
2πiz
. Th e function θ
Λ
is called the theta
function of the quadratic m odule Λ. It is holomorphic on H.
Theorem 6.7.3. 1. The dim ension n of V is divisible by 8.
2. Also, the function θ
Λ
is a modular form of weight n/2.
Proof. We prove th e identity
θ
Λ
(1/z) = (iz)
n/2
θ
Λ
(z). (6.36)
Since the two sides are analytic in z, it suffices to prove this formula when z = it with
t > 0 real. We have
θ
Λ
(it) =
xΛ
e
πtx,x
= Θ
Λ
(t).
Similarly, θ
Λ
(1/it) = Θ
Λ
(t
1
). Formula (6.36) res u l t s thus from Proposition 6 .7 . 2 , takin g
into accou nt that ν = 1 and Λ = Λ
.
Now, to prove th e first assertion, su p pose that n is not divisibl e by 8; replacing Λ, if
necessary, by Λ Λ or Λ Λ Λ, we may suppose that n 4 (mod 8). Formula (6.36)
can th en be written
θ
Λ
(1/z) = (1)
n/4
z
m/2
θ
Λ
(z) = z
n/2
θ
Λ
(z).
If we put ω(z) = θ
Λ
(z) dz
n/4
, we see that the differential form ω is transformed into ω by
S : z → 1/z. Since ω is invariant by T : z → z + 1, we see that ST tra n sfor m s ω into
ω, which is absu r d because (ST )
3
= 1.
For (2), since n is divisible by 8, we can rewrite ( 6 .3 6 ) in the form
θ
Λ
(1/z) = z
n/2
θ
Λ
(z) (6.37)
which shows that θ
Λ
is a modular form of weight
n
2
.
146
Corollary 6.7.4. There exists a cusp form f
Λ
of weight
n
2
such that θ
Λ
= E
k
+ f
Λ
, where
k =
n
4
.
Proof. This follows from th e fact that θ
Λ
() = 1, hence that θ
Λ
E
k
is a cusp form.
Corollary 6.7.5. We have |Λ
m
| =
4k
B
k
σ
2k1
(m) + O(m
k
), where k =
n
4
.
Proof. This follows from Corol la r y 6.7.4, equation (6.28) and Theorem 6.6.6.
Note t h at the ‘error term’ f
Λ
is in general nonzero. However, Siegel has proved th at the
weighted mean of the f
Λ
is zero. More preci sel y, let C
n
be the set of classes (up to isomor-
phism) of even unimodular lattices Λ, and denote by g
Λ
the order of the automorphism
group of Λ C
n
. One has:
ΛC
n
1
g
Λ
· f
Λ
= 0, (6.38)
or equi valently
ΛC
n
1
g
Λ
· θ
Λ
= M
n
E
k
, wh er e M
n
=
ΛC
n
1
g
Λ
. (6.39)
Note that this is also equivalent to saying that the weighted me an o f the θ
Λ
is an eigen-
function of the Hecke operator T (n) (for a proof see [ 16 5 ] ).
Example 6.7.6 (The case n = 8). Every cusp form of weight
n
2
= 4 is zero. Corollary
6.7.4 then shows that θ
Λ
= E
2
, in other words, |Λ
m
| = 240σ
3
(m) for all integers m 1.
This ap p l i es to the lattice Q
E
8
constructed in S ect i on 4.6.
Example 6.7.7 (The case n = 16). For the same reason as above, we have θ
Λ
= E
4
=
1 + 480
m
σ
7
(m)q
m
. Here one may take Λ = Q
E
8
Q
E
8
or Λ = Q
E
16
; even though these
two lattices are not isomorphic, they have the same theta function , i. e., they represent
each integer the same number of t i m es. Note that the function θ associated to the lattice
Q
E
8
Q
E
8
is the square of the function θ of Q
E
16
; we recover thus the identity:
1 + 240
m1
σ
3
(m)q
m
2
= 1 + 480
m1
σ
7
(m)q
m
.
Example 6.7.8 (The case n = 24). The space of modular forms of weight 12 is of dimension
2. It has for basis the two functions:
E
6
= 1 +
65520
691
σ
11
(m)q
m
,
and
η = (2π)
12
= q
m1
(1 q
m
)
24
=
m1
τ( m)q
m
.
147
The theta fun c ti o n associated wit h the lattice Λ can t hus be written θ
Λ
= E
6
+ c
Λ
η, with
c
Λ
Q. We have
|Λ
m
| =
65520
691
σ
11
(m) + c
Λ
τ( m), for m 1.
The coefficient c
Λ
is deter m i n ed by putting m = 1:
c
Λ
= |Λ
m
|
65520
691
, for m 1.
For exampl e:
The Leech lattice Λ
24
is such th a t |
24
)
1
| = 0
24
has no 1-norm elements). Hence,
c
Λ
24
=
65520
691
.
The lattice Q
E
8
Q
E
8
Q
E
8
is such that |(Q
E
8
)
1
| = 3240. Hence, c
Q
E
8
=
432000
691
.
148
Chapter 7
Conway-Norton fundamental
conjecture
This chapter is a preamble to the proof given by Borcherds. In fact, we d evelop some
techniques and tools which appear in the Conway-Norton conjecture. We i n i t i al ly give
some ba ckground theory on Hecke operators. Then we descr i be som e properties of the
congruence su b gr oups of SL
2
(Z) and their normalizers. Fi n al l y, following [38], we mention
various moonshine properties that leaded Conway and Norton to formulate their conjectu r e,
including the r ep l i cat i on formulas, useful in the proof of Moo n sh i n e conjecture.
7.1 Hecke operators
Definition 7.1.1. Let E be a set and let X
E
be the free abelian grou p generated by E.
A correspondence on E (with integer coefficients) is a homomorphism T : X
E
X
E
. We
can give T by its values o n the elements x of E:
T (x) =
yE
n
y
(x)y, where n
y
(x) Z, (7.1)
the n
y
(x) being ze r o for almost all y.
Let F be a numerical valued function on E. By Z-linearity, it extends to a function (again
denoted F ), on X
E
. The transform of F by T , denoted T F, is the restriction to E of th e
function F T . With the notations of (7.1), we have
T F(x) = F (T(x)) =
yE
= n
y
(x)F (y). ( 7. 2)
Example 7.1.2 (The operators T (n)). Let L be the set of lattices of C (c. f. Section 6.3) .
Let n 1 be an integer. We denote by T (n) the correspondence on L which transforms
a latti ce to the sum (in X
L
) of its sublattices of index n. Thus we have:
T (n =
[Λ:Λ
]
Λ
, i f Λ L . (7.3)
The sum on the right-hand side of (7.3) is finite. Indeed, the latti ces Λ
all contain nΛ and
their number is also the number of subgroups of order n of Λ/nΛ = (Z/nZ)
2
. If n is prime,
149
one sees in fact that thi s number is equal to n + 1 (number of points of the projective lin e
over a fiel d with n elements). We also use the homothety operato r s R
λ
(λ C
×
), defined
by
R
λ
Λ = λΛ, if Λ L . ( 7. 4)
Definition 7.1.3. The operators T (n) defined above are usually called Hecke operators.
It makes sense to compose the cor r espondences T (n) an d R
λ
, since they are endomorphisms
of the abelian group X.
Proposition 7.1.4. The correspondences T (n) and R
λ
verify the identities
1. R
λ
R
µ
= R
λµ
, for all λ, µ C
×
.
2. R
λ
T (n) = T (n)R
λ
, for all n 1 and all λ C
×
.
3. T (m)T (n) = T (mn), if (m, n) = 1.
4. T (p
n
)T (p) = T (p
n+1
) + pT (p
n1
)R
p
, for p prime, n 1.
Proof. Statements (1) and (2) are trivial. Note that (3) is equivalent to the following
assertion: Let m, n 1 be two relati vely prime integers, and let Λ
′′
be a su b l at t i ce of a
lattice Λ of index mn; there exists a unique sublatt i ce Λ
of Λ, containi n g Λ
′′
, su ch that
: Λ
] = n and
: Λ
′′
] = m. This assertion follows itself from the fact that the group
Λ/Λ
′′
, which is of order mn, decomposes uniquely into a direct sum of a group of order
m and a group of order n (Bezout’s theorem). To prove (4), let r be a lattice. Then
T (p
n
)T (p)Λ, T (p
n+1
)rΛ and T (p
n1
)R
p
Λ are linear combinatio n s of latt i ces contained in
Λ and of index p
n+1
in Λ (note that R
p
Λ is of index p
2
in Λ ) . L et Λ
′′
be such a lattice. In
the above lin ea r combinations it appear s with coefficients a, b, c, say; we have to show that
a = b + pc, i. e., that a = 1 + pc since b is equ al to 1.
We have two cases: (i) Λ
′′
is not contained in pΛ. Then c = 0 and a is the numbe r of
lattices Λ
, intermediate between Λ and Λ
′′
, and of index p i n Λ; such a lattice Λ
contains
pΛ. In Λ/pΛ the image of Λ
is of index p and it contains the image of Λ
′′
, which is of order
p (hence also of index p because Λ/pΛ is of order p
2
); hence there is only one Λ
which does
the trick. This gives a = 1 and the formula a = 1 + pc is valid. (ii) Λ
′′
pΛ. We have
c = 1; any lattice Λ
of index p in Λ contains pΛ, thus a fortiori Λ
′′
. This gives a = p + 1
and a = 1 + pc is again valid.
Corollary 7.1.5. For n > 1, the T (pn) are polynomials in T (p) and R
p
.
This foll ows from (4) by induction on n. Moreover,
Corollary 7.1.6. The algebra generated by the R
λ
and the T (p), p prime, is commutative;
it contains all the T (n).
150
We will study now the act i on of T (n) on the lattice functions of weight 2k. Let F be a
function on L of weight 2k (recall Section 6.3). By definition
R
λ
F = λ
2k
F, for al l λ C
×
. (7.5)
Let n be an integer. Property (2) in Propositi on 7.1.4 shows that
R
λ
(T (n)F) = T (n)(R
λ
F ) = λ
2k
T (n)F,
in other words, T (n)F is also of weight 2k. Assertions (3) and (4) on same proposition
give:
T (m)T(n)F = T (mn)F, if (m, n) = 1; (7.6)
T (p)T(p
n
)F = T (p
n+1
)F + p
12k
T (p
n1
)F, for p prime, n 1. (7.7)
Let Λ be a lattice with basis {ω
1
, ω
2
} and let n 1 be an integer. The following lemma
gives all the sublattices of Λ of ind ex n:
Lemma 7.1.7. Let S
n
be the set of integer matrices (
a b
0 d
), with ad = n, a 1, 0 b < d.
If σ = (
a b
0 d
) is contained in S
n
, let Λ
σ
be the sublattice of Λ having for basis
ω
1
=
1
+
2
, ω
2
=
2
.
The map σ → Λ
σ
is a bijection of S
n
onto the set Λ(n) of sublattices of index n in Λ.
Proof. The fact th at Λ
σ
belongs to Λ(n) foll ows from th e fact t h a t det σ = n. Conversely
let Λ
Λ(n) . We put
Y
1
= Λ/
+ Zω
2
) and Y
2
= Zω
2
/
Zω
2
).
These are cyclic groups generated respectively by the i m ag es of ω
1
and ω
2
. Let a a n d d be
their o r d er s. The exact sequence
0 Y
2
Λ/Λ
Y
1
0
shows that ad = n. If ω
2
=
2
, then ω
2
Λ
. On the other hand, there exists ω
1
Λ
such tha t
ω
1
1
(mod Zω
2
).
It is clear that ω
1
and ω
2
form a basis of Λ
. Moreover , we can write ω
1
in th e form
ω
1
=
1
+
2
, wit h b Z,
where b i s uniquely determined modul o d. If we impose on b the inequ a li ty 0 b < d, this
fixes b, and also ω
1
. Thus, we have associated to every Λ
Λ(n) a matrix σ
) S
n
, an d
one checks that the maps σ → Λ
σ
and Λ
→ σ
) are inverses to each other; the lemma
follows.
Example 7.1.8. If p is a prime, the elements of S
p
are the matrix
p 0
0 1
, and the matrices
1 b
0 p
, with 0 b < p.
151
7.2 Action of T (n) on modular functi o ns
Let k b e an integer, and let f be a weakly modular function of weight 2k. As we saw in
Section 6.3, f correspon d s to a function F of weight 2k on L , satisfying equation (6.7)
F (Λ(ω
1
, ω
2
)) = ω
2k
2
f(ω
1
2
). (7 .8 )
We define T (n)f as the function on H associated to the funct i on n
2k1
T (n)F on L . (Note
that the numerical coefficient n
2k1
gives formulas ‘without denom i n at or s ’ in what follows.)
Thus by definition:
T (n)f(z) = n
2k1
T (n)F(Λ(z, 1)),
or else by Lemma 7.1.7
T (n)f(z) = n
2k1
ad=n, 0b<d
1
d
2k
f
az + b
d
. (7.9)
Proposition 7.2.1. The function T (n)f is weakly modular of weight 2k. It is holomorphic
on H if f is. We have:
T (m)T(n)f = T (mn), if (m, n) = 1, (7.10)
T (p)T(p
n
)f = T (p
n+1
)f + p
2k1
T (p
n1
)f, if p is prime, n 1. (7.11)
Proof. Formula (7.9) shows that T (n)f i s meromorphic on H, thus weakly m odular; if in
addition f is holomorphic, so is T (n)f. Formulas ( 7. 10) and (7. 11) foll ow from formulas
(7.6) and (7.7) taking int o account the numerical coefficient n
2k1
incorporated into t h e
definition of T (n)f.
Now suppose tha t f is a modular function (of weight 0), with q-ex p an s io n of the form
f(z) =
mZ
H
m
q
m
, q = e
2πiz
.
Theorem 7.2.2. T (n)f is also a modular function with q-expansion
T (n)f(z) =
mZ
s|(m,n)
s
1
H
mn/s
2
q
m
.
Proof. By definition, we have
T (n)f(z) = n
1
ad=n, 0b<d
µZ
H
µ
e
2π(az+b)/d
.
The sum
0b<d
e
2πiµb/d
=
0, µ ≡ 0 (mod d);
d, µ 0 (mod d);
152
since it is the sum 1
µ
+ζ
µ
2
+. . .+ζ
µ
d
over all d-roots of un i ty. Hence the sum is over multiples
of d. Puttin g µ = td,
T (n)f(z) = n
1
ad=n, µ=td
dH
td
e
2πiaµz/d
= n
1
ad=n, tZ
dH
td
q
at
.
Collecting powers of t and putting m = at, this gives
T (n)f(z) =
mZ
q
m
a|(m,n)
d
n
H
m
a
·
n
a
=
mZ
a|(m,n)
a
1
H
mn/a
2
q
m
.
Since f s m er om or phic at infinity, there exists an integer N 0 such that c
n
= 0 for all
m N. The c
md/a
are thus zero for all m nN, which shows that T (n)f is also
meromorphic at infinity. Since it is weakly modular (by Proposition 7.2.1), it is also a
modular function.
7.3 Congruence subgroups of SL
2
(R)
We shal l describe in this section the genus 0 subgroups G SL
2
(R) which appear in t he
Conway-Norton co n ject u r e (see Chapter 1).
Let N be a positive integer an d let
Γ
0
(N) =
a b
c d
Γ : c 0 (mod N )
.
This is called a congruence subgroup of SL
2
(Z). We shall describe the normalizer
N
SL
2
(R)
0
(N)),
i. e., the largest subgroup of SL
2
(R) in which Γ
0
(N) is normal. In describing this norm al -
izer, the divisors of 24 play a crucial role. Let h be the largest divisor of 24 such t h a t h
2
divides N, and consider the factorization
N = hn.
Here, n is a positive integer d i v is ible by h. Let T be the subg ro u p of SL
2
(R) given by
T =
a b/h
cn d
SL
2
(R) : a, b, c, d Z, ad bcn/h = 1
.
It is readily seen that T i s a subgroup of SL
2
(R) containing Γ
0
(N) and that Γ
0
(N) is
normal in T . Hence
T N
SL
2
(R)
0
(N)).
153
In fact, T is conjugate to Γ
0
(n/h) in SL
2
(R) sinc e we have
h
1/2
0
0 h
1/2
a b/h
cn d
h
1/2
0
0 h
1/2
=
a b
cn/h d
.
Now, T is not in general th e full normalizer of Γ
0
(N). However, T is normal in this
normalizer and the factor group is an e lem e ntary abelian 2-group. This can be understood
as follows.
Definition 7.3.1. A Hall divisor of n/h is a divisor e such that ( e, n/he) = 1.
The number of Hall divisors of n/h is a power of 2. The Hall divisors form an elementary
abelian group Z
2
× . . . × Z
2
under the composition e f = g, where
g =
e
(e, f)
·
f
(e, f)
.
For each Hall divisor e of n/h we may describe a coset of T in the norma l ize r of Γ
0
(N),
which corresponds to e under the above isomor phism. This coset consists of all mat r i ces of
the form
ae
1/2
(b/h)e
1/2
cne
1/2
de
1/2
,
for a, b, c, d Z and ade bc(n/h)e
1
= 1. Moreover, N
SL
2
(R)
0
(N)) is the union of these
cosets for all Hall d i v is or s e of n/h (see [38]). Thus, the quotient of the norm a li zer by its
normal subgroup T is isomor p h i c to the group of Ha l l divisors of n/h. Thus we have
N
SL
2
(R)
0
(N))/T
=
Z
2
× . . . × Z
2
,
and th e number of factors on the right-hand side is the number of distinct prime divisors
of n/h. We can now st a t e a more precise form of the Conway-Norton conjecture 1.7.1
Conjecture 7.3.2 (Moonshine Conjecture). (Conway-Norton, 1979). Let g be an ele-
ment of the Monster group M. Then, the McKay-Thompson series T
g
(z) is the normalized
Hauptmodul
J
G
(z) :
H/G CP
1
for some genus 0 subgroup G of SL
2
(R) satisfying
T G N
SL
2
(R)
0
(N)),
for some N.
We call a discrete subgroup G of SL
2
(R) a sub gr ou p of moonshine-type if it contains some
congruence subgroup Γ
0
(N), and obeys
1 t
0 1
G t Z. (7.12)
154
Of course, not all subgroups G satisfying the above condition will have genus 0. Thompson
[173] proved that there are only finitely many modular grou p s of moonshine-type in each
genus. Cumm ins [42] has found all of th es e of genus 0 and 1. In particular , t h er e are
precisely 6,486 genus 0 m oonshine-type group s. Exactly 616 of these h ave Hauptmoduls
with rational (in fact integral) coeffici ents; the remainder have cyclotomic integer coeffi-
cients. There are some natural equivalences (for exampl e a Galois action) which collapse
this number to 371, 310 of which have integral Hauptmoduls.
We i ll u s t ra t e this situation by considering the special case in which N is prime. So, let
N = p be a prime, and then we have n = p and h = 1. It follows that T = Γ
0
(p) an d
[N
SL
2
(R)
0
(p)) : T ] = 2.
Now, it has been shown by Ogg (see [158]) that N
SL
2
(R)
0
(p)) h as genus 0 if an d only if
p {2, 3, 5, 7, 11, 13, 17, 19, 23, 29, 31, 41, 47, 59, 71 },
the pr i m e divisors of the order of the Monster group M (recall Chapter 1).
As mentioned in the introductory chapter, th e central stru ct u r e in th e attem p t t o un der-
stand equations (1. 7 ) is an in fi n i t e- d i m en si o n al Z-graded module for the Monster, V
=
V
1
V
1
V
2
. . ., with graded dimension J(z) (see (1.8)). If we let ρ
n
denote the n-th small-
est irreducible M-module, with dimension d
n
numbered as in Table 1.2, then t h e first few
subspaces will be V
0
= ρ
0
, V
1
= {0}, V
2
= ρ
0
ρ
1
, V
3
= ρ
0
ρ
1
ρ
2
, V
4
= 2ρ
0
2ρ
1
ρ
2
ρ
3
and so on. As we know from the rep r esentation theor y for finite groups, a dime n si on can
(and should) be replaced with the character. This gives us the graded traces (1.9)
T
g
(z) =
nZ
tr(g|V
n
)q
n
.
or McKay-Thompson series for this module V . Of course, T
e
= J.
We write c
g
(n) to be the coefficient of q
n
in McKay-Thompson series T
g
, tha t is
T
g
(z) = q
1
+
n1
c
g
(n)q
n
, wit h q = e
2πiz
.
Conway-Norton conjecture (7.3.2) states th at for each element g of the Monster M, the
McKay-Thompson seri es T
g
is the Hauptmodul J
G
g
(z) for a genus 0 g ro u p G
g
Γ of
moonshine-type (recall ( 7. 12)). These groups each contain Γ
0
(N) as a normal subgroup ,
for some N dividing o(g) ·(24, o(g)), o(g) th e order of g, and the quotient grou p G
g
/Γ
0
(N)
have exponent 2. So for each n the map g → c
g
(n) is a character tr(g|V
n
) of M. Conway
and Nor t o n [38] explicitly identify each of the groups G
g
. The fir st 50 coefficients c
g
(n) of
each T
g
are given in [148]. Together with the recursions given in Section 7.5 below, this
allows one to effectively com pute arbitrarily many coefficients c
g
(n) of the Hauptmoduls.
155
It is also this that uniquely defines V
, up to equivalence, as a graded M-module.
There are around 8×10
53
elements in the Monster, so naively we may ex pect about 8×10
53
different Hauptmoduls T
g
. However, a character evaluated at g and at any of his conjugat es
hgh
1
will always be equal, so T
g
= T
hgh
1
. Hence there can be at most 19 4 distinct T
g
(one for each conjugacy class). All coefficients c
g
(n) are integers ( a s are in fact most entries
of the charact er table of M). This implies that T
g
= T
h
whenever the cyclic subgroups
g and h are equal. In fact, the total numbe r of distinct McKay-Thompson series T
g
arising in Monstrous Moonshine turns out to b e only 171. Of those many redundan ci es
among the T
g
, only one is unexpect ed —and unexplained—: the McKay-Thompson series
of two unrelated classes of o rder 27, nam e ly 27A and 27B (i n Atlas notation), are equal.
It would be interesting to understand what gen er al phenomenon, if any, is res ponsible for
T
27A
= T
27B
. But as we know from the theory of vertex algebras, the McKay-Thompson
series T
g
(z) are actu al l y specialisations of 1-point fu n ct i ons and as such are fun ct i on s of not
only z but of all M- i nvariant vectors v V
. What we call T
g
(z) is really the spe cia l isa t io n
T
g
(z, 1) of this function T
g
(z, v).
Not all subgroups G of genus 0 satisfyi n g T G N
SL
2
(R)
0
(N)), for some N, correspon d
to McKay-Thompson series. If h = 1 almost all of them do, but there a r e three exceptions
(c. f. [38]). It has recently shown by Conway (see [105]) that, apar t from these exceptions
when h = 1, there is a fairly simpl e characterization of the groups ar i si n g as G
g
in Monstrous
Moonshine:
Proposition 7.3.3. A subgroup G of SL
2
(R) equals one of the modular groups G
g
appear-
ing in Conjecture 7.3.2 if, and only if
1. G is genus 0;
2. G has the form Γ
0
(n/h) + e, f, g, . . .;
3. the quoti ent of G by Γ
0
(nh) is a group of exponent 2;
4. each cusp Q {i∞} can be mapped to {i∞} by an element of SL
2
(R) that conjugates
the group to one containing Γ
0
(nh).
This is an observation mad e by examining the possible cases, although the signific an ce of
this condition is not yet underst ood. Also, n o t at i on Γ
0
(n/h) + e, f, g, . . . in (2) corresponds
to the full normaliser Γ
0
(N) in P SL
2
(R, obtained vy adjoinig to the group Γ
0
(n/h) its
Atkin-Lehner involutions. See [38] or [105] for more details.
The subgroup G corresponding to a McKay-Thompson series T
g
(z) was con jec t u r ed expl i c-
itly by Conway and Norton for each g M. The su bgroup G is specified by giving the
integer N and a s u b set of Hall divisors of n/h. In fact, N arises as the least positive integer
156
such tha t
1 0
N 1
T
g
(z) = T
g
(z),
and n arises as the order o f g. Then, n divides N and the quotient h = N/n divides 24. In
fact, h
2
divides N. The subgroup G is given by
G =
a b
c d
SL
2
(R) :
a b
c d
T
g
(z) = ζT
g
(z) for some ζ C with ζ
h
= 1
.
We give a few e xam ples to illustrate the situation.
Example 7.3.4. M has 5 conjugacy c la sses of elements of order 10 . They all ar i se from
the congruence subgroup with N = 10. We have h = 1, n = 10, n/h = 10 and the Hall
divisors of n/h form a group isomorphic to Z
2
× Z
2
. The five subgroups of this group are:
{1, 2, 5, 10}, {1, 2},
{1, 5}, {1, 1 0}, {1}.
There are 5 corresponding subgroup s G o f t h e n or m a li zer of Γ
0
(10) giving the 5 conjugacy
classes of M or order 10.
Example 7.3.5. M has 6 conjugacy classes of elements of order 6. Five of them arise from
the congruence subgroup with N = 6. In this case, we have h = 1, n = 6, n/h = 6 and the
Hall divisors of n/h form a gr ou p isomorphic to Z
2
×Z
2
. The fi ve subgroups of this group
are:
{1, 2, 3, 6}, {1, 2},
{1, 3}, {1, 6 }, {1}.
There are 5 corresponding subgroups G of the normal i zer of Γ
0
(6). The remaining conjugacy
class of elements of order 6 arises from the congruence subgroup with N = 18. In this case,
we have h = 3, n = 6, n/h = 2, The Hall divisors of n/h form a group iso m or p h i c to
Z
2
. The unit subgroup {1} is the one giving the required subgroup G of the n or m al i zer of
Γ
0
(18).
Example 7.3.6. M has 2 conjugacy classes of elements of order 78. They arise from the
congruence subgroup with N = 78. In this case we have h = 1, n = 78, n/h = 78 and the
Hall divisors of n/h form a group isomorphic to Z
2
× Z
2
× Z
2
. The two subgrou p s which
we require here are:
{1, 2, 3, 13, 6, 26, 39, 78},
{1, 6, 26, 39}.
These give rise to the two required subgroups G of the normalizer of Γ
0
(78).
These exampl es show some of the subtlety of the correspondences between subgroups of
SL
2
(R) and conjugacy classes of M. The full cor r espondence can be found in [38, Table 2].
157
7.4 Replicable functions
A conjecture in [38] that played an important role in proving the main Conway-Norton
conjecture involves the replication formulae. Conway and Norton init ia l ly thought of the
Hauptmoduls T
g
as being intimately con n ect ed with M; if so, then the grou p structure of
M should somehow directly relate different T
g
. Considering the power map g → g
n
leads
to th e following.
It was well known classically the following
Proposition 7.4.1. The function J(z) (equiv alently, j(z)) has the property that
K(z) = J(pz) + J
z
p
+ J
z+1
p
+ . . . + J
z+p1
p
(7.13)
is a polynomial in J(z), for any prime p.
Proof. The proof i s straightforward, and is b as ed on the principle that the easiest way to
construct a function invariant with r espect to some group G is by averaging it over the
group:
gG
f(gx). Here f(x) is J(pz) and G is the modul ar group P SL
2
(Z), a n d we
average over finitely many cosets rather than infinitely many elements.
First, writing Γ for P SL
2
(Z), n o t e that
Γ
p 0
0 1
Γ =
p 0
0 1
Γ
p1
i=0
1 i
0 p
Γ = {A GL
2
(Z) : det A = p} = S
p
. (7.14)
Here, we have used the fact that we know all the matrices in the group S
p
, according to
Example 7.1.8, for p prime.
Now, applying the Hecke operator T (p) to the modular form J(z) (with weight 2k = 0),
by equat i on (7.9) we have
T (p)J(z) = p
2k1
ad=p, 0b<d
1
d
2k
J
az + b
d
=
1
p
GS
p
J(Gz)
=
1
p
J(pz) + J
z
p
+ J
z+1
p
+ . . . + J
z+p1
p
=
1
p
K(z).
Thus, K(z) = pT (p)J(z). Since J is a modular fo r m , Proposition 7.2.1 implies th at K(z) is
also a m odular form, hence a rational function on
Q(J(z))
P (J(z))
, by Theorem 6.5.2. Since the only
poles of J(z) are at the cusps, th e same applies t o K(z). This implies that the denominat o r
polynomial P (J(z)) must be trivial (recall that J(H) = C). Thus, K(z) is a polynomial in
J(z).
158
In particular, because K(z) is a modu la r invariant (it is co n st a nt on orbits of H/Γ, it must
satisfy K(z) = K( z + 1) (see also Proposition 6.2.2). Thus, K(z) must have a q-expansion
K(z) =
nZ
κ
n
q
n
, wh er e q = e
2πiz
.
More gen er al l y, the same argument says th a t
ad=n, 0b<d
J
az + b
d
= Q
n
(J(z)), (7.15)
is a polynomial in J(z). In fact, Q
n
is the unique polynomial for which q
n
+ Q
n
(J(z))
has a q-e xp a n si on with only strictly positive powers of q (see Appendix A for more details).
For example, Q
1
(x) = x, Q
2
(x) = x
2
2c
1
, Q
3
(x) = x
3
3c
1
x3c
2
, where J(z) =
n
c
n
q
n
.
In fact, thes e equations (7.15) can be rewritten into recursio n s such as a
4
=
a
1
2
+ a
3
, or
can be collected together into the remarkable remarkable identity originall y due to Zagier
[187], but discovered indepen d ently by Borcherds and ot h er s
1
:
p
1
m1, nZ
(1 p
m
q
n
)
c
mn
= J( y) J(z), (7.16)
with p = e
2πiy
, q = e
2πiz
and the powers c
mn
are the coefficients of t h e q-expansion of the
modular fu nction J(z). This is directly used in the proof of the Monstrous Moonshine
conjecture.
Conway and Norton conjectured in [38] tha t these formulae have an analogue for any
McKay-Thompson series T
g
. In particular, (7.15) becomes
ad=n, 0b<d
T
g
a
az + b
d
= Q
n,g
T
g
(z)
, (7.17)
where th e Q
n,g
plays t h e same role as Q
n
plays for J in (7.15). For example
T
g
2
(2z) + T
g
z
2
+ T
g
z+1
2
= T
g
(z)
2
2c
g
(1),
T
g
3
(3z) + T
g
z
3
+ T
g
z+1
3
+ T
g
z+2
3
= T
g
(z)
3
3c
g
(1)T
g
(z) 3c
g
(2).
These are called the replication formulae. Again, these yield recursi ons like c
g
(4) = (c
g
(1)
2
c
g
2
(1))/2 + c
g
(3), or can be collected into the exp r essi on
p
1
exp
k>0
m1, nZ
c
g
k
(mn)
p
mk
q
nk
k
= T
g
(y) T
g
(z). (7.18)
1
D. Zagier (personal communication) has pointed that he was not the first one who proved this ide ntity.
This kind of relations was already known by many other authors, and also appeared in [38]. In de ed , he
proved this formula and other similar relations in [187].
159
This looks a lot more complicated than (7.16), but we can glimpse the Taylor expansion
of log(1 p
m
q
n
) there. In fact, taking g = 1, th e identity element of M, equation (7.18)
reduces to (7.16). Equation (7.1 8 ) is called the Borcherds’ identity, and we will give its
proof on Chapt er 8.
Axiomatising (7.17) leads t o Conway and Norton’s notion of replicable fun ct i on .
Definition 7.4.2. Let f be any function on H of the form f(z) = q
1
+
n1
a
n
q
n
, and
write f
(1)
= f and a
(1)
n
= a
n
. Let X
n
= X
n
(f) the unique monic polynomial of degree n
such tha t the q-expansion of q
n
+ X
n
(f(z)) has only positive powers of q. Use
ad=n, 0b<d
f
(a)
az + b
d
= X
n
f
(1)
(z)
,
to define recursively each f
(s)
, for s 2. In each f
(s)
has a q- expansion of the form
f
(s)
(z) = q
1
+
n1
a
(s)
n
q
n
—that is, n o fractional power s of q arise—, then we call f a
replicable function.
The r ead e r can see Appendix A to get a bett er idea of these functions. Just as a matter of
information, we have the following characterization of replicable functions:
Proposition 7.4.3. Let f be a function of the form q
1
+
n1
a
n
q
n
, and define X
n
(f)
as in Definition 7.4.2. Then, f is replicable if and only if H
m,n
= H
r,s
holds whenever
mn = rs and (m, n) = (r, s).
Proof. The proof is not difficult. Taking the expansion form of the functions X
n
(f). If f
is repl i cable, with replicates f
(s)
= q
1
+
n1
a
(s)
n
q
n
, then
H
m,n
=
s|(m,n)
1
s
a
(s)
mn/s
2
,
(by Theorem 7.2.2), and the H
m,n
= H
r,s
property man i fest s. The converse follows in a
similar way.
Equation (7.17) conjectures that the McKay-Thompson series are replicable. In parti cu l ar ,
we have (T
g
)
(s)
= T
g
s
. Cummins and Norton [44] proved that the Hauptmodul of any genus
0 mod u l ar group of moonshine-type is replicable, provided it s coefficients are rational. In-
cidentally, if the coefficients a
n
are irrational, the n Definition 7.4.2 should be modified to
include Galois au t om or phisms (see [40]).
Any function f obeyin g the replicable functions (7.17 ) will also obey modular equations,
i. e., a certain type of 2-variable polinomial identities sat i sfied by f(x) and f(nx). The
160
simplest examples come from t h e exponential and cosine function s: note that for any n 1,
exp(nx) = (exp(x))
n
and cos(nx) = T
n
(cos(x)), where T
n
is a Tchebychev polynomi al . It
was known clasically that j (and hence J) satisfy a modular equation, for example, the
invariant quantity in (7.13). Note that this property of J depends crucially on it being a
Hauptmodul. Conversely, does the existen ce of modular equ at i ons force the Hauptmodul
property? Unfort unately not; both the exponential and cosine trivially obey modular
equations for each n (use Tchebychev polynomials for cos(nz)). However, we have the
following remarkable fact [124]: The only functions f(z) = q
1
+ a
1
q + a
2
q
2
+ . . . which
obey modular equations for all n ar e J and the ‘modular ficti ons’ q
1
and and q
1
±q (which
are essentially exp, cos, and sin). More generally, Cummings [43] p r oved the following.
Theorem 7.4.4. A function B(q) = q
1
+
n1
a
n
q
n
which obeys a modular equation for
all n 1 (mod N), will either be of t he form B(q) = q
1
+ a
1
q, or will be a Hauptmodul
for a modular group of Moonshine-type.
The converse is also true [43]. This theorem is the desired al geb r a ic interpretation of the
genus 0 property. The denominator id entity argum ent in Chapte r 8 will tells us that each
T
g
obeys a modular equation for each n 1 modulo t he order of g, so Theorem 7.4.4 then
would con cl u de the proof of Monstrous Moonshine. Moreover, Norton has conjectured
Conjecture 7.4.5. Any replicable function with rational coefficients is either a Hauptmodul
for a genus 0 modular group of moonshine-type, or is one of the ‘modular fictions’ f(z) =
q
1
= e
2πiz
, f(z) = q
1
+ q = 2 cos(2πz), f(z) = q
1
q = 2i sin(2πz).
This con ject u r e seems difficult and is still open.
7.5 The replication formulae
Let g be and elem e nt of the Monster group M, and consider
T
g
(z) =
nZ
tr(g|V
n
)q
n
, wit h q = e
2πiz
,
the McKay-Thompson seri es of g. Let G be the genus 0 sub g ro u p of SL
2
(R) associated to
the con ju gacy class of g by Conway and Nort on . Let
T
g
(z) :
H/G CP
1
,
be t he canonical isomorph i sm of Riemann surfaces obtained from the action of G on the
upper half-plane H. The n , the Moonshine conjecture asserts that T
g
(z) = T
g
(z).
Conway and Nort on noticed that the coefficients of T
g
(z) satisfy certain recurrence formulas,
called replication formulae. Let
T
g
(z)
= q
1
+
n>1
c
n
q
n
, wit h q = e
2πiz
.
161
Then, the replica t io n formulas express certain coefficients c
n
in terms of sm al l er coefficients
c
k
, related either to g or g
2
. In the case g = 1, such replication formulas for th e coefficients
of J(z) had been obta i n ed by Mahle r [?]. In the case of arbitrary g M, the following
replication formulae were proved by Koike [118]. We write c
g
(n) to be the coefficie nt of q
n
in T
g
(z), tha t is
T
g
(z)
= q
1
+
n>1
c
g
(n)q
n
, wit h q = e
2πiz
.
We then have four replication formulas:
c
g
(4k) = c
g
(2k + 1) +
1
2
c
g
(k)
2
1
2
c
g
2
(k) +
k1
j=1
c
g
(j)c
g
(2k j); (7.19)
c
g
(4k + 1) = c
g
(2k + 3) c
g
(2)c
g
(2k) +
1
2
c
g
(2k)
2
+
1
2
c
g
2
(2k) +
+
1
2
c
g
(k + 1)
2
1
2
c
g
2
(k + 1) +
k
j=1
c
g
(j)c
g
(2k + 2 j) +
+
k1
j=1
c
g
2
(j)c
g
(4k 4j) +
2k1
j=1
(1)
j
c
g
(j)c
g
(4k j); (7.20)
c
g
(4k + 2) = c
g
(2k + 2) +
k
j=1
c
g
(j)c
g
(2k + 1 j); (7.21)
c
g
(4k + 3) = c
g
(2k + 4) c
g
(2)c
g
(2k + 1)
1
2
c
g
(2k + 1)
2
+
1
2
c
g
2
(2k + 1) +
+
k+1
j=1
c
g
(j)c
g
(2k + 3 j) +
k
j=1
c
g
2
(j)c
g
(4k + 2 4j) +
+
2k
j=1
(1)
j
c
g
(j)c
g
(4k + 2 j). (7.22)
We note in particular that
c
g
(4) = c
g
(3) +
1
2
c
g
(1)
2
1
2
c
g
2
(1), (7 .2 3)
but t h e second replication formula with k = 1 gives simply
c
g
(5) = c
g
(5).
The idea for obtaining such replication for mulae is explained in [38, Section 8] . These
replication formulae can be used to express c
g
(n), for all n, in terms onl y of c
h
(1), c
h
(2),
162
c
h
(3), c
h
(5), for various elements h M, which are powers of g.
This fact gave Borcherds his strategy for proving the Moonsh i n e conjecture. If the coeffi-
cients of t he McKay-Thompson series T
g
(z) could be shown t o satisfy th e same replicati on
formulae, and if their coefficients c
g
(1), c
g
(2), c
g
(3) and c
g
(5) agree with those of T
g
(z), then
it would follow that T
g
(z) = T
g
(z). In fact, Borcherds was able to obtain such replicat i on
formulae for T
g
(z) by me an s of the theor y of infin i t e dimensional Lie algebras, as we shall
explain in th e next chapter.
163
Chapter 8
Borcherds ’ proof of Conway-Norton
conjecture
Borcherds’ proof of Moonshine conjecture makes use of the properties of a Lie alge b r a
called the Monster Lie algebra [10]. This is an exampl e of what is known as gener al iz ed
Kac-Moody algebras, or Borcherds algebras [9, 11]. We shall first descri be the properties
of Borcherds algebras and, subsequently concentrate on the Monste r Lie algebra.
8.1 Borcherds Lie algebras
Definition 8.1.1. A Lie algebra g over R is called a Borcherds algebra if it satisfies the
following axioms:
(i) g =
iZ
g
i
has a Z-grad i ng such that dim g
i
is finite, for all i = 0 (dim g
0
not need to
be fi n i t e) .
(ii) There exists a linear map ω : g g such that
ω
2
= 1, the identity map on g;
ω(g
i
) = g
i
, for all i Z;
ω = 1 on g
0
.
(iii) g has an invariant bilinear form ·, · : g × g R, such that
x, y = 0, if x g
i
, y g
j
and i + j = 0;
ωx, ωy = x, y, for all x, y g;
−x, ωx > 0, if x g
i
, i = 0, x = 0.
These axioms imply that g
0
is abelian and that the scalar product ·, ·
0
: g × g R,
defined by
x, y
0
= −x, ωy
is positive definite on g
i
, for all i = 0. This ·, ·
0
is called the contravariant bilinear form
on g. We shall now give some examples of Borcherds algebras.
165
Example 8.1 .2. Let a = [a
ij
], i, j I, be a symmetr i c mat r i x wit h a
ij
R. The index set
I need not nec essar i l y be to finite —we assume it is either finite or countably infinite—.
Thus, our matrix a may be an infinite matrix. We assume that this matrix satisfies the
conditions
a
ij
0, if i = j;
if a
ii
> 0, then 2
a
ij
a
ii
Z, for all j I.
There is a Borcherds algebra g associated to the matrix a which is defined by gener at or s
and rel at i o n s as follows. g is generate d by elements
e
i
, f
j
, h
ij
for i, j I,
subject to the rela t io n s
[e
i
, f
j
] = h
ij
,
[h
ij
, h
kℓ
] = 0,
[h
ij
, e
k
] = δ
ij
a
ik
e
k
,
[h
ij
, f
k
] = δ
ij
a
ik
f
k
,
if a
ii
> 0 and i = j, then
(ad e
i
)
n
e
j
= 0, (ad f
i
)
n
f
j
= 0, where n = 1 2
a
ij
a
ii
,
if a
ii
0, a
jj
0 and a
ij
= 0, then
[e
i
, e
j
] = 0, [f
i
, f
j
] = 0.
This Li e algebra g can be graded by the condition
deg e
i
= n
i
, d eg f
i
= n
i
,
for some n
i
Z
+
. Ther e is an involution ω : g g satisfying
ω(e
i
) = f
i
, ω(f
i
) = e
i
.
There is also an invariant bilinear for m on g uniquely determined by the condit io n e
i
, f
i
=
1, for all i I. We wri t e h
i
= h
ii
. Then, e
i
, f
i
= h
i
and
h
i
, h
j
= [e
i
, f
i
], h
j
= e
i
, [f
i
, h
j
] = e
i
, a
ii
f
i
= a
ij
,
for all i = j. Thus, h
i
, h
j
= a
ij
, for all i, j I.
We therefore see that th e Lie algebra g satisfies the axioms of a Borcherds algebra (Defi-
nition 8.1. 1). It is called the universal Borcherds algebra associated with the matrix [a
ij
].
166
The grading on g can be chosen in many ways, depending on the choice of the positive
integers n
i
.
We next observe that any sy m m et r i sable Kac-Moody algebra over R (recall Sect io n 4.3)
gives rise to a universal Borcherds algebra. For that, let g be the Kac-Moody algebra over
R with sym m et r i sab l e generalized Car t an matrix A = [A
ij
]. Thus, there exists a diagonal
matrix
D =
d
1
.
.
.
d
n
,
with ea ch d
i
Z
+
, such that the matrix DA i s symmetric. Let a = [a
ij
] be given by
a
ij
=
d
i
A
ij
2
, for all i, j.
Then we have a
ji
= a
ij
and a
ii
= d
i
. Thus a
ij
0 if i = j and a
ii
is a positive integer.
Also,
2
a
ij
a
ii
= A
ij
Z.
Thus, the symmetric matrix [a
ij
] satisfies the conditions needed to construct a Borcherds
algebra, and the universal Borcherds algebra with symmetric matri x [a
ij
] coincides with
the sub a lg eb r a of the Kac-Moody algebra g obtained by generators and relations prior to
the adjunction of the commutative algebra of outer derivations (see Definition 4.3.3). The
difference between a symmetrisable Kac-Moody algebr a and a uni versal Borcherds algebra
is tha t , in a Borcherds algebra:
1. The index se t I may be countably infin it e rather that finite;
2. The a
ii
’s may not be positive and need not lie in Z;
3. 2
a
ij
a
ii
is onl y assumed to lie in Z when a
ii
> 0.
The center of a u n i versal Borcherds algebr a g lies in th e abelian subalgebra generated by
the elements h
ij
and contains a ll h
ij
with i = j. In fact, it can be se en that h
ij
= 0 unless
the i-t h and j-th colu m n s of a are identical. If we factor out an ideal I of g which lies in
the center, then g/I retains the struct u r e of a Borcherds algebra. If we then adjoi n to g/I
an abelian Lie algebr a a of outer derivations to g ive the Lie algebra
g
= (g/I) · a,
where a (g
)
0
and [e
i
, x] Re
i
, [f
i
, x] Rf
i
, for all x a, then g
retains the structure
of a Borcherds algebra.
The converse is also true. G iven any Borcherds algebra g, there is a unique universal
Borcherds al gebra g
U
and a homomorphi sm f : g
U
g (not necessarily unique), such that
167
Ker f lies in the center of g
U
,
Im f is an ideal of g,
g is the semi d i r ect product of Im f with a commutative Lie al geb r a of outer derivati ons
lying in the 0-graded component of g and pr eser vi ng all subspaces Re
i
and Rf
i
.
The hom o m or p h i sm f preserves the grading, involution and bi l inear form.
8.2 The Borcherds character formula
Let g be a universal Borcherds algebra. Recall from Section 4.3 that the root lattice Q of g
is the free abelian group with basis r
i
, for i I, with symmetric bilinear form Q ×Q R,
given by
(r
i
, r
j
) → r
i
, r
j
= a
ij
.
The basi s elements r
i
are call ed the simple roots. We have a grading
g =
αQ
g
α
,
determined by e
i
g
r
i
, f
i
g
r
i
. We have seen in Sect i on 4.3 that an element α Q
is called a root of g if α = 0 and g
α
= 0. The root α is ca ll ed a positive root if α
is a sum of si m p l e roots. For any root α, ei t h er α or α is positive. Let Φ = Φ
+
Φ
be the set of roots of g. We say that α Φ is real if α, α > 0, and imaginary if α, α 0.
Remember also from Section 3.5 that the Weyl group W o f g is the group of isometries of
the r oot l at t i ce Q generated by the reflections w
i
corresponding to the simple real roots.
We have from (3.5)
w
i
(r
j
) = r
j
2
r
i
, r
j
r
i
, r
i
r
i
= r
j
2
a
ij
a
ii
r
i
.
We recall that 2
a
ij
a
ii
Z, since a
ii
> 0. Let h be the abelian suba lg eb r a o f g generated by
the elements h
ij
, for all i, j I. We have a map Q h under which r
i
maps to h
i
, which
is a homomorphism of abelian grou p s and preserves the scalar product. However, this map
need n ot be injective.
If g is any Borcherds algebra, the root system and Weyl group of g i s defined to be that of
the corr esponding universal Borcherds algebra.
We now introduce certain irreducible modules for a Borcherds algebra. Recal l from Section
4.3 that if g is a finite dimensional simple Lie algebra over C, the irredu ci b l e finite dimen-
sional g-mod u l es are in 1-1 correspondence with do m i n ant integral weights. Remember
that a weight λ h
is dominant and integral if, and on l y if, λ(h
i
) 0 and λ(h
i
) Z, for
all i I. The weight λ arises as the h i ghest weight of this module, where λ, µ h
satisfy
168
λ µ if and on l y if λ µ is a sum of simple roots.
Now these finite di m ensional irreducible g-m odules are also in 1-1 corresponde n ce with
antidominant integral weights, i. e ., weights λ h
satisfying λ(h
i
) 0 and λ(h
i
) Z, for
all i I. For there is a un i q u e lowest weight for the module, and th i s is antidominant and
integral. In the case of Borcherds algebras, i t is most convenient to consider lowest weight
modules rather than hig h est weight modules.
Recall that if g is a fi n i t e dimen si onal sim p l e Lie algebr a and λ is an antidominant inte-
gral weight, the corresponding finite dimensional irreducibl e lowest weight module M
λ
has
character g iven by the Weyl’s character formula (4.16)
(char M
λ
)e
ρ
αΦ
+
(1 e
α
) =
wW
ǫ(w)w(e
λ+ρ
),
where ρ =
i
ω
i
(this is an alternative way of writing (4.16), making use of th e denomi-
nator i d entity (4.17)).
Next, suppose that g is a symmetrisable Kac-Moody algebra and λ h
is a weight sat-
isfying λ(h
i
) 0, λ(h
i
) Z, for all i I. Then, g has a corresponding irreducible lowest
module M
λ
whose character is given by Kac’ s character formula (4.18)
(char M
λ
)e
ρ
αΦ
(1 e
α
)
mult α
=
wW
ǫ(w)w(e
λ+ρ
), (8.1)
where ρ h
is any element satisfying ρ(h
i
) = 1, for all i I. This time the sum and
product may be infi n i t e.
Finally suppose that g is a Borcherds algebra. Let λ Q R be a weight satisfying
λ, r
i
0, for all i I;
2
λ,r
i
r
i
,r
i
Z, for all i for which r
i
, r
i
> 0.
Then, there is a corresponding irreducible lowest weight module M
λ
whose character is
given by Borcherds character formula
(char M
λ
)e
ρ
αΦ
+
(1 e
α
)
mult α
=
wW
ǫ(w)w
e
λ+ρ
αQ
ǫ(α)e
α
, (8.2)
where ρ Q R sati sfi e s
ρ, r
i
=
1
2
r
i
, r
i
, for all i with r
i
, r
i
> 0,
and ǫ(α) = (1)
k
if α Q is a sum of k orthogonal simpl e imagi n ar y r oots all orthogonal
to λ, and ǫ(α) = 0 otherwise. Th i s formula reduces to Kac’s character formula (8.1) in
169
the case of a symmetrisable Kac-Moody algebra, since in this case there are no simpl e
imaginary roots and so
αQ
ǫ(α)e
α
= 1.
In the special case λ = 0, the modul e M
λ
is the tri vi al 1-dimensional module and Borcherds’
character for mula becomes
e
ρ
αΦ
+
(1 e
α
)
mult α
=
wW
ǫ(w)w
e
s
ρ
αQ
ǫ(α)e
α
. (8.3)
This is called Borcherds denominator identity. It general i zes Kac’s deno m inator id entity
(4.19), which was itself a generalization of Weyl’s den om i n a t or identity (4.17). As we
shall see, Borcherds’ denominator identity plays a key role in the proof of the Moonshine
conjecture.
8.3 The monster Lie algebra
We now consider an example of a Borcherds algebra, called the Monster Lie algebra [1 0 ],
which is our second main object of study, and plays an important role in the proof of the
Moonshine conjecture.
We start with the Monster vertex algebra V
, constructed in Chapter 5. Re cal l that V
has
a conformal vector of central charge 24. We will r ep l ace it with a vertex operator algeb r a of
central charge 26. Let Π be the lattice of rank 2 whose symmetric bilinear scalar product
Π × Π Z is given by
b
1
, b
1
= 0, b
1
, b
2
= 1, b
2
, b
2
= 0,
where b
1
, b
2
is a basis of Π. Thus we have
mb
1
+ nb
2
, m
b
1
+ n
b
2
= mn
m
n.
There is a vertex a lg eb r a V
Π
associated with the lattice Π as in Section 4.5, and V
Π
has a
conformal vector of central charge 2.
The tensor product V
V
Π
also has the structure of a vertex operator algebra. This vertex
algebra has a conformal vector
ω 1 + 1 ω
Π
of central charge 26, where ω and ω
Π
are conformal vectors of V
and V
Π
, respectively.
Both vertex algebras V
, V
Π
have symmetric bilinear forms, and these define a symmetric
170
bilinear form on V
V
Π
. We define the subspaces
P
1
= {v V
V
Π
: L
0
(v) = v, L
i
(v) = 0 for i 1}; (8.4)
P
0
= {v V
V
Π
: L
0
(v) = 0, L
i
(v) = 0 for i 1}. (8.5)
Now recall from Remark 4.1.3 that the quoti ent (V
V
Π
)/T (V
V
Π
) has the structure
of a Lie algebra. The space P
1
/T (V
V
Π
) P
1
can be identified with a L i e subal geb r a of
(V
V
Π
)/T (V
V
Π
). In fact, we have T P
0
P
1
and
T P
0
= T (V
V
Π
) P
1
.
Thus, P
1
/DP
0
has the structure of a Lie algebra ( see [10]).
The symmetric bilinear form on V
V
Π
induces such a form on P
1
, and T P
0
lies in the
radical of this form. Thus, we obtain a symmetric bilinear form on the Lie algeb r a P
1
/T P
0
.
Let M be the quotient of t h i s Lie algebra by the rad i ca l of the bilinear form, that is,
M =
P
1
/T P
0
Rad·, ·
.
In fact, M is itself a Li e algebra.
Definition 8.3.1. The Lie algebra M is called the Monster Lie algebra.
Now the vertex algebra V
Π
has a grading by elements of th e lattice Π and this induces
gradings of V
V
Π
, and then of its subquotient M by elements of Π. We write
M =
m,nZ
M
(m,n)
, (8.6)
where (m, n) is the graded component corresponding to mb
1
+ nb
2
Π. It was realized by
Borcherds that a theorem fr om string theory, known as the no-ghost theorem, applies to
this si tuation.
8.3.1 The no-ghost theorem
By the remarkable impor t an c e of applyi n g this th eor em in th e proof of Moonshine conjec-
ture, we mention a slight version of the no-ghost theorem (used by Borcherds in [12]). The
idea of using the no-ghost theorem to prove results about Kac-Moody algebras appeared
in Frenkel’s paper [65], which als o contains a proof of the no-ghost theorem. The original
proof of Goddard and Thor n [76] works for the cases we need with only trivial modi fi ca -
tions. For conve n i en ce we give a sketch of their proof.
171
Theorem 8.3.2 (The no-ghost t h eo re m ) . Suppose that V is a vector space with a non-
singular bilinear form ·, ·, and suppose that V is acted on by the Virasoro algebra of
Definition 4.2.2 in such a way that
the adjoin t of L
i
is L
i
,
the central element of the Virasoro algebra acts as multiplication by 24,
any vector of V is a sum of ei genvectors of L
0
with nonnegative integral eige nvalues,
and all the eigenspaces of L
0
are finite dimensional.
We let V
i1
the subspace of V on which L
0
with has eigenvalue i. Assume that V is acted
on by a group G which preserves all this structure. We let V
Π
be the vertex algebra of the
two dimensional even lattice Π (so that V
Π
is Π-graded, has a bilinear form ·, ·, and is
acted on by the Virasoro algebra). We let P
1
be the subspace of V V
Π
as defined in (8.4),
and we let P
1
α
be the subspace of P
1
of degree α Π. All these spaces inherit an action of
G from the action of G on V and the trivial action of G on V
Π
and R
2
.
Then, the quotient of P
1
α
by the nullspace of its bilinear form is naturally isomorphic, as a
G-module with an invariant bilinear form, to
V
−α,α/2
if α = 0
V
0
R
2
if α = 0
.
The nam e ‘no-ghost theorem’ comes from the fact that in the original statement of the the-
orem in [76], V was part of the underlying vector space of the vertex algebra of a positive
definite lattice, so the inner product on V
i1
was positive definite, and thus, P
1
α
had no
ghosts ( i . e., vectors of neg at i ve norm) for α = 0.
We give a sketch of proof taken from [12] and [76]. Fix some nonzero α Π and some
norm 0 vector w Π, with α, w = 0. We use the following operators. We have an action
of the Virasoro algebra on V V
Π
generated by its conformal vector. The operat or s L
i
of
the Vira sor o algebra satisfy the relations (4.13)
[L
i
, L
j
] = (i j)L
i+j
+
1
2
i + 1
3
δ
i+j,0
26,
and t h e adjoint of L
i
is L
i
(the 2 6 comes from the 24 in (4. 13 ) plus the dimension of Π).
We d efi ne operators K
i
, for i Z, by K
i
= v
i1
, where v is the element e
w
2
e
w
of the vertex
algebra o f Π, and e
w
is an element of the group ring R[Π], corresponding to w Π, and
e
w
is its inverse. These operators satisfy the relations
[L
i
, K
j
] = jK
i+j
, [ K
i
, K
j
] = 0,
since w has norm 0 and the ad joi nt of K
i
is K
i
.
We define the following subspaces of V V
Π
:
172
H is the subspace of V V
Π
, of degree α Π. H
1
is its subspace of vectors h with
L
0
(h) = h.
P is the subsp ace of H of all vectors h wit h L
i
(h) = 0, for all i > 0. P
1
= H
1
P .
S, the space of spurious vectors, is the sub sp ac e of H of vectors perpendicular to P .
S
1
= H
1
S.
N = S P is the r adical of the bilinear form of P , and N
1
= H
1
N.
T , the transverse space, is the subspace of P ann i h i l at ed by all the operators K
i
, for
i > 0, and T
1
= H
1
T .
K is the space generat ed by the action of the operator s K
i
, i > 0.
V e
α
is the subspace V e
α
of H.
We have the following inclusions of subspaces of H:
S P K
N
__
@
@
@
@
@
@
@
>>
~
~
~
~
~
~
~
T
__
@
@
@
@
@
@
@
>>
~
~
~
~
~
~
~
V e
α
bb
D
D
D
D
D
D
D
D
and we construct the isomorphism from V
−α,α/2
to P
1
/N P
1
by zigzagging up and down
this diagram; more p re cis el y we show th at V e
α
and T are both isomorphic to K modulo
its nullspace, and then we show that T
1
is isomorphic to P
1
modulo its nullspace P
1
N.
In fact, the no-ghost theorem follows immediately from the next sequence of lemmas [12]:
Lemma 8.3.3. If f is a vector of nonzero norm in T , then the vectors of the form
L
m
1
L
m
2
. . . K
n
1
K
n
2
. . . (f)
for all sequences of integers with 0 > m
1
m
2
. . ., 0 > n
1
n
2
. . ., are linearly
independent and span a space invariant under the operators K
i
and L
i
on which the bilinear
form is nonsingular.
Lemma 8.3.4. Th bilinear form on T is nonsingular, and K is the direct sum of T an d
the nullspace of K.
Lemma 8.3.5. V e
α
is naturally isomorphic to T .
Lemma 8.3.6. The associative algebra generated by the elements L
i
, for i < 0, is generated
by elements mapping S
1
into S.
Lemma 8.3.7. P
1
is the direct sum of T
1
and N
1
.
173
Recall from (8.6) that
M =
m,nZ
M
(m,n)
,
where ( m, n) is the graded component corresponding to mb
1
+ nb
2
Π. Applying the
no-ghost theorem t o this vertex algebra, this theorem asserts th at for α Π,
M
α
is isomo r p h i c to V
−α,α/2
, i f α = 0,
where V
i
= {v V
: L
0
(v) = (i + 1)v}. Let α = mb
1
+ nb
2
Π. Since α, α = 2mn,
we then have
M
(m,n)
=
V
mn
, i f α = (0, 0).
The no- gh o st theorem also asserts in this situation that
M
(0,0)
=
R
2
.
The graded components of the monster Lie algebra M can ther efor e be shown in the fol-
lowing t able:
.
.
.
0 0 0 0 0 V
4
V
8
V
12
V
16
0 0 0 0 0 V
3
V
6
V
9
V
12
0 0 0 0 0 V
2
V
4
V
6
V
8
0 0 0 V
1
0 V
1
V
2
V
3
V
4
··· 0 0 0 0 R
2
0 0 0 0 ···
V
4
V
3
V
2
V
1
0 V
1
0 0 0
V
8
V
6
V
4
V
2
0 0 0 0 0
V
12
V
9
V
6
V
3
0 0 0 0 0
V
16
V
12
V
8
V
4
0 0 0 0 0
.
.
.
Table 8.1: Grad ed compon ents of the Monster Lie algeb r a M.
The grou p ring R[Π] of the lattice Π has an involution defined by
e
α
→ (1)
α,α/2
e
α
, for α Π.
This gives rise to an involution on the vertex algebra V
Π
= S(
˜
h
) R[Π]. This in turn,
gives rise to an involution on the vertex algebra V
V
Π
, which acts trivially on V
. This
involution acts on the subquotient M of V
V
Π
, giving a map ω : M M such that
174
ω
2
= 1,
ωM
(m,n)
= M
(m,n)
,
ω = 1 on M
(0,0)
,
ωx, ωy = x, y;
where · , · is the invariant b i l i n ear form on M. Moreover, the contravariant form
x, y
0
= −x, ωy, for x, y M,
is positive d efi nite on M
(m,n)
, for all (m, n) = (0, 0).
We can give M a Z-grad ing by means of the formula
deg M
(m,n)
= 2m + n.
The Z-g ra d ed compone nts are:
··· 5 4 3 2 1 0 1 2 3 4 5 ···
··· V
2
V
3
V
2
V
1
0 V
1
R
2
V
1
0 V
1
V
2
V
2
V
3
···
Table 8.2: Z-gr a d ed components of the Monster Lie al geb r a M.
Hence, M satisfies the axioms for a Borcherds algebra, having the necessary Z-grading,
involution, and invariant bilinear form giving rise to a positive definite contravari ant form
on non-zero graded components.
Let Q be th e root lattic e of the Borcherds algebra M, and let h be the Cartan subalgebra
of M. Then , h = M
(0,0)
= R
2
and we have a map Q h as in Section 8.1, under which
each simple root r
i
Q maps to h
i
h, and preserving the scalar product. The elements
of h may be written in the form mb
1
+ nb
2
, for m, n Z. The elements of h which a ri se as
images of simple roots in Q are those with (m, n) equal to
(1, 1), (1, 1) , (1, 2), (1, 3), (1, 4), . . .
Thus, M has i n fi n i t el y may simple root s (note that they are not linearly independent).
Since
mb
1
+ nb
2
, mb
1
+ nb
2
= 2mn,
we see that (1, 1) gives a real simple r oot and that all th e other simple roots (1, n), for
n 1, are imaginary.
175
We have pointed out in Section 8.1 that the map Q h need not be injective, and in
the present si t u at i on it is far from injective. Thus, there can be several simple roots in Q
mapping t o the same element b
1
+nb
2
of h. The number of simp le roots mapping to a given
element b
1
+ nb
2
is called the multiplicity of (1, n). This multiplicity is
dim M
(1,n)
= dim V
n
= c
n
, (8.7)
where c
n
is the coefficient of q
n
in th e expansion of the normalized Hauptmodul (1.4)
J(z) = q
1
+
n1
c
n
q
n
, wit h q = e
2πiz
.
Thus,
(1, 1) has multiplicity 1,
(1, 1) has multiplicity 196,884,
(1, 2) has multiplicity 21,493,760;
. . .
and the sum of th e simpl e roo t spaces in M is isomorphic to the Moonshi n e module V
.
Hence, the Monster Lie algebra M contains within it the Monster vertex algebra V
as the
sum of its root spaces corresponding to the simple roots.
The symmetri c matrix [a
ij
] corresponding to the Borcherds algebra M is thus a countable
matrix with many repeated rows and columns (see Table 8.3). It has the following form:
Since M has onl y one simple real root, its Weyl group W has or d er 2. Any root of M maps
to an element mb
1
+ nb
2
h such that M
(0,0)
= 0 and (m, n) = (0, 0). The multiplicity of
(m, n) is then
dim M
(m,n)
= dim V
mn
= c
mn
.
8.4 Denominator identities
In this section we will describe Borcherds’ denom i n a to r identity and twisted denominator
identity for the monster Lie algebra M. First , we obtain the Borcherds’ deno m i n at or iden-
tity for th e monster Lie algebra, supposing only we know completely the simple roots of M.
Remember from Sect i on 8.1 the Borcherds identity (8.3)
e
ρ
αΦ
+
(1 e
α
)
mult α
=
wW
ǫ(w)w
e
ρ
αQ
ǫ(α)e
α
,
176
(1, 1) (1, 1) ··· (1, 1) (1, 2) ··· (1, 2) (1, 3) ···
(1, 1) 2 0 ··· 0 -1 ··· -1 -2 ···
(1, 1) 0 -2 ··· -2 -3 ··· -3 -4 ···
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
(1, 1) 0 -2 ··· -2 -3 ··· -3 -4 ···
(1, 2) -1 -3 ··· -3 -4 ··· - 4 -5 ···
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
(1, 2) -1 -3 ··· -3 -4 ··· - 4 -5 ···
(1, 3) -2 -4 ··· -4 -5 ··· - 5 -6 ···
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
Table 8.3: Cartan matrix for the monster Lie algebra M.
where ρ Q R is any vector satisfying
ρ, r
i
=
1
2
r
i
, r
i
, i I.
Consider Q h the hom o m or p hism described in Section 8.1, map p i n g si m p l e roots in Q
to h. By abuse of language, we will also call these elements in h of roots. In the case of the
monster Lie algebra M, simple root s in h are the elem ents mb
1
+ nb
2
h, with m, n Z
and mn > 0 or mn = 1. If there are k roots in Q mapping to the same root in h, we will
say that this root in h has multiplicity k. Also, we identify the root mb
1
+ nb
2
with the
pair (m, n) Z
2
. So, we know that the simple roots of M are
(1, 1), (1, 1) , (1, 2), (1, 3), (1, 4), . . .
and we could take ρ = (1, 0), because
(1, 0), (1, n) = n,
(1, n), (1, n) = 2n, for all n,
so that ρ, r
i
=
1
2
r
i
, r
i
for each simple root r
i
. Also, we know that root (m, n) has
multiplicity exactly c
mn
.
Let p = e
(1,0)
and q = e
(0,1)
. We have, e
ρ
= e
(1,0)
= p
1
, so left-hand side of Borcherds’
identity ( 8 .3) is
p
1
m>0, nZ
(1 p
m
q
n
)
c
mn
,
177
(because e
(m,n)
= e
m(1,0)+n(0,1)
= (e
(1,0)
)
m
(e
(0,1)
)
n
= p
m
q
n
). Remember al so that for α Q,
ǫ(α) = (1)
k
, when α is the sum of k imaginary orthogonal simple root s , and ǫ(α) = 0 in
other ca se.
For the monster Lie algebra M, there are no two imaginary orthogonal simple roots, because
(1, m), (1, n) = m n < 0, for all m, n. Thus, the elements α Q contributing to sum
α
ǫ(α)e
α
are α = 0 with ǫ(α) = 1 and all the imaginary simple roots (1, n) h, n Z
+
.
Since there are precisely c
n
of these roots in Q (mapping to (1, n)) and all of them have
ǫ(α) = 1, we have
αQ
ǫ(α)e
α
= 1
n>0
c
n
pq
n
. (8.8)
Also, |W | = 2 and W = {1, s}, where s(p) = q and s(q) = p. Thus, right-h an d side of
Borcherds’ identity is
wW
ǫ(w)w
e
ρ
αQ
ǫ(α)e
α
=
wW
ǫ(w)w
p
1
1
n>0
c
n
pq
n
= p
1
1
n>0
c
n
pq
n
q
1
1
n>0
c
n
qp
n
=
p
1
n>0
c
n
p
n
q
1
n>0
c
n
q
n
= j(p) j(q). (8.9)
Combining (8.8) and (8.9), the denominator identity for the monster Lie algebra M estab-
lishes Zagier’s id entity (7.16)
p
1
m>0, nZ
(1 p
m
q
n
)
c
mn
= j(p) j(q). (8.10)
In fact, this identity were first proved by Borcherds in terms of the M structure, and then
were used for prove that the simple roots of M are (1, 1) , (1, 1), (1, 2), (1, 3), . . .
8.5 The twisted denominator identity
In order to complete the proof of Moonshine conjecture, we shall need a more general form
of id entity (8.10), named th e twisted denominator identity. To explain it, we will give an
outline of proof of Zagier’s identity (8.10), and then we wil l generalize it.
178
Definition 8.5.1. Let U be a finite-dimensional r ea l vector space, with a graded decompo-
sition U =
αL
U
α
, for some lattice L. We define the graded dimension of U as the element
in R[L] given by
gr dim U =
αL
(dim U
α
)e
α
,
where R[L] is the group algebra of L with basis e
α
, for α L.
Let
0
U,
1
U,
2
U, . . . be the exterior powers of U, that is
k
U = {k-linear alternate forms ω : U × . . . × U R},
and
0
U = R. We have a simple for mula for the alternate sum
k0
(1)
k
gr dim
k
U,
given by
k0
(1)
k
gr dim
k
U =
αL
(1 e
α
)
dim U
α
. (8.11)
Note th at the right-hand side can also be writ t en as exp
k>0
1
k
αL
(dim U
α
)e
, since
exp
αL
(dim U
α
)
k>0
1
k
e
= exp
αL
(dim U
α
) log(1 e
α
)
=
αL
(1 e
α
)
dim U
α
.
Suppose now that U is a G-module, for some finite group G, and that G a ct s on each
graded component U
α
.
Definition 8.5.2. The graded character of U, as the map gr char U : G R[L] given by
g −
α
tr(g|U
α
)e
α
.
Then, t h e alternate sum
k0
(1)
k
gr dim
k
U is given by th e map
g − exp
k>0
1
k
α
tr(g
k
|U
α
)e
.
Observe that when g = 1 , this map r ed uces to the alternat e sum in (8.11 ) . If U is an
infinite dimensi on a l vector space and U =
αL
U
α
is a graded decomposition of U, such
179
that each component U
α
is fini te dimension al , then the formulas are still valid (changing
the sum s for infinite sums).
Suppose now that g is a Borcherds al gebra with triangular decomposi t i on
g = n
+
h n
,
where n
+
=
αΦ
+
g
α
and n
+
=
αΦ
g
α
(note that n
+
can be infinite dimens io n al , but
each g
α
must be finit e dimension al ) . Consider the exterior powers
0
U,
1
U,
2
U, . . .
We have a sequence
. . .
d
4
3
n
+
d
3
2
n
+
d
2
1
n
+
d
1
0
n
+
d
0
0
with h o m ol og y groups
H
0
n
+
=
Ker d
0
Im d
1
, H
1
n
+
=
Ker d
1
Im d
2
, H
2
n
+
=
Ker d
2
Im d
3
, . . .
Observe that
k
n
+
and H
k
n
+
are graded vector spaces with finite dimensional graded
components. By definition of the homology groups H
k
n
+
, we have
k0
(1)
k
gr dim
k
n
+
=
k0
(1)
k
gr dim H
k
n
+
.
Garland and Lepowsky [73] p r oved that for Kac-Moody algebras (and also it is verified for
Borcherds algebras), H
k
n
+
can be identified with a subspace of
k
n
+
as follows:
H
k
n
+
α
=
(
k
n
+
)
α
if α + ρ, α + ρ = ρ, ρ
0 in other case
,
and this holds for each α in the root lattice of g.
Now, we specialize for the case g = M, the monster Lie algebra. The formula (8.11) gives
k0
(1)
k
gr dim
k
M
+
=
(m,n), m>0
(1 p
m
q
n
)
c
mn
,
where p = e
(1,0)
, q = e
(0,1)
and M
+
=
αΦ
+
M
α
.
180
We also compute the alternate sum
k0
(1)
k
gr dim H
k
m
+
. We have, gr dim
0
M
+
= e
0
,
i. e.,
0
M
+
= R is 1-dimension a l with weight e
0
. Since 0 + ρ, 0 + ρ = ρ, ρ, we also have
gr dim H
0
M
+
= e
0
. Next we have
gr dim
1
M
+
= gr dim M
+
=
(m,n), m>0
c
mn
p
m
q
n
.
Now we consider gr dim H
1
M
+
. We have ρ = (1, 0). Let α = (m, n) be a root with
m > 0. Then α + ρ, α + ρ = 2(m 1)n and ρ, ρ = 0, so we get α + ρ, α + ρ = ρ, ρ
if and only if m = 1 or n = 0. Since M
+
has no roots with n = 0 and the roots (m, n) wi t h
m = 1 are precisely the sim p l e roots, we obtain
gr dim H
1
M
+
=
(1,n)
c
n
pq
n
= p
nZ
c
n
q
n
.
Now, the weights of
2
M
+
are sums of two distinct weights in M
+
. Since all weights of
M
+
are of the form (m, n) with m 1 then there are no weights (m, n) in
2
M
+
with
m = 1. However, there are some weights (m, n) wit h n = 0. These are precisely o f the
form (1 , 1) + (m 1, 1), wh e re m 1 1. It follows that
gr dim H
2
M
+
=
m2
c
m1
p
m
.
Note that for
3
M
+
the weights are sums of three distinct weights of M
+
. None has the
form (m, n) with m = 1 or n = 0. Then, H
3
M
+
= 0. Simi l ar l y, H
k
M
+
= 0 for all k 3.
Thus we have
k0
(1)
k
gr dim H
k
M
+
= e
0
p
nZ
c
n
q
n
+
m2
c
m1
p
m
= e
0
+ p
m1
c
m
p
m
p
nZ
c
n
q
n
= e
0
+ p
mZ
c
m
p
m
p
1
p
nZ
c
n
q
n
= p
j(p) j(q)
.
Therefore, we have derived the denominator identity (8.10)
(m,n), m>0
(1 p
m
q
n
)
c
mn
= p
j(p) j(q)
.
181
In order to obtain the twisted denominator identity, we consi d e r M
+
as an M-module for
the Mon st er group M. Then, each root space (M
+
)
α
is an M-modu le. From that we have
k0
(1)
k
gr dim
k
M
+
=
k0
(1)
k
gr dim
k
H
k
M
+
,
(observe that each sub sp a ce (H
k
M
+
)
α
is also an M-m odule). Moreover, the left-hand side
k
(1)
k
gr dim
k
M
+
is the map
g − exp
k>0
1
k
αΦ
+
tr(g
k
|(M
+
)
α
)e
, (8.12)
for g M. If we replace all dimensions for characters in the formulas above involving
H
0
M
+
, H
1
M
+
and H
2
M
+
, we get that the su m
k
(1)
k
gr char H
k
M
+
is the map
g − p
nZ
tr(g|V
n
)p
n
nZ
tr(g|V
n
)q
n
. ( 8. 13 )
Comparing (8.12) and (8.13 ) we deduce that
p
1
exp
k>0
1
k
(m,n), m>0
tr(g
k
|V
mn
)p
mk
q
nk
=
nZ
tr(g|V
n
)p
n
nZ
tr(g|V
n
)q
n
. (8.14)
This is what we have called the twisted denominator identity for the monster L i e algebra M.
Observe that if we wr i t e c
g
(n) = tr(g|V
n
) for the coefficient of q
n
in the graded character
nZ
tr(g|V
n
)q
n
, then equation (8.14) is just
p
1
exp
k>0
(m,n), m>0
1
k
c
g
k
(mn)p
mk
q
nk
=
nZ
c
g
(n)p
n
nZ
c
g
(n)q
n
,
or simply
p
1
exp
k>0
(m,n), m>0
1
k
c
g
k
(mn)p
mk
q
nk
= T
g
(y) T
g
(z),
where y = e
2πip
, z = e
2πiq
. That is, exactly the form of equat i on (7.18).
8.6 Replication formulae aga in, an d proof ’s end
The twisted denominator identity for the monster Lie al g eb r a (8.14) is just what is needed
to obtain the replicat i on formulae (7.19)-(7.22). By comparing the coefficients of p
2
and p
4
182
in this identity and applying some elementary algebra, Borcherds d er i ved the replication
formulae. We omit the proof of this fact here, because the abundance of calculations. The
complete derivation of the replication formulae is made in Appendix A.
This is alm ost sufficient for the proof of the Conway-Norton conjecture. In order to complete
the proof, i t rema ins to show that the coefficients c
g
(1), c
g
(2), c
g
(3) and c
g
(5) of the
McKay-Thompson series T
g
(z), agree with those of T
g
(z). These coefficients for T
g
(z) were
completely obtained by Conway and Norton in [38]. In order to obtain the coefficients for
the graded characters T
g
(z), it is sufficient to know how the modules V
1
, V
2
, V
3
and V
5
of
the Monster M, with dimensions c
1
, c
2
, c
3
and c
5
, respectively, decompose into irreducib l e
modules. Observe that the only irreducible characters of M less or equal than c
5
= dim V
5
are
χ
0
, χ
1
, χ
2
, χ
3
, χ
4
, χ
5
, χ
6
(see Table 1.2), t hus these are the only possibl e irreducible components of V
1
, V
2
, V
3
and
V
5
. Borcherds was able to p roof th at
dim V
1
= χ
0
+ χ
1
,
dim V
2
= χ
0
+ χ
1
+ χ
2
,
dim V
3
= 2χ
0
+ 2χ
1
+ χ
2
+ χ
3
,
dim V
5
= 4χ
0
+ 5χ
1
+ 3χ
2
+ 2χ
3
+ χ
4
+ χ
5
+ χ
6
, (8.15)
where as u su a l, χ
0
, χ
1
, . . . , χ
6
are the rst 7 irreduc ible characters of M (denoted by d
i
in Chapter 1 ) . This was proved by finding 7 elements g
1
, g
2
, . . . , g
7
of M for which the
7 × 7-matrix [χ
i
(g
j
)] is non-singular and by showing that the above equations (8.15) hold
when eval u a t ed at each g
i
. They must then hold for all g M.
We mention in conclu si o n , th a t the proof of the Conway-Norton conjecture for th e Monster
is by no means the sole achievement of Borcherds’ work in [12]. Other spor a d i c simple
groups are also discussed, including the Baby Monster B, th e Conway group Co
1
, th e
Fischer group F i
24
, the Harada-Norton group HN, the Held group He, and the Mathieu
group M
12
, and denominator identities for all th ese groups are also obtained. Thus, the
topic of Monstrous Moonshine is by no me an s confined to the Monster M.
183
Chapter 9
Concluding Remarks
We give in this chapter a quick sketch of further development s and conjectures. As can be
seen, Moonshine is an area where it is much easier to conjecture than to prove.
9.1 Orbifolds
In string t h eor y, the m o st tractable way to introduce singularities is by quotienting (‘gaug-
ing’) by a finite group. This construction plays a fundamental role for CFT and vertex
operator al geb r a s; it is the physics underlying what Norton calls generalized Moonshine.
This is where finite group theory touches CFT. Let M be a manifold and G a finite grou p
of symmet r i es of M. The set M/G of G-orbits inherits a topology from M, a n d forms a
manifold-like space called an orbifold. Fixed points become conical singularities. For ex-
ample, 1} acts on M = R by multiplication. The orbifold R/1} can be identified with
the interval [0 , ). Th e fixed point at x = 0 becomes a sing u l ar point on the orbifold, that
is, a point where locally the orbifold d oes not look like some open n-ball. Orbifolds were
introduced into geometry in the 1950’s as spaces wit h certain kin d of singular i t ie s. They
were introduced into string theory in [49], which greatly increased the class of background
space-times in which the string could live and still be amenable to calculation. This section
briefly sketches the corresponding construction for CFT; our purpose is to moti vate some
generalization of Moonshine conjecture.
About a third of the McKay-Thompson series T
g
have some negative coefficients. We shall
see Borcherds int er p r et them as dimensions of superspaces (which come with signs). In
the important announcement [155], on a par with [38], Norton proposed that, although
T
g
(1/z) will not usually be another McKay-Thompson series, it will always have non-
negative integer q-coefficients, and th ese can be interpreted as ordin ar y dimension s . In
the process, he extended the g → T
g
assignment to commuting pairs (g, h) M × M. In
particular
Conjecture 9.1.1 (Norton). To each such pair g, h M, with gh = hg, we have a function
N
(g,h)
(z), such that
N
(g
a
h
c
,g
b
h
d
)
(z) = αN
g, h;
az+b
cz+d
, for all
a b
c d
SL
2
(Z), (9.1)
185
for some root of unity α (of order dividing 24, and depending on g, h, a, b, c, d). N
(g,h)
(z) is
either constant, or generates the modular functions for a genus 0 subgroup of SL
2
(Z) con-
taining some Γ(N) (but otherwise not necessarily of moonshine-type). Constant N
(g,h)
(z)
arise when all elements of the form g
a
h
b
(with (a, b) = 1) are ‘non-Fricke’. Each N
(g,h)
(z)
has a q
1/N
-expansion for that N; the coefficients of this expansion (not necessarily integers)
are characters ev aluated at h of some central extension of the centralizer C
M
(g). Simulta-
neous conjugation of g, h leaves the Norton series unchanged: N
(aga
1
,aha
1
)
(z) = N
(g,h)
(z).
We call N
(g,h)
(z) the Norton series associated to (g, h). An element g M is called Fricke
if the group G
g
contains an element sending 0 to i —the id e ntity 1 i s Fricke, as are 120
of the 171 G
g
—. For example, when g, h
=
C
2
×C
2
and g, h, gh are all in class 2A, then
N
(g,h)
(z) = q
1/2
492q
1/2
22590q
3/2
+. . ., while N
(g,g)
(z) = q
1/2
+4372q
1/2
96256q+. . ..
The McKay-Thompson series are recovered by taking g = 1: N
(1,g)
= T
g
. This action (9. 1 )
of SL
2
(Z) is related to its natural action on the fundamental group Z
2
of the torus, as well
as a natural action of the braid group, a s we shall see in the next s ect i on . Norton arrived
at hi s conjecture empirically, by studying the data of Queen (Section 9.3).
The basic tool we have for approaching Moonshine conjectures is the theory of vertex op-
erator algebras, so we need to understand Norton’s suggestion from that point of view. For
reasons of space, we limit our discussion to V
, bu t it generalizes. Given any automor-
phism g Aut(V
), we can define g-twisted modules in a straightforward way [51]. Th en
for each g M, there is a un i que g-twisted module, call it V
(g),. More generally, given
any a u t om o r p h i sm h Aut(V
) commuting with g, h will yield an automorphism of V
(g),
so we can perform Thompson’s graded characters (1.9) a n d define
Z
(g,h)
(z) = q
c/24
tr
V
(g)
hq
L
0
. (9.2)
These Z
(g,h)
can be thou ght of as the building b l ocks of the graded dimensions of various
eigenspaces in V
(g); for example if h has order m, then the subspace of V
(g) fixed by
automorphism h will have graded dimension m
1
m
i=1
Z
(g,h
i
)
. In the case of the Monster
considered here, we have Z
(g,h)
= N
(g,h)
.
At t h e l evel of algebr a, this orbifold theory is analogous to t h e construction of twisted
affine algebras from nontwisted ones. At the level of mod u l ar forms, it involves twists an d
shifts much like how θ
4
(z) =
(1)
n
q
n
2
/2
and θ
2
(z) =
q
(n+1/2)
2
/2 are obtained from
θ
3
(z). Far from an esoteric technical development, orb ifo lds are central to the whole theory,
and a crucial aspect of Moonshine. The important p aper [51] proves that, whenever the
subgroup g, h generated by g and h is cyclic, then N
(g,h)
will be a Hauptmod u l satisfying
(9.1). One way this will happen of course is whenever the orders of g and h are coprime.
Extending [51] to all commuting pairs ( g, h) is one of the most pressing tasks in Moonshine.
At least some aspects of orbifolds are more tractable in the subfactor framework (see for
example [55] , so further investigations in that direction should be fruitful. This orbifold
construction is the same as was use d to construct V
from V
Λ
24
; V
is the sum of the ι-
invariant subspace V
+
of V
Λ
24
with the ι-invariant subspac e V
of the unique 1-twisted
186
module for V
Λ
24
, where ι Aut(Λ
24
) is some involution. The graded dimensions of V
±
are
2
1
(Z(±1, 1) + Z(±1, ι)), respectively, and these sum to J.
The orbifold construction is al so involved in an interesting reformulation of the Hau p t m odul
property, due to Tuite [177]. Assume the following
Conjecture 9.1.2 (Uni qu eness of V
). V
is the only vertex operator algebra with graded
dimension J.
Tuite argues from th i s that, for each g M, T
g
will be a Haup t m odul if and onl y if
the only orbifolds of V
are V
Λ
24
and V
itself. In [104], this analysis is extended to
some of Norton’s series N
(g,h)
, where the subgroup g, h is not cyclic (thus going beyond
[51]), although again assuming the uniqueness conjecture. Recently [24], [25] (and other i s
preparation), Carn a h an has outlined an approach to the generalized mooshine conjecture
by usin g Borcherds’ products.
9.2 Why the Monster?
In the work [189], Zhu introduced a particular algebra with special features that provided
some results about t h e modularity of some functions arisi n g on vertex operator algebras.
The fact that M is associated with modular functions can be explained by it being the au-
tomorphism group of the Moonshine vertex operator algebra V
and the foll owing theorem
Theorem 9.2.1 (Zhu’s Theorem). Suppose V is a C
2
-cofinite weakly rational VOA (see
[189] for the definitions), and let Φ(V ) be the finite set of irreducible V -modules. Then,
there is a representation π of SL
2
(Z) by complex matrices π(A) indexed by V -modules
M, N Φ(V ), such that the one-point functions
χ
M
(z, v) = tr
M
o(v)q
L
0
c/24
= q
c/24
n0
tr
M
h+n
o(v)q
h+n
,
obey
χ
M
az + b
cz + d
, v
= (cz + d)
n
NΦ(V )
π
a b
c d
MN
χ
N
(z, v),
for any v V obeying L
0
v = nv, for some n N.
What is so speci al about this gro u p M that these modular functions T
g
and N
(g,h)
should
be Hauptmod uls? Thi s is stil l open. One approach is due to Norton, a n d was first stated
in [155]: the Monster is probably the largest (in a sense) group with the 6-tran sposition
property. Recall that a k-transposition group G is one generated by a conjugacy class K
of involut i on s , where the product gh of any two elements of K has order at most k. For
example, taking K to be the transpositions in the symmetric group G = S
n
, we find t h a t
S
n
is 3-tr ansposition.
187
A transitive action of Γ = P SL
2
(Z) on a finite set X with one distingui sh ed point x
0
X,
is equivalent to specifying a finite index subgroup Γ
0
of Γ. In particular, Γ
0
is the stab i l i zer
{g Γ : g·x
0
= x
0
} of x
0
, X can be identified with the cosets Γ
0
\Γ, and x
0
with the coset Γ
0
.
To su ch an action, we can associat e an interesting triangu la t io n of the closed surface Γ
0
\
H,
called a (modular) quilt. The definition, origin al l y due to Norton and further developed
by Parker, Co nway, an d Hsu, is somewhat involved and will be avoided h er e (but you can
see [98]) . It is so-named because there is a polygonal ‘patch’ covering every cusp of Γ
0
\H,
and the closed surface is formed by sewin g together the patches along their edges ‘seams’
(Figure 9.1).
Figure 9.1: A ‘friendly’ process of compactifying and sewing a 4-p unctured pl an e .
There are a tot al of 2n triangles and n seams in the triangulation , where n is the index
188
|Γ
0
\H| = |X|. The boun d ar y of each patch has an even number of edge s, namely the
double of the correspondi n g cusp width. The familiar formula
γ =
n
12
n
2
4
n
3
3
n
2
+ 1
for the genus γ of Γ
0
\H in term s of the index n an d the numbers n
i
of Γ
0
-orbits of fixed
points of order i, can be interpreted in terms of the data of the quilt ( see [98]), and we find
in particular that i f every patch of the quilt has at most six sides, then the genus will be 0
or 1, and genus 1 only exceptionally. In particular, we are interested in one class of these
Γ-actions (actuall y an SL
2
(Z)-action). For example, it is well known that the braid g ro u p
B
3
has pr es entation
σ
1
, σ
2
| σ
1
σ
2
σ
1
= σ
2
σ
1
σ
2
,
and center Z = (σ
1
σ
2
σ
1
)
2
(see [5]). It is related to t h e modular group by
B
3
/Z
=
P SL
2
(Z), B
3
/(σ
1
σ
2
σ
1
)
4
=
SL
2
(Z).
Fix a finite group G (we are most interested in the choice G = M). We can define a right
action of B
3
on tri p l es (g
1
, g
2
, g
3
) G
3
by
(g
1
, g
2
, g
3
)σ
1
= (g
1
g
2
g
1
1
, g
1
, g
3
), (g
1
, g
2
, g
3
)σ
2
= (g
1
, g
2
g
3
g
1
2
, g
2
). (9.3)
We will be interested in this action on the subset of G
3
where all g
i
G are involu-
tions. The action (9.3) is equi valent to a reduced version, where we replace (g
1
, g
2
, g
3
) with
(g
1
g
2
, g
2
g
3
) G
2
. Then (9.3) becomes
(g, h)σ
1
= (g, gh), (g, h)σ
2
= (gh
1
, h). (9.4)
These B
3
actions come from specializations of th e Bu ra u a n d r educed Burau representa-
tions respectively [5], [114], and genera li ze to actions of B
n
on G
n
and G
n1
. We can get
an acti on of SL
2
(Z) from the B
3
action (9.4) in two ways: either
(i) by restricting to commuting pairs (g, h); or
(ii) by identifying each pair (g, h) with all its conjugates (aga
1
, aha
1
).
Norton’s SL
2
(Z) action of (9.1) arise s from the B
3
action (9.4) when we perform both (i)
and (i i ) .
The quilt picture was designed for this SL
2
(Z) action. The po int of this construction is
that the number of sid es in each patch is determin ed by the order s of the correspond ing
elements g, h. If G is, say, a 6-tr a n sposition group ( su ch as the Monster), and we take the
involutions g
i
from 2A, then each patch will have 6 sides, and the corresponding genus
will be 0 (usually) or 1 (exceptionally). In this way we can relate the Monst er wi t h a genus
0 property.
Based on the actions (9.3) and (9.4), Norton anticipates some analogue of Moonshine
valid for noncommuting pairs. Alth ou g h they alway s seem to be modular functions, they
189
will no longer always be Haup t m oduls and their fixing groups won’t always contain a
Γ(N). Conformal field theory considerations ( ‘ h i gher genus orbifolds’) alluded in Section
9.1 suggest t h a t more natural should be for example quadruples (g, g
, h, h
) M
4
obeying
[g, h] = [h
, g
].
An important question is, how much does Monstrous Moonshine determine the Monster?
How much of the M structure can be deduced from, for example, McKay’s
ˆ
E
8
Dynkin
diagram observation , and/or the (complete) replicability of the McKay-Thompson series
T
g
, and/or Norton’s conjectures 9.1.1, and/or Modu l ar Moonshine in Section 9.4 below?
A small start toward this is taken in [157], where some control on the subgroups of M
isomorphic to C
p
×C
p
(p prime) was obtained, using only the properties of t he series N
(g,h)
.
For related work, see [98].
9.3 Other finite groups: Mini-Moonshine
It is natural to ask about Moonshine for other grou p s. For example, the Hauptmodul for
Γ
0
(2)
+
looks like
J
Γ
0
(2)
+
(q) = q
1
+ 4, 372q + 96, 256q
2
+ 1, 240, 002q
3
+ . . . (9.5)
and we find the relations 4, 372 = 4, 371+1, 96, 256 = 96, 255+1, 1, 240, 002 = 1, 139, 374+
4, 371 + 2 ·1, where 1, 4371, 96255, and 113 937 4 ar e all d im e n si on s of irreducible represen-
tations of the Baby Monster B. Thus we find ‘Moonshine’ for B.
Of course any subgroup of M automatica l ly inher i t s Moonshin e by restriction, but this
is not at all interesting. A more accessible sporadi c is M
24
(see for example[39]). Most
constructions o f the Leech lattice start with M
24
, and most constructions of th e Monster
involve the Leech lattice. Thus we are led to the following natu ra l hierarchy of (most )
sporadics:
1. M
24
(from wh i ch we can get M
11
, M
12
, M
22
, M
23
); which leads to
2. Co
0
=
2.Co
1
(from wh i ch we get HJ, HS, McL, Suz, Co
3
, Co
2
); which leads to
3. M (from whi ch we get He, Fi
22
, Fi
23
, Fi’
24
, HN, Th, B).
It can thus be argued that we coul d approach problems in Monstrous Moonshine by first
addressing in order M
24
and Co
1
, which should be much simpler. Indeed, the full vertex
operator algebra orbifold t h e or y (the compl et e analogue of Section 9. 1 ) for M
24
has been
established in [52] (the relevant seri es Z
(g,h)
had already been constructed in [ 14 0]). The
orbifold theory for Co
1
though seems out of reach at present. Remarkab l y, that for the
Baby Monst e r B is much m o r e straightforward and has been worked out by ohn [93].
190
Queen [159] established Moonshine for the following grou p s (all essentially centralizers of
elements of M): Co
0
, Th, 3.2. S u z, 2.HJ, HN, 2.A
7
, He, M
12
. In particular , to each ele-
ment g of these gr ou ps, there corresponds a series Q
g
(z) = q
1
+
n0
a
n
(g)q
n
, which is a
Hauptmodul for some modular group of Moon s h i n e- type, and where each g → a
n
(g) is a
virtual character. For example, Q u ee n ’ s series Q
g
for Co
0
is the Hauptmodul (1.10) for the
genus 0 group Γ
0
(2). For Th, HN, He and M
12
it is a proper character. Other differences
with Monstrous Moonshine are that there can be a preferred nonzero value for the constant
term a
0
, and that although Γ
0
(N) will be a subgrou p of th e fixing gr ou p, it will not neces-
sarily be normal. We will return to these results i n the next section, wher e we will see that
many seem to come ou t of the Moonshine for M. About half of. Queen’s Hauptmodul s
Q
g
for Co
0
do not arise as a McKay-Thompson series for M. Norton conjecture 9.1.1 are a
reinterpretation and extension of Queen ’ s work.
Queen never reached B because of its size. However, the Moonshine (9.5) for B falls into
her and Nor t on ’ s scheme because equation (9.5) is the McKay-Thompson series associated
to class 2A of M, and the centralizer of an element in 2A is a double cover of B.
There can not be a vertex operator al geb r a V =
V
n
with graded dimension (9.5) and
automorphisms in B, because for example the B-module V
3
does not contain V
2
as a sub-
module. However, ohn deepened t h e analogy between M and B by constructing a vertex
operator superalgebra V B
of rank c = 23
1
2
, called the shorter Moonshine module, closely
related to V
(see for example [94]). Its automorphism group is C
2
× B. Just as M is the
automorphism group of the Griess algebra B = V
2
, so is B the automorphism group of the
algebra (V B
)
2
. Just as V
is associated to the L eech la t ti ce Λ
24
, so is V B
associated to
the shorter Leech lattice O
23
, the unique 23-dim en si onal positive-definite self-dual lattice
with no vectors of l en g t h 2 or 1 (see for example [39]). The automorphism group of O
23
is C
2
× Co
2
. A si m i l ar theory has recently appeared for Co
1
in [58]. There has been no
interesting Moonshine for the remaining six sporadics (the pariahs J
1
, J
3
, Ru, O’N, Ly, J
4
).
There will be so m e sort of Moonshine for any group which is an automorphism group of
a vertex operator algebra (so this means any finite group [53]). Many finite groups of Lie
type should arise as automorphism groups of vertex operator algebras associated to affine
algebras except defined over fi n i t e fields, but apparently all known examples of genus 0
Moonshine are l im i t ed to the groups involved with M.
Lattices are r el at e d to groups through their automorphism grou p s, which are always finite
for positive-definite lattices. The automorphism grou p Aut Λ
24
= Co
0
of the Leech lattice
has ord er abou t 8 × 10
18
, and is a central exten s io n by Z
2
of Conway’s simple group
Co
1
. Several other sporadic groups are also involved in Co
0
, as we have seen. To each
automorphism α Co
0
, let θ
α
denote the theta function of the sublattice of Λ
24
fixed
by α. Conway and Norton also associate with each automorphism α a certain function
η
α
(z) of the form
i
η(a
i
z) /
j
η(b
j
z) built out of the Dedekind ’ s η function (Example
1.4.2). Both θ
α
and η
α
are constant on each conjugacy class in Co
0
, of which there are
167. [38] remarks that t h e ratio θ
α
α
always seems to equal some McKay-Thompson
191
series T
g(α)
. It turns out that this observat i on is not correct [126] . For each autom or p hism
α Co
0
, the subgroup of SL
2
(R) that fixes θ
α
α
is indeed always genus 0, but for exactly
15 conjugacy classes in Co
0
, θ
α
α
is not the Hau p t m odul. Nevertheless, this construction
proved useful for establishi ng Moonshine for M
24
[140]. Similarly, one can ask this for the
E
8
root lattice, whose automorphism group is the Weyl group of t he Lie algebra E
8
(of
order 696,729,600). The automorphisms of the lattice E
8
that yield a Hauptmodul were
classified in [162]. On t h e other hand, Koike estab l i sh ed a Moonshine of this kind for the
groups P SL
2
(F
7
), P SL
2
(F
5
)
=
Alt
5
and P SL
2
(F
3
), of o r d er 168, 60 and 12, respectively
[119, 1 21 , 120, 122, 123].
9.4 Mod ul ar Moonsh ine
Consider an element g M. We expect from [ 15 5], [159], [51], that there is a Moonsh i n e for
the centralizer C
M
(g) of g in M, governed by the g-twisted module V
(g). Unfortuna t ely,
V
(g) is not usually itself a vertex operator algebra, so the analogy wit h M is not perfect.
Ryba and Borcherds [161], [19], [16] found it interesting that, for g M of prim e order p,
the Norton series N
(g,h)
can be transformed into a McKay-Thompson series (and has all the
associated n i ce properties) whenever h is p-regular (that is, h ha s order coprime to p). This
special behavior of p-regular elements suggested to him to look at modul ar representations.
The basics of modular representations and Brauer characters are discussed in sufficient
detail in [45].
A modular representation π of a group G is a repres entation defined over a field of positive
characteristic p dividing the order |G|. Such representations possess many special features.
For one thing, they are no longer completely redu ci b l e (so the role of irreducible modules
as direct summands will be replaced with their role as compo si t i on factors). For another,
the usual notion of character (the trace of representation matrices) loses its usefulness and
is replaced by the more subt l e Brauer character β( π): a complex-valued class function on
M which is onl y well-defin ed on the p-regular elements of G. We have, for exampl e (see
[16], [19], [161]).
Theorem 9.4.1. Let g M be any element of prime order p, for any p dividing |M|.
Then, there is a vertex operator superalgebra
g
V =
nZ
g
V
n
defined over the finite field F
p
and acted on by the centralizer C
M
(g). If h C
M
(g) is p-regular, then the graded Brauer
character
R
(g,h)
(z) = q
1
nZ
β(
g
V
n
)(h)q
n
equals the McKay-Thompson series T
gh
(z). Moreover, for g belonging to any conjugacy
class in M except 2B, 3B, 5B, 7B, or 13B, this is in fact an ordinary vertex operator algebra
(that is the ‘odd’ part vanishes), while in the remaining cases the graded Brauer characters
of both the odd and e ven parts can separate ly be expressed using McKay-Thompson series.
192
By a vertex operator superalgebra, we m ea n t h er e is a Z
2
-grading into even and odd
subspaces, and for u, v both odd, th e com mutator in the locality axiom of Theorem 4.1.2 is
replaced by an anticommutator. In the proof, the superspaces arise as cohomology groups,
which naturally form an alternating sum. The centralizers C
M
(g) in the theorem ar e quite
nice; for examp le for g in classes 2A, 2B, 3A, 3B, 3C, 5A, 5B, 7A, 11A, r espectively,
these involve the sporadi c groups B, Co
1
, Fi’
24
, Suz, Th, HN, HJ, He, and M
12
. The
proof for p = 2 is not complete at the present time. The conjectures in [161] concerning
modular analogu es of the Griess algebra for several sporadics follow from Theorem 9. 4. 1.
Can these m odular
g
V vertex operator algebras be interpreted as a red u c t io n (mod p)
of (su per)algebras in characteristic 0? Also, what about elements g of composite order?
Borcherds has stated the following in [16]
Conjecture 9.4.2 (Borcherds). Choose any g M and let n denote its order. Then, there
is a
1
n
Z-graded superspace
g
ˆ
V =
i(1/n)Z
g
ˆ
V
i
over the ring of cyclotomic integers Z[e
2πi/n
].
It is often (but probably not always) a vertex operator superalgebra; in particular,
1
ˆ
V is an
integral form of the Moonshine module V
. Each
g
ˆ
V carries a representati on of a central
extension of C
M
(g) by C
n
. Define the graded trace
B
(g,h)
(z) = q
1
i
1
n
Z
tr(h|
g
ˆ
V
i
)q
i
.
If g, h M commute and have coprime orders, then B
(g,h)
(z) = T
gh
(z). If all q-coefficient s
of T
g
are non-negative, then the ‘odd’ part of
g
ˆ
V vanishes, and
g
ˆ
V is the g-twisted module
V
(g). If g has prime order p, then the reduction (mod p) of
g
ˆ
V is the modular vertex
operator superalgebra
g
V of Theorem 9.4.1.
When we say
1
ˆ
V is an integral form for V
, we mean that
1
ˆ
V has the same structure as a
vertex operator algebra, with everything defined over Z, and tensoring it with C recovers
V
. This remarkable conjecture, which tries to explain Theorem 9.4.1, i s completely open.
9.5 The geometry of Moonshine
Algebra is the mathematics of stru c t u r e, and so of course it has a profound relationship
with every area of mathematics. Therefore the trick for findin g possib le fingerpr i nts of
Moonshine in, say, geometry is to look there for modular functions. That sear ch quickly
leads t o the elliptic genus.
For details see for example [9 5 ], [1 64 ] , or [170]. All manifold s here are compact, oriented
and differentiable. In Thom’s cobordism ring Ω, elements are equivalen ce classes of cobor-
dant manifolds, addition is connected sum , and multiplication is Cartesian produ ct . The
universal ell i p t i c genus φ(M) is a ring homomorphism from Q to the ring of power series
in q, which sends n-dimensional manifolds with spin connections to a weight
n
2
modular
193
form of Γ
0
(2) with integer coefficients. Several variations and generalizations have been in-
troduced, e. g., the Witten genus assigns to spin manifolds with vanishing first Pontryagin
class a weight
n
2
modular form of SL
2
(Z) wit h integer coefficients.
We have noticed several deep relationships between elliptic genera Moonshine. For instance,
the i m portant rigidity property of the Witten genus with respect t o any compact Lie group
action on the manifold, is a consequence of the modula ri ty of the characters of affine alge-
bras [135 ]. The elliptic genus of a manifold M has been interpreted as the gr ad ed dimension
of a vertex operat o r superalgebra constructed from M [169]. Seeming ly related to this, [21]
recovered the elliptic genus of a Calab i -Yau manifold X from the sh ea f of vertex algebras
in the chiral de Rham complex MSV[139] at t ached to X. Un e xpectedly, the ellip t i c genus
of even-dimensional projective spaces P
2n
has non - n ega t ive coefficients and in fact equal s
the graded dimension of some vertex al gebra [138]; this suggests interesting representation
theoretic questions in the spirit of Mon st r ous Moonsh ine. In physics, elliptic genera arise
as partition functions of N = 2 superconformal fi el d theories [183]. Mason’s construc-
tions [140] associated to Moo n shine for the Mathieu group M
24
have been interpreted as
providing a geometric model (elliptic system) for elliptic cohomology Ell
(BM24) of the
classifying space of M
24
[170], [54]. Th e Witten genus (norma li zed by η
8
) of the Milnor-
Kervaire manifold M
8
0
, an 8-dimensional manifold buil t from the E
8
diagram, equals j
1/3
[95] ( r eca ll (4.32)).
Another interesting fact is that a Borcherds algebr a can be associated with any even
Lorentzian lattice, and also with any Calabi-Yau manifold [92]. Of course it is a broad
enough class that almost all of them will be uninteresting; an intriguing approach to ide n -
tifying the interesting ones is by considering the so called automorphic product s [14, 15].
Hirzebruch’s ‘prize question’ [95] asks for the construction of a 2 4- d i m en si onal manifold
M with Witten genus J (aft er being normali zed by η
24
). We would like M to act on M
by diffeomorphisms, and the twisted Witten genera to be the McKay-Thompson series T
g
.
It would also be nice to associate Norton series N
(g,h)
to this Moonshine manifold. Con-
structing such a manifold is perhaps the remaining ‘Holy Grail’ of Monstrous Moonshine.
Hirzebruch’s question was partially answered by Mahowal d and Hopkins [137], who con-
structed a manifold with Witten genus J, but could n ot show that it would support an
effective act i on of the Monster. Related work is by Aschbacher [1], who constructed several
actions of M on, for exam ple, 24-dimensional manifolds (but none of which could have
genus J), and Kultze [125 ] , who showed that the graded dimensions of the subspaces V
±
of the Moonshine module are twisted
ˆ
A-genera of the Milnor-Kervaire manifol d M
8
0
(the
ˆ
A-genus is the specialization of elliptic genus to t he cusp i).
There has been a second conjectured relationsh i p between geometry and Monstrous Moon-
shine. Mirror symmetry s ays that most Calabi-Yau manifolds come in closely related
pairs. Consider a 1-parameter famil y X
z
of Calabi-Yau manifolds, with mirror X
given
by the resolution of an orbifold X/G for G finite and abelian. Then the Hodge num-
194
bers h
1,1
(X) and h
2,1
(X
) will be equal, and more precisely the moduli space of (com -
plexified) ahler structures on X will be locally isomet r i c to the moduli space of com-
plex structures on X. The ‘mirror map’ z(q), which can be d efi ned using the Picard-
Fuchs equation [151], gives a canonical map between those moduli spaces. For example,
x1
4
+ x
4
2
+ x
4
3
+ x
4
4
+ z
1/4
x
1
x
2
x
3
x
4
= 0 is such a family of K3 surfaces (that is Calabi-Yau
2-folds), where G = C
4
× C
4
. Its mirror map is given by
z( q) = q 104q
2
+ 6, 444q
3
311, 744q
4
+ 13, 018, 830q
5
493, 025, 760q
6
+ . . . . (9.6)
Lian and Yau [133] noticed th at the reciprocal
1
z(q)
of the mirror map in (9.6) equals the
McKay-Thompson series T
g
(z) + 104 for g in class 2A of M. After looking at several other
examples with simi la r conclusions, they proposed their Mirror-Moonshine conjecture: The
reciprocal 1/z of the mirror map z of a 1-parameter family of K3 surfaces with an orbifold
mirror will be a McKay-Thompson series (up to an addit i ve constant).
A counterexample (and more examples) are given in [180]. In particu l ar , although there are
relations between mirror symmetry and modular functions (see for exam p l e [89] and [92]),
there does not seem to be any special relation with the Monster. Doran [56] demystifies
the Mirror-Moonshine ph enomenon by finding n eces sar y and sufficient cond i t io n s for 1/z
to be a modular function for a modular group commensurable with SL
2
(Z).
9.6 Moo ns hi ne and physics
The physical side (perturbative string theory, or equivalently conformal field theory) of
Moonshine was noticed early on, and has pr ofoundly i n fl u e n ced the development of Moon-
shine an d vertex operator algebras. This is a very rich subject , which we can only super-
ficially touch on. The book [64], with its ext en si ve bibliography, provides an introductio n
but will be difficult rea d i n g for many mathema t ic ia n s (as will this section). The treatme nt
in [70] is m o r e accessib le and shows how naturally vertex operator algebras arise from the
physics. This effectiveness of physical interpretations is not magic; it m er el y tell s us that
many of our finite-dimensi on a l objects are seen much more clearly when studied thr ough
infinite-dimensional structures. Of course Moonshine, which teaches us t o stud y the finit e
group M via its infinite-dimension a l module V
, fits perfectly into this picture.
A confor mal field theory i s a quantum fiel d theory on 2-dimensional space-time, whose sym-
metries incl u de the conformal transformations. In string theory the basi c objects are finite
curves —called strings rather than points (particles), and the conformal field theory lives
on the surface traced by the strings as they evolve (colliding and separating ) through time.
Each conformal fiel d theory is associated with a pair V
L
, V
R
of mutually commuting vertex
operator a lg eb r as , called its chiral algebras [4]. For example, strings living on a compact
Lie group manifold (the so-called Wess-Zumino-Witten model) will have chiral algebras
given by affine algebra vertex operator algebras. The space H of stat es for the conformal
field the or y carries a representation of V
L
V
R
, and many authors have (somewhat op t i -
mistically) concluded that the study of conformal field theories redu ces to that of vertex
195
operator algebra r epresentation theory. Rational vertex operator algebras correspond to
the important cl ass of rational conformal field theories, where H decomposes into a finite
sum M
L
M
R
of irreducible modules. The Virasoro algebra of S ect i on (4.2) arises natu-
rally in conformal field theory t h r ough infinitesimal conformal transformations. Th e vertex
operator Y (φ, z), for the space-time parameter z = e
t+ix
, is the quantum field which cre-
ates from the vacuum |0 H the state |φ H at time t = −∞ : |φ = lim
z0
Y (φ, z)|0.
In particular, Borcherds’ definition [8, 13] of vertex operator algebras can be interpreted
as an ax io m at i sat i on of the notion of chiral algebra in conform al field theory, and for this
reason a lo n e is important.
In conformal fiel d theory, t he Hauptmodul property of Moonshine is hard to interpret, and
a less direct formulation, li ke that in [177], is needed. However, both the statement and
proof of Zhu’s Theorem are natural from the conformal field theory framework (see [70]);
for example, the modularity of the series T
g
and N
(g,h)
are automatic in conformal field
theory. This modularity arises in conformal field t h e or y through the equivalen ce of the
Hamiltonian formulation, which d escr i bes concretely the graded spaces we take traces on
(and hence the coefficients of our q-expansions), and the Feynman path formalism, which
interprets these graded traces as sections over moduli spaces (and hence makes modularity
manifest). Beautiful revi ews are sketched in [184], [183]. More explicitly, the Virasoro
action on moduli spaces discussed in section 4.2 of [71] gives rise to a syst em of partial dif-
ferential eq u at i ons (the Knizhnik-Zamolodchikov equat io n s) . According to conformal field
theory, the vertex operator algebra characters will satisfy those eq u at i on s for a torus with
one puncture, an d their modularity (that is, Zhu’s Theorem) arises from the monodromy
of those equations.
Because V
is so mathematically special, it may be expected that it corresponds to in-
teresting physics. Certainly it has bee n the subject of some speculati on. There wil l be a
c = 24 rational conformal field theor y whose chiral algebra V
L
and state space H are both
V
, while V
R
is trivial ( t his is possible because V
is holomorphic). This conform al field
theory is nicely d escr i bed in [48]; see also [50]. The Monster is the symmet r y of that con-
formal fi e ld theory, but the Bimonster M C
2
will be the symmetry of a ra t io n al conformal
field theory with H = V
V
. The paper [41] finds a fami l y of D-branes for the latter
theory which are in one-to-one correspondence with the elements of M, and their ‘overlaps’
g||q
1
2
(L
0
+
L
0
+c/24)
||h equal the McKay-Thompson series T
g
1
h
. However, we still lack any
explanation a s to why a conformal field theory involving V
should yield interesting physics.
Almost every facet of Moonshine finds a natural formulation in conform al field theory,
where it often was discovered first . For example, the no-ghost theorem (Theorem 8.3.2)
of Brower-Goddard-Thorn was used to great effect in [12] to understand the structu r e of
the Monster Lie algebra M. On a fin i t e- d i m en si onal manifold M, the index of the Di r ac
operator D in the hea t kernel interpretation is a path integral in supersymmetric quantum
mechanics, that is, an integral over the free loop sp ace LM = {γ : S
1
M}; the string
196
theory version of this is that the in dex of the Dirac operator on LM should be an integral
over L(LM), that is over s m ooth maps of tori into M, and this is just th e elliptic genus,
and explains why it should be modular. The orbifold construction of [51] comes straight
from conformal field theory (although const r uction of V
in [66] predates conformal field
theory orbifolds by a year and in fact influen ced their development in physics). That said,
the translation process from physics to mathematics of cours e is never easy; Borcherds’
definition [8] i s a prime example.
From this standpoint, what is most excit ing is what has not yet been fully exploited. String
theory tells us t h a t conformal field th eor y can live on any su r face Σ. The vertex oper at or
algebras, including the geometri c vertex operator algebras of [99], capture conformal field
theory in genus 0. Th e gr a d ed dimensions and traces considered above co n cer n conformal
field theory quantities (conformal blocks) at genus 1: z → e
2
πiz maps H onto a cylinder,
and the tr ace identifies the two ends. There are analogues of all this at higher genus [188 ]
(though the formulas can rapidly become awkward). For example, the graded dimension
of the V
conformal field theory in genus 2 is computed i n [178], and involves for instance
Siegel theta functions. The orbifold theory in Section 9.1 is genus 1: each ‘sector’ (g, h)
corresponds to a homomorphism from the fundamental group Z
2
of the tor u s into the orb-
ifold group G (for example G = M); g and h are t h e targets of the two generator s of Z
2
and
hence must commute. More generally, the sectors will correspond to each homomorph i sm
ϕ : π
1
(Σ) G, and to each we will get a high er genus trace Z(ϕ), wh i ch will be a fu n ct i on
on the Teichm¨uller space T
g
(generalizing the up per half-plane H for genus 1) . The action
of SL
2
(Z) on the N
(g,h)
generalizes to the action of the mapping class group on π
1
(Σ) and
T
g
. See for examp le [6] for s om e thoughts in this direction . Recently, Witten [186] has also
related the Monster with 3- d i m en si onal gravity in black holes, although the paper is still
abundant i n conjectures.
9.7 Conclusion
There are different basic aspects to Mon st r ou s Moonshine: (i) why mod u l ar i ty enters at all;
(ii) why in particular we have genus 0; and (iii) what it has t o do with the Monster. Today,
we understand (i) best. There will be a Moonshine-like relations between any (subgroup of
the) automorphism group of any ratio n al vertex operator algebra, and th e characters χ
M
,
and the same can be expected to hold of the orbifold characters Z in Section 9.1.
To prove the genus 0 property of the McKay-Thompson series T
g
, we needed recursions
obtained one way or another from the Monster L i e algebra M, and from these we apply
Theorem 7.4.4. These r ecu r si o n s are very special , but so presumably is the 0 genus prop-
erty. The suggestion of [28], though, is th at we may be able to considerably simplify this
part of the argument.
Every group known to have ri ch Moonshine properties is contained in the Monster. To what
extent can we deri ve M from Monstrous Moonshine? The understanding of this seeming l y
central r ol e of M is the poorest of those three aspects.
197
The central role that vertex operator algebras play in our current understanding of Moon-
shine should be clear from this review (that is basically Zhu’s Theorem). The excellent
review [54] makes this point even m or e forceful l y. However, it can be (and has been) ques-
tioned whether the full and diffi cu l t machinery of vertex operator algebras is really needed
to understand this; that is, whether we really have isola t ed the key conjunction of prop-
erties needed for Moonshin e to arise. Conformal field theory has been an invaluable guide
thus far, but perhaps we are a little too steeped in its lore.
In particular, what it is really needed is a second in d e pendent proof of the Moonshine con-
jectures. One t em p t i n g possi b i l i ty is the heat kernel; its general role in modularity concerns
is emphasized in [109] , and is also th e central ingredient in the Knizhnik-Zamolo dchikov
equations for affine algebras [96]. A heat kern el probably plays an analogous role in the
Knizhnik-Zamolodchikov equations associated to V
, but does it relate to the genus 0
property? Another pos si b i li ty is the braid group B
3
, whose fingerprints are all over the
mathematics and physics of Moonshine.
Moonshine (in its more general sense) is a relation between algebra and number theory,
and its impact on algebra ha s been dramatic (for example: vertex operator algeb r a s, V
,
Borcherds-Kac-Moody algebras). Its impact on numb er t h eo r y has been far less so. This
may merely be a temporary acci d ent due to the backgrounds of m os t researchers (i n cl u d i ng
the mathematical physicists) working to date in the area. Gannon [71] has suggested t h a t
the most exciting prospects for the future of Moonshine ar e in the direction of number
theory. Hints of this fut u r e can be found in for example [15], [47], [57], [90], [150], [185]
and [18]. Other recent advances and discussions are [14 4 ] , [63], [37], [146], and [143].
198
References
[1] M. G. Aschbacher. Finite groups acting on homology manifolds. Pa cific J. Math., Special Issue:3–36,
1997.
[2] A. L. Atkin and J. N. L’Brien. On divis ib il ity properties of the coefficients of the j-function. Trans.
Amer. Math. Soc., 126, 1967.
[3] A. L. Atkin and J. Lehner. Hecke operators on Γ
0
(m). Math. Ann., 185:134–160, 1970.
[4] A. Belavin, A. Polyakov, and A. Zamolodchikov. Infinite conformal symmetries in two dimensional
quantum field theory. Nucl . Phys., B241:333–380, 1984.
[5] J. S. Birman. Braids, links and mapping class groups. Princ et on University Press, Princeton, 1974.
[6] P. antay. Higher genus moonshine. Moonshine, the Monster, and related topics. Contemp. Math.,
193:1–8, 1996.
[7] R. E. Borcherds. Vertex algebras. Preprint.
[8] R. E. Borcherds. Vertex algebras, Kac-Moody algebras and the Monster. Proc. Natl. Acad. Sci.
USA, 83:3068–3071, 1986.
[9] R. E. Borcherds. Generalized Kac-Moody algebras. J. Algebra, 115:501–512, 1988.
[10] R. E. Borcherds. The monster Lie algebra. Adv. Math., 83:30–47, 1990.
[11] R. E. Borcherds. Central extensions of generalized Kac-Moody algebras. J. Algebra, 140:330–335,
1991.
[12] R. E. Bor cherds. Monstrous moonshine and monstrous Lie superalgeb r as. Invent. Math., 109:405–
444, 1992.
[13] R. E. Borcherds. Sporadic groups and string theory. In First European Congress of Math. (Paris,
1992), I, pages 411–421, Base l, 1994. Birkh¨auser.
[14] R. E. Borcherds. Automorphic forms on O
(s+2,2)
(R) and generalized Kac-Moody algebras. In Pro-
ceedings of the International Congress of Mathematicians (Z¨urich, 1994), pages 744–752, Basel, 1995.
Birkh¨auser.
[15] R. E. Borcherds. Automorphic forms with singularities on Grassmannians. Invent. Math., 132:491–
562, 1998.
[16] R. E. Borcherds. Modular moonshine III. Duke Math. J., 93:129–154, 1998.
[17] R. E. Borch er d s . What is moonshine? In Proceedings of the International Co ngress of Mathematicians
(Berlin, 1998), pages 607–615, Bielefeld, 1998. Documenta Mathematica.
[18] R. E. Borcherds. Problems in Moonshine. In First International Congress of Chinese Mathematics
(Beijing, 1998), pages 3–10, Providence, 2001. American Mathematical Society.
[19] R. E. Borcherds and A. J. E. Ryba. Modular mo ons h in e II. Duke Math. J., 83:435–459, 1996.
[20] A. Borel. Seminar on Complex Multiplication. Lecture Notes in M ath emat ics , 21. Springer-Verlag,
1966.
[21] L. A. Borisov and A. Libgober. Elliptic genera of toric varieties and appl icati ons to mirror symmetry.
Invent. Math., 140:453–485, 2000.
199
[22] T. Br¨ocker and T. Dieck. Representations of Compact Lie groups. Springer-Ver lag, Berlin, 1985.
[23] W. Burnside. Theory of groups of finite order. Cambridge University Press, Cambridge, 1911.
[24] S. Carnahan. Generalized Moonshine I: genus zer o functions. arXiv:0812.3440v2, 2008.
[25] S. Carnahan. Generalized Moonshine II: Borcherds pr oducts. arXiv:0908.4223v2, 2009.
[26] A. A. Castro. Curso de equa¸oes diferenciais ordin´arias. preprint, Rio de Janeiro, 2009.
[27] C. Chevalley. Sur certains groupes simple s. Tohoku Math. J., 7:14–66, 1955.
[28] H. Cohn and J. McKay. Spontaneous gener ati on of modular invariants. Ma th. Comp., 65:1295–1309,
1996.
[29] J. H. Conway. A perfect group of order 8,315,553,613,086,720,000 and the sporadic simple groups .
Proc. Natl. Acad. Sci. USA, 61:398–400, 1968.
[30] J. H. Conway. A characterization of Leech’s lattice. Invent. Mat h., 7:137–142, 1969.
[31] J. H. Conway. A group of order 8,315,553,613,086,720,000. Bull. London Math. Soc., 1:79–88, 1969.
[32] J. H. Conway. Three lectures on exceptional groups. In Finite Simple Groups, pages 215–247, 1971.
[33] J. H. Conway. The miracle octad generator. In Topics in group theory and computation, pages 62–68,
1977.
[34] J. H. Conway. A simple construction for the Fischer-Grie ss monster group. Invent. Math., 79:513–540,
1985.
[35] J. H. Conway. The monster group and its 196884-dimensional sp ace. In Sphere Packings, Lattices
and Groups, pages 555–567, 1988.
[36] J. H. Conway, R. T. Curtis, S . P. Norton, and et. al. Atlas of finite groups. Clarendon Pre s s, Oxford,
1985.
[37] J. H. Conway, J. McKay, and A. Sebbar. On the discrete groups of Mo ons h ine . Proc. Amer. Math.
Soc., 132:2233–2240, 2004.
[38] J. H. Conway and S. P. Norton. Monstrous Moonshine. Bull. London Math. Soc., 11:308–339, 1979.
[39] J. H. Conway and N. J. A. Sloane. Sphere Packings, Lattices and Groups. Springer-Verlag, New
York, 1988.
[40] A. Coste and T. Gannon. Remarks on Galois in rational conformal field theories. Phys. Lett.,
B323:316–321, 1994.
[41] B. Craps, M. R. Gaberdiel, and J. A. Harvey. Monstrous branes. Comm. Math. Phys., 234:229–251,
2003.
[42] C. J. Cummins. Congr u en ce subgr oup s of groups commensurable with P SL
2
(Z) of genus 0 and 1.
Experiment. Math., 13:361–382, 2004.
[43] C. J . Cummins and T. Gannon. Modular equations and the genus zero pr operty of moonshine
functions. Invent. Math., 129:413–443, 1997.
[44] C. J. Cummins and S. P. Norton. R ational Hauptmodul are replicable. Canad. J. Math., 47:1201–
1218, 1995.
[45] C. W. Curtis and I. Reiner. Methods of representation theory with applications to finite groups and
orders, I. Wiley, New York, 1981.
200
[46] R. T. Curtis. A new combintori al approach to M
24
. Math. Proc. Cambridge Philos. Soc., 79:25–42,
1976.
[47] R. Dijkgraaf. The mathematics of fivebranes. In Proc. Intern. Congr. M ath . (Berlin, 1998), III,
pages 133–142, Bielefeld, 1998. Documenta Mathematica.
[48] L. Dixon, P. Ginsparg, and J. A. Harvey. Beauty and the beast: Superc onfor mal symmetry in a
monster module. Comm. Math. Phys., 119:221–241, 1988.
[49] L. Dixon, J. A. Harvey, C. Vafa, and E. Witten. Strings on orbifolds. Nucl. Phys., B261:678–686,
1985.
[50] L. Dolan, P. Goddard, and P. Montague. Conformal fi el d theory of twisted vertex operators. Nucl.
Phys., B338:529–601, 1990.
[51] C. Dong, H. Li, and G. M ason . Mo d ul ar invariance of trace functions in orbifold theory and gener-
alized moonshine. Comm. Math. Phys., 214:1–56, 2000.
[52] C. Dong and G. Mason. An orbifold theor y of genus zero associated to the sporadic group M
24
.
Comm. Math. Phys., 164:87–104, 1994.
[53] C. Dong and G. Mason. Nonabelian orbifolds and the boson-fermion correspondence. Comm. Math.
Phys., 163:523–559, 1994.
[54] C. Dong and G. Mason. Vertex operator algebras and moonshine: A survey. Progress in Algebraic
Combinatorics. Adv. Stud. Pure Math., 24:101–136, 1996.
[55] C. Dong and F. Xu. Conformal nets associated with lattices and their orbifolds. ar Xi v:ma th. Q A/0 41.
[56] C. F. Doran. Picard-Fuchs uniformization and modularity of the mirror map. Comm. Math. Phys.,
212:625–647, 2000.
[57] V. G. Drinfeld. On quasitriangular quas i-Hopf algebras on a group that is closely related with
Gal Q/Q. Leningrad. Math. J., 2:829–860, 1991.
[58] J. F. Duncan. Super-Moonshine for Conway’s largest sporad ic group. arXiv:math.RT/0502267.
[59] K. Erdmann and M. J. Wildon. Introduction to Lie algebras. Springer-Verlag, London, 2006.
[60] G. Faber.
¨
Uber polynomische Entwicklungen. Math. Ann., 57:389–408, 1903.
[61] W. Feit and J. Thompson. Solvability of groups of o dd order. Pacific J. Math., 13:775–1029, 1963.
[62] B. Fischer, D. Livingstone, and M. P. Thorne. The characters of the mons t er simple group. Birm-
ingham, 1978.
[63] D. J. Ford and J. McKay. Monstrous Moonshine - Problems Arising I, Tate Characters. preprint,
2002.
[64] P. Di Francesco, P. Mathieu, and D. S´en´echal. Conformal field theory. Springer, New York, 1997.
[65] I. Frenkel. Representations of K ac-M oody algebras and dual r e sonan ce models. Lect. Appl. Math.
AMS, 21:325–353, 1985.
[66] I. Frenkel, J. Lepowsky, and A. M eu r man. A natur al representation of the Fischer-Griess monster
with the modular function J as character. Proc. Natl. Acad. Sci. USA, 81:3256–3260, 1984.
[67] I. Frenkel, J. Lepowsky, and A. Meurman. Vertex Operator Algebras and the Monster. Academic
Press, New York, 1988.
[68] R. Fricke. Die Elliptis che Funktionen un Ihre Anwendungen. Teubner, Leipzig, 1922.
201
[69] L. Fuchs.
¨
Uber eine Klasse von Funktionen nehrerer Variabeln welche durch Umkehrun der Integrale
von osungen der linearen Differenetialgleichungen mit Rationalen Coefficienten entstehen. J. Reine
Angew., 89:151–169, 1880.
[70] M. R. Gaberdiel and P. Goddard. Axiomatic conformal field theory. Comm. Math. Phys., 209:549–
594, 2000.
[71] T. Gannon. Monstrous Moonshine: the first twenty-five years. Bull. London Math. Soc., 38:1–33,
2006.
[72] T. Gannon. Moonshine Beyond the Monster. Cambridge University Press, Cambridge, 2006.
[73] H. Garland and J. Lepowsky. Lie algebra homology and the Macdonald-Kac fornulas. Invent. Math.,
34:37–76, 1976.
[74] G. Glauberman and S. P. Norton. On McKay’s connection between the affine E
8
diagram and the
Monster. In Proc. on Moonshine and related topics, pages 37–42, 2001.
[75] P. Goddard and D . Olive. Kac-Moody and Virasoro algebras: a reprint volume for physicists. Scien-
tific World, Singapore, 1988.
[76] P. Goddard and C. B. Thorn. Compatibility of the dual Pomeron with unitary and the absence of
ghosts in the dual resonance model. Phys. Lett. B, 40:235–238, 1972.
[77] D. Gorenstein. Finite s im ple groups: an introduction to their classification. Plenum Press, New
York, 1982.
[78] R. L. Griess. The structure of the ‘Monster’ simple group. In Proc. of the Conference on Finite
Groups, pages 113–118, New York, 1976. Acad. Press.
[79] R. L. Griess. A construction of F
1
as automorphisms of a 196,993 d imen s ional algebra. Proc. Natl.
Acad. Sci., 78:689–691, 1981.
[80] R. L. Griess. The friendly giant. Invent. Math., 69:1–102, 1982.
[81] R. L. Griess. Code Loops. J. of Algebra, 100:224–234, 1986.
[82] R. L. Griess. The Monster and its nonassociative algebra. Contemp. Math., 45:121–157, 1986.
[83] R. L. Griess. Sp or adic groups, code loops and nonvanishing cohomology. J. of pure and appl. Algebra,
44:191–214, 1987.
[84] R. L. Griess. A M oufang loop, the exceptional Jordan algebra and a cubic form in 27 variables. J.
of Algebra, 131:281–293, 1990.
[85] R. L. Griess. Codes, loops and p-locals. In Proc. of the Monster Bach Conference, Columbus, 1993.
Ohio U.
[86] R. L. Griess. Twelve sporadic groups. Springer-Verlag, Berlin, 1998.
[87] R. L. G r ie ss , U. Meierfrankenfeld, and Y. Segev. A uniqueness proof for the Monster. Annals of
Math., 130:567–602, 1989.
[88] P. A. Grillet. Abtract Algebra. Springer-Verlag, GTM 242 series, 2007.
[89] V. A. Gritsenko and V. V. Ni ku lin . The arithmetic mir r or symmetry and Calabi-Yau manifolds.
Comm. Math. Phys., 210:1–11, 2000.
[90] S. Gu kov and C. Vafa. Rational conformal field theories and complex multiplication. Comm. Math.
Phys., 246:181–210, 2004.
[91] B. C. Hall. Lie groups, Lie algebras and representations. Springer-Verlag, GTM 222 series, 2003.
202
[92] J. A. Harvey and G. Moore. Algebras, BPS states, and strings. Nucl. Phys., B463:315–368, 1996.
[93] G. ohn. Generalized moonshine for the Baby Monster. Prepint.
[94] G. ohn. The group of symmetries of the shorter Moonshine module. arXiv:math.QA/0210076.
[95] F. Hirzebruch, T. Berger, and R. Jung. Manifolds and modular fo rm s . Friedr. Vieweg und Sohn,
Berlin, 1991.
[96] N. Hitchin. Flat connections and geometric quantization. Comm. Math. Phys., 131:347–380, 1990.
[97] K. Hoffman and R. Kunze. Linear Algebra. Prentice-Hall, 1971.
[98] T. Hsu. Quilts: Central extensions, braid actions, and finite groups. Lecture Notes in Mathematics,
1731, 2000.
[99] Y.-Z. Huang. Two-dimensional conformal geometr y and vertex operator algebras. Birkh¨auser, Boston,
1997.
[100] J. Humphreys. Introduction to Lie al gebras and representation theory. Springer-Verlag, GTM 9
series, 1972.
[101] K. Ireland and M . Rosen. A Classical introduction to Modern Number Theory. Springer-Verlag,
GTM 84 series, 1982.
[102] A. A. Ivanov. Geometric presentations of groups with an application to the Monster. In Proc. Intern.
Congr. Math. (Kyoto, 1 990 ) , pages 1443–1453, Hong Kong, 1990. Springer.
[103] A. A. Ivanov. Y-groups via transitive extension. J. Algebra, 218:412–435, 1999.
[104] R. Ivanov and M. P. Tuite. Some irrational generalized Moonshine from orb if olds . Nucl. Phys .,
B635:473–491, 2002.
[105] J. McKay J. H. Conway and A. Sebbar. On the discrete groups of moonshine. Proc. Amer. Math.
Soc., 132:2233–2240, 2004.
[106] C. G. Jacobi. Fundamenta nova theoriae functionum ellipticarum. onigsberg, 1829.
[107] N. Jacobson. Basic Algebra I. W. H. Freeman and Company, San Francisco, 1974.
[108] N. Jacobson. Basic Algebra II. W. H. Freeman and Company, San Francisco, 1980.
[109] J. Jorgenson and S . Lang. The ubiquitous heat kernel. In Mathematics unlimited - 2001 and beyond,
pages 655–683, Berlin, 2001. Springer-Verlag.
[110] V. Kac. Simple irreducib le graded Lie algebras of finite growth. Math. USSR-Iz v., 2:1271–1311, 1968.
[111] V. Kac. An elucidation of: infinite-dimensional algebras, Dedekind’s η-function, classical obius
function and the very strange formula E
(1)
8
and the cube root of the modular invariant j. Adv.
Math., 35:264–273, 1980.
[112] V. Kac. Infinite dimensional Lie algebras. Cambridge U. Press, Cambridge, 1991.
[113] V. Kac. Vertex algebras for Beginners. American Mathematial Society, University Lecture Series 10,
Providence, 1997.
[114] C. Kassel and V. Turaev. Braid Groups . Springer-Verlag, G TM 247 series, 2008.
[115] F. Kle in. Zur Theorie der elliptischen Modulfunktionen. Gesammente mathematische Abhandlungen,
3:13–75, 1923.
[116] A. W. Knapp. Lie Groups: Beyond an introduction. Birkhauser, Berlin, 1996.
203
[117] N. Koblitz. Introduction to Elliptic Curves and Modular Forms. Springer-Verlag, GTM 97 series,
1984.
[118] M. Koike. On replication formulae and Hecke operators. Nagoya University, preprint.
[119] M. Koike. On McKay’s conjecture. Nagoya Math. J., 95:85–89, 1984.
[120] M. Koike. Mathieu group M
24
and modular forms. Nagoya Math. J., 99:147–157, 1985.
[121] M. Koike. Moonshine for P SL
2
(F
7
). Adv. Stud. Pure Math., 7:103–111, 1985.
[122] M. Koike. Moonshines of P SL
2
(F
q
) and the automorphism group of Leech lattice. Japan J. M ath .,
12:283–323, 1986.
[123] M. Koike. Modular forms and the automorphism group of Leech lattice. Nagoya Math. J., 112:63–79,
1988.
[124] D. N. Kozlov. On completely replicable funct ions and extremal poset theory. MSc thesis, University
of Lund, Sweden, 1994.
[125] R. Kultze. Elliptic genera and the moonshine module. Math. Z., 223:463–471, 1996.
[126] M. L. Lang. On a ques ti on raised by Conway-Norton. J. M a th. Soc. Japan, 41:263–284, 1989.
[127] S. Lang. Algebra. Benjamin/Cummings, California, 1984.
[128] J. M. Lee. Introduction to Topological Manifolds. Springer-Verlag, GTM 202 series, 2000.
[129] J. M. Lee. Introduction to Smooth Manifolds. Springer-Verlag, GTM 218 series, 2003.
[130] J. Leech. Some sphere packings in higher space. Canadian J. Math, 16:657–682, 1964.
[131] J. Leech. Notes on sphere packings. Canadian J. Math., 19:251–267, 1967.
[132] J. Lepowsky. Euclidean Lie algebras and the modular function j. In The Santa Cruz Conference
on Finite Groups (Santa Cruz, 1979), Proc. Sympos. Pure Math., pages 567–570, Providence, 1980.
American M athe matic al Society.
[133] B. H. Lian and S. T. Yau. Arithmetic properties of mirror map and quantum coupli ng. Comm.
Math. Phys., 176:163–191, 1996.
[134] Elon Lages Lima. Variedades Diferenci´aveis. Publica¸oes matem´aticas, Impa, Rio de J ane ir o, 2008.
[135] K. Liu. On modular invariance and rigidity theorems. J. Differential Geom., 41:343–396, 1995.
[136] F. J. MacW i lli ams and N. J. A. Sloane. The theory of error correcting codes. North-Holland, New
York, 1977.
[137] M. Mahowald and M. Hopkins. The structure of 24 dime ns ional manifolds having normal bundles
which lift to BO[8]. Recent Progres s in Homotopy Theory. Contemp. Math., 293:89–110, 2002.
[138] F. Malikov and V. Schechtman. Deformations of vertex algebr as, quantum cohomology of toric
varieties, and elliptic genus. Comm. Math. Ph ys ., 234:77–100, 2003.
[139] F. Malikov, V. Sch echtman, and A. Vaintrob. Chiral de Rham complex. Comm. Math. Phys.,
204:439–473, 1999.
[140] G. Mason. On a system of elliptic modular forms attached to the large Mathieu group. Nagoya
Math. J., 118, 1990.
[141] E. Mathieu. M´emoire sur l’´etu de des functions de plusieures quantit´es, sur la mani`ere de les formes
et sur les substitutions qui les laissent invariables. Crelle J., 6:241–323, 1861.
204
[142] J. McKay. Grou ps , singularities and finite groups. In The Santa Cruz conferenceon Finite Groups,
pages 183–186, 1980.
[143] J. McKay, J. P. Harnad, and P. Winternitz. Groups and Symmetries. From Neol it hic Scots to John
McKay. CRM Proceedings and Lecture Notes, 47, Am. Math. Soc. Providence, 2009.
[144] J. McKay and A. Sebbar. Proceedings on Moonshine and Related Topics. CRM Proceedings and
Lecture Note s , 30, Providence, 2001.
[145] J. McKay and A. Sebbar. Replicable functions: An introduction. In Frontiers in Number Theory,
Physics, and Geometry, II, pages 373–386, Berlin, 2007. Springer-Ver lag.
[146] J. McKay and D. Sevilla. Aplicaci´on de la de s composici´on racional univariada a monstrous moon-
shine. arXiv:0805.2311, 2008.
[147] J. McKay and D. Sevilla. Decomposing re pl icab le functions. LMS J. Comput. Math., 11:146–171,
2008.
[148] J. McKay and H. Strauss. The q-se r ies of monstrous moonshine and the decomposition of the head
characters. Com mu n. Alg., 18:253–278, 1990.
[149] R. V. Moody. Lie algebras associated with generalized Cartan matrices. Bull. Amer. Math. Soc.,
73:217–221, 1968.
[150] G. W. Moore. Les Houches lectures on strings and arithmetic. ar Xi v: hepth/0401049.
[151] D. R. Morrison. Picard-F u chs equations and mirror maps for hypersurfaces. In Essays on mirror
manifolds, pages 1241–264, Hong Kong, 1992. Intern. Press.
[152] H. Movasati. On differential modu lar forms and some analytic relations between Eisenstein series.
Ramanujan J., 17:53–76, 2008.
[153] H. V. Niemeier. Definite quadratische Formen der Dimension 24 und Diskr imi nante 1. J. Number
Theory, 5:142–178, 1973.
[154] S. P. Norton. More on moons hi ne . In Computational group theory, pages 185–193, Cambridge, 1984.
Cambridge University Press.
[155] S. P. Norton. Generalized moonshine. Arcata Conference on Representations of Finite Groups . Proc.
Sympos. Pure Math., 47:208–209, 1987.
[156] S. P. Norton. Constructing the Monster. In Groups, Combinatorics and Geometry, pages 63–76,
Cambridge, 1992. Cambridge U. Press.
[157] S. P. Norton. From moonshine to the Monster. Proc. on Moonshine and related topics, Amer. Math.
Soc, Providence:163–171, 2001.
[158] A. P. Ogg. Automorphismes des cour bes modulaires. Seminaire Delange-Pisot-Poi tou , 16e ann´ee,
7, 1975.
[159] L. Queen. Modular functions arising from some finite groups. Math. Comp., 37:547–580, 1981.
[160] M. Ronan. Symmetry and the Monster. Oxford University Press, Oxfor d, 2006.
[161] A. J. E. Ryba. Modular moonshine? M oonshine, t he Monster, and related topics. Contemp. Math.,
193:307–336, 1996.
[162] M. L. Lang S. P. Chan and C. H. Lim. Some modular fun ct ions asso c iate d to Lie algebra E
8
. Math.
Z., 211:223–246, 1992.
[163] L. Schwartz. Th´eorie des distributions. Hermann, Paris, 1973.
205
[164] G. Segal. Elliptic cohomology. Seminaire Bourbaki 1987-88 no. 695, 161-162:187–201, 1988.
[165] J.-P. Serre. A course in arithmetic. Springer-Verlag, GTM 7 ser ie s, 1973.
[166] J.-P. Serre. Linear representations of finite groups. Springer-Verlag, GTM 42 series, 1977.
[167] J. H. Silverman. The Arithmetic of Elliptic Curves. Springer-Verlag, GTM 106 series, 1986.
[168] J. Sotomayor. Li¸oes de equa¸oes diferenciais ordin´arias. Pr ojeto Euclides, Impa, Rio de Janeiro,
1979.
[169] H. Tamanoi. Elliptic genera and vertex operator superalgebr as. Lecture Notes in Mathematics, 1704,
1999.
[170] C. B. Thomas. Elliptic cohomology. Kluwer, New York, 1999.
[171] J. Thompson. Finite groups and mod u lar functions. Bull. London Math. Soc., 11:347–351, 1979.
[172] J. Thompson. Some numerology between the Fischer-Griess monster and modular functions. Bull.
London Math. Soc., 11:352–353, 1979.
[173] J. G. Thompson. A finiteness theorem for subgroups of P SL
2
(R) which are commens u r able with
P SL
2
(Z). In Santa Cruz Conference on F inite Groups, pages 533–555. Proc. Sympos. Pure Math.
37, 1980.
[174] J. Tits. esum´e de cours. Annuaire du C oll`ege de France, 1982-1983, pages 89–102, 1983.
[175] J. Tits. On R. Griess’ friendly giant. Invent. Math., 78:491–499, 1984.
[176] J. Tits. Le monstre, Seminaire Bourbaki, expos´e no. 620. Ast´eris que, 121-122:105–122, 1985.
[177] M. P. Tuite. On the relationship between Monstrous Moon sh in e and t he u ni qu en es s of the Moonshine
module. Comm. Math. Phys ., 166:495–532, 1995.
[178] M. P. Tuite. Genus two meromorphic conformal field theories. In Proc. on Moonshine and related
topics, pages 231–251, Providence, 2001. Amer. Math. Soc.
[179] S. Varadarajan. Lie groups, Lie algebras and their representations. Springer-Verlag, GTM 102 series,
1974.
[180] H. Verrill and N. Yui. Thompson series and the mirror maps of pencils of K3 surfaces. In The
arithmetic and geometry of algebraic cycles, pages 399–432, Paris, 2000. Centre Res. Math. Proc.
and Lectur e Notes 24.
[181] M. Waldschmidt, P. Moussa, J. M. Luck, and C. Itzykson. From Number Theory to Physics. Springer
Verlag, Berlin, 1992.
[182] Z.-X. Wan. Introduction to Kac-Moody algebras. World Scientific, Singapore, 1991.
[183] E. Witten. Elliptic genera and quantum field theory. C omm . Math. Phys., 109, 1987.
[184] E. Wi tt en . Physics and geometry. In Proc. Intern. Congr. Math. (B er keley, 1986), pages 267–301,
Providence, 1987. Amer. Math. Soc.
[185] E. Witten. Quantum field theory, Grassmannians, and algebraic cur ves. Comm. Math. Phys.,
113:529–600, 1988.
[186] E. Witten. Three-Dimension al Gravity Reconsidered. arXiv:0706.3359v1, 2007.
[187] D. Zagier. Traces of singular moduli. In Motives, Polylogarithms and Hodge Theory, Part I: Motives
and Polylogarithms, pages 211–244, Sommerville, 2002. International Press.
[188] Y. Zhu. Global vertex operators on Rie mann surfaces. Comm. Math. Phys., 165:485–531, 1994.
[189] Y. Zhu. Modular invariance of characters of vertex operator algebras. J. Amer. Math. Soc., 9:237–302,
1996.
206
Appendix A
Proo f o f t he Repl ica t io n Formulae
This part is a continuation of Section 7.5. Our purpose is to give a detailed proof of the
recursion formulas ( 7. 19 ) - ( 7. 22 ) .
A.1 Faber polynomials
The Faber polynomials [60] originated in approximation theory in 1903 and are central to
the theory of replicable functions. We define them in a formal way. The reader interested
in anal y t ica l aspect s of these polynomials can consult [60] or sectio n 3 and 12 of [145 ] .
Let f : H C be a function having q-expansion
f(z) =
1
q
+
n0
H
n
q
n
,
where we take q = e
2πiz
, for z H, the upper half-plane. Throug h out, we interpret
derivatives of f with respect to q. We initially assume that the coefficients H
n
of f are
in C and we choose the con s t ant term to be zero. For each n Z
+
, th e re exists a uni q u e
monic polynomial P
n
in f , such that
P
n
(f) =
1
q
n
+ O(q) as q 0,
(or equivalently, P
n
(f) q
n
(mod qZ[q])). In fact, P
n
= P
n
(f) depends on the coefficients
of f, but we denote i t simply by P
n
when there is no confusio n . The polynomial P
n
is called
the n-th Faber polynomial associ at ed with f. It can be shown that the Faber polynomials
are given by the generating series
qf
(q)
z f(q)
=
n0
P
n
(z)q
n
,
with P
0
(z) = 1, P
1
(z) = z, P
2
(z) = z
2
2H
1
, P
3
(z) = z
3
3H
1
3H
2
, and more generally:
P
n
(z) = det(zI A
n
),
207
where
A
n
=
H
0
1
2H
1
H
0
1
.
.
.
.
.
.
.
.
.
.
.
.
(n 2)H
n3
H
n4
H
n5
. . . 1
(n 1)H
n2
H
n3
H
n4
. . . H
0
1
nH
n1
H
n2
H
n3
. . . H
1
H
0
.
It is useful to note that the Faber poly n om i al s satisfy a Newton type recurrence relation
of the form
P
n+1
(z) = zP
n
(z)
n1
k=1
H
nk
P
k
(z) (n + 1)H
n
, (A.1)
for all n 1.
Another useful way to see the Faber polynomials according to Norton [ 15 4 ] is the following.
Consider f with the q-expa n si on
f(z) =
1
q
+
n0
H
n
q
n
, q = e
2πiz
.
We define some coefficients H
m,n
by the formula
F (y, z) = log
f(y) f(z)
= log(p
1
+ q
1
)
m,n1
H
m,n
q
m
p
n
,
(the bivarial transformation of f), wher e p = e
2πiy
. Then X
n
(f) =
1
n
q
n
+
m
H
m,n
q
m
is
the coefficient o f p
n
in th e expansion
log p log
f(y) f(z)
,
so that it is a polynomial in f. In fact, X
n
(f) =
1
n
P
n
(f), and we can see X
n
as a (non-
monic) n-th Fabe r polynom i al associ at ed to f.
Example A.1.1. If f(z) is a modular form of weight 2k on SL
2
(Z), then the Hecke
operators T
n
, n 1 act on f as
T
n
(f)(z) = n
k1
ad=n, 0b<d
1
d
k
f
az + b
d
,
for all n 1. See Chapter 6 and S ect io n 7.1, or the first five chapter of [18 1 ] (especially
Zagier’s article) for background details. When k = 0 and f (z) is the j-function, we have
T
n
(j)(z) =
1
n
ad=n, 0b<d
j
az + b
d
.
208
The generators of SL
2
(Z) permute the lin ear fractional transformations in the sum, hence
T
n
(j) is invaria nt under SL
2
(Z). Since T
n
(j) has no poles in the upper half-plane H, it
follows t h at it is a polynomial in j (see Section 7.1 ) . We find that,
T
n
(f)(z) =
1
q
n
+ O(q) as q 0,
for all n 1, and so T
n
(j) =
1
n
P
n
(j). Thus the Hecke operator of the j-function are
examples of Faber polynom i al s.
A.2 Replicable functions
Let f be a function having q-expansion
f(z) =
1
q
+
n0
H
n
q
n
,
where q = e
2πiz
. Such a function is called replicable if ther e exist a sequence {f
(s)
}
sZ
+
of
functions f
(s)
: H C, such that for all n 1, the expression
P
n
(f) =
ad=n, 0b<d
f
(a)
az + b
d
. (A.2)
The fun ct i on f
(s)
introduced above is called the s-th replication power of f.
Thus, we can think this polynomial P
n
(f) as the action of a generalized Hecke oper at or .
Following Norton [154], let us consider the coefficients {H
m,n
}
m,n1
introduced in previous
section by
X
n
(f) =
1
n
q
n
+
m1
H
m,n
q
m
, (A.3)
for n 1 (note in particu la r that H
n,1
= H
n
), an d the slightly modified ones
h
m,n
= (m + n)H
m,n
.
It follows from equation (A.1) that this h
m,n
are given recursively by
h
m,n
= (m + n)H
m+n1
+
m1
i=1
n1
j=1
H
i+j1
h
mi,nj
.
A useful characterization for replicable functions proved by Norton in [15 4 ] is th e following:
Theorem A.2.1. The function f is replicable if, and only if, H
m,n
= H
r,s
for all positive
integers m, n, r, s satisfying mn = rs and (m, n) = (r, s).
209
Let H
(s)
n
be the coeffici ent of q
n
in th e s-th replication power of f, that is
f
(s)
(z) =
1
q
+
n0
H
(s)
n
q
n
,
for all n, s 1 (i n p a r t ic u l ar , H
(1)
n
= H
n
). According t o Pr oposition 7.4.3 and Theorem
7.2.2, equation (A.2) can be written as
H
m,n
=
s|(m,n)
1
s
H
(s)
mn/s
2
. (A.4)
Applying the obius inversion formula to equation, then we have
H
(s)
n
= s
d|s
µ(d)H
dsn,
s
d
, (A.5)
where µ is the obius function.
Another result obt a ined by Norton [154] is:
Theorem A.2.2. Suppose that f is replicable. For any k Z
+
, as s ranges over divisors
of k, the following are equivalent:
1. f
(s)
is replicable.
2. The bivarial transform of f
(s)
is the generating function of H
(s)
m,n
.
3. In addition to condition (1), the t-th replication power of f
(s)
is f
(st)
.
Other r esu l t s about replicable functions can be found in [124] and [147].
A.3 Deduction of the replication formula e
Let f (z) be given by the series
f(z) =
1
q
+
n1
H
n
q
n
,
where q = e
2πiz
. D efi ne X
n
(f) as in (A.3) (observe in particular that X
1
(f) =
1
q
+
H
n,1
q
n
= f (z). We will write X
1
= f .
We have already mentioned that X
n
is the coefficient of p
n
in the expansion of log p
1
log
f(y) f(z)
, where p = e
2πiy
. Writin g
log p
1
n1
X
n
p
n
= log
f(y) f(z)
,
210
it foll ows that
p
1
exp
n1
X
n
p
n
= p
1
+
n1
H
n
p
n
q
1
n1
H
n
q
n
,
so
exp
n1
X
n
p
n
= 1 + p
q
1
n1
H
n
q
n
+
n1
H
n
p
n+1
. (A.6)
Using the Taylor expansion exp
X
n
p
n
=
k0
1
k!
X
n
p
n
k
, and comparing the
coefficients of p
2
, p
3
and p
4
in (A.6 ) , we obtain
H
1
=
1
2
2X
2
+ X
2
1
=
1
2
2X
2
+ f
2
; (A.7)
H
2
=
1
6
6X
3
+ 6X
2
f f
3
; (A.8)
H
3
=
1
24
24X
4
+ 24X
3
f + 12X
2
2
12X
2
f
2
+ f
4
. (A.9)
Consider f
(s)
=
1
q
+
n1
H
(s)
n
q
n
, the replication powers of f. Also, define the linear operator
U
n
, n 1, such that for any series of the form
Z
a
q
, we have
Z
a
q
U
n
= n
Z
a
nℓ
q
.
Then, Koike [118] proved the following formulas called 2-plication and 4-plication, respec-
tively:
X
2
(f) =
1
2
f|
U
2
+ f
(2)
(2z)
; (A.10)
X
4
(f) =
1
4
f|
U
4
+ f
(2)
|
U
2
(2z) + f
(4)
(4z)
. (A.11)
In fact , both these for mulas can be obtained from the recursion equation (A.4) as follows.
From (A.4 ) we have
H
n,2
=
H
2n
+
1
2
H
(2)
n/2
, n 0 (mod 2);
H
2n
, n 1 (m od 2);
and
H
n,4
=
H
4n
+
1
2
H
(2)
n
+
1
4
H
(4)
n/4
, n 0 (mod 4);
H
4n
+
1
2
H
(2)
n
, n 2 (mod 4);
H
4n
, n 1, 3 (mod 4).
211
Hence,
1
2
f|
U
2
+ f
(2)
(2z)
=
n1
H
2n
q
n
+
1
2
1
q
2
+
n1
H
(2)
n
q
2n
=
1
2
q
2
+
n1
H
2n
q
n
+
n1
1
2
H
(2)
n
q
2n
=
1
2
q
2
+
n1
H
n,2
q
n
= X
2
(f),
and
1
4
f|
U
4
+ f
(2)
|
U
2
(2z) + f
(4)
(4z)
=
n1
H
4n
q
n
+
1
2
n1
H
(2)
2n
q
2n
+
1
4
1
q
4
+
n1
H
(4)
n
q
4n
=
1
4
q
4
+
n1
H
4n
q
n
+
n1
1
2
H
(2)
2n
q
2n
+
n1
1
4
H
(4)
n
q
4n
=
1
4
q
4
+
n1
H
n,4
q
n
= X
4
(f).
Using t h ese Koike’s formulas, from (A.7) and (A.10) we obtain
H
1
=
1
2
f
2
f|
U
2
f
(2)
(2z)
; (A.12)
and from (A.7)-( A. 11) :
H
3
+
1
2
H
2
1
=
1
24
f
4
1
2
X
2
f
2
+
1
2
X
2
2
+ X
3
f X
4
+
1
8
f
2
f|
U
2
f
(2)
(2z)
2
=
1
24
f
4
1
4
f|
U
2
f
2
1
4
f
(2)
(2z)f
2
+
1
8
(f|
U
2
)
2
+
1
4
f|
U
2
f
(2)
(2z) +
+
1
8
(f
(2)
(2z))
2
+ X
3
f
1
4
f|
U
4
1
4
f
(2)
|
U
2
(2z)
1
4
f
(4)
(4z)
+
1
8
f
4
1
4
f|
U
2
f
2
1
4
f
(2)
(2z)f
2
+
1
8
(f|
U
2
)
2
+
1
4
f|
U
2
f
(2)
(2z) +
1
8
(f
(2)
(2z))
2
=
1
6
f
4
X
2
f
2
+ X
3
f
1
4
f|
U
4
1
4
f
(2)
|
U
2
(2z)
1
4
f
(4)
(4z) +
+
1
4
(f|
U
2
)
2
+
1
2
f|
U
2
f
(2)
(2z) +
1
4
(f
(2)
(2z))
2
= H
2
f
1
4
f|
U
4
+
1
2
f|
U
2
f
(2)
(2z) +
1
4
(f|
U
2
)
2
+
+
1
4
(f
(2)
(2z))
2
f
(2)
|
U
2
(2z) f
(4)
(4z)
= H
2
f
1
4
f|
U
4
+
1
2
f|
U
2
f
(2)
(2z) +
1
4
(f|
U
2
)
2
+
1
2
H
(2)
1
,
212
thus, we have
H
3
+
1
2
H
2
1
1
2
H
(2)
1
=
1
4
(f|
U
2
)
2
+
1
2
f|
U
2
f
(2)
(2z) H
2
f
1
4
f|
U
4
. ( A. 13)
If we compare the c oefficients of q
2k
and q
2k+1
(for k 1) of both sides in (A.12), and carry
out som e calculation , we find:
H
1
=
1
2
f
2
f|
U
2
f
(2)
(2z)
=
1
2
1
q
+
n1
H
n
q
n
2
1
2
2
n1
H
2n
q
n
1
2
1
q
2
+
n1
H
(2)
n
q
2n
=
1
2q
2
+
1
q
n1
H
n
q
n
+
1
2
n1
n1
j=1
H
j
H
nj
q
n
n1
H
2n
q
n
1
2q
2
1
2
n1
H
(2)
n
q
2n
= H
1
+
n1
H
n+1
+
1
2
1j<n
H
j
H
nj
H
2n
q
n
1
2
n1
H
(2)
n
q
2n
.
Thus, for n = 2k and n = 2k + 1 we have t h e following relations
n = 2k : H
2k+1
+
1
2
1j<2k
H
j
H
2kj
H
4k
1
2
H
(2)
k
= 0;
n = 2k + 1 : H
2k+2
+
1
2
1j2k
H
j
H
2kj+1
H
4k+2
= 0.
Since
1j<2k
H
j
H
2kj
= 2
1j<k
H
j
H
2kj
+ H
2
k
and
1j2k
H
j
H
2kj+1
= 2
1jk
H
j
H
2kj
, in
particular, we obt ai n our first two replicable formulas
H
4k
= H
2k+1
+
1j<k
H
j
H
2kj
+
1
2
H
2
k
H
(2)
k
. (A.14)
H
4k+2
= H
2k+2
+
1jk
H
j
H
2kj+1
. (A.15)
Develop the other two remaining replication for mulas is a bit more demand i n g. Applying
(A.14) to 2k, k + 1 and 2k + 1 in place of k we ob t a i n
H
8k
= H
4k+1
+
1j<2k
H
j
H
4kj
+
1
2
H
2
2k
H
(2)
2k
, (A.16)
H
4k+4
= H
2k+3
+
1jk
H
j
H
2kj+2
+
1
2
H
2
k+1
H
(2)
k+1
, (A.17)
H
8k+4
= H
4k+3
+
1j2k
H
j
H
4kj+2
+
1
2
H
2
2k+1
H
(2)
2k+1
; (A.18)
213
and replacing k + 1 for k in (A.15), we also have
H
4k+6
= H
2k+4
+
1jk
H
j
H
2kj+3
. (A.19)
Now, comparing the coefficients of q
2k
and q
2k+1
(for k 1) of both sides in (A.13), and
doing some calcula t io n s, we find:
H
3
+
1
2
H
2
1
1
2
H
(2)
1
=
1
4
(f|
U
2
)
2
+
1
2
f|
U
2
f
(2)
(2z) H
2
f
1
4
f|
U
4
=
1
4
2
n1
H
2n
q
n
2
+
1
2
2
n1
H
2n
q
n

1
q
2
+
n1
H
(2)
n
q
2n
H
2
1
q
+
n1
H
n
q
n
1
4
4
n1
H
4n
q
n
=
n1
n1
j=1
H
2j
H
2n2j
q
n
+
1
q
2
n1
H
2n
q
n
+
k1
k1
j=0
H
4j+2
H
(2)
kj
q
2k+1
+
+
k1
k2
j=0
H
4j+4
H
(2)
kj+1
q
2k
1
q
H
2
n1
H
2
H
n
q
n
n1
H
4n
q
n
= H
4
+
n1
H
2n+4
+
1j<n
H
2j
H
2n2j
H
2
H
n
H
4n
q
n
+
+
k1
k1
j=0
H
4j+2
H
(2)
kj
q
2k+1
+
k1
k2
j=0
H
4j+4
H
(2)
kj+1
q
2k
Thus, taki n g n = 2k and n = 2k + 1 above, we have the following relations
n = 2k : H
4k+4
+
1j<2k
H
2j
H
4k2j
+
k2
j=0
H
4j+4
H
(2)
kj+1
H
2
H
2k
H
8k
= 0;
n = 2k + 1 : H
4k+6
+
1j2k
H
2j
H
4k2j+2
+
k1
k1
j=0
H
4j+2
H
(2)
kj
H
2
H
2k+1
H
8k+4
= 0.
Replacing (A.16) and (A.17) in the expression for 2k above, we have
H
2k+3
+
1jk
H
j
H
2kj+2
+
1
2
H
2
k+1
H
(2)
k+1
+
1j<2k
H
2j
H
4k2j
+
+
k2
j=0
H
4j+4
H
(2)
kj+1
H
2
H
2k
H
4k+1
1j<2k
H
j
H
4kj
1
2
H
2
2k
H
(2)
2k
= 0.
214
Thus,
H
4k+1
= H
2k+3
H
2
H
2k
1
2
H
2
2k
H
(2)
2k
+
1
2
H
2
k+1
H
(2)
k+1
+
+H
2
2k
+ 2
1j<k
H
2j
H
4k2j
+
1jk
H
j
H
2kj+2
+
1j<k
H
4k4j
H
(2)
j
1j<2k
H
j
H
4kj
= H
2k+3
H
2
H
2k
+
1
2
H
2
2k
+ H
(2)
2k
+
1
2
H
2
k+1
H
(2)
k+1
+
1jk
H
j
H
2kj+2
+
1j<k
H
4k4j
H
(2)
j
+
2
1j<k
H
2j
H
4k2j
1j<2k
H
j
H
4kj
= H
2k+3
H
2
H
2k
+
1
2
H
2
2k
+ H
(2)
2k
+
1
2
H
2
k+1
H
(2)
k+1
+
1jk
H
j
H
2kj+2
+
1j<k
H
4k4j
H
(2)
j
+
1j<2k
(1)
j
H
j
H
4kj
.
Similarly, replacing (A.18) and (A.1 9 ) in the expression for 2k + 1 above, we obtain
H
2k+4
+
1jk+1
H
j
H
2kj+3
+
1j2k
H
2j
H
4k2j+2
+
0j<k
H
4j+2
H
(2)
kj
H
2
H
2k+1
H
4k+3
1j2k
H
j
H
4kj+2
1
2
H
2
2k+1
H
(2)
2k+1
= 0.
Thus,
H
4k+3
= H
2k+4
H
2
H
2k+1
1
2
H
2
2k+1
H
(2)
2k+1
+
1jk+1
H
j
H
2kj+3
+
+
0j<k
H
4j+2
H
(2)
kj
+
1j2k
H
2j
H
4k2j+2
1j2k
H
j
H
4kj+2
= H
2k+4
H
2
H
2k+1
1
2
H
2
2k+1
H
(2)
2k+1
+
1jk+1
H
j
H
2kj+3
+
+
1jk
H
4k4j+2
H
(2)
j
+
2
1jk
H
2j
H
4k2j+2
1j2k
H
j
H
4kj+2
= H
2k+4
H
2
H
2k+1
1
2
H
2
2k+1
H
(2)
2k+1
+
1jk+1
H
j
H
2kj+3
+
+
1jk
H
4k4j+2
H
(2)
j
+
1j2k
(1)
j
H
j
H
4kj+2
,
215
and we have
H
4k+1
= H
2k+3
H
2
H
2k
+
1
2
H
2
2k
+ H
(2)
2k
+
1
2
H
2
k+1
H
(2)
k+1
+
+
1jk
H
j
H
2kj+2
+
1j<k
H
4k4j
H
(2)
j
+
1j<2k
(1)
j
H
j
H
4kj
. (A.20)
and
H
4k+3
= H
2k+4
H
2
H
2k+1
1
2
H
2
2k+1
H
(2)
2k+1
+
1jk+1
H
j
H
2kj+3
+
+
1jk
H
4k4j+2
H
(2)
j
+
1j2k
(1)
j
H
j
H
4kj+2
. (A.21)
Hence, we have obtained the following four rep l i cat i on formulas (A.14), (A.20), (A.15),
(A.21):
H
4k
= H
2k+1
+
1j<k
H
j
H
2kj
+
1
2
H
2
k
H
(2)
k
;
H
4k+1
= H
2k+3
H
2
H
2k
+
1
2
H
2
2k
+ H
(2)
2k
+
1
2
H
2
k+1
H
(2)
k+1
+
+
1jk
H
j
H
2kj+2
+
1j<k
H
4k4j
H
(2)
j
+
1j<2k
(1)
j
H
j
H
4kj
;
H
4k+2
= H
2k+2
+
1jk
H
j
H
2kj+1
;
H
4k+3
= H
2k+4
H
2
H
2k+1
1
2
H
2
2k+1
H
(2)
2k+1
+
1jk+1
H
j
H
2kj+3
+
+
1jk
H
4k4j+2
H
(2)
j
+
1j2k
(1)
j
H
j
H
4kj+2
.
To obtain the set of equations (7.19)-(7.22), it is just a change of notation. Recall that
we write c
g
(n) to be the coefficient of q
n
in the McKay-Thompson series T
g
(z)
of g M,
that is T
g
(z)
= q
1
+
n1
c
g
(n)q
n
, with q = e
2πiz
. Also, r e cal l Conway and Norton
already proved in [38] that the s-th replication associated to T
g
(z)
is exactly T
g
s
(z)
, the
McKay-Thompson series associated to g
s
. Hence, taking f = T
g
, we have the following
dictionary
H
n
= c
g
(n) and H
(2)
n
= c
g
2
(n),
from wh er e the replication formulae (7.19)-(7.22) directly appear.
Also, note that when comparin g coefficients of q
2k
and q
2k+1
in (A.13), we have obtained
in add i t i on
H
4
= H
3
+
1
2
H
2
1
1
2
H
(2)
1
,
that i s, equation (7.23), using notation c
g
(n).
216
Index
adjoining the derivation, 36
adjoint map, 28, 51
adjoint representation, 51
algebra
affine, 46
extended, 46
extended twisted, 48
twisted, 47
untwisted, 46
associative, 27
commutative, 27
graded, 38
nonassociative, 27
quotient, 34
Atlas of finite groups, 4, 5
automorphism, 28
of vertex algebras, 93
B, see Griess algebra
Baby Monster, 3, 5, 6
base
of a root system, 63
Bernoulli numbers, 140
bihomomorphism, 38
Bimonster, 7
Borcherds
algebra, 165
character formula, 169
denominator identity, 170
identity, 160
C
24
, see Golay code
Cartan matrix , 88
generalized, 89
of affine type, 90
of finite type, 90
of indefin it e type, 90
symmetrisable, 90
Cartan sub algebr a, 57
center
of a Lie algebra, 35
character, 66, 88
table, 5
Chevalley basis, 100
chiral algebra, 195
class func ti on, 66
co-root, 59
real, 67
code, 105
of type I, 105
of type II, 105
self-dual, 106
code loops, 125
commensurable, 15
commutator ideal, 35
compact
matrix Lie group, 25
real form, 56
complete reducibility property, 53
completely reducible, 39, 53
complexification
of a Lie algebra, 32
of a vector space, 32
conformal
field th eor y, 195
vector, 84
vertex algebra, 85
congruence subgroup, 153
connected
matrix Lie group, 25
contravariant bilinear form, 165
convergence
in matrix groups, 22
Conway group
Co
0
, 110
Co
1
, 110
Conway sporadic group
Co
1
, 3, 5
Co
2
, 3, 5
Co
3
, 3–5
correspondence, 149
Coxeter graph, 7
associated group to, 7
cusp, 10
cusp form, 130
Dedekind’s η f un ct ion, 10
degree operator, 36
denominator identity
twisted, 182
derivation, 28
direct produ ct
217
of Lie algebras, 35
direct s u m
of modules, 39
discriminant, 134
dual
code, 105
lattice, 94
root system, 63
Dynkin diagram, 98
Eisenstein series, 10, 133, 142
endomorphism, 28
equivalence
of Dynk in diagrams, 98
of extens ions , 35
of modules, 34
of repr es e ntations, 50
exponencial matr ix , 29
extension
central, 35
of Lie algebras, 35
trivial, 35
Faber polynomial, 207
finite gr oups
classification of, 2, 4
Fischer sporadic group
Fi
22
, 3, 5
Fi
23
, 3, 5
Fi’
24
, 3, 5
Fricke element, 186
fundamental functions, 10
genus
of a group, 15
Golay cod e, 4, 106
G
pqr
, see Coxeter graph
graded
algebra, 38
character, 179
dimension, 179
module, 39, 123
subspace, 36
vector space, 36
graded component, 36
graded dimen si on, 13
grading
of a tensor product, 38
grading pr es er v in g, 36
Griess algeb r a, 4, 116
Griess module, 114
group
alternating, 2
Chevalley, 2
classification project , 2
Coxeter, 7
cyclic, 2
Monster, see Monster
of Lie type, 2
of Monster - typ e , 125
simple, 2
sporadic, 2, 3
group algebr a, 41
Haar measure , 54
Hall divi sor , 154
Hall-Janko sporadic group HJ, 3
Hamming code, 106
Happy family, 4
Haputmodul
of a group, 11
Harada-Norton spor adic group HN, 3, 5
Hauptmodul, 10
normalized, 11, 15
Hecke operator, 150
Heisenberg group, 24
Held sporadic group He, 3
higher, 65
highest weight, 65
Higman-Sims spor adic group HS, 3
holomorphic at infinity, 130
homogeneous
space of degree α, 36
linear map, 36
subspace, 36
homomorphism, 28
of Lie algebras, 28
of modules, 34
ideal, 34
central, 35
left, 34
right, 34
indecomposable
module, 34
induced
module, 40, 41, 43
integral
dominant integral, 64
integral element, 63, 67
intertwining map, 50
invariant
218
bilinear map, 45
subspace, 34, 49
involution, 47
irreducible, 49
module, 34
isometry, 94
isomorphism, 28
of lattices, 94
of modules, 34
j function, 1, 139
Janko sp or adic group
J
1
, 3, 5, 6
J
3
, 3–5
J
4
, 3, 4
Jordan-H¨older theorem, 2
Kac-Moody algebra, 89
Λ
24
, see Leech lattice
lattice, 93
even, 94
integral, 94
non degener ate, 94
positive definite, 94
self-dual, 94
Leech lattice, 4–6, 108
shorter, 191
Lie algebra, 28
commutative, 28
of a Lie group, 29
quotient, 34
reductive, 55
semisimple, 55
Lie group, 26
linear c ode, 105
linear gr oup
general, 22
special, 22
Lorentzian space R
25,1
, 4
Lyons spor adic group Ly, 3, 4
M, see Monster
Mathieu group
M
24
, 3, 5
Mathieu sporadi c group
M
11
, 3
M
12
, 3
M
22
, 3, 5
M
23
, 3, 5
M
24
, 4
Matieu group M
24
, 107
matrix Lie group, 22
McKay-Thompson series, 14
McLaughlin sp oradi c group McL, 3
meromorphic at infinity, 130
mirror symmetry, 194
Mirror-Moonshine conjecture, 195
modular
equation, 160
form, 10, 130
function, 10, 130
lattice function, 132
weakly, 130
modular group, 1, 8, 127
fundamental domain, 128
module, 33, 41
Monster, 1, 2, 5, 6
irreducible character degrees, 4
Lie algebras , 171
representation of, 7
size of, 5
Monster vertex algebr a, 114
Moonshine
generalized, 185
Moonshine conjec tu r e, 1
Moonshine module, 1, 113
moonshine-type, 154
multiplicity, 176
of weights, 64
Niemeier lattice, 111
of type A
24
1
, 111
no-ghost t he ore m, 171
normal ord er e d product, 84
Norton se r ie s , 186
O’Nahn s poradic group O’N, 3, 4
octad, 106
orbifold, 185
order
of f at p, 135
orthogonal gr oup , 23
complex speci al, 23
special, 23
orthonormal group
complex, 23
pariah, the, 4
product
in an algebra, 27
219
quilt, 188
quotient
module, 39
Ramanujan’s τ function, 143
rank
of a Lie algebra, 58
replicable function, 160
replication formulae, 158, 159, 161
replication power, 209
representation
adjoint, 34
basic, 91
complex, 49
faithful, 49
left re gular , 41
modular, 192
of a group, 41
of Lie algebras, 33
of modules, 33
real, 49
trivial, 41
root, 58, 168
imaginary, 90, 168
negative, 63
positive, 63, 168
real, 66, 90, 168
simple, 63, 168
system, 62
root space, 58
root vector, 58
Rudvalis sporadic group Ru, 3, 4
semidirect pro d uc t
of Lie algebras, 35
semisimple
module, 39
sextet, 106
shorter Moonshi ne module, 191
simple
module, 34
simple Li e algebra, 34
simply c onne ct ed
matrix Lie group, 26
skew-symmetry, 28
SL
2
(Z), s ee modular group
sporadic groups
list of all, 3
standard representation
of a Lie group, 50
of Lie algebras, 51
string, 195
subalgebra, 28
submodule, 34
Suzuki sporadi c group Suz, 3
symmetric algebra, 42
symmetric p owers, 43
symplectic group
compact, 24
complex, 24
real, 24
tensor algebra, 42
tensor map, 37, 38
tensor multiplication, 37
tensor produ ct, 41
module, 39
of vector spaces, 37
theta f un ct ion, 96, 146
Third gen er ati on, 4
Thompson
order for mula, 5
sporadic group Th, 3, 5
totally singular, 106
transform, 149
unimodular
lattice, 94
unimodular Lie group, 54
unitary matrix, 23
universal Borcherds algebra, 166
universal enveloping algebra, 43
V
, see Moonshine module
vertex algebra, 81
vertex operator, 82
vertex operator algebra, 1, 85
Virasoro algebr a, 86
Weierstrass function, 134
weight, 64, 76
fundamental, 78
of a code, 105
real, 67
space, 64
vector, 64, 76
weight distribution, 106
Weyl
character formula, 67
group, 168
Weyl group, 60
220
Livros Grátis
( http://www.livrosgratis.com.br )
Milhares de Livros para Download:
Baixar livros de Administração
Baixar livros de Agronomia
Baixar livros de Arquitetura
Baixar livros de Artes
Baixar livros de Astronomia
Baixar livros de Biologia Geral
Baixar livros de Ciência da Computação
Baixar livros de Ciência da Informação
Baixar livros de Ciência Política
Baixar livros de Ciências da Saúde
Baixar livros de Comunicação
Baixar livros do Conselho Nacional de Educação - CNE
Baixar livros de Defesa civil
Baixar livros de Direito
Baixar livros de Direitos humanos
Baixar livros de Economia
Baixar livros de Economia Doméstica
Baixar livros de Educação
Baixar livros de Educação - Trânsito
Baixar livros de Educação Física
Baixar livros de Engenharia Aeroespacial
Baixar livros de Farmácia
Baixar livros de Filosofia
Baixar livros de Física
Baixar livros de Geociências
Baixar livros de Geografia
Baixar livros de História
Baixar livros de Línguas
Baixar livros de Literatura
Baixar livros de Literatura de Cordel
Baixar livros de Literatura Infantil
Baixar livros de Matemática
Baixar livros de Medicina
Baixar livros de Medicina Veterinária
Baixar livros de Meio Ambiente
Baixar livros de Meteorologia
Baixar Monografias e TCC
Baixar livros Multidisciplinar
Baixar livros de Música
Baixar livros de Psicologia
Baixar livros de Química
Baixar livros de Saúde Coletiva
Baixar livros de Serviço Social
Baixar livros de Sociologia
Baixar livros de Teologia
Baixar livros de Trabalho
Baixar livros de Turismo