You are on page 1of 710

Lie Algebras With

Triangular Decompositions
CANADIAN MATHEMATICAL SOCIETY SERIES OF
MONOGRAPHS AND ADVANCED TEXTS

Monographies et Etudes de la Société Mathématique
du Canada

EDITORIAL BOARD

Frederick V. Atkinson, Bernhard Banaschewski, Colin W. Clark,


Erwin O. Kreyszig (Chairman) and John B. Walsh

Frank H. Clarke
^Optimization and Nonsmooth Analysis
Erwin Klein and Anthony C, Thompson
^Theory of Correspondences: Including Applications to Mathematical Economics
/. Gohbergy P. Lancaster, and L. Rodman
Invariant Subspaces of Matrices with Applications
Jonathan Borwein and Peter Borwein
Pi and the AGM—A Study in Analytic Number Theory and
Computational Complexity
John H. Berglund, Hugo D. JUnghenn, and Paul Milne
"^Analysis of Semigroups: Function Spaces, Compactifications, Representation
Subhashis Nag
The Complex Analytic Theory of TeichmüUer Spaces
Manfred Kracht and Erwin Kreyszig
^Methods of Complex Analysis in Partial Differential Equations with Applications
Ernest J. Kani and Robert A. Smith
The Collected Papers of Hans Arnold Heilbronn
Victor P. Snaith
"^Topological Methods in Galois Representation Theory
Kalathoor Varadarajan
The Finiteness Obstruction of C.TC. Wall
G. Watson
"^Statistics on Spheres
F. Arthur Sherk
Kaleidoscopes: Selected Writings ofH. 5. M. Coexeter

*Indicates an out-of-print title


Lie Algebras With
Triangular
Decompositions
ROBERT V. MOODY
ARTURO PIANZOLA
University o f Alberta
Edmonton, Canada

A Wiley-Interscience Publication
JOHN WILEY & SONS
New York Chichester • Brisbane • Toronto • Singapore
Photograph on title page by M. Goretz

This text is printed on acid-free paper.


Copyright © 1995 by John Wiley & Sons, Inc.
All rights reserved. Published simultaneously in Canada.
Reproduction or translation of any part of this work beyond
that permitted by Section 107 or 108 of the 1976 United
States Copyright Act without the permission of the copyright
owner is unlawful. Requests for permission or further
information should be addressed to the Permissions Department,
John Wiley & Sons, Inc., 605 Third Avenue, New York, NY
10158-0012.

Library o f Congress Cataloging in Publication Data:


Moody, R. V., 1941-
Lie algebras with triangular decompositions / by Robert V. Moody,
Arturo Pianzola.
p. cm. — (Canadian Mathematical Society series of monographs
and advanced texts)
Includes bibliographical references.
ISBN 0-471-63304-6 (alk. paper)
1. Lie algebras. 2. Decomposition (Mathematics) I. Pianzola,
Arturo, 1955- . II. Title. III. Series.
QA252.3.M66 1995
512'.55— dc20 92-46890
Printed in the United States of America
10 9 8 7 6 5 4 3 2 1
Our formula fo r happiness: a yea, a nay, a straight line, a goal.

F. Nietzche
Contents

Introduction XI

How to Read This Book XV

Course Outlines xix

Chapter 1 Lie Algebras 1

1.1 Basic Definitions


1.2 Tensor, Symmetric, and Exterior Algebras 9
1.3 Gradings 15
1.4 Virasoro and Heisenberg Algebras 20
1.5 Derivations 23
1.6 Representations 28
1.7 Invariant Bilinear Forms 35
1.8 Universal Enveloping Algebras 38
1.9 Central Extensions 49
1.10 Free Lie Algebras 59
1.11 The Campbell-Baker-Hausdorff Formula 65
1.12 Extensions of Modules 71
Exercises 80

Chapter 2 Lie Algebras Adm itting Triangular Decompositions 90

2.1 Triangular and Weight Space Decompositions 91


2.2 Highest Weight Modules 103
2.3 Verma Modules 106
2.4 §l2(IK)-Theorem 116
2.5 Characters 128
2.6 The Category & 141
2.7 The Radical 160
2.8 The Shapovalov form 168
2.9 Jantzen Filtrations 175
2.10 Bernstein-Gel’fand-Gerfand Duality 184
2.11 Embeddings of Verma Modules 200
2.12 Decomposition of Modules in Category Û 204
Exercises 209
vüi Contents

Chapter 3 Lattices and Root Systems 216

3.1 Lattices 217


3.2 Finite Root Systems 229
3.3 Bases for Finite Root Systems 239
3.4 Graphs and Coxèter-Dynkin Diagrams 247
3.5 Classification of Cartan Matrices and FiniteRoot Systems 258
3.6 The Perron-Frobenius Theorem and Its Consequences 272
3.7 Constructing Lie Algebras from Lattices 280
3.8 Central Extensions of Lattices 289
Exercises 302

Chapter 4 Contragredient Lie Algebras 309

4.1 Contragredient Lie algebras 310


4.2 Realizations of Contragredient Lie Algebras 330
4.3 Embeddings, Field Extensions, and Decomposability 342
4.4 Invariant Bilinear Forms 355
4.5 Casimir-Kac Operators 367
4.6 The Radical Theorem 375
4.7 Hermitian Contragredient Forms 382
Exercises 385

Chapter 5 The Weyl Group and Its Geometry 395

5.1 Root Data 396


5.2 The Length Function 408
5.3 Coxeter Groups and the Exchange Condition 418
5.4 The Bruhat Ordering 425
5.5 Morphisms of Root Data: Subroot Systems 430
5.6 The Geometry of a Set of Root Data 436
5.7 Subroot Systems 457
5.8 Imaginary Roots 464
5.9 Conjugacy of Bases 472
Exercises 476

Chapter 6 Category 0 for Kac-Moody Algebras 483

6.1 Integrable Modules 484


6.2 Weight Systems 508
6.3 The Triangular Decomposition of G 516
6.4 The Formulas of Weyl-Macdonald-Kac 531
6.5 Complete Reducibility 542
6.6 Shapóvalov Determinant Formula for Kac-Moody Algebras 545
6.7 The BGG Theorem and Generalization 556
Contents

6.8 Translation Functors and the Generalized


Character Formula 564
Exercises 574

Chapter 7 Conjugacy Theorems 586

7.1 Locally Finite Endomorphisms and Jordan-Chevalley


Decompositions 586
7.2 Locally Finite Elements in Kac-Moody Algebras 606
7.3 The Kost2mt Cone 612
7.4 Conjugacy of Split Cartan Subalgebras 623
Exercises 639

Appendix .4\*)-An Extended Example 647

Bibliography 675

Index 681
Introduction

One of the great achievements of nineteenth-century mathematicians was the


formalization of the notion of symmetry through the introduction of groups
and their representations. Now, in the late twentieth century, we can see the
pervasive way in which group theory has entered almost every area of
mathematics. The Lie groups in particular, which are those that permit
infinitesimal motions, have turned out to be of fundamental significance in
numerous areas including differential equations, differential geometry, alge­
braic geometry, quantum mechanics, particle physics, special functions, alge­
braic topology, combinatorics, and probability theory, not to mention their
role within group theory itself.
The entire local structure of a Lie group is codified in a much simpler
and purely algebraic structure called its Lie algebra. One of the early
accomplishments of the theory of Lie groups was the classification by
W. Killing and E. Cartan of the simple and semisimple Lie groups via the
classification of their Lie algebras. At the heart of this classification lie some
combinatorial objects, the finite root systems, and finite Weyl groups, which
are of amazing beauty and to which innumerable problems involving semisim­
ple Lie groups finally come to rest.
All of these combinatorial objects admit natural infinite-dimensional gen­
eralizations, and it is possible to develop from them an infinite-dimensional
generalization of the semisimple theory that parallels it to a large degree.
This generalization is far from complete. In particular the Lie group side of it
is still rather modest. By contrast, the Lie algebra side has proved to be very
successful and has already turned out to have a variety of applications in
other parts of mathematics, notably in differential equations, combinatorics,
the theory of modular forms, singularity theory, and string models and
conformal field theories in physics. This book is intended as an introduction
to the theory of these infinite-dimensional Lie algebras. They have gone by a
number of names in the literature, but now they are universally called
Kac-Moody algebras.
Our objective is to present a self-contained development of the algebraic
theory of the Kac-Moody algebras, their representations, and their close
relatives, the Virasoro and Heisenberg algebras. We have tried to make the
exposition accessible to anyone with a reasonable background in linear
Introduction

algebra. There are a few exceptions, but these are clearly signaled in advance
and are not critical to reading the book as a whole. This permits a graduate
student who is just beginning Lie theory to get quickly into recent areas in
which there are still plenty of accessible open problems.
Since the important monograph by V. Kac in 1983 [Ka5], there have been
a number of notable developments in Kac-Moody algebras, often at the
hands of Kac and his collaborators. We have been able to include the
axiomatic description of root data, the structure of Verma modules, and
the conjugacy theorems. In recent years there have also been developments
in the theory of the Virasoro algebra and its representations. A common
feature of both of these algebras, as well as of the Heisenberg algebra and
the contragedient algebras, is the existence of triangular decompositions. It
was a paper by Rocha-Caridi and Wallach [RC-W] that led us to think about
developing the book from this point of view. Indeed one of the most
satisfying results of writing this book has been the realization of how much of
the theory depends only on this concept and how much unity and economy
are achieved by systematically adopting it at the outset.
As the scope of the present work became increasingly apparent, we
realized the futility of trying to encompass so much material into one volume
while remaining true to the style and level of presentation that we hoped to
achieve. Thus there is a notable absence of some standard theorems from
finite-dimensional Lie algebras (though some of the theory is covered in the
exercises). Rather we have concentrated on developing the theory of triangu­
lar decompositions and the part of the Kac-Moody theory not specific to the
affine case. In fact many results of the affine case are special instances of
results here; there is an extended example of the affine Lie algebra in
the appendix that serves to highlight how the affine case looks and to
exemplify almost everything that we discuss in the body of the text. For a
guide to developments in the affine case, we recommend [Ka5] and [KMPS].
Another important omission is the remarkable representation theory of the
Virasoro algebra. For more on this, the reader can consult [GO].
A secondary theme running through the book is the. subject of lattices
(discrete subgroups of R" carrying an integral-valued symmetric bilinear
form). Root lattices and weight lattices have long been part of the finite­
dimensional theory. They are equally important in the infinite-dimensional
theory. We have tried to emphasize Lie algebra-lattice connections by
introducing lattices early and keeping them in mind in subsequent sections.
Since some readers will not be familiar with the finite-dimensional semisim­
ple theory, we devote considerable attention to finite root systems and their
Weyl groups before getting into the infinite-dimensional theory. After con­
structing the root lattices of types A, D, E, we use the cocycle method of
Garland-Frenkel-Lepowsky to construct the finite-dimensional Lie algebras
of types A, D, E, complete with triangular decompositions. These serve as
motivating material for the contragredient and Kac-Moody algebras.
Depending on their background, readers will approach this book in
different ways. Some suggestions are given in “How to Use This Book.” In
Introduction xiii

OUI bibliography we list only works that we have referred to in the text. We
apologize to the many researchers in the field whose work has not been
quoted. There is a comprehensive bibliography of Kac-Moody theory by
Géorgie Benkart [Bk] that covers the literature up until about 1985. Other
good sources are [LMS], [FLM], and [Hu2]. Together these references
provide a good view of the various ways in which the theory has developed
and a good departure point for further study.
We have received the help of many people in writing this book. In
particular we would like to mention the continued encouragement of Stephen
Berman, who never gave up hope that it would be finished, and of John Bliss,
Nicole Lemire, Chen Liang, Marc Fabbri, Liu Keqin, Alejandra Premat,
Jorge Valencia, and Shi Zhiyong, who patiently read through many drafts as
it reached its final form. We also thank Rolf Farnsteiner for several sugges­
tions. A number of typists labored through the preparation of the manuscript.
We would particularly like to thank Marion Benedict for her Tex-nical skills
and her tireless and cheerful efforts.
R. V. M o o d y
A . PlANZOLA

Burro Alley Cafe


Santa Fe, NM
April 1994
How to Read This Book

The book can be read from cover to cover in the order of presentation. But
such an approach is not necessarily the best when background and pedagogi­
cal needs or time constraints are taken into consideration. This is particularly
true of Chapter 2 which develops the theory of Lie algebras with triangular
decomposition in complete generality. Many readers will probably be impa­
tient to see this theory applied to the Kac-Moody situation where it takes on
a life of its own. Those readers should therefore read enough of Chapter 2 to
acquire the background needed for Chapter 4, returning to Chapter 2 ás
required.
Chapter 3 could have been placed almost anywhere before Chapter 5.
Because of the impressive and beautiful combinatorial structure of finite root
systems and their logical independence from Lie algebras, the material of
Chapter 3 offers an excellent beginning for a course in Lie theory. The first
section of Chapter 7 provides a self-contained account of the Jordan-
Chevalley decomposition of an endomorphism both in the finite-dimensional
and the locally finite (infinite-dimensional) setting.
To accommodate the varying needs of the readers, we have divided the
contents of the book into “blocks.” Below we show the logical order in which
these blocks interact, and we outline different “courses” that can be pursued.

Block A: Basics o f root systems and Lie theory— Chapters 1 and 3


Sections 1.1-1.11 Basic Lie algebra theory
{ Lattices
Finite root systems and their
Chapter 3 classification, construction
of simple lie algebras of
types A, D, and E

Block B: Basic Kac-Moody theory up to the character formula-


Chapters 2, 4, 6, and the Appendix
Section 2.1 Triangular decompositions
Section 2.2 Highest weight modules
xvi How to Read This Book

Section 2.3 Verma modules


Section 2.4 § 12-theory
Section 2.5 Characters
Section 2.6 Category ^
Section 4.1 Contragredient Lie algebras
Section 4.2 Existence of contragedient Lie algebras
Section 4.3 Embedding, etc.
Section 4.4 Invariant bilinear forms
Section 4.5 Casimir-Kac operators
Section 4.6 Gabber-Kac theorem
Section 4.7 Hermitian forms
Section 6.1 Integrable modules
Section 6.2 Weight systems
Section 6.4 The character formula
Section 6.5 Complete reducibility
Appendix An Extended example

Block C: Root data and geometry o f chambers— Chapter 5


Section 5.1 Root data
Section 5.2 The length function
Section 5.3 Coxeter groups
Section 5.4 Bruhat ordering
Section 5.5 Morphisms
Section 5.6 Geometry of root data
Section 5.7 Subroot systems
Section 5.8 Imaginary roots (uses Section 3.6)
Section 5.9 Conjugacy of bases

Block c: Basics of root data theory and geometry o f chambers— Chapter 5


Section 5.1 Root data
Section 5.2 The length function
Section 5.5 Geometry of root data
Section 5.6 Subroot systems

Block D: Structure o f Verma modules— Chapters 1, 2, and 6


Section 1.12 Extensions of modules
Section 2.8 Shapovalov form
Section 2.9 Jantzen filtrations
Section 2.10 BGG duality
Section 2.11 Embeddings of Verma modules
Section 2.12 Decomposition of modules in category Û
How to Read This Book

Section 6.6 Shapovalov determinants


Section 6.7 BGG theorem
Section 6.8 Generalized character formula

Block E: Conjugacy theorems— Chapters 6 and 7


Section 6.3 Triangular decomposition of the group
Sections 7.1-7.4 All of the chapter

Logical dependence of the blocks

Basics of Lie algebras

Basic Kac - Moody theory Basic root data theory

C Root data theory

Conjugacy theorems
Course Outlines

Course 1
Contains: Block A, Sections 3.7 and 3.8
Description: A first course in Lie theory directed towards the Kac-Moody
approach
Comments: This course falls short of providing the classical background
in Lie algebras (e.g., Lie’s theorem and Cartan’s criterion are
not present). These deficiencies can be corrected by the
instructor in a one-term course.
Suggestions: After block A is completed, Sections 3.7 and 3.8 might be
used to show the relationship between lattices and Lie alge­
bras. Assignments should be designed to incorporate as much
of the classical theory as possible into the course.

Course 2

Contains: Block B
Description: Basic Kac-Moody theory up to the character formula
Comments: This course is intended for students who have already covered
either Block A or a course in classical Lie algebra theory
(e.g., a full-year course using Humphrey’s book).
Suggestions: Some of Chapter 3 could be assigned, especially Sections 3.5,
3.7, and 3.8. The example at the end of the book might
be used as a guide for the new concepts introduced. How the
classical split semisimple theory fits into this bigger picture
should be made clear.

Course 3

Contains: Blocks B, C, and D


Description: Basic Kac-Moody theory, geometry of root systems, and struc­
ture of Verma modules
Comments: This is a full-year course. It presupposes familiarity with the
material in block A or a solid background in classical split
semisimple theory.
XX Course Outlines

Suggestions: Exploit the background of students to cover block B in one


semester. Block C could prove too time-consuming, in which
case block c is adequate.

Course 4

Contains: Chapters 3 and 4


Description: Foundations of finite and infinite root systems
Comments: This course is a good complement to Course 2. It is primarily
for those interested in the combinatorial aspects of Lie the­
ory.
Suggestions: Section 5.1 should be presented in a Lie algebra free form,
using examples and assignments to show connections with Lie
algebras.
Lie Algebras With
Triangular Decompositions
Chapter One

Lie Algebras
There is no sea innavigable, no land uninhabitable
—Robert Thorne, Merchant and Geographer, 1625

This chapter introduces the basic definitions and concepts of Lie algebra
theory. Because it is essentially written as a primer for the subsequent
material, it covers topics such as graded Lie algebras, universal enveloping
algebras, central extensions, free Lie algebras, and the Campbell-Baker-
Hausdorif formula, which often appear much later in a book on Lie theory.
For the same reason it omits standard material that is historically and
mathematically important but unnecessary for our purposes. Readers new to
this subject might want to broaden their knowledge by consulting standard
references such as [Hu] for Lie algebra theory and [Bo2] [Vj] for the Lie
group/Lie algebra connection.

LI BASIC DEFINITIONS

Let K be a field of characteristic 0. By a K-algebra (or algebra over K) we


mean a vector space A over K together with a mapping (x, y) xy from
A X A into A satisfying

x(y+z)=xy+xz, (x+y)z=xz+yz,
c(xy) = (cx)y =x( cy)

for all x, y, z ^ A and c g IK.


The mapping ( x, y) xy is called the multiplication or composition law
of the algebra A, and jcy is called the product of x and y. Almost always we
suppress (as we have just done) all mention of the field IK since it is
understood. The dimension of A is the dimension of ^ as a vector space.
Much of this beginning material, especially the definitions, uses only the
fact that IK is a commutative ring with identity and that ^ is a IK-module (see
2 Lie Algebras

the end of this section). Very occasionally we will talk about algebras over Z
in which the underlying Z-module is free. However, this does not involve
anything in the least bit profound. To keep the material here as straightfor­
ward as possible, we simply assume that IK is a field of characteristic 0. For
the reader who is not an expert in field theory, it is always safe to assume that
K is the field of complex numbers C. Lie algebras over arbitrary commutative
rings will be treated at the end of this section.
An algebra A is said to be associative if x(yz) = (xy)z for all x , y , z ^ A
and is said to be commutative if xy = yx for all x, y ^ A, There can exist at
most one element 1 = ^ A with the property x l = x = l x for all x ^ A.
If A (0) and A has such an element, then A is called an algebra with
identity element and 1 is called the identity element of A.
Although we will have occasion to deal with various types of algebras, the
Lie algebras (definition below) are at the center of our study. It is a
long-established custom to denote the multiplication in the Lie algebra by
[*, • ] and to call the expression [jc, y] the bracket or commutator of x and y.
It is also customary to use lowercase old German letters^ to denote Lie
algebras, and we will adhere to this convention except where common usage
has established otherwise.
An algebra g with a composition law [*, *] is called a Lie algebra if it
satisfies the following two identities for all x , y , z in g:
LAI [ x , y ] + [ y , x ] = 0,
LA2 [ x, [ y, z ] ] + [y,[z,jc]] + [z ,[j:,y ]] = 0.
Notice that (LAI) implies that
(LA3) [x, x] = 0 for all X e g,
and conversely (LA3) implies (LAI). The identity of (LA2) is called the
Jacobi identity and we shall denote its left-hand side by Jac(jc, y, z). Because
of (LAI), (LA2) is equivalent to [[x, y], z] + [[y, z], x] + [[z, j c ] , y] = 0. It is
useful to note that to establish that an algebra A is a. Lie algebra it suffices to
verify that (LAI) and (LA2) hold for all jc , y, z in some linear basis of A.
A consequence of LAI is that a Lie algebra is commutative if and only if
[ jc , y] = 0 for all JC, y ^ A . The custom is to call a Lie algebra abelian rather
than commutative. Clearly any vector space V over IK can trivially be made
into a Lie algebra. No Lie algebra g can have an identity element. Nor can a
Lie algebra be associative except under extreme circumstances.
If g is a Lie algebra and jc e g, then we define the linear mapping ad j c :
9 9 by
(1) ad x( y) := [ jc , y] for all y e g.
We will have a lot more to say about “ad” in Section 1.6.

^Gótica textura quarata, to be precise.


1.1 Basic Definitions

If {jCy}, j e J, is a basis for an algebra A, then its multiplication is


completely determined by the equations

( 2) XiXj= E
k^J
The scalars are called the structure constants of A (relative to the given
basis). Of course it is part of the definition of a basis that all the vectors in
the space A are finite linear combinations of basis elements. Thus for a given
i and j only finitely many will be nonzero.
Generally speaking, structure constants are not very useful for studying
algebras, especially since they depend on the basis chosen. Having said this, it
is only fair to point out that we will actually construct some Lie algebras
directly from their structure constants.
A subset A' of a K-algebra A is called a subalgebra of A if it is a
K-subspace and is closed under the multiplication, that is, if xy ^ A when­
ever X, y ey l'. For example, if g is a Lie algebra and x e g, then g^ —
{y e g| [y, x] = 0} is a subalgebra called the centralizer of x in g. If A is
either an associative algebra or a Lie algebra, then the centre of A defined by

3 ( ^ ) '•= {x ^ A \x y = yx for all y ^ A)

is a subalgebra of In the case of a Lie algebra we have equivalently

Z ( A ) = (x e A\[x, y] = 0 for all y ^ A ) ,

A mapping f : A ^ B from one algebra into another is called a homomor­


phism if it is linear, and for all x , y ^ A, f(xy) = /( x ) /( y ) . One defines
monomorphisms (= injective homomorphism), epimorphism (= surjective
homomorphism), isomorphism (= bijective homomorphism), endomorphism
(= homomorphism of an algebra into itself), and automorphism (= bijective
endomorphism) in the usual way. We will also have need of antiautomor­
phisms of an algebra A, that is, bijective linear mappings a: A -> A
satisfying cr(xy) = o-(y)a(x) for all x , y ^ A. For example, it is a conse­
quence of (LAI) that for Lie algebras the mapping x -> - x is an antiauto­
morphism (of period 2).
A subspace J of an algebra ^ is a left ideal (respectively a right ideal) of
A if for all X e / and for all y ^ A, yx E: J (resp. xy g /). A subspace that is
both a left ideal and a right ideal is called a two-sided ideal, or simply an
ideal. An ideal J (left, right, or two-sided) of an algebra A is called proper if
J ¥=A, For Lie algebras the identity (LAI) shows that the distinction between
left and right is superfluous, so there are only ideals. In the case of a Lie
algebra g it is easy to see that the centre of g is an ideal of g.
Let A be an algebra, and let X and Y be subsets of A. Define X Y (or
[X,Y] if ^ is a Lie algebra) to be the set of all finite sums of products xy
4 Líe Algebras

where x and y e y. If ^ or y is empty, then X V •= (0), If AT or y is a


subspace, then X V is clearly also a subspace. Notice that AA (or [A, A]) is
always an ideal of A. When ^4 is a Lie algebra [A, A] is called the derived
algebra of A and is denoted by D (^).
Let A be an algebra and J an ideal of A, Then the quotient space
A/J + / U ^ A } can be made naturally into an algebra by defining
xy ■•=xy, where : A ^ A / J is the natural quotient map. The resulting
algebra is called the quotient algebra of A by 7. The natural mapping is an
epimorphism of algebras. If A is an associative algebra (resp. a Lie algebra),
then so is A /J,
We have just seen how an ideal J of an algebra A gives rise to a quotient
algebra and a natural homomorphism of A onto A/ J . Conversely, given any
homomorphism f . A - ^ B o i algebras, we retrieve an ideal in the form of
k er(/) := {x ^ A \ f ( x ) = 0}. The relationship between ideals and homomor­
phism is given in the standard type of way:

Proposition 1
Let f:A B be a homomorphism o f algebras, and let J be an ideal o f A and
:A A / J the natural mapping. I f J ^ ker{f), then f factors through J. In
other words, there exists a homomorphism g : A / J B that makes the follow­
ing diagram commute:

A B

A /J
The mapping g is unique and its kernel is (k e r(/))//. In particular, if
J = ker(/), then g is a monomorphism. □

If {Aj\j e J} is a family of algebras then the K-space ® j ^ j A j may be


given the structure of a IK-algebra by defining a multiplication via
( == for all Oj, bj ^ A', j ^ J
(each sum contains only finitely many nonzero terms). This algebra is called
the direct product of the algebras Aj and is denoted by (or
A^ X ••• X if J = {1,..., n}). Note that this definition does not coincide
with the categorical definition of the direct product Tlj^jAj, which we will
not need. Many authors use the notation ®j ^ j A j and call it the “direct sum
of the algebras.” We prefer not to use this notation because it is not
uncommon to be in a situation in which we have a direct sum (in the ordinary
vector space sense) of subalgebras of an algebra that is not a direct product.
An algebra A ¥= (0) is usually called simple if it has no ideals except (0)
and itself. In the case of Lie algebras the trivial one-dimensional Lie algebra
is excluded, and the definition can be restated: g is simple if it has no ideals
1.1 Basic Definitions

different from (0) and itself and if [g, g] = G- We will say that a Lie algebra g
is semisimple if g is isomorphic to a direct product of simple Lie algebras.
Throughout the rest of this book we will make the following assumptions
on associative algebras:
AAl. Any space referred to as an associative algebra is assumed to have
an identity element.
AA2. Any subalgebra of an associative algebra A is assumed to have an
identity element, and this identity element is the same as that of A,
AA3. All homomorphisms between associative algebras are assumed to
carry identity elements to identity elements.
According to AA2, proper ideals are not subalgebras [an ideal (left, right, or
two-sided) of an algebra A that contains the identity element of >1 is ^
itself]. When we speak of the subalgebra of A generated by a subset 5 of
we mean the smallest subalgebra of A containing S and 1^. Notice that
because of AAl, any associative algebra has a copy of the base field K as a
subalgebra.
If A is an associative algebra over IK then an A-module is a vector space
M over K and a bilinear mapping A X M ^ M, {a,m) ^ a - m, satisfying
{ab) • m = a {b • m)
1'm =m
for all a,b Ei A, m E: M.
Example 1 {n X n matrices) Let MJiK) denote the set of n X n matrices
with entries in the field K. Then M„(IK) is a vector space over IK in the usual
way, and it becomes an associative IK-algebra with identity when multiplica­
tions is taken as the usual matrix multiplication.
Example 2 (Algebra of endomorphisms) The vector space Endj^(F) of
IK-linear transformations of a vector space V over IK into itself becomes an
associative K-algebra by defining fg = g (composition of mappings) for all
f , g ^ End|,^(K). If V is of finite dimension n, then Endj^(F) and M„(IK) can
be identified by fixing a basis of V.

Example 3 (Lie algebra of an associative algebra) Let A be an associative


K-algebra. Consider the mapping [*,•]: A X A A defined by [x,y] xy
- yx, where xy is the product of x and y in A. Then A, taken as a vector
space together with this new composition law, becomes a Lie algebra. Only
the Jacobi identity is not immediately clear. We have
Jac(o:,y,z) = [ x, y z - zy] + [ y, z x - xz] + [ z , xy - yx]
= xyz —xzy — yzx + zyx + yzx — yxz — zxy + xzy
-\-zxy - zyx - xyz + yxz = 0.
This algebra is denoted by Li&(A) and is called the Lie algebra of the
associative algebra A.
6 Lie Algebras

Example 4 (General linear algebra) If in Example 3 we set A = End(F)


for some vector space V over K, then Lie(^) is commonly denoted by gI(F)
and called the general linear algebra of the vector space V, Any Lie
subalgebra of- qI(V) is called a Lie algebra of linear transformations (of V).
If dim(K) = Az is finite and a basis of V is chosen, then we can identify
End(F) with M„(1K). We will denote Lie(M„((K)) by gI„(IK).

Example 5 (Special linear algebra) Let ^I„(1K) = [X ^ M„(IK)|tr(A!') = 0}.


It is well-known (and easy to verify) that if X and Y belong to M„(K), then
tr(AT) = tx(YXl Thus tT([X,Y]) = t i (XY - YX) = 0. As a consequence
§I„(IK) is a subalgebra of gI„(IK); indeed [gI„(lK), gI„(IK)] c §I„(1K). Let 1„
denote the identity of Then for all X e M„(IK) we have

Z = i(tr(Z ))l„ + | z-^ tr(Jf)l„J

so that gI„(lK) = IK1„ © §I„(1K); in fact gI„(IK) = X §I„(IK) (direct prod­


uct of Lie algebras). Clearly dim §I„(IK) = - 1. The Lie algebras of type
§I„(1K) are called special linear algebras.

Notice in Example 5 that though ^I„(IK) is closed under brackets, it is not


closed under usual matrix multiplication whenever n >2. In particular
^I„([K) is generally not a subalgebra of M„(1K) even if we disregard our
convention about existence of identities in subalgebras. This already shows
that if we have an associative algebra A, we cannot expect the Lie substruc­
ture of Lie ( ^ ) to be a simple reflection of its associative substructure. Every
subalgebra of an associative algebra gives rise to a Lie subalgebra of the
associated Lie algebra, but not vice versa. However, we will see in Section 1.8
that any Lie algebra is a subalgebra of Lie {A) for some associative alge­
bra A.

Example 6 (§I2(IK)) The three-dimensional Lie algebra ^l2(IK) plays a


fundamental role in Lie theory. We fix the following standard basis for
^l2(K):

‘ “ (2 i)’ o)' '■"(J -W


Simple calculations give

[/ze] = 2e,
[hf] = - 2f ,
[ef] = h.
1.1 Basic Definitions 7

Notice in particular that e, f , h are eigenvectors for ad h with eigenvalues


2, - 2,0 and thus that

^I2(1K) = IK /0 Kh 0 Ke

is an eigenspace decomposition of §I2(IK). Using this, it is straightforward to


show that is simple. In fact §I„(IK) is simple for all n >2. This is also
not hard to prove directly, but we will see it later in a wider context.

Example 7 (Quaternions) Let H be the usual real quaternion algebra:

H = Kl 0 Ri 0 Uj 0 Uk

with ij = k = —ji, jk = i = —kj\ ki = j = —ik, H is well-known as the


simplest example of a noncommutative associative algebra which is a division
algebra (every nonzero element has a multiplicative inverse). Observing that
for Lie(H) we have

L L
2 ’2
k
2’
L i
2 ’2
i
2’
k i
2 ’2
J
2’

and

[1,R/ 0 Ry 0 = 0,

we see that u — R/ -h R; -h RA: is a subalgebra of Lie(lHI) isomorphic to R^


with the standard vector (cross) product and

Lie([H) = R1 X u.

Example 8 (Geometric description of ^l4(R)) Let Si,S2, S2 be an or­


thonormal basis for the euclidean 3-space E, and let

A == i Sj\i # yj c + Zc2 ^^3*

These 12 vectors define the vertices of a cube-octahedron as shown below.


The edges join those pairs of vertices {y, 5} for which y • S = 1. One may
observe that for such a pair, y — 8 (and 5 —y) e A also. Thus 6 = (5 - y)
+ y is the sum of two elements of A. In fact every pair {a, 0} such that
a + /3 e A is of this form:

a 0 2 = ( a + j 8 ) - ( a + j 3 ) = a * a = ) 3 - j 8 = > a - j 3 = —1.
Lie Algebras

Then a • (a + p) = 1 and ((a + j8) - a) + a = a + )8. In this case we de­


fine

(1 if the edge from a to a + /3 is positively


oriented,

- 1 otherwise.
We construct a Lie algebra g as follows (see Figure 1.1): As a vector
space, g is a 15-dimensional vector space with basis S2, £3, and a set of 12
linearly independent vectors indexed by the a e A. Identify E with the
span of £1, 82, ^3 in g. The Lie bracket is defined by

( 3) [ e , ,e j = 0 fo ra lli,/,

( 4) [e,,Xj = = isra)X,
j'sgn(a,/3)A'^+^ ifa + ^ e A ,
( 5) [ X^ , X^ ] = l - a ^ E ifi3=-a,
\0 otherwise.

The reader need not worry about verifying the Jacobi identity. We will do it
more generally later on. In fact g = §14(1^). The point of the example is to
see how remarkably the Lie algebra is related to an intrinsically beautiful
geometric object. This is not a fortuitous accident. There are whole families
of similar examples which we will take up beginning in Chapter 3.
Remark 1 There is no compelling reason for restricting the definition of
algebras to K-spaces. For example, in Section 3.4 we will encounter Lie
algebras over Z.
1.2 Tensor, Symmetric, and Exterior Algebras 9

In general a Lie algebra over a commutative ring A is an ^-module M


together with a bilinear map
[•, ]: M X M ^ M
satisfying conditions (LA2) and (LAS) above. [Note that (LAI) and (LAS)
need not to be equivalent.] The concept of an arbitrary algebra over A is
defined along the same lines
Another abstract construction that we will encounter is that of extension
of the base ring. Let M be a left ^-module, and let B be a ring containing
A. Consider B as a right ^-module and form the tensor product
B (S)^ M.
This has a natural left 5-module structure satisfying
X ' { y ^ m ) = xy ^ m ,

which is called the 5-module obtained from M by extension of the base ring
from A to 5.
If M is an ^-algebra and 5 is commutative, then the 5-module B <S>^M
has a natural 5-algebra structure satisfying
( j c 0 m ) ( y 0 n ) = xy (S> m n .

The most common occurrence of this is when ^4 is a field and 5 is a field


extension of In this case if is a basis of M over A , then we can
think of 5 as the 5-space with the same basis.
Finally, if A and 5 are as above and M is a 5-module (resp. 5-algebra),
then by restriction of scalars from 5 to >1 we give M an ^-module (resp.
y4-algebra) structure.

i.2 TENSOR, SYMMETRIC, AND EXTERIOR ALGEBRAS

Let us begin by constructing, in an obvious way, the polynomial ring in a


number of noncommuting variables. To this effect let X = be a set of
symbols. By a monomial in X we will understand an ordered sequence or
string Xj^Xj^ • • • Xj^ of (not necessarily distinct) elements of X. We let
M = M i x ) be the set of all monomials in X and make the following
conventions:

1. The empty monomial (i.e., the string with no elements) belongs to M.


This monomial is denoted by 1.
2. If equal consecutive symbols appear in a monomial, then these can be
grouped together according to the usual exponent notation; for exam­
ple,
^ ’
we will write jc,J \ jcfJ l jc,j \ jc,J 3 instead of x,J l jc,J 2 X:J 2 jcJ,2 X:J] X:J 3 (but X:j \ jc,J 2 jc,j \ ^
^ hV J 2 if a
10 Lie Algebras

We now define a multiplication • on M by juxtaposition. Thus given monomi-


als m = jc,J l X:J 2 XjJp and m' = XfJ l XfJ l x,f, we can obtain a new monomial
m m = XjXj^ XjXjfXy^ • • • Xy/ called the product of m and m'. This
multiplication is obviously associative, but not commutative if card J > 1.
The empty word 1 acts like an identity in the sense that m • 1 = m = 1 • m
for all m e Af.
M i x ) is called the free monoid on X, If N is any monoid and f : X - ^ N \ s
any map, then there exists a unique extension of / to a monoid homomor­
phism /: M i x ) N, namely fixj^ • • • Xj) = fixj^) • • • f i x j ) .
Now let A i X ) be the K-space admitting M as a basis. Then every element
of A i X ) can be written uniquely as a finite linear combination of the form

where the a,^’s e IK and the m^’5 g M. We next define a multiplication on


by bilinear extension of the multiplication on M; in other words,

( E «,«>,) • (E&yn»>) = E«,Am< ■"*;•


Thus defined, A i X ) is an associative algebra (the element 1 is its identity).
^(AO is called the free associative algebra generated by X. Intuitively we
may think of A i X ) as being built out of linear combinations of products of
elements of X subject to no other constraints than the resulting structure be
an associative algebra. The precise definition of a free associative algebra is
the following:
Let AT be a nonempty set. A free associative algebra on X over IK is an
associative algebra A together with a mapping i:X A such that for every
associative algebra B over K and every mapping f :X B there exists a
unique homomorphism f : A - ^ B such that In other words, the
diagram

A / /
B
commutes.
Note Recall that our conventions on associative algebras are in force.

A well-known example of a free associative algebra is the polynomial ring


K[x] in one variable. If A is any ring containing IK (which amounts to saying
that A is an algebra over IK) and a ^ A is arbitrary, then there exists a
unique homomorphism IK[x] A such that x ^ a, namely the one given by

i;c ,x ' E c ,a '.


1.2 Tensor, Symmetric, and Exterior Algebras U

It is easy to see that the algebra A{X), as constructed above, together


with the identity map id: X A ( X ) is a free associative algebra on X. If A
is any associative_ algebra and f : X A is any map, then there is a unique
linear mapping / : A ( X ) -* A defined by

• • ■^ 0 = f ( ^ h ) • • • f ( ^ 0 /(1 ) = 1-

It is trivial to see that / is a homomorphism of algebras and that /« i d = / .


Now we will show that (>l(Ar),id) is unique in the following sense: If
{A', i') is another free associative algebra on X , then there exists a unique
isomorphism
g:A{X)

such that the diagram

A{X)

commutes.
The argument used to prove this is a very standard one that follows more
from the form of the definition of freeness rather than the particular context
of associative algebras. (More precisely it is a result about universal objects in
categories.) We write down the argument here to serve as a model for all
future occasions when universal objects are defined.
Let {A, ¿X ( A , V) be any two free associative algebras over X. Using the
definition of freeness on A and A in turn, we find unique homomorphisms
g: A A , g': A ^ A such that both inner triangles in the diagram

commute. Thus the outer triangle of mappings

X
\ g ^8
is also commutative. On the other hand, the diagram

id

commutes. The definition of freeness says that only one homomorphism k:


A ^ A exists, making the diagram

X — ^ A

commutative. Thus g' °g = id^. Likewise g^ g' = id^r, and we have g' = g~^,
proving that g is an isomorphism.
Return to the initial construction of A(X). For each n e we let M%X )
denote the set of monomials involving precisely n symbols; for example,

X: Xi X: X J
J\ J2 Jl J3
M \X ), w hilel e

For each n e Z define

m|in € M " (A r),a , if « > 0,


A%X) =
1 ( 0) if n < 0.

Then
A{X) = © A \ X ) ,

and evidently
A^{ X) A^{ X)
for all m and n in Z. We will see in Section 1.3 that this means that A{ X) is
graded by Z.
We will make frequent use of free associative algebras in the sequel, often
in the following way: Let V ¥= (ff) be a IK-space with basis [Vj}j^j, and
construct the algebra A ( X ) as above. We can now identify K as a vector
subspace of A ( X ) via the injective linear map defined by Uj •-> Xj. In this way
we can think of A{ X) as being an associative algebra constructed out of V.
We denote this algebra by T{V) in order to emphasize the identification
between V and the K-span of Xy’s. If we do so then T{V) can be thought as
being the space whose elements are finite linear combinations of the form
1.2 Tensor, Symmetric, and Exterior Algebras 13

where the belong to K. The algebra T(V) is called the tensor


algebra of V. For all n ^ Z the linear span of the monomials involving
precisely n elements of V form a subspace T%V) =A%X) . Of course the
way in which we have constructed T(V) is basis dependent. It is not hard to
see that a different choice of basis leads to a naturally isomorphic object.
(See below for a basis-free constructing of T(V).)
The construction of the free associative commutative algebra K[X] on a
set X = {Xj}j^j is completely analogous to that of A ( X \ but we now allow
the X:’s to commute. For example, x,- x,- x,- and x f x f x : will now be one
and the same monomial. (Formally M is now the free commutative monoid in
X and K[X] its algebra over K.) Of course K[X] is nothing but the familiar
polynomial algebra over IK in the (commuting) variables Xj, j e J. If < is a
total-ordering in J, then the set of monomials

{ X:J l X:J 2 X j \ j \ < j 2 ^ ■■■ <j „;n

is a basis of the IK-space K[X], The universal nature of K[X] is formally


expressed as follows: Given any map / from X into an associative commuta­
tive algebra A, there exists a unique homomorphism / from IK[A"] into A
such that the diagram

X
f

commutes. Moreover K[X] is, up to isomorphism, the unique associative


commutative algebra with this property.
The homogeneous monomials of degree n generate a subspace K[ XT of
K[ Xl Clearly

Finally, notice that K[X] can be thought as a quotient of A(X): More


precisely K[X] - A ( X ) / J , where J is the ideal of A ( X ) generated by all
elements of the form XjXj^ - Xj^Xj^ with j\ and j'2 in J.
Suppose now that V =5^ (0) is a IK-space and that j ¡s a basis of K As
before, we can identify V with a subspace of IK[A"]. If we do so, then the
elements of 1K[A!'] can be thought of as linear combinations of the form

Jl,J2y-yJn
but where now UjUj^ = VjVj^ for all ^ J* When emphasizing this corre­
spondence, we will write 5(F ) instead of 1K[A"], and call this algebra the
symmetric algebra of F. As before, one can easily see that up to isomorphism
this construction is independent of the choice of basis.
14 Lie Algebras

The descriptions we have given of T{V) and S(V) are sufficient for most
of our purposes. We will, however, assume at certain points in the text that
the reader is familiar with the concept of balanced maps and tensor products
of modules over noncommutative rings. (We have already used this assump­
tion in Remark 1.) In this language we have the following (basic free)
constructions of the tensor, symmetric, and exterior algebras.
If ^ is a commutative ring and M is an ^-module, we let,

T ^ ( M) = M ® ‘ 0 M, p-times, p > 0,

T ( M) = © r ^ ( M ) .
p > 0

This last is the tensor algebra of M where the algebra structure is given via
the canonical isomorphism

T ^ ( M) 0 T^ { M) =

Let J be the two-sided ideal of T(M) generated by all elements of the form
X 0 y - y 0 jc, jc, y e M. This is a graded ideal and hence

S{ M) : = T { M ) / J = © 5^(M ),
p > 0

where

S^{M) ^ T ^ { M ) / J f l T ^ ( M)

(see Section 1.3 for gradings). 5(M) is the symmetric algebra of M.


Finally, if I is the ideal of T(M) generated by all elements x 0 x with
X e M, then I is graded and the exterior algebra A(M) of M is defined by

A (M ) = T ( M ) / L

Then

A ( M ) = © AP(M)
p^O

with A^i M) = Tf>(M)/U n r^(M )).


1.3 Gradings 15

1.3 GRADINGS

In this section we introduce the concept of graded algebras. Gradings appear


in every aspect of this work and are fundamental to the subject. In fact, once
we move away from the traditional finite-dimensional theory of Lie algebras,
there are few established ways of pursuing the subject. Either one assumes
that the Lie algebra carries some topological structure, or one assumes that it
carries a grading or at least a filtration. In this book we depend on gradings,
usually with the degree spaces being of finite dimension.
Let K be a K-space and Q an abelian additive group. By a grading of V
by Q we will understand a family eq subspaces of V such that
V= © F “.
a^Q
Given such a grading, we will say that V is graded by Q or Q-graded. For
each Of e 0 we call F “ the degree subspace^ of degree a. If i; e Ei;“ ^ F,
i;" e F “, then is called the homogeneous component of v of degree a. An
element lying in a degree subspace F “ is said to be homogeneous (of degree
a). Notice that in this sense 0 is homogeneous of every degree, whereas a
nonzero element of F can be in at most one degree subspace. The fact that
(2 is a group comes into the picture when we consider graded algebras.
Let A be an algebra over K, and let {A^}^^ q be a ^-grading of A (as a
vector space). We say that the grading is compatible with the algebra
structure of A and that ^4 is a Q-graded algebra if
A^A^ (z A ^^^ for all a ,p ^ Q .
Assume that A is an algebra graded by Q and that F is an ^-module that
is also (2-graded. These gradings are said to be compatible if
c for all a ,p ^ Q .
In this case we say that F is a graded ^-module.
Referring to Section 1.2, we see that
A(X) = © A \ X )

T{V) = © r"(K )
neZ

K[AT] = © K'-iA']
neZ

5 ( F ) = © 5 " (F )
neZ
are all Z-graded algebras. These gradings are called the total gradings.

^We avoid using the terminology “ homogeneous subspace” because this concept has an entirely
different meaning elsewhere in mathematics.
16 Líe Algebras

In the case of A{ X) or K[X] there is another obvious grading available.


Let X = and let

Zj := © Z,-

be the direct sum of card(J) copies of Z. Let aj e Zj be the element whose


yth component is 1 and all other components are 0. Given a monomial

"* ■■■
we assign to it a degree:

deg(m) =aj^ + +aj^ e 2 j.

Evidently for all m, n e M(AT),

deg(ni • n) = deg(m) + deg(n).

It follows that if we define A “( X) to be the linear span of all the monomials


of degree a, then

A { X ) = 0 ^ “(AT)
aeZj

and A{ X ) is Zj-graded. In precisely the same way IKlAf] is Zj-graded.


However, T(V ) and 5(F ) cannot carry a natural Zj-grading since they are, in
principle, free of any particular basis.
If A is any <2-graded algebra and 0 is the identity element of Q, then for
all a ^ Q ,

( 1)

In particular A^A^ c A^. If A is associative with identity element 1, then


\ EiA^ (Exercise 1.4). Thus in all cases A^ is a subalgebra of A.
Let V and W be (2“graded IK-spaces where Q is an abelian group. Let
A e g . An element / g Homn^(K, is called homogeneous of degree A if

/ ( F “) c W A + a

for all a e g . Define Homn^CF, W Y to be the K-subspace of Hom,K(F, W )


consisting of homogeneous elements of degree A and

grHomK(F,IF) := L HomK(F,IF)" cHomK(F,IF).

This sum is direct. Indeed, if e Homu^CF, W)^ and Exeg/A = 0 (sum with
13 Gradings 17

finite support), then for e F" of arbitrary degree a e ¡2 we have

0 = L /x(i^a)-
A

But e so that = 0 for all A e g . Thus

grHomK(K,»^) = e H o m ^ (F ,lT ) \

but notice that in general grHoni|,^(F, PF) c Honi|,^(F, PF) if V is infinite


dimensional.
If /: F PF is homogeneous of degree 0, we say that / is a graded
homomorphism.

Example 1 Let F be a 0-graded K-space. View K as a (2-graded K-space


via
= K
K“ = (0) if a ^ 0.
Then
grHomo^(F, IK) c F*
and
grHom„^(F, K)
= { / G F * |/ ( F “) = (0) for all but a finite number of a e 2}.

It is customary to denote grHom„^(F, K) by or g rF * and call it the


restricted dual of F. Of course

^es = e Kts
Ae0
where
F4 = H om j,(F \K ) = (F")*.

Example 2 Let F be 2-graded. Let gr E n d F ):= gr Horno^(F, F). If


/ e End,^(F)'>' and g g EndK(F)^ then for all a e 0

/g ( F “) c / ( F “ ‘^®) c r)

so that fg G Endi^jiF)’"^®. This shows that grEndi^íF) is a 0-graded alge­


bra. Similarly [/, g] = fg - g f ^ End^(F)’'+® so that Lie (grEnd^íF)) ~
gr qK F) is a 2-graded Lie algebra.
18 Lie Algebras

Let be a subspace of a Q-graded space A. For each a ^ Q define


C\ B. We say that 5 is a graded subspace of ^ if B (the
sum is necessarily direct). To say that B is a graded subspace of A is
equivalent to saying

(2) whenever a = ^ ^ B,
a^Q
then each homogeneous component E: B,

That is, if B is graded and a = e B, then a can be written in the form


Eb“, where e B°^ = A^ C\ B. Thus e B for all a. Conversely if
(1.9) is true, then e A"^ C\ B a e EB“. Thus B = EB “.
If i? is a graded subspace of A , then the quotient space A / B has an
inherited grading with

{ A / B Y := {A^ ^ B ) / B = {a +B\a e A ^ ) / B.

The point is that if Eiz“ B = B, then E a“ e B, so each e B by (2).


Thus

A/B = © {A/By,

In addition the natural mapping

A^ ^ {A^ + B ) / B

has kernel A°" r \ B = J?"; thus

A ^ / B ^ =- {A^ + B ) / B = { A / B Y ,

Let A = © ^ g g ^ “ b e a graded algebra. A 2-sided ideal / of ^ is called a


graded ideal if / is a graded subspace of If / is a graded ideal of A, then
the quotient space A / J = ®^^sq ( ^ / / ) “ is graded in a way compatible with
its algebra structure:

a+/3
{A /J)\A /jY ^{A /J)

If 5 is any subset of homogeneous elements of A, then both the


subalgebra and the ideal of A generated by S are graded. We leave this as an
exercise (also see the next example).
1.3 Gradings 19

Example 3 Consider the free associative algebra A ( X ) with its total


grading. Let J be the ideal of A ( X ) generated by all the commutators
x^Xj - XjX^, i, 7 e J. A typical element of / is a finite sum

s= Y. aij{XiXj-XjXi)bij,

where and ^ A ( X ) . Evidently j;,0Cy - XjXf is homogeneous of degree


2. Let Oij = ^ A ( X ) " , and similarly let = T,bjp. Then we
can further decompose 5 as a sum of elements

- XjXi)bjf,

which are homogeneous of degree m + n + 2 and which are also in J. This


shows that J = so / is a graded subspace, hence a graded ideal.
Then A { X ) / J = IK[Ar] inherits the Z-grading with

{ A{ X) / j y = { A \X ) + /)//.

One ej^ects that K[X] will end up with its standard Z-grading. Indeed it
does: Let : A ( X ) A ( X ) / J be the natural quotient homomorphisni. Since
A^^(X) is spanned by the monomials m = ♦ Xy, j \ , . • •, Jn ^
{ A { X ) / J Y is spanned by the monomials Xj^ • ** Xy, which are precisely the
elements of the free commutative associative algebra that generate the
subspace of degree n.

Let F be a (2‘graded space. We give the tensor algebra T(V) of F a


¡2-graded algebra structure as follows: Given a ^ Q define

T ^ ( V) := 52 F “* ® ® F “" = © ® • • • <8) F “"


a, + ••• +a„=a Of, + ••• +a-=a

(where by convention the right-hand side is IK if n = 0). Then the sum of the
r “(F ) is direct and

r ( F ) = © r “( F )
a^Q

is the desired j2-graded algebra structure on T(F). This is said to be induced


from F.
20 Lie Algebras

The defining ideals of both the symmetric and exterior algebras are
homogeneous, so we obtain ¡2-gradings

S ( V) = © 5 “( F )
a^Q

and

A ( V ) = © A "(F).
a^Q

1.4 VIRASORO AND HEISENBERG ALGEBRAS

In this section we introduce the Virasoro algebra, which is a unique object


(for a given field IK), and the Heisenberg algebras, which are a special class of
nilpotent Lie algebras. These algebras are quite important in contemporary
physics and provide excellent examples of graded Lie algebras. To make the
definitions seem less arbitrary, and also to see how Lie algebras appear
naturally as algebras of operators, we give explicit realizations of the centre­
less Virasoro and the finite-dimensional Heisenberg algebras.
Let X = be an infinite-dimensional vector space over IK with
basis {L„}; « e Z. Make X into an algebra by defining multiplication through

= { n - m)L„ for all m^n ^ Ж.

It is trivial to check the skew-symmetry and the Jacobi identity on this basis,
so AT is a Lie algebra. We call this algebra the Witt algebra or centreless
Virasoro algebra and denote it by SB.
The Virasoro algebra S3 is a one-dimensional central extension of SB (see
Section 1.9 for central extensions).
Define

ss = a: 0 iKc,

and define multiplication by

(la) [L ^ , L„] = (n - m )L„+„ +

(lb) [c ,L „ ] = [ L „ ,c ] = 0 f o r a llm ,n e Z .

Here d is the Kronecker symbol. If m - n , then bracket of and L„ is


exactly as it was before. The mysterious 1/12 is irrelevant at this point. By
using a basis ^ where a is some nonzero scalar, one can
replace the 1/12 in (la) by any other nonzero constant.
1.4 Virasoro and Heisenberg Algebras 21

The verification that 83 is a Lie algebra is straightforward. For example, if


m + n + /? = 0 and m, n, p 0, then

Jac{L^, L„, Lp) = ( p - n)[L„,L„^p] + (m - p)[L„, Lp+„]


+ (n - m)[Lp,L„+„]
= { ( p — n) ( n p — m) + ( m —p ) ( p m — n)
+ (n —m) { m + n - p ) } L q

+ ^ { ( P - n){m^ - m) + ( m - p)(n^ ~ n)

+ (n - m )(p^ - P ) ] c
= 0.
Although SB is a subspace of S3, SB is not a subalgebra of S3. However,
we have an epimorphism
7t : S3 ^ SB
defined by
H^n)=Ln, tr(c)=0.

This gives rise to the exact sequence of Lie algebras

(2) 0 Kc ^ S3 ^ SB 0.

Obviously IKc is in the center S3. In fact it is the center. That is, if
La„L^ + ac is central, then

0 = [L^,La„L„ + ac] = La„{n - “ >n)c

for all m e Z, and it is clear then that all the a„’s are 0. From the point of
view of the representation theory it is S3 and not SB that arises. S3 and SB
display obvious Z-gradings, namely

SB" = KL„ for all n,


IKL„ if Az ^ 0,
S3" =
IKLq 0 IKc if n = 0.

It is interesting to see how the Witt algebra arises naturally when one
considers the space SB (over C) of complex vector fields on the unit circle
U := {e'^\6 0 U}. Each vector field is a linear operator fd/dO on the space
22
Lie Algebras

S (t/) of complex valued C*-functions on i/. is a subalgebra of 9l(^(i/))


since

d d (fdg/dd - gdf/dd)d
( 3)

Indeed, if e §((/), we obtain

d d d I dh\

dg dh d^h
r _ __
f de
jn de
jtk dS^ ’
and similarly

d d d f dh d^h
^de^de ^de de de^'

Then (3) follows from subtracting these two equalities.


Let SB be the subspace of vector fields fd/dO for which / has a finite
Fourier series expansion e C, A: e Z. Then SB has a basis

and

[ L „ L j ] = - i{ e ‘V ^ d / d e - e‘^‘>ke“‘^) —

= {j - k)L^^j ,

showing that SB is indeed a Witt algebra.


Next we introduce the concept of Heisenberg algebras. Suppose that a is
a Lie algebra over K admitting a one-dimensional subspace c with the
following two properties

( 4) [ a , a ] = c,

( 5) [ a ,c ] = (0).

Then if {c} is a basis of c, we have an associated skew-symmetric bilinear


form iff on a defined by

( 6) [x, y] = y)c for all X, y e a.


1.5 Derivations 23

Clearly c lies in the radical of </i, and hence we obtain an induced skew-sym­
metric bilinear form iff on a/c. The definition of e/r, and hence of i/r, depends
on the choice of c, so these forms are unique only up to scalar multiples.
A Heisenberg Lie algebra is a Lie algebra satisfying (4) and (5) above such
that its associated skew-symmetric bilinear form e/f on a /c is nondegenerate.
One can reverse the process above as follows: Suppose we are given a
nondegenerate skew-symmetric bilinear form i/i on a IK^^space Qq. Let a == Qq
© Kc be a one-dimensional extension of Qq. Extend if/ to 2l bilinear form {¡f
on all of a by setting c) = 0 = if/ic, x) for all x e a. Then defining
[x, y] = ij/ix, y)c determines a multiplication on a for which a is a Lie
algebra. Indeed the Jacobi identity is completely trivial. We have a /c = Qq,
and the associated bilinear form is ij/.
Heisenberg algebras arise naturally in the following context: Let S =
IK[jCy]ygy be the polynomial algebra over IK in some set of commuting
variables Xj,j e J. Consider the following linear operators on S:

l(Xj), multiplication by Xj,


d/dXj, the partial derivative with respect to Xj,
1, the identity operator.

The linear span a of all these operators is a subalgebra of gl(5). In fact we


need only note that

VI

to see that a is a Heisenberg algebra of dimension 2(card / ) + 1.


Every finite-dimensional Heisenberg algebra may be realized in this way,
since any finite-dimensional vector space Uq with a nondegenerate skew-sym­
metric bilinear form permits a polarization Uq = b ® c into lagrangian
(= maximal totally isotropic) subspaces.
The Heisenberg algebras we have just constructed offer any number of
gradings. For instance, if for each j, dj is some integer, and if we define the
degree of Xj to be dj, then S = IK[xy]yey is graded by Z, and the operators
l(xj) and d/dxj are homogeneous of degrees dj and - d j , respectively. This
gives a an inherited grading.

1.5 DERIVATIONS

Let ^ be a K-algebra. A linear map 8: A A called a derivation of A if


it satisfies the following familiar rule:

( 1) 5(;0') = x 5 (y ) + 5 (x )y for all x, y e yl.


24 Lie Algebras

Notice that if 5 e End(^) satisfies (1) on all pairs (x, y) where x and y run
through a basis of A, then 8 must be a derivation.
It is simple to see that the composition of two derivations is not necessarily
a derivation. However, as the following computation shows, their commutator
always is

[5 ,5 '](ry ) = 5 (5 '(j^ )) - 5'(5(jry))


= 5(A:5'(y) + S'(;c)y) - 5 '(x 5 (y ) + 8 ( x ) y )
= x8{d' {y)) + 8( x) 8' ( y) + 5'(;c)5(y) + 5 (5 '(^ ))y
- A;5'(5(y)) - 8' {x) 8( y) - 8 ( x ) 8 \ y ) - 8' {8( x) ) y
= a:[5,5'](y) + [5 ,S '](x )y .

Thus the space Der(/1) of all derivations of ^ is a subalgebra of the Lie


algebra Qi(A).
If an algebra A is anticommutative [i.e., if (LA3) holds], then the Jacobi
identity (LA2) holds in A (and A is therefore a Lie algebra) if and only if for
all X e ^ the map 1/. y xy (i.e., left multiplication by a:) is a derivation of
A. Thus, if g is a Lie algebra, and if for each j: e g we define ad a:: g g by

ad a:: y y],

then ad a : is a derivation of g. Moreover the map ad: a : -» ad a: is a Lie


algebra homomorphism from g into Der(g) (we will return to ad in Section
1.6).

Example 1 Let A — 0^^g.<4“ b e a graded algebra, and \e,t d: Q -* K any


group homomorphism (of additive groups). Define 6: A ^ A by linear
extension of 5: ac„ •-> d{a)x^ for all ac„ e A “, a ^ Q. Then 5 is a derivation
of A. Indeed, if a:„ e A “ and x^ e A^, then x^x^ e Therefore

= ( d ( a ) + d(p))x„Xp
= d{a)x^Xp + d (^)x „ x p

= x^{d{P)Xp) + {d(a)x^)xp
= x„8{xp) + 8(x„)xp

Example 2 Let A be an associative algebra. For each y ^ A the map A^:


A ^ A given by Ayi x ^ yx —xy is a derivation of A. Derivations of this
form are called inner derivations of A.
1.5 Derivations 25

We recall for later use the following result:

Proposition 1
Let Vbe a K-space, Т(У) its tensor algebra and S(V) its symmetric algebra.
Then

(i) given any linear map f: V ^ T(V), there exists a unique derivation
8= of T(V) making commutative the diagram

T { V) T { V)

(ii) statement (i) holds if T(V ) is replaced by S{V).

Proof. Let {vj)j^j be a basis of V. Then recall that T(V) can be thought of
as the polynomial algebra IK[f;]; ey in the noncommuting variables Vj.
Recall also that the set of monomials on the v/s (including the empty
monomial 1 e K) form a basis for the IK-space T(V). Hence the map
defined on these monomials by

( 2) 5/(1) = 0,
( 3) 5/( Vj) = / ( Vj) for all j e J,
n
(4) ••• О = Г t'y. ••• • ^Jr,
k= l

for all

admits a unique linear extension (also denoted by 5y:) to an endomorphism of


T(V). It is clear that 8^ is a derivation that satisfies the required property. As
to its uniqueness, notice that for any x ¥= 0, 8jr(x) = 5yr(lx) = 8f(x) +
5y^(l)x => 6y^(l) = 0. On the other hand, if the diagram above is to commute,
then necessarily 8f(vj) = f(vj). This leaves no choice, but (4) if 8jr is to be a
derivation. This concludes the proof of (i).
To establish (ii), we observe that 8jr stabilizes the defining ideal J (ZT(,V)
of 5(K). Thus 8f induces a derivation of 5(K) with the required properties.

Remark 1 Let {Vj)j^j be a basis for a K-space V. Fix an index e /, and


consider the functional / e F* given by linear extension of /: Vj^ 1,
26 Líe Algebras

/: Vj ^ 0, if j ^ Jq. Proposition 1 then shows that the usual partial derivative


is a derivation of the (commuting or noncommuting) polynomial
algebra

An endomorphism / of a K-space V is called locally nilpotent if for each


V Ei V there is an n = n{v) e N such that /"(i;) = 0. If / is locally nilpotent,
Ei V, and n := maxinCfi),. . . , «(¿;^)}, then /" ( f ) = 0 for all v in
the linear span of Thus if dim F < oo, a locally nilpotent endomor­
phism of V is actually nilpotent: /" ( F ) = {0} for some n.
If / e End(F) is locally nilpotent, then

00 fn
E —

is a well-defined endomorphism of F. It is called the exponential of /.

Proposition 2
Let f, g E End(F) be locally nilpotent, and suppose that fg = gf. Then f + g is
locally nilpotent and

ex p (/ + g) = exp / exp g.

Proof. Let V e V, and find N e H such that f^{v) = 0 = g ^ { v ) . Then

since in each term of the sum either ; > or 2A/^ - ; - 1 > AT. Thus / + g
is locally nilpotent, and we have

2 N - 2 ( f ^ g ) ^

ex p (/ + g ) ( y ) = £ — - — (v)
n= 0 ”•
2 N - 2 f j g k

= „ =0
i:
j + k = n L !<' •

N - l f j N - \ g k

= L ^ E fr(^)
j=0 L A:= 0

= e x p /e x p g (z ;) □
1.5 Derivations 27

The conclusion of Proposition 2 is false if / and g do not commute. The


question of whether we can write

exp / exp g = exp( /z)

for some h is difficult and fundamentally related to Lie theory. Formally the
answer is yes, h being determined by the Campbell-Baker-Hansdorff formula
(see Section 1.11). This gives /z as a series

/ + g + ^ [/, g ] + ** higher commutators products of / and g,

which unfortunately need not converge in End(K) ([Bo2] Ch. 2.7).


In the preceding proof we have explicitly mentioned the index N = N(v)
for which f ^ ( u ) = 0. This index will be omitted in the sequel. The reader
ought to keep in mind that the sum T!^^Qf\u)/n\ is finite for all z; e K
whenever / is locally nilpotent.

Proposition 3
Let 8 be a locally nilpotent derivation o f an algebra A, Then exp(5) is an
automorphism o f A.

Proof It is fairly simple to show, by induction on n, that 8 satisfies the usual


Leibnitz rule
^ 8\x)
;=0 j'

for all x , y ^ A and n . We now have

8’’( x ) ' y
exp S(x)exp 8( y) = \ £
[P = 0 ..=0 /
^ /
= E
n =0 \;=o J'- (« -/)!)
8"(xy)
00

= E (Leibnitz)
« = 0 n\
= exp 5(xy).

We conclude that exp 8 is an endomorphism of To show that it is


invertible, we produce its inverse, namely exp(-S) (see Proposition 2). □

Now let g be a Lie algebra. We say that a derivation 5 of g is inner if


5 = ad X for some x e g. (Recall that ad x e Der(g) for all x e g). We call
28 Líe Algebras

an element x ^ Q ad-nilpotent (resp. locally ad-nilpotent) in case ad jc is a


nilpotent (resp. locally nilpotent) endomorphism of g. By Proposition 3, if
X e g is locally ad-nilpotent exp(ad ;c) belongs to the group Aut(g) of all
automorphisms of g. The set

E := {exp(ad x)\x ^ q; x locally ad-nilpotent}

generates a subgroup A u t/g ) of Aut(g). The elements of Aut^(g) are called


elementary automorphisms of g. Notice that E is closed under inversion
(exp(ad x)~^ = exp(-ad x)) but that E is not generally closed under multi­
plication.

1.6 REPRESENTATIONS

We have previously seen several examples of Lie algebras “represented” as


Lie algebras of linear operators on some vector space V. For example, the
Heisenberg algebra in Section 1.4 was made to act as operators on a
symmetric algebra. Again the mapping x ^ ad x makes a Lie algebra act on
itself as operators.
With both of these examples we have started with an abstract Lie algebra
and represented it as linear operators. In applied mathematics or physics the
situation is normally reversed—Lie algebras arise in the first place as Lie
algebras of operators. Yet each of the Lie algebras that we study has many
very different representations. As a simple example consider the case of
§I2(1K). By definition, this is a Lie algebra of operators in a two-dimensional
vector space. On the other hand, its adjoint representation jc ad x makes
it act on a three-dimensional space. We will see later (Section 2.4) that these
two representations are the two simplest nontrivial irreducible finite-dimen­
sional representations of ^12(^)1 in fact they are the two first members of an
infinite family of such representations.
Let g be a Lie algebra, and let F be a IK-space. By a representation of g
on V, we will understand a Lie algebra homomorphism tt: g ^ gI(F). In
other words, tt is a linear map from g into End(F) satisfying

(1)

for all JC, y G g and v ^ V, Through tt, g is made to “act” on V as the linear
operator 7r(jc), jc e g.
By an action of a Lie algebra g on a K-space V, we will understand a
bilinear mapping

•g X F
( x , v ) X ' V
1.6 Representations 29

satisfying
[x, y ] ‘ V = X ' y u —y x
Given an action, we say that g acts on V and that V is then called a
g-module (relative to this action). These two concepts (modules and repre­
sentations) are essentially the same. If tt is a representation of g on F, then
g acts on V via
X ' V = 7 r(x )(f) for all jc e g and v El V.
On the other hand, if g acts on F, then v ^ x • v is a linear mapping of F
into itself. Defining tt: g ^ End(F) by
'7t ( a: ) ( í;) = X • u

determines a representation of g on F. It is simple to verify that the two


processes we have just described are inverses of each other.
Although g-modules and representations of g are just different ways of
looking at the same mathematical situation, we will find both concepts useful.
(Notice that our correspondence is between representations of g and left
g-modules. It is easy to turn this into right action by defining u ' x = —x - v ) .
Given a homomorphism tt: g ^ g l ( F ) , we say that the representation tt
is finite (resp. infinite) dimensional and that F is a finite (resp. infinite)
dimensional g-module if the dimension of F over K is finite (resp. infinite).
Let F be a g-module. A subspace F ' c F is said to be a submodule of F if
F' is itself a g-module. In symbols, if
X ' v' ^ V' for all jc e g and v' e F '.
A submodule F ' of F is called trivial if V' = (0) or F ' = F, and it is
called nontrivial otherwise. If F ' ¥= F, we say that F ' is proper. A module
F ^ (0) is said to be simple, or irreducible, if F does not have any nontrivial
submodules. The equivalent concepts for subrepresentations and irreducible
representations are defined in a similar way.
If F is a g-module, then the set Ann(F) = {jc e g|jc • i; = 0 for all z; e V)
is an ideal of g called the annihilator of F. Of course, if tt: g -> g I(F )is the
representation corresponding to F, then Ann(F) = ker(7r). We say that a
g-module F (resp. a representation tt of g) is faithful if its annihilator (resp.
kernel) is trivial. If Ann(F) = g so that jc • z; = 0 for all x e g, then the
module (or corresponding representation) is called trivial.
If F is graded space and g is a graded Lie algebra, both graded by the
same group Q, and if F is a g-module with the property that g“ • F^ c F “"^^
for all a, p E: Q then we say that F is a graded g-module.

Example 1 (Adjoint representation) Let g be a Lie algebra, and recall the


linear map ad: g ^ End(g) is given by
ad ;«:(y) = [ a:, y].
30 Lie Algebras

Then (1) is immediately satisfied because of the Jacobi identity. Therefore ad


is a representation of g into gl(g). It is called the actjoint representation of
g. The corresponding g-module is g itself where the action is given by left
multiplication. The kernel of the adjoint representation is the center of g. It
follows that this representation is faithful whenever g is centreless, in
particular whenever g is simple.

Example 2 Any graded Lie algebra can be viewed as a graded module over
itself via the adjoint representation.

Example 3 Recall the Heisenberg Lie algebra a as constructed in Section


1.4. At that time we “realized” a as an algebra of operators 5. This actually
constitutes a faithful and irreducible representation of a. If we now view S as
an a-module, the grading of S is compatible with that of a (ibid) and 5 is a
graded a-module.

Example 4 Let F be a vector space. Then gI(F) and §I(F ) act on V by


their very definition, and we obtain the natural representations of gI(F ) and
§I(F) on V,
We list next some basic operations that can be performed with modules
and representations of a Lie algebra g. We will intentionally switch back and
forth between these two concepts in order to emphasize their equivalence.

/ QUOTIENT MODULES
Let F be a g-module, and let F ' c F be a submodule of F. The quotient
space F / F ' can be given a natural g-module structure by defining

x ' V = x - V for all jc e g , and i; e F,

where : F ^ F / F ' is the canonical map. This module is called the quotient
module of F by F'.

2 DIRECT SUMS
Let iTj-: g gl(l^), ; e /, be a family of representations of a Lie algebra g.
Then the space F = has a natural g-module structure by defining

^ •{L = E ‘^j (x)(vj) for all e q ,Vj e Vj,j e J.


'^j£j ' j ej

The representation tt of g in F obtained in this way is called the direct sum


of the ir/s. We denote this by tt =
1.6 Representations 31

3 TENSOR PRODUCTS
Let TTyi g Qi(Vj), ; = 1 ,..., n, be a finite family of representations of g.
Set F = ® • • • (8) and define tt: g End(F) to be the unique linear
map satisfying
n
tt{ x ) { v ^ (8> • • • ® ¿;„) = ^ 0 *• • 0 7Ty(A:)([;y) 0 • • * 0
; =i

for all X e g whenever Vj e l / , for all ; = We claim that tt is a


representation of g in V. Indeed, if x and y e g and if 0 ••• 0 is as
above, then

7r(x)7r(y)(i;i 0 • • • 0 z;„)

= u-(j:)i E Di 0 ••• 0Tr*(y)(y;t) 0 ••• 0y„


\k = l

n I k —\
= L E I^i 0 ••• ®Trj{x){Vj) 0 ••• 0 ir* (y )(i;* ) 0 ••• 0 y „
*= 1\ ;= 1

+ Ui 0 • • • 0 ir*(j:)ir*(y)(i;*) ® ®v„

+ E ® ••• 0 0 • • • 0 TTy(A:)(:;y) 0 • • • 0 D„|.


j=fe+i /

By interchanging the roles of a: and y, we have

(7r(;c)'n-(y) - 17(y)l7(:i))(D i ® ■■■ ® V „ )

n
= E ® ••• ® - ‘^ j ( y ) ' ^ j ( x ) ) ( V j ) 0 ••• 0
y=i
n
= E ® • • • ® ^;([^. y])(i^;) ® ® v„
7= 1

= ® ••• 0 y„).

and hence ir is a representation of g as claimed. This representation is


called the tensor product of the iTy’s and we write ir = tti 0 • • • 0

4 TENSOR REPRESENTATION
Let it: g ^ gI(F ) be a representation of g, and let T(V) =
the tensor algebra of the space V. By 3 above we obtain a family of
representations ir„: g ^ g I(r"(F )) for all n e (where we take ttq: g -♦ IK
32 Lie Algebras

to be identically zero). Then itj = a representation of g in T(F),


thanks to construction 2. This representation is called the tensor representa­
tion of the module V. Notice that by Proposition 1.5.1 and the definition of
7T„, we can characterize TTjixXx e g, as the unique derivation of T iV \
extending the linear map tt{ x ): V V.

5 EXTERIOR REPRESENTATION
Let K be a g-module and consider the tensor g-module T(V), Recall the
exterior algebra of V

A ( V ) = T ( V ) / ( v ® v\v e V).

If jc e g, we have

X ‘ (u u) = X'V<S>V-\-U<S>X-U
= ( x ^ V + v) <S>( x ' V v) - X ' V ^ X ' V - V ® V ,

This, together with the fact that x acts as a deviation on T { V \ implies that
the defining ideal of A(F) in T(K) is a g-module and hence that A(F) has
an induced g-module structure that satisfies
n
X• f\ • • ■ f \ v ^ = ^ ■ / \ x • Vi ^ • f\v^
1= 1

for all a e g and y j,. . . , ii„ e F. The g-module A(g) is called the exterior
module of V. Note that each A"(F) is a submodule of A(F).

6 HO M JV, W) AND g rH O M jV , W)
Let V and W be g-modules. Define a map

g X Hom^CF, W ) ^ H om ^iF, W )

by

(x,f) -^x-f,

where

( x - f ) ( v ) ■■=x-f(v) - f ( x ■v)

for all X G g, f e Hom|,^(V, W), and y e V. This map is clearly bilinear.


1.6 Representations 33

Moreover, if y g g, we have

y] * / ) ( i ’) = [ x, y] - f i v ) - f ( [ x , y ] ■v)
= x - y - f { v) - y - x - f { v ) - f i x - y v ) + f ( y - X ■v)
= ( x - y - f - y -X - / ) ( y ) .

This shows that the map in question gives Hom^iF, W) a g-module struc-
ture. A particularly interesting case is when W = K viewed as a trivial
g-module. Then K* has a g-module structure via

( x - f ) v = - f ( x • v) for all ;c G g , / e K*,z; G K.

Next suppose that both V and W are Q-graded g-modules. Recall the
subspace

grH o m ^(F ,iF ) = © H o m J F ,iF ) " c H o m K (F ,iF ).

If jc G g", / G Hom(|^(K, WY , and v g we have

( x - f ) v = x - f ( u ) - f ( x • u) e

Thus X - / g HomQ^CK, and hence grHomu^CK, IT) is a graded g-


module.
Let g be a Lie algebra, and V and V' two g-modules. A map f : V ^ V ' is
called a homomorphism of g-modules if / is linear and if f ( x • l>) = x • f ( v)
for all X G g and i; g K. The concepts of endomorphism, isomorphism, and
so on, of g-modules are defined in the usual way (see Section 1.1).

Proposition 1 (Schur^s lemma)


Let 7t: g ^ oMV) be an irreducible representation o f a Lie algebra % on a
space V. Then

(i) Every endomorphism o f the %-module V is either trivial or an automor­


phism,
(ii) Let a be an endomorphism o f the K-space V. Suppose that cr com­
mutes with all o f 7r(g). I f there exists an eigenvector o f a in V then a
is a scalar.

Proof If /: K F is a g-module homomorphism, then both k er(/) =


{¿; e V\ f(v) = 0} and im (/) = {f(v)\v g V) are g-submodules of V. Part (i)
follows from the simplicity of V.
If 0-: F F is any K-linear map such that Tr(x)a = cnr(x) for all x g g,
then each of the maps (o- —A • 1), A g K, is an endomorphism of the
34 Líe Algebras

g-module K Suppose there exist v and Aq e K such that aiv) = X^v. It


follows that (o- - Aq ‘ 1) is not an automorphism and hence that it is trivial.
Thus cr{u) = \ qU for all v ^ V, and hence (ii) holds. □

Remark 1 The existence of eigenvectors for cr in part (ii) of Schur’s lemma


is ensured whenever V is finite dimensional and IK algebraically closed.

Let A be an associative algebra and V a IK-space. By a representation of


A on F, we will understand a homomorphism tt: A End(F). In other
words, 77 is a linear map from A into End(F) satisfying for all x , y in A and
V in V:
(2) iriAi'Ki') = 'rr(x)(ir(y)(u)) = (i7-(x)ir(y))(i;),
(3) i7(l)(£ ;) = y.

The concept of representation of an associative algebra is analogous to that


of a Lie algebra except for the (obvious) fact that we now view End(F) as an
associative algebra instead of a Lie algebra. Condition (3) follows from our
convention on homomorphisms of associative algebras (see Section 1.1).
Similarly we define an ^-module to be a IK-space V together with a bilinear
map (a,u) ^ a ‘ V from A X V into V and satisfying
( 4) (ab) ' V = a ‘ {b • v) for all ¿z, 6 g i; g V.
Let F be a module for an associative algebra A, and let M be a
submodule of F. A submodule of F is called a supplement of M if
F = M 0 iV
as an y4-module.

Proposition 2 [Ja2].
Let V be a module for an associative algebra A. The following conditions are
equivalent:

(i) There exist simple submodules (M ,),^/ o f V such that V =


(ii) There exist simple submodules (M,),=/ o f Vsuch that V =
(iii) Every submodule o f V admits a supplement.

A module satisfying the equivalent conditions of Proposition 2 is called


semisimple, or completely reducible. One easily sees the following proposi­
tion:

Proposition 3
Let V be a semisimple module o f an associative algebra A. Then every
submodule of V is semisimple.
J.7 Invariant Bilinear Forms 35

Remark 2 The concept of complete reducibility also applies to g-modules.


We will see later (in Section 1.8) that there exists a one-to-one correspon­
dence between g-modules and left modules of the so-called universal en­
veloping algebra U(g) of g. The algebra U(g) is associative, and hence the
whole theory of modules over associative algebras can be put to work to
obtain information about g-modules.

We finish this section by introducing the concept of absolutely irreducible


representations. We do this for modules over Lie algebras, though a similar
concept exists for modules over associative algebras.
Let g be a Lie algebra over K, and let M be a g-module. Let IK' be a field
extension of IK. By extension of the base field from K to K' (Section 1.1),
consider the K'-Lie algebra
g' := IK' 0 g

and the K'-space


M' = K' 0 M.

It is clear that M' has a natural g'-module structure satisfying


a ^ X ' b <S > m = a b <^ X' m for all ¿z, e IK', x e g, m e M.

Now if M is irreducible as a g-module, then M ' need not be irreducible as a


g'-module (see Example 5). If M' is irreducible for all field extensions IK' of
IK, we say that M is an absolutely irreducible g-module.

Example 5 Let g = IR (one-dimensional real abelian Lie algebra). Let


^ Let p: g gI(M ) be the unique representation
satisfying p(l) = Then p is irreducible. On the other hand, the
corresponding complex representation of C g = C on is not.

IJ INVARIANT BILINEAR FORMS

The concept of a group acting as isometries on a quadratic space is a familiar


one. One has a space V, a bilinear form ( I ) on F and a group G acting on V
so that for all g e G, for all v,w ^ V (g • v\g • w) = (i;|w). The correspond­
ing concept for Lie algebras is as follows:
Let 77: g ^ gI(F ) be a representation of g on a vector space V over IK. A
bilinear form ( I ) on V is said to be g-invariant (relative to ir) if for all
jc e g and for all y, w e F

(1) ( 7 r { x ) v \ w ) + ( v \tt( x ) w ) = 0 .
36 Lie Algebras

In the particular case that V = q and tt is the adjoint representation, (1)


reads

( 2) ([j^ ])lz ) + (y l[^z]) = 0

for all X, y, z e g, and we then say that (*|*) is an invariant bilinear form
on g.
There are minor variations of the preceding definition. When IK = C,
hermitian forms are often more appropriate, especially in physical situations.
In many cases straight invariance is too strong, so instead we have an
antiautomorphism a of period 2 on g and replace (1) by

( 3) (7t( x ) i;1vv) = (v\Tr{(r{x))w).

Rather than trying to discuss all of these variations at once, we will discuss
only g-invariant bilinear forms here and bring up others as they arise. The
general way in which they are used is always much the same. Likewise,
although one could conceive of arbitrary g-invariant bilinear forms, in
practice the only ones occurring are either symmetric or skew-symmetric, so
we will restrict ourselves to these two types.
Suppose that ( I ) is a g-invariant symmetric or skew-symmetric bilinear
form on a g-module V. If IT is a g-submodule, then the invariance of ( I )
implies that the subspace W orthogonal to W defined by

W (i; e V\{u\w) = 0 for all w e W)

is also a g-submodule. If one knows that W is nonsingular (e.g., if (I*) is


positive definite), then W O W = (0); if in addition we assume that V is
finite dimensional, then V = W ^ W . This type of reasoning is standard in
showing that a finite-dimensional g-module is completely reducible. In deal­
ing with infinite-dimensional g-modules, we will have to proceed with more
caution.
The preceding argument applied to the adjoint representation shows that
if a is an ideal of g, then so is a . In particular g is an ideal of g (called
the radical of the form). We conclude that if g is simple, then ( 1 ) is either
trivial or nondegenerate.

Example 1 Define (-I ) on gI„(IK) (or gI(K) for finite-dimensional V) by

( X\ Y ) = i r ( X Y ) ( = tr(M i)).

This is symmetric and invariant:

t r{[ XY] Z) = tr(ATZ - YXZ) = tr(ATZ) - ít(YXZ)


= tr(yZA^) - tr(yXZ) = - t i { Y [ X Z ] ) .
1.7 Invariant Bilinear Forms 37

Since K1 is an ideal of 9l„([K), so is (IKl)-^ = and we arrive again at


gI„(lK) = IKl X (Example 1.1.5). The trace form on ^I„(IK) must be
nondegenerate because ^I„([K) is simple (Exercise 4.1).

Let g be a finite-dimensional Lie algebra over K. The Killing form k :


9 X 9 9 is defined by

k ( x , y) = tr(ad X ad y) for all x, y e g.

Just as in Example 1 the Killing form is symmetric and invariant.

Example 2 (The Killing form on ^ I2([K)). We use the basis e , h , f of


Example 1.1.6. Then the matrices of ad e, ad h, and ad / are

0 -2 0] /2 0 0] 0 0 0
0 0 1 , ad = 0 0 0 , ad/ = -1 0 0
0 0 oj \o 0 -2j 0 2 0

and, for instance,

/2 0 0\
K( e, f ) = tr(ad e a d / ) = tr 0 2 0 = 4.
\0 0 0^

The matrix for k relative to the basis e , h , f is

In particular det(K) = —128 so that k is nondegenerate.

The Killing form is a cornerstone in the conventional treatment of finite­


dimensional Lie algebras. Traces rarely make sense for infinite-dimensional
Lie algebras, and we usually have to construct invariant bilinear forms in a
very explicit way.

Proposition 1
Let (•!*) be an invariant bilinear form on a Lie algebra g. Then every
elementary automorphism o f g is an isometry o / (• | *).

Proof Let X, y e g. The invariance condition reads (ad u{x)\y) =


-(x\ad(u)y) for all w e g. Now suppose that u is locally ad-nilpotent, and
38 Lie Algebras

set <T = expiad u). Then


(a-x\y) = ((expad M)x|y)
1
= D 77 ((ad «)'x |y)
y=0 J'-

= L ( - l ) l 7 7 (^l(adM)'y)
;=0 \J' J

= (j:|e x p (-a d w)y) = (^|cr"^y).


We conclude that (x|y) = (x\o-~^ay) = {(xx\cry), and hence that a is an
isometry of (*| *). Since, by definition, every elementary automorphism of g
is a product of automorphisms of the form cr = as above, the result now
follows. □

L8 UNIVERSAL ENVELOPING ALGEBRAS

Suppose that we are given a representation tt; g gI(K)ofa Lie algebra g.


Through 77 we can realize g as a Lie algebra of transformations on V. Since
7r(g) is a subspace of the associative algebra End(K), we consider the
associative envelope A ^ of 77(g), that is, the (associative) subalgebra of
End(K) generated by 7 7 ( g ) . The nature of the representation 77 is completely
refiected in the associated representation of (g-submodules of V corre­
spond to ^^-submodules of K, etc.). The close connection between g and
A ^ gives rise to the question of what we can say about associative algebras A
that are related to a Lie algebra g in the following way:

1. there is a Lie algebra homomorphism 7: g ^ Lie(^),


2. y(g) generates A as an associative algebra.

We call such an algebra an associative enveloping algebra of g.


In this section we show that there is a universal algebra of this type—uni­
versal in the sense that every other associative enveloping algebra of g is a
homomorphic image of this one. We make considerable use of these univer­
sal enveloping algebras in the sequel.
Let g be a Lie algebra. By a universal enveloping algebra of g we will
understand a pair (U, /) composed of an associative algebra U together with
a map /: g U satisfying the following conditions.
UEl The map i is a Lie algebra homomorphism from g into Lie ( U ) ;
that is, i is linear, and
i([^>y]) = - i {y)i {x) for all x , y e g.
1.8 Universal Enveloping Algebras 39

UE2 Given any associative algebra A and any Lie algebra homomorphism
/: 9 ^ Lic(A), there exists a unique algebra homomorphism /:
Vi ^ A such that the following diagram commutes:

In other words, f = f ' °

The standard type of argument for universal objects (e.g,, Section 1.2)
shows that if g has universal enveloping algebras (U ,0 and then
Vi ^V i' in the strong sense that an isomorphism / between these algebras
exists such that f °i = V. As for the existence of this object we have

Proposition 1
Let q be a Lie algebra. Then g admits an universal enveloping algebra.

Proof. Let T = r( g) = tensor algebra of the space g. We


let u be the two-sided ideal of T generated by all elements of the form
X ^ y - y ® X - [x,y]. Symbolically

u <x ® y - y ® JC - [jc, yjlx, y e 9>t(9).

Let U:= r / u and denote by tt the canonical map from T into U. The
algebra U is clearly associative [the multiplication in U, which will be
denoted simply by juxtaposition, is associative since that is the case in T, and
7t(1) is its identity. Notice also that U # (0) since u c ©">iT"]. Define
/: g U by i = ttIt’i; in other words, i is the restriction of tt to r^(g) = g.
We claim that (U, i) is a universal enveloping algebra of g. To begin with, if
X and y are in g, then i([x, y]) = tt([x , y]) = tt( x <8>y —y 0 x), since [x, y]
= j c 0 y —y 0 x mod u and u = ker tt. On the other hand, tt( x 0 y —y 0
x) = 7r(x)7r(y) — 7r(y)'7r(jc) = /(x)/(y) —z(y)/(jc), and hence U El holds.
Now let / : g ^ ^ be any linear map of g into an associative algebra A
satisfying

(1) f{[x, y\)=f{x)f{y)-f{y)f{x).

By the universal nature of r(g ) (see Section 1.2) there exists a unique
algebra homomorphism f . T ^ A satisfying /(1) = 1 and /( x ) = /(jc) for all
X e g [we are using here the identification g = T^g)]. Because of (1), /
factors through U (see Proposition 1.1). In other words, we obtain an induced
40 Lie Algebras

homomorphism f : U A such that the diagram

commutes. Since f ( x ) = f {x) for all x e g, we obtain the desired commuta­


tive diagram (UE2). □

If 9i and 02 Lie algebras over K and /: 82 is a homomorphism,


then it follows from the definitions that there is a unique homomorphism
U (/): U(Qi ) 11(92) such that

9i ---------> 92

'■'1 U(/) 1'^


U(gi) ^ U(92)

commutes. A direct consequence is that we can view U as a functor from the


category of Lie algebras and Lie algebra homomorphisms (over K) to the
category of associative algebras and their homomorphisms (over K). In the
sequel we will denote the universal enveloping algebra of g constructed
above by U(g), the map i being understood. Notice that U(g) is generated
(as an associative algebra) by the elements i{x), e g.
Let g be a Lie algebra and ir a representation of g into a IK-space V.
Thus TT is a Lie algebra homomorphism from g into gI(F) = Lie(End(F)).
It follows that V extends to an algebra homomorphism, also denoted by v ,
from U(g) into End(F) where U(g) is the universal enveloping algebra of g.
In other words, V becomes a left U(g)-module with the action u ■v = tt( uX v )
for all Me U(g) and y e F. Conversely, if F is a left U(g)-module, we
retrieve a representation ir of g in F by defining v i x X v ) ^ where
i: g -» U. It is simple to verify that these two procedures are inverses of each
other, and hence that the categories of representations of fl and of left
U(g)-modules are isomorphic. °

Note that if F is a g-module and v V, then U(g)y is the smallest


g-submodule of F containing v.
1.8 Universal Enveloping Algebras 41

Our construction seems to give us very little information about U(g).


When g is abelian, however, we have the following:

Proposition 2
The universal enveloping algebra U(f)) o f an abelian Lie algebra is isomorphic
to the symmetric algebra 5(1^) o f

Proof Consider the tensor algebra r(l^) of the space 1^. The defining ideal u
of U(]&) is then given by

n = ( x ( S > y - y ( S f x\x, y e f)>7'(i))

since 'tj is abelian. But T(ij)/n is then the symmetric algebra Sfi)) (see
Section 1.2). □

The next result, which is of utmost importance for all that follows, tells us
how to construct a basis of U(g) from a basis of g. The injectivity of the map
i: g U(g) will be an immediate consequence of it.

Theorem 1
(Poincare-Birkhoff-Witt) Let q be a Lie algebra with a K-basis {xj\j e /}
indexed by some totally ordered set J. Let (U, /) be the universal enveloping
algebra of g. Then the family o f elements

with j\ <j'2 ^ ‘ j'ny ri ^ together with 1 form a basis of the


space U.

Proof Let U = U(g) be the universal enveloping algebra of g as constructed


above. The first part of the proof shows that the displayed family spans U.
The second, and considerably more difficult part, shows linear independence
of these elements by constructing a suitable representation of g in which
they clearly act as linearly independent operators.
Let M be the set of finite ordered sequences of elements of J. Thus

M= < ••• < y „ , n e N } .

The empty sequence, assumed to belong to M, will be denoted by 0 (it


corresponds to the case n = Oin the above notation). If m = (j\, . . . , ; „ ) g M,
we define n to be the length of m and write /(m) = n. We also define

••• Xj^^
42 Lie Algebras

By convention
= 1G = T’®=
Notice that if J is empty so that g = (0), then our present theorem states that
U = IK • 1. This statement is true, as can be seen from Proposition 2. We will
henceforth restrict our attention to the case where J is not empty. We
proceed in several steps:

(a) The family {7r(x„)|m e M) spans U.

Let U' be the span of this set. Clearly and ttCTO are in U', since in
this case the elements in question are either scalars or linear combinations of
the Xy’s with ; e J and both 0 and (;) are in M. We now reason by
induction on n to show that ttCT") c U' for all n. Because tt is linear, it
suffices to show that if jc e T" is of the form x = X: 0 • • • 0 x , , then
7t( jc) g u '. If 7i < * * < we are done. Otherwise, there exists 1 < k < n
such that jj^ > Then
7t{ x ) = TT{Xj^ 0
• • • ®

= tt( xj^ ® • • • ® ® (^ 4 .. ® + [^ 4 ’ ’^ h J ) ® • • • ®
= tt( xj^ ® ® 1 0 X :Jk-\ 1 0 JCJk: 0 • • ®^y„) +
where x' e T^~^, By the induction assumption, 7t( jc') e U'. If the sequence
in the first summand is not increasing, we repeat this procedure until it is.
We conclude that ttCT") c U' for all n. Since T = and 7r(D = U, step
(a) follows.
We make now a technical pause to see that there is no loss of generality in
assuming that the total ordering in J is actually a well ordering. From (a) we
see that if the result fails then there exists elements mj, N M and
nonzero scalars E: IK such that
N

( 2) E = 0-
n= \

Now the indeces appearing on these m are finite in number, say <
j 2 < **' < Jk (where N < l(m^) + • • • +/(m^)). By Zorn’s lemma there
exists a well ordering ^ on J such that j\ < j'2 ^ ' ** Jk - Now (2) shows
that it is sufficient to establish the Theorem for well ordered sets. We assume
henceforth this to be the case for the given total ordering of J.
(b) Let u = ^ ^et of symbols, and let V be the IK-space
constructed with i; as a basis. We intend to give V a g-module structure and
later use the universal nature of U to show the independence of the 7t( x„).
We begin with some notational conventions. If m = (j\, e M and
; e J, we write ; < m in case j < j\. We also agree that j < 0 for all j e J.
If j and m are as above and j < m, we write j m for the sequence
( j , j i , j 2’ -' yJn) ^ Clearly l(jm) = 1 + /(m). Finally, for all « e Kl, we
let denote the subspace of V spanned by all the such that /(m) < n.
1.8 Universal Enveloping Algebras 43

(c) There exists a bilinear map (x, -> x • v from g X K into V such that
for all y e / and for all m e M:

(0 if j < m,
(ii) Xj - - if / > m, m = A:m',
(iii) Xj ■v„ G Ki

We reason by induction on n to show that there exists a map

with the desired properties. If n = 0, then Vq is spanned by U0, and we


define
• ^0 =
which is clearly as desired after linear extension to g. Let n > 0. Then
has < /1} as a basis. Since we already have a bilinear map

Q X V n -i^ V n ,

we define Xj • for all ; e J and m e M with /(m) = n. We do this by


induction on the order of J. If j'q is the first element of J, we set
X:Jo ' = U^:0™

Now let y e J be arbitrary, and suppose that • ¿;^ has been defined for all
k < y. If y < m, we set

Otherwise, m = (y^,. . . , y„) with j > j^. Set m' = (j'2, . . . , ;„), and define
Xj ■ = Xj^ ■X j ■v„, + [xj, • i;„-.

The right-hand side is already defined. Indeed, since /(m') < /(m), Xj • is
already defined. Moreover

Xj-v^= £ c„u„.
l{n)<n

Now is defined, since <j. Also [xy, Xy^]eg, and we


already have the map

Now by bilinear extension we get a map

as desired.
44 Lie Algebras

Before continuing we make an observation: If m = A:m' and r > k, then

(3) Jfr • + y
for some n with /(n) = /(m'), some h where r > h > k, and some v e Vi^^y
Indeed, if r < m', set /i = r and y = 0. Otherwise, m' = hwl' where r > h >
k, and we use step (c)(ii).

(d) The bilinear map constructed in step c makes V into a g-module.


We must verify that x • y ' v —y ' x • v = [x, y] • v for all x, y e g and
V EiV.W suffices to show that the equation

( ;, k; Xj Xj ■ = [xj, x^] • v„.

holds for all ;, A: e J and m e M. We use induction on /(m). Assume,


without loss of generality, that j > k [both sides of { j , k \ v ^ are skew-sym­
metric in j and k].
If A: < m, then (;, k; i;^) is an immediate consequence of (i) and (ii). We
can therefore restrict our attention to the case where m = rm' and j > k > r.
Now (j, k;Uj^) reads

( ;, k, r) Xj - x ^ - x , - - x^ • Xj ■x, ■v„, = [xj, x^] • x , • v^,.

Of course we do not know whether (;, A:, r) holds. But we do know that both
(A:, r, j) and (r, j, k) hold. For instance, (A:, r, j) reads

{ k, r , j ) X^- X^- {Xj ■ v„,) - X , - X k - (Xj ■ = [x^, X, ] ■Xj ■v^.

But by (3), Xj • + u' for some n with /(n) = /(m'), v' e and
j > h > r. Thus both (A:, r; and (k, r; v') hold [the former because r < h,
the latter by induction on Km)]. It follows that (k, r, j) = (k, r; v^^) +
(k, r;u' ) also holds.
Since the induction assumption implies that for all x , y in g,

x-yv^,-yx v„, = [ x , y ] - v ^ ,

the right-hand side of (;, A:, r) can be rewritten as follows:

[ x j , x ^ ] ■X , ■ = j:, • [ X j , Xk\ ■ + [[xj, x^], x,] ■

= •Xj ■X,, ■ - x , - x y Xj ■ + [ [ X p J c * ], x, ] ■v„,,

and similarly for i k , r , j ) and (r,j,k). Now compute ( j , k , r ) + ( k , r , j ) +


(r, j, k) to obtain an identity of the form 1 = 1 + Jac(j:y, x^., x^) • v^, which
obviously holds. Thus, since {k, r, j) and (r, j, k) are known to be true, so is
(;, k, r).
1.8 Universal Enveloping Algebras 45

(e) We now finish the proof of Theorem 1 That the elements of the form

with j\ < • • • < « e Z+ together with 1 span U over K, is an immediate


consequence of step (a), since these elements are precisely the family
{irix jlm e M}. To see that they are linearly independent, we view the
g-module V constructed in step (d) as a left U(g)-module. Let us show that

= for all m e M.

Again use induction on /(m). If /(m) = 0, then m = 0 , = 1 and 1 • ¿;0 =


i;0, so our assertion is clear. If /(m) > 0, we write m = ;m', where j < m' and
Km!) = Km) - 1. Then = 7r(xj)7r(x^.) and therefore tt( x J • =
Tr(Xj) • TT(x^r) • U0 = 7t(Xj) • = Uj^r = v^.
Suppose now that we have a linear combination of the form

L = 0,
meM

where the and all but a finite number of these are zero. Then by
letting this element act on u^, we obtain

meM meM

and hence all c„ = 0, since the y„’s form a basis of V. □

We list below some important consequences to the Poincare-Birkhoff-Witt


(PBW) theorem.

Corollary 1
The images {i(xj)}j^j in U(g) of the basis {xj}j^j o f g form a linearly
independent set. In particular

i: g U(g) is injective.

With this Corollary we can (and will henceforth) identify g with a


subspace of U(g). If Ji e g, we write x e U(g) [instead of i{x) e U(g)].
Notice then that tt{x ^ ® • • • ®x„) = x ^ . . . x ^ for all Xj,. . . , j(:„ e g.

Corollary 2
Let b be a subalgebra o f the Lie algebra g. Suppose that C {y*}*eK ^
a basis o f g such that { x j \ ^ j is a basis o f b and both J and K are totally
46 Lie Algebras

ordered. Then

(i) the injection 6 -» g -> U(g) can be lifted to an injection o f U(b) into
U(9),
(ii) U(g) is a free left (resp. right) U.(b)-module admitting the family
{y*, • • • ^ ^ a basis.

Proof From b -» g ^ U(g) we obtain a homomorphism U(b) U(g)- We


totally order the set JU K (disjoint union) by declaring that j < k for all
y e J and A: e K. The family

{ X,-Jl . . . X,
Jn
y*. • • • ^ ■■■ ...k„,n,m

is then a K-basis of the space U(g). Moreover the subfamily that does not
involve the (i.e., when m = 0) corresponds to a basis of U(b). This clearly
implies (i) and (ii). The statements on right U(b)-modules is proved similarly.

Corollary 3
Let 9i , . . . , be Lie algebras, and s e t Q = X • X {direct product of
Lie algebras). Then

U(9) - U ( 9 i ) ® ® ll(9n)

as K-spaces.

Proof Exercise ^

Proposition 3
Let 8 be a derivation o f a Lie algebra g. Then there exists a unique derivation
Sy o f U(g) extending 8.

Proof If such an extension exists, it is necessarily unique, since g generates


U(g) as an algebra. As for its existence we know that 8 extends to a
derivation 8r of the tensor algebra T(g) of g [Proposition 1 of Section 1.5
together with the canonical injection of g in T(g)]. Now let U be the ideal of
r(g ) defining U(g). Recall that U is generated by all elements of the form

p( x , y) = X ® y —y 0 X — [x, y], ^ ,y ^ 9 -
1.8 Universal Enveloping Algebras 47

By straightforward computation we verify that

= ^ + ^(-*^) ® y - y ® 5 ( x ) - 5 ( y ) ® X

- [ ^ , 5 ( y ) ] - [5(^),y]
= p ( x , 5 ( y ) ) + p { 8{ x ) , y ) .

Hence dj- stabilizes the linear span of the p(x, y), and therefore it also
stabilizes U (use the fact that 8j is a derivation). The induced map 5u:
jj(g) U(g), U(g) = T (g )/u , is then a derivation of U(g) extending 8. □

Proposition 4
l£ t % be a Lie algebra. The adjoint representation extends uniquely to a
representation {denoted by adu and also called the adjoint representation) o f
g on U(g) by which g acts as inner derivations on U(g). More precisely for all
X e g, u e U(g),

adu(x)(«) = XU - ux.

Proof Let X e g. By Proposition 3 the derivation ad x of g extends uniquely


to a derivation adjjX of U(g)- Consider the inner derivation Aj.. m —>xm —ux
of U(g) (see Section 1.5). Evidently 8 = ady x —A^. is a derivation of U(g),
which annihilates g. Thus 8 annihilates the associative algebra generated by
g that is, U(g). This proves that ad„ x = A^. The mapping x A^ is a Lie
homomorphism of g into gI(U(g)), showing that ad„ affords a representa­
tion of g. ^
Remark 1 Recall that in Section 1.6 we introduced the tensor representa­
tion of a given g-module. In the case of the adjoint representation this leads
to a representation ad^ of g on T(g) in which

adj.(x )( yi® ®y„) = ¿ y i ® ••• ® [^>y;]® ••• ®y«

for all X, y j , . . . , y„ ^ 9-

On the other hand, in U(g) we have

a d „(*)(y,...y.)- Z y i - U . y , l - y- ........

from which we see that the mapping tt: T(g) ^ U(9) is a g-module map.
48 Lie Algebras

Remark 2 One ought to be aware that U(fl) has two natural g-module
structures which are quite different.

1. The preceding adjoint representation in which


n

; =i

2. The restriction to g of the natural left U(g)-module structure of U(g)


as an associative algebra in which

^ • (y i •••>’/,) = •••>'«•

Let g = © „ e c 9 “ he a graded Lie algebra. We can then find a basis


e j of the space g consisting of homogeneous elements, say, Xj ^ Q ' for
all y e J where aj e Q (of course the aj need not all be distinct). The tensor
algebra T(g) is then naturally <2-graded by declaring that each monomial
;c ® • • • ® ;c is of degree a, + • • • + «, • The defining ideal u of U(g) is
graded since

® ® ^22] ^ 9“'^ “'-

It follows (see Section 1.3) that the quotient algebra U(g) = T(g)/u inherits
a (2-grading called the grading of U(g) inherited from g. Thus

U(g) = © U(g)“

and
a+/3
U (g )“U (g)^ cU (g) for all a, P ^ Q-

Remark 3 The spaces U(g)“ are independent of the choice of basis [xj}j^j
of g as above. Notice also that if jc e g, then the expression x is homoge­
neous of degree a ” is independent of whether we view x as an element of g
or as an element of U(9> via the identification g ^ U(g).

Proposition 5
Let Q be a Lie algebra. Then U(g) « « domain [i.e., the condition ah = 0 -
a = 0, orb = 0 holds for all a and b in U(g)J.

Proof Let f x } , be a basis indexed by a totally ordered set J U t X =


Proof. Let 1 ^ „ e N} be the basis of U(g) prescribed by the
li - "
1.9 Central Extensions 49

Poincaré-Birkhoff-Witt theorem. Define subspaces U„, n e N, of U(g) as


follows:
Uo = D<,
Uj = IK + g,

U„ = IK + g + gg + • • • + s".

In other words, U„ is the subspace of U(g) spanned by all x : X such that


l(x) < n. Clearly U. q C U i C c U„ c ... and U„U„ c U„ for all
n, m € N. The family (U„)„eN therefore a filtration of U(g).
For each n e N, define IK-spaces U" by U" == where by con­
vention U_i = (0). Let

gr(u) = e U".

The map U„ x \X„ given by multiplication induces a well-defined


bilinear map U" X U” ^ U"+'". We can then extend this to a multiplication
in gr(U). The algebra gifU) thus obtained is called the graded algebra
associated with the above filtration of U(g).
If X = Xj^ . X: G X, we let X e be defined hy x ■= x If is
clear that
(4) The family is a basis of gi<U), and that gr(U) is generated, as
an algebra, by 1 together with the Xj.
Now - XjX^^ = x^x^^ - Xj^Xj^ + Ui = 0, since XjXj^ - Xj^Xj^ =
[Xj^, Xj^] e Uj. It follows that the Xj mutually commute. By (4) gr(U) is
commutative, and we can see that gr(U) - IK[xy]ye j» the polynomial algebra
in the commuting variables Xj. Since polynomial algebras are domains (very
easy to show), it follows that gr(U) is a domain.
We now finish the proof of our proposition. Let a and b be two nonzero
elements of U(g). Then there exist two unique natural numbers n and m
such that a e U„ and b e but a ^ U„_i and b ^ Thus a = a -\-
U„_i and & = ft + are well-defined nonzero homogeneous elements of
gr(U) of degree n and m, respectively. Since gr(U) is a domain, ab is nonzero
and homogeneous of degree n m. But ab = ab + and therefore
ab ^ particular, ab ¥= 0, and hence U(g) is a domain. □

1.9 CENTRAL EXTENSIONS

We have already had occasion to extend the centre of a Lie algebra in the
construction of the Virasoro algebra from the Witt algebra. Basically a Lie
algebra e is a central extension of the Lie algebra g if e has a subalgebra c
50 Lie Algebras

contained in its centre such that e/c = g. The simplest example is e = g X c,


where c is an abelian Lie algebra. Somewhat less trivial is to let c (0) be a
commuting subspace of derivations of g and then give the space e = g 0 c a
Lie algebra structure by defining for all c g c and x e g , [c, x] = c(x) =
- [ x , c], and [c, c] = (0). The Lie algebra e is then a central extension of g;
moreover g is an ideal of e and [e, e] c g. The Virasoro algebra, however, is
not like this: The Witt algebra 93 is not a subalgebra of 93 and [93 , 93] = 93.
We are concerned here with central extensions of this second and more
subtle kind, which are called coverings.
It is not obvious at this point that there is any necessity for making central
extensions, and we do not try to motivate the concept any further except to
say that the representation theory of a covering e of a Lie algebra g may be
vastly richer than that of g itself and that in practice such objects arise
completely naturally. The purpose of this section is to present the concepts of
covering algebras and central extensions in a coherent way and to find
necessary and sufficient conditions for the existence of universal covering
algebras. Finally, as an example, we establish that the Virasoro algebra is a
universal covering of the Witt algebra. The basic theory in this section is
based on [Gr].
Let g and c be Lie algebras over K. By a central extension of g by c we
will understand an exact sequence of Lie algebras:

(CE) 0 0

such that c is in the centre of e. In other words, we have a surjective Lie


algebra homomorphism tt: e g and an injective homomorphism /: c e
such that c = ker(7r) is in the centre of e.

Given another central extension

(CE') 0 c' -U e' 0

of g, then by a morphism of the central extension (CE) to the central


extension (CE') we will understand a pair (<#>, </>o) of Lio algebra homomor-
phisms such that the diagram

‘ "1

is commutative. The central extensions together with their morphisms form a


category.
1.9 Central Extensions 51

A Lie algebra I is said to be perfect if I equals its derived algebra, that is,
if I = [1,1]. The central extension (CE) is said to be a covering of g in case e
is perfect.

Proposition 1
If the central extension {CE) o f % is a covering, then there exists at most one
morphism o f (CE) to a second central extension of g.

Proof Let (</), (/>o) and (</>', be morphisms from (CE) to (CE'). We then
have a commutative diagram analogous to the diagram above by replacing (¡)
and <t>Q by (j)' and </>'o, respectively. It follows that for all z e e , 7r'(/)(z) =
7t( z ) = Tr'(t>'(z), and therefore that

(1) 4>{z) - 4>'{z) G ker(i7') = i'(c').

Now let X and y belong to e. We have

(<l> - <A')([^>y]) = H[ x > y ] ) - <t>'([x,y])

= [4>(x),(f>{y)] - [<!>'(x),<l>'{y)]
= [4>(x) - <i>'(x),<l>(y)] + [<!>'(x),<l>(y) ~(f>'(y)]
= 0,

where the last equality follows from (CE). We conclude that 0 and (¡>'
coincide on [e, e], and therefore on e, since e is perfect by assumption. Thus
(f) = and therefore <f)Q= (this last result by the injectivity of i and /').

A covering of g is said to be universal if for every central extension of g
there exists a unique morphism from the covering to the central extension.
Remark 1 To verify that a covering is universal, it suffices to show the
existence of a morphism from the covering to any central extension. The
uniqueness of each such morphism follows then by Proposition 1.
Remark 2 It is obvious from the definition that any two universal coverings
are isomorphic (in the sense of central extensions). Moreover by Proposition
1 this isomorphism is unique.

We now try to decide which Lie algebras admit a universal covering. The
final answer is surprisingly elegant: All perfect Lie algebras admit universal
coverings. As motivation for the construction needed to establish this result,
we look at the general situation of central extensions in a little more detail.
52 Líe Algebras

Beginning with a central extension

0 9

let us choose any subspace g' of e such that g' g (isomorphism of


K-spaces). Such a subspace can easily be constructed: Let be a
IK-basis of g, and let x'j g e be chosen so that irixj) = Xj. Then the IK-span
g' of the family j is as desired. We l e t g g' be the inverse map of
77. Then for all x, y e g,

'n-{[x,y] - [x',y']) = 0,
and therefore

^ ( x , y) — [x, y j - [x', y'] e c.

Our choice of g' has thus led us to a bilinear map

(Co) a: g X g ^ c

satisfying the following properties for all x, y, z g g:

(Col) a ( x , x) = 0
and hence
a( x , y) = - a ( y , x ) ,
(Co2) a ( x , [ y , z ] ) + a ( y , [ z , x ] ) + a ( z , [ x , y ] ) = 0.

[These properties are immediate consequences of (LAI) and of the Jacobi


identity.]
A bilinear mapping a from g X g into a K-space c is called a 2-cocycle of
g with coefficients in c if it satisfies (Col) and (Co2). We do not intend to
discuss the cohomology of Lie algebras from which this terminology arises
[Ja 3]. Important for us is the fact that any 2-cocycle of g with coefficients
in c gives rise to a central extension

0 - » c - ^ e - > g -^ 0

of g where /: c e is the inclusion map. This is done as follows: Let

e := g 0 c (direct sum of vector spaces).

Define a multiplication [*,*]:,« on e by

[x + a, y + fe]* = [ x , y ] + a ( x , y ) f o r a l l x , y Ge ; a , f c Gc .
1.9 Central Extensions 53

Properties (Col) and (Co2) of the given 2-cocycle a guarantee that e,


together with [•,•]*, is a Lie algebra. We also see that the projection tt:
e ^ 9 given by
X a^ X

is then a Lie epimorphism and that its kernel c is included in the centre of c.
Remark 3 The construction of our 2-cocycle a from (CE) depends upon
the choice of the subspace g'. Different subspaces will in general lead to
different 2-cocycles. On the other hand, different 2-cocycles may lead us to
isomorphic central extensions. These questions are again in the domain of
cohomology of Lie algebras and need not concern us here.
Our intention is to construct universal coverings by means of “universal”
2-cocycles. For this we need to recall the second exterior power A^(g) of a
vector space g.
Let T\ %) = g <S) g. (This is just a tensor product of K-spaces. If we think
of the tensor algebra T(g) = a noncommutative polynomial
algebra, then T^(g) can be thought of as the linear span of all the monomials
of degree 2.) Let J be the subspace of 7^(g) spanned by all elements of the
form X 0 X, X e g, and define A^(g) = T \ q ) / J , We let r^(g ) A^(g)
be the canonical map and for all x and y in g write x A y instead of x 0 y.
Notice that x A x = 0 and x A y = —y A x for all x, y e g.

Proposition 2
For a Lie algebra g to admit a universal covering, it is necessary and sufficient
that g be perfect.

Proof If g is admits a universal covering then (by definition of covering) g is


the homomorphic image of a perfect algebra and hence is perfect.
For the converse let I be the subspace of A^(g) spanned by all elements of
the form

X A [ y ,z ] + y A [ z , a:] + z A [jc,y], x , y , z ^ g.

Define V = A^(g)/I, and for all x and y in g, let a(x, y) denote the image
of X A y e A^(g) in V under the canonical map. Evidently a is a 2-cocycle of
g with coefficients in V. Set

e = g ® E.

Define a composition law [ , on e by

[^ + y, y + w]j = [x, y] -I- a{x, y) for all x, y ^ q ; v, w e V.


54 Lie Algebras

As we said earlier, this makes e into a Lie algebra. For completeness we fill
in the details. The bilinearity of [ •, • \ is clear. The anticommutativity follows
from the fact that for all^:, yin g,[jr,y]= - [ y , a:] anda(x, y) = - a ( y , x ) . As
for the Jacobi identity consider three elements where i = 1,2,3,
e g, and e K of e. Straightforward computation shows that

Jac( X 2, X^) = Jac(A:i, X2, ^^3)

+ («(^l»[^2»^3]) + + «(^3>[^1.^2]))
= 0.

Indeed Jac(;ci, X2, x^) = 0 because g is a Lie algebra, and by definition the
remaining term in the sum is the image in A^(g)/I of the element x^ A
[ x 2 , x^] X 2 /\ [ ^ 3, ^ 1] + A [ x j , X 2 ] of A^(g), which is also zero.
Now let 7t: e ^ g be the canonical surjection; that is, tt: a : + r; jc for all
a : e g and v El V, Then tt is a Lie algebra epimorphism, and since V clearly

lies in the centre of e, we obtain the central extension

( 2) V- 9^0.

We intend to show that there exists a morphism of this central extension of g


to any other.
Suppose that

( 3)

is any central extension of g. We will denote the bracket of e' by [•, • ]j-. Let
us write e' as a direct sum of subspaces g' ® i'(F'), where g' is a preimage of
g under tt'. Identifying g with the subspace g' of e' and V with i(F'), we
write e' = g © V . Then we have a 2-cocycle f : Q X Q - * V with

[x,y], . = [ x, y ] + / ( x , y ) .

Since / satisfies the identities of (Co), there is an induced map /: A^(g) V'
vanishing on the subspace I c A^(g) defined above. This allows us to define a
linear map V ^ V' by

<f>o{a(x,y)) = f { x , y ) x,y^Q.

Recall that c = g 0 F. Consider then the linear map <l>: c c' given by
(f>(x v) = X <I}q( v ) for all a : e g and v ^ V, We claim that ((f>, </>q) is a
morphism of (2) to (3). Since <t>o is the restriction of <f>to F, it will suffice to
1.9 Central Extensions 55

show that

(a) the diagram

is commutative, and
(b) </> is a Lie algebra homomorphism.

Let j:, y e g and y, w e K. Then

Tr'(<f>(x + v)) = tt\ x + <f>o{v)) = X = tt( x + v)

so that (a) is clear. Next

(4) <^([x: + y ,y + iv]e) = <^([x,y]e)


= </»([^> y] + a ( x , y))
= [ x, y ] + if>oa(x,y)

= [^> y] + f { x , y),
(5) <l>([x + v ,y + wle) = [x, y]e- = [x + ^
= [<f>(x + v),<j>{y + w )],,,

and hence (b) holds [Equations (4) and (5) follow from the fact that V and V'
are included on the centers of e and c', respectively.] This establishes the
claim that there exists a morphism from (2) to (3).
We cannot conclude that (2) is the desired universal covering of g, since c
need not be perfect. This problem is corrected by defining

e' = [ e , e ] , ,

the derived algebra of e. Since g is perfect, it is clear that 7r(e") = g, and


hence that e = + F. Thus

e" = [e, c] = [e^ + F, e" + F ] = [e^ , e^],

since F is in the centre and therefore c" is perfect. Now let = e" n F.
We contend that the central extension

(6) 0 9^0,
56 Líe Algebras

where tt" is the restriction of tt to , is universal is surjective because,


as remarked earlier, c = + V), Indeed, given any other central extension
(3) of 9, we know the existence of a morphism (</>, <^q) from (2) to (3). If we
now define <f)" and (¡)q to be the restrictions of (f> and <f>Q to and c^,
respectively, it is clear that (</>", ) is a morphism of (6) to (3). Since e" is
perfect; (6) is indeed a universal covering of g. □

Our last Proposition makes the following definition meaningful:


A perfect Lie algebra e is said to be centrally closed if e is its own
universal covering.

We then have

Proposition 3
(i) The universal cover o f a perfect Lie algebra is centrally closed,
(ii) Every central extension o f a centrally closed Lie algebra splits,
(iii) I f c is centrally closed and c is a subspace o f Z(g), then
0 ->c-^e->e/c^ 0

is a universal central extension of e/c.

Proof, (i) Suppose that e is the universal cover of a perfect Lie algebra g and
that the extension is given by

0 c-U e ^ g 0.
Suppose, on the other hand, that

0 - > b - *^ f A" c 0
is a covering extension of e.
Let i be the kernel of tt o f ^ g. If jc e i, then fi(x) e kerirr) = /(c),
so (0) = [M-i), e] = f] and [x, f] c ;(b ). But [oi, f] = [ac, [f, f]] c
[[ji, f ], f ]] <= f ] = (0). We conclude that

0 i ^ / -----*Q ->■ 0
is a central extension of g. Accordingly there exist homomorphisms v: c f
and vq: c t such that the diagram

-i 'i
TTofJL
' - r f
1.9 Central Extensions 57

commutes. We then have

f = [f>f] = [»'(e) + f, v(e) + i] = v[e, e] = v(e).

Finally, let w := o v: e e. Since tt ° to(x) = v i x ) for all a: e e, co(x) = x


+ c(x), where c: e c is some linear mapping. From (a[x, y] = [&)(ac), 6>(y)]
we find that c vanishes on [ e, e ] = e. Thus ¡x°v = id ^, and we have the
commuting diagram

0
i f

Since V is surjective we see that ¡i is injective and finally that f = e and


b = (0).
(ii) Next let

(7) 0 b —^ f —> e 0

be any central extension of e. One sees that f - = [ f , f ] is perfect, and


replacing f by f ' in (7) if necessary, we may assume that f is perfect. Then (7)
is a covering of e and part (i) of the proposition gives with v : e ^ f,
¡jLop = idg. Thus (7) splits.
(iii) The proof of this part is left as an exercise. □

A very important example of a central extension occurs in connection with


constructing the covering algebra of the loop algebra The reader is
referred to the appendix at the end of the book for a discussion of As a
second example we prove the following proposition:

Proposition 4
(i) The Virasoro algebra is the universal covering o f the Witt algebra.
(ii) Virasoro algebras are centrally closed.

Proof. Let 2B = the Witt algebra (notation of Section 1.4).


Since SB is perfect, SB has a universal covering algebra e = SB 0 F (vector
space sum) with corresponding 2-cocycle a. Using property (Co2), we have

(8) 0 = a ( L „ [L„,, L J ) + a ( L „ , [ L „ , L j ) + a { L „ , { L „ L j )
= ( n - m ) a ( L ^ , L„+„) + (/: - n ) a{ L„, L„^*)
+ {m - k)a{L„,Li^^„) for all A:, m, « e Z.
58 Lie Algebras

Setting A: = 0, we obtain

(9) 0 = (n - m)a{Lo, L„+„) - na{L„, L„) + ma{L„, L„)


= (« - m)a(Lo, ¿ m + J - (m + n ) a { L ^ , L„).

If /c + m + n = 0, we can eliminate n in (8) to obtain

0 = ( - * - 2m)a(L*, L_*.) + (2A: + m ) a ( L „ , L_„)


+ (m —k ) a [ L

Setting A: = 1, this becomes

(10) (m - 1) « ( L „ + i , L - ( „ +d) = ( - 1 - 2m ) a ( L i , L _ i )
+ (m + 2) a( L„, L _ ^ ) .

We now set about to “adjust” the L„’s by elements of the centre V. We


have

[^0? ^(-^0? ^n) •


Set

for all n ¥= 0,

and

^0 ~ ^0
Clearly

[¿0. L„] = [ L q, L„] = nL„ for all n.

We next see that

(11) [ l „,L„] = ( n - m ) L ^ ^ „
for all m , n 0 with m + n ¥= 0,
In fact
1.10 Free Líe Algebras 59

whereas

i n —m \
(n - m) L„^„ = (« - ^m+n)-

Therefore (11) now follows from (9). Now

[¿m. ¿-m] = [■i'm- ^-m ] + ¿'-m) = ~ 2mLo + ¿-m )

= -2mLo - m a ( L i , L _ i ) + a ( L „ , L _ „ ) .

Set P(L„, L_„) = - m a ( L i , L_,) + a(L „ , L_„) e F for all m e Z, and


extend /3 bilinearly to the entire linear span of the elements L„ by defining
P(L„, L„) = 0 whenever m + n # 0. Then, for all m, n e Z,

The mapping )8 is a 2-cocycle with L_j ) = 0. Since equation (10) is


valid for any 2-cocycle, we have

(m — + ^ - ( m + l)) ”

By induction it is easy to see that

^(L„,L_„) = - m )/3(L2,L_2) for all m > 2.

Setting c = 2p ( L 2, L _ 2\ we arrive at

[ l „, L„\ = ( n - m)L„+„ +

We now let 33 be the K-span of the L„, n ^ Z, and c in e. Then 33 is a


subalgebra of e isomorphic to the Virasoro algebra. Finally, c = [c,e] =
[33 + F, 33 + F] = [33,S] = 33. This proves part (i). Part (ii) follows from
Proposition 3(i). □

1.10 FREE L IE ALGEBRAS

In this section we investigate what sort of K-space a set X of symbols will


generate if they are allowed to freely combine subject only to the condition
that the resulting object be a Lie algebra. Consider at first a Lie algebra q
and a subset X of g. Form the set 33(^) of all products of elements of X.
60 Líe Algebras

Thus 93 may be constructed recursively by

(la) A^c93(AT),

(lb) if p y , p 2 ^ ‘^ { X ) , then [ p i , p 2] e 93(-^)-

The set of all (finite) linear combinations of elements of 93(AT) is clearly a


Lie subalgebra [X]^ of g, called the subalgebra of g generated by X. In
particular g is generated by if g = [X]^. Similarly let Q(AT) be defined
recursively by

(2a) X^£i(X),

(2b) if Pi e g, ^ Q(-^), then [pi,p2l ^

The linear span of £i(X) is called the ideal of g generated by X and is


denoted by I ik(A0.

Example 1 Consider the Lie algebra gj j of 3 X 3 matrices

a, b, c G K,

IK-E12 + K jE23 ^^13

where E^j denotes the matrix with 1 in the (/,;) position and 0 elsewhere.
Since [£12, £23! = ^13j 91,1 is generated by £12 and £23. Since a single
element in a Lie algebra can only generate the space it spans, it is clear that
no fewer than two elements could generate gj j.
It is immediate that [£i2, [ £ i 2> £23]] = 0 —[£23,[£ i2> £ 2311* In general,
what can one say about a Lie algebra g generated by two elements x, y
satisfying the “relations” [x,[x, y]] = 0 and [y,[x, y]] = 0? Since any prod­
uct involving two or more multiplications is 0, g = Kx Ky + lK[x, y], and
it is easy to check that £12 x, £23 y gives rise to surjective homomor­
phism (f> (in which £13 -> [jc, y]) of gj 1 g. There is no reason why (¡>
should be injective; for instance, g could have been abelian so that [x, y] = 0.
However, we have established that in some sense gj 1 is a “largest” model
for Lie algebras with two generators satisfying the above two relations.
This observation leads us to the natural question of presentations of Lie
algebras: Given a set X and a set of identities (relations) involving sums and
1.10 Free Lie Algebras 61

commutators of the elements of what sort of Lie algebras are there that
can be generated by X and satisfy the given identities? Is there some “largest
one” that satisfies only these identities and the Lie identities? The answer is
yes, and “largest” means that every other Lie algebra generated by a set X
satisfying the same relations is a homomorphic image of it. Free Lie algebras
are Lie algebras on which no relations are imposed (other than those for a
Lie algebra: skew-symmetry and the Jacobi identity).
We will make a number of important constructions of Lie algebras by
presentations in this book. Always the problem with this method is to get
some understanding of what the resulting Lie algebra is really like. For
instance, consider the two presentations:

9i 3: x , y : [A:,[A:,y]] = 0 = [y, [y, [y, [y, Ji]]]],


82,2: x , y : [Ai,[jc,[-»:,y]]] = 0 = [y, [y, [y, Ji]]] •

It is not obvious that 3 is six dimensional and that 82,2 ¡s infinite


dimensional. Clarifying such problems almost always involves ad hoc con­
struction of explicit representations of the Lie algebras.
Free Lie algebras are defined by their mapping properties:
Let be a set. A Lie algebra ^ is said to be free on AT if AT c g , and for
any Lie algebra g and any map / from X into g there exists unique Lie
algebra homomorphism / ': g g for which the following diagram com­
mutes:

Proposition 1
Let X 0 be a set. Then there is a free Lie algebra on X , and it is unique up to
isomorphism.

Proof. The uniqueness question is established by the usual method (see


Section 1.2) and is left as an exercise. Let A ( X ) be the free associative
algebra on X over K. Inside Lie(^4(AO) let g = S ( ^ ) be the Lie subalgebra
[X]^ generated by X. We show that g is free on X.
h&t g be any Lie algebra over K, and let /: A!' -> g be any map. Let U(g)
be the universal enveloping algebra of g. Then there exists a unique homo­
morphism
/ : A ( X ) ^ U ( q)
62 Lie Algebras

of associative algebras extending / (we are assuming that g is embedded in


U(g) in the usual way). We claim that / ' == / |g is the required Lie algebra
homomorphism.
First we have

/'([^.y]) =f(xy - y x ) = /( x ) /( y ) - f i y ) f ( x )
= [f'ix),f'(y)\ forallx,yeg,

so / ' is a Lie homomorphism. This shows that if / ' ( x ) , / ' ( y ) s g, then also
f'i[x, y]) e g. Since f ' i X ) = /(AT) c g, it now follows from the definition
that /'([Afl^) c g [see (la) and (lb)]. Thus / ' makes the above diagram
commutative. It is clearly unique since X generates □

Remark 1 The free Lie algebra on the set X is denoted by ^(A"). When it
is convenient, we will use the model constructed above without further
comment. A Lie algebra g is said to be free if it is isomorphic to §(AO for
some X.

Proposition 2
In the notation above, the free associative algebra A ( X ) is the universal
enveloping algebra o f g(AT).

Proof Let A be any associative algebra over IK, and let / : %{X) Lie(A)
be any Lie homomorphism. By the definition of A(X), f \ x lifts uniquely to a
homomorphism / ': A ( X ) ->A. Now f'l^^x) is a Lie homomorphism, and it
extends fix'- X -» Lie(.<4). Thus, by the definition of free Lie algebras, it
must coincide with /:

A(X)

Since / ' is the only possible lifting of f i x , it is the only lifting of /. This
proves that A ( X ) has the mapping properties of a universal enveloping
algebra of g( AO. □

The free associative algebra A ( X ) (which is also the tensor algebra of the
K-span V of AO is Z-graded by the total grading. Each element of AT is
homogeneous of degree 1, and g(AO = [AOk- With the notation of (la) and
1.10 Free Lie Algebras
63

(lb), ©(A") consists of homogeneous elements of A(X), and hence § (A ) is


the K-span of homogeneous elements. It follows that § ( A ) is graded bv 7- Jr.
fact ’ "

S(A)= eg(A )",


n= l

where
g ( A ) " : = g ( A ) DA' ^i X) .

Example 2 X = {x, y}. Bases for g"(A ), 1 < « < 5:


g ^ ( A ) : X, y;

g ^ (A ) :[ ; c, y ] ;
g3(A): [ji,[jc,y]],[y,[y,A:]],
3 '‘(A ): [Ji, [Ji,[ji:,y]]],[y, [y,[y,J:]]],
[x,[y,[x,y]]],[y,[x,[x,y]]Y,
[x,[x,[x,[x,y]]]]; [y,[:c,[x,[x,y]]]]; [y,[y,[x,[x,y]]]];
[y>[y>[^.[^»y]]]]; [[^,y],[jc,[Ai,y]]]; [[jc,y],[y,[jc,y]]].
Suppose that A = {xj,X2, . . . } (possibly finite), and let g = Z e Z ® • • •
[number of factors = card(A)]. Then A ( X ) is graded by Q: For n =
(«i, ti2, . . . ) e (2 (all n, > 0) ^ " ( A ) is the K-span of the monomials y ^ . .. y*,
where k = En„ and a:, occurs among the y,’s exactly n, times. Precisely the
same argument as above shows that 5 (A ) inherits the grading

S(A) = e 5"(A),

where
5"(A ) ==5(A) n ^ " ( A ) .

Example 3 In the example above,

g^i-o^A) = K x; 5^‘’’*^(^) = Ky;


g(2.0) ( A ) = 0; 5<'’'>(A) = K[x,y];
5 (^’i>(A) = K [ x , [ x , [ x , y ] ] ] ;
g<2.2)( A ) = K[y, [x, [x, y]] + K [ x , [ y , [x, y ] ] ] ;
g(3.2)(A) = lK [y ,[ x , [ x ,[ x , y ] ] ] ] + K [ [ x , y ] , [ x , [ x , y ] ] ] .
64 Lie Algebras

There is a considerable amount of literature on determining bases of free


Lie algebras. The most famous of these are the Hall bases [Bo 2] and the
Lyndon bases [Lth].
Let be a nonempty set, and let g = g(Ai) be the free Lie algebra on X
over K, Let i? c g be any subset of elements, and let l(R) be the ideal of g
generated by R. Then g == ^ / l ( R ) is called the Lie algebra defined by the
generators X and the relations R.
Remark 2 The terminology seems to suggest that AT is a subset of g, which
it is not. Moreover it is quite common to say that ^ / l ( R ) is the algebra
generated by the generators X and the relations = 0, thus both thinking of
the natural image of X in as X itself (even though X need not be
mapped injectively into the quotient) and anticipating that in the quotient the
relations R appear equal to zero.

Example 4 X = {x, y}, R = {[x,[x, y]], [y,[x, y]]}. The ideal generated by
R contains all products of three or more elements of g. Let g == ^ / l ( R) .
Then g = Kx + IKy + K[x, y] and g = g^^ of Example 1.

Let g be a Lie algebra over K. By a presentation of g we will understand


a pair (X; R) consisting of a set X and a subset R of the free Lie algebra
S(AO on X such that g(A ')/I(i?) = g, where l(R) is the ideal of
generated by R.
Remark 3 Every Lie algebra has an infinite number of distinct presenta­
tions. For instance, we can take X to be any subset of g that generates it as a
Lie algebra and take R = ker(g(AO ^ g).

Evidently, if two Lie algebras have the same presentations then they are
isomorphic and we may therefore speak of the Lie algebra with the presenta­
tion (X;R).
Let {Qjlyej family of Lie algebras over K. For each j s. J, let
{Xj-; Rj) be a presentation of g^. Let ^(uATy) be the free Lie algebra on the
symbols u Xj (disjoint union), and identify ^ (Xj ) as a subalgebra of g(uA(y)
in the obvious way. The Lie algebra f with the presentation ( u X j i U R j ) is
called a free product of the family [with respect to the presentations
(Xji Rj)l
The natural mappings vf. g^ f defined by gy = %{Xj)/\{Rj)
^(uA y)/I(ui?y) = f are injective (see below) homomorphisms called the
natural embeddings of gy in f.

Proposition 4
{Qy}ye J ^ family o f Lie algebras over K. For each j e J, let (Xj; Rj) be
a presentation of gy, and let f be the free product o f the family {gy}ye j ^dh
1.11 Campbell-Baker-Hausdorff Formula 65

respect to these presentations:

(i) The natural embeddings Vj\ ^ f are injective,


(ii) I f a is any Lie algebra over IK and Qj ^ d are homomorphisms,
then there is a unique homomorphism tt: f a such that for all; e J
9;

f a
commutes.
(iii) I f f' with embeddings v'j is another free product o f the family {9y}yej?
then there exists an isomorphism (f>: f ^ f' such that (f>°Vj = Vj for
all j.

Proof (i) We maintain the previous notation. Fix any i e J. Then there
is a homomorphism Qj such that 7t,( jc) = x for all x ^
TTj(Xj) = 0 for j i. Since = {0}, defines a homomorphism
f g^, and evidently ° = id. Thus is injective. We leave (ii) and (iii)
for the reader to check. □

In view of part (iii) of Proposition 4 we can speak of the free product of a


family of Lie algebras. We denote this by j9y or simply by g^ * ♦ * g„
if J = {1,..., n} is a finite set.

Example 4 The free product of a family of one-dimensional Lie


algebras is the free Lie algebra on the set j. This can be seen by using
the presentations ({xy}; 0 ) of the IKxy’s.

We can use free products to combine Lie algebras together with cer­
tain relations imposed between them. We use this in a significant way in
Chapter 4.

Example 5 (Heisenberg algebras) Let a+, a_, and c be abelian Lie


algebras with linear bases {x^\i e J}, [y.\i e J}, and {c}, respectively. Let
/ = a + * c * a _ , and let R be the set of relations
- dijC,[Xi,c],[yi,c]\iJ e j}.
Then f / l ( R ) = a+ 0 c 0 a_ is a Heisenberg algebra.

1.11 CAMPBELL-BAKER-HAUSDORFF FORMULA

We begin with a nonempty set X and form the free monoid M = M{ X) on


X (see Section 1.2). A typical word in M has the form
W = X. • • • JCi,
66 Lie Algebras

where . . . , X/^ ^ X, k > 0. We call k the length of w. M is finitely


generated if X is finite.
Let [K be a field of characteristic 0. (Much of what we say only requires
that IKbe a commutative ring.) We can then form the free associative algebra
A{ X) on X which has basis M and multiplication defined from that of M
(see Section 1.2). ^4(J^) is graded with A ( X T spanned by the elements of M
of length n.
We denote by L ( X ) the free Lie algebra on X over K (see Section 1.10
where it was denoted by S(A")). According to Proposition 1.10.2, L ( X ) may
be viewed as the subalgebra of Lie(^(AT)) generated by X, and then ^(A") is
the universal enveloping algebra of X,
The ring of formal power series (in the noncommuting variables AO over
K is the set K ^ of all K-valued functions on M with algebraic operations of
scalar multiplication, addition and multiplication defined by

(flcr)( w) = aa{w),
(c r + t )( w ) = (t { x ) + t ( w ),

(crr)(w) = £ cr{ y) T( z) for all cr, T e w, y, z ^ M; a


yz = w

Intuitively a- G K ^ may be thought of as the infinite series We


have a natural embedding of ^(AO in K ^ which realizes ^(AO as the space
of functions on M with finite support (i.e., cr g K ^ such that cr{x) = 0 for
almost all x ^ M).
Set
= (w G Af|lengthw = n)
and

K^ = G K^\( t \mp = 0 if p ^ n} for all n ^ N.

If O' G K^, we define o-„ by

( 1) ’J mp = 0 if p # n.

We will find it useful to generalize this to the case of K^ where A/" is a


direct product of finitely many free monoids. The most important case is
N = M X M where M is a free monoid, as above. Then K ^ ® K ^ ^ K ^ ^ ^
by identifying o- ® r with the function (cr 0 rXx, y) = o-(x)T(y).
If = M 1 X • • • X Af^, then the length of jc = (x^, . . . , x ^ ) ^ N is
length (^¿). Let A/' be a direct product of finitely many free monoids. A
subset 5 of K^ is locally finite if for each ;c g AT, {cr g 5|o-(ji:) 0} is finite.
1.11 Campbell-Baker-Hausdorff Formula 67

If S is locally finite, then the sum defined by

( E = E o-(^)
^ o-e5
makes sense. For example, if o- e then 5 •= {o-„|« is locally finite
and

( 2) O’ = E o-„ = E o-„.
<r„^S n= \

Again, for each x e AT, define

K,

Sx(y) = ^ x . y

For any choice of e K, x e N, the set

G N}

is locally finite. Using the definition of sums given above, we then have

(3) O’ = E o-( a:)5^ for all o- G


xsN
A third important example of a locally finite set is

{o-ln = 0, 1, 2, . . . } ,

where o- g == { t g IK^|t(1) = 0}. Thus for a g IKi",

( 4) exp 0-— 1 + 0 - + - — +

and
0-2 cr^
(5) log(l + (t ) ■■=<T - — + — -

are well-defined sums. It is a straightforward exercise to show that

exp: -> 1 + (K+,


log: 1 +

are inverses of one another.


68 Lie Algebras

If O’, T e K+ and СГТ = ТСГ, then


(6) exp cr exp r = exp(o- + r ) ,
(7) log[(l + o-)(l 4- r)] = Iog(l + O’) + log(l + r ) ,
as one sees directly by expanding both sides. Even if ат Ф та,
exp a exp r = exp[log(exp a exp r ) ] .

Our aim is to give some information on the nature of log(exp a exp r). Our
exposition follows [Lth]. As a preliminary step we are going to establish
another interesting result about free Lie algebras.
Let ЛГ be a nonempty set, as before, and define
(8) Д: A ( X ) - ^ A ( X ) ^ A { X )
to be the unique homomorphism of associative algebras extending the map
x»->x<8)l + l0 jc , X ^ X.
We observe that Д is a graded homomorphism relative to the grading of
A ( X ) <8>ЖАГ) for which

(A(X) А ( Х ) У = Ф A ( x y ® A ( X y ~ ‘.
¿=0
As mentioned above, we view L ( X ) as the subspace of A ( X ) generated by X
where A ( X ) is viewed as a Lie algebra.

Theorem (Friedrichs)
F o r w ^ A (,X ),
w G: L ( X ) «=> Ди' = 1 ^ 0 1 + 1 0 ^ .
Proof. If = w 0 l + l 0 ] v and Д 2 = г 0 1 + 1 0 г , then
A[w, z] = [Aw, Дг] = [н ^ 0 1 + 1 0 н ^, z 0 l + l 0 z ]
= [w, z] 0 1 + 1 0 [w, z],
which shows immediately that the elements of L ( X ) satisfy the diagonal
condition.
Conversely, let w satisfy the diagonal condition. We may assume that w is
homogeneous of degree m > 1. Let {уДе/ be a totally ordered basis of
L(J^). Then by the Poincare-Birkhoff-Witt theorem,

< ¿2 < • • • < iV, r > 0, ^1, . . . e Z+}

is a basis of A(X).
1.11 Campbeli-Baker-HausdorfP Formula 69

Let w = where each 21s above, and ^


Then ’
w <8> 1 + 1 0 )v = Aiv =

= L«j,k(Ay„) ‘ • • • (Ay,.J
Consider one of these summands Aw„ which, for simplicity, we write as
A(y^i *• * y^"), where 5^,..., e Z+. This expands as

(yi 0 1 + 1 0 yi)*‘ ••• (y„ ® 1 + 1


= yi'yl' • • • y^» 0 1 + Siyi‘"V2" ■ • • y„" ® yi
+ S2yi'y|""‘ ••• y^" ® y2+ ••• + i„ y i'y !' ••• y^"“ ‘ 0 y „ + r(n ,s ),
where r(n,s) e E,2.2-^(A!’)'"“' 0 A i X Y and m = + ■■■ +s„.
Suppose that m >2. Since the expression
w 0 1 + 1 0 IV ^ A ( X ) 0 IK + IK ® A ( X )
and

( A ( X ) ® A { X ) ) " = ( A i X ) " 0 K) ® (K ^ A i X ) " )

® { A { X ) ' ”~^ 0 v 4 ( J f ) ‘)

® ( ®A{xy]

and also {yj/e/ c : A ( x y is independent, we see that each nonzero summand

W/ yp~^ ••• «>y/


must cancel with some similar summand in some other Ah^^. However, the
pair (n, s) can be recovered from and hence cannot appear in any other
AWj^. This shows that in fact m = 1, and hence w is in the linear span of the
set {yj, e /; that is, w e L(X). □

We return to the formal power series rings where N is some finite


product of free monoids.
Let Ni, N 2 be finite direct products of finitely generated free monoids. An
algebra homomorphism /: -> IK^2 jg continuous if whenever 5 is a
locally finite set in IK^^, then /(5 ) is locally finite in and =

If /: ^ IK^2 is continuous, then /(IK^O c for if cr e and


/(cr) = + T, T e IK+2^ a 0, then {{a\ + r)"}„eN is not locally finite.
Now, with and N2 as in the definition, let us see that any monoid
homomorphism /: ^ IK^2 (^^2 being considered multiplicatively) for
70 Líe Algebras

which f ( Ni \ {1}) c IK+2 extends uniquely to a continuous algebra homomor­


phism /: ^ IK^2 If we recall that ^ requires that we identify x
and 5^ and that any cr e IK^^ has the expression (3), we see that the obvious
way to extend / to IK^» is by

( 9) f{a) = D o-(x ) / ( a: ) .

A moment’s reflection shows that fiN^) is locally finite, and hence (9) makes
sense. We leave it as an exercise to show that / is continuous and that (9) is
the only way to extend / continuously to an algebra homomorphism.
As an example consider the map

A: M ( X ) ^ IK^®
(8) IK-' ^ IK^^^

This will extend to a continuous homomorphism

(10) A : I K ^ ^ IK^^^,

which extends (8) and satisfies

Ao- = £ A(T„
n= l

for all a fM [see (2)].

Proposition (Campbell-Baker-Hausdorff)
Let X be a finite set, and let x, y e M{X). Then log(exp x exp y) is a Lie
element o f the form

X + y + j[x,y] + •••

where the remaining terms are composed of higher (3 or more factors) commu­
tators of X and y.

Proof Set M = M i x ) and consider the continuous mapping (10). We note


that A ( M \ {!}) c IK^^^^. According to Friedrichs’ theorem log(exp x exp y)
will be a Lie element if

(11) A log(exp X exp y)


= log(exp X exp y) ® 1 + 1 0 log(exp x exp y)
= log(exp X exp y 0 1) + log(l 0 exp x exp y ).
1.12 Extensions of Modules 71

Since exp x exp y ® 1 and 1 ® exp x exp y commute, and since A is continu­
ous, we can use (6) and (7) to rewrite (11) as

(12) log A(exp jc exp y) = log(exp x exp y (S>exp x exp y).

However,

A(exp X exp y) = A exp x A exp y


= exp Ax exp Ay
= exp(x <8> 1 + 1 0 x)exp(y 0 1 + 1 0 y)
= (exp X 0 exp x)(exp y 0 exp y)
= exp X exp y 0 exp x exp y,

from which (12) is immediate. Finally,

exp X exp y = l + x H ---- - +


x\

= 1 + X + y + ^(x^ + 2xy + y^)


H- terms of degree > 3 in x and y
= 1 + x + y + | ( x + y ) ( x + y) + |( xy - y x ) + •••

= exp(x + y + ^ [ x , y] + ••• ).

Since log(exp x exp y) is a Lie element, the higher terms are also commuta­
tors of X and y (involving at least three letters). □

In the exercises we give an explicit formula for computing further terms of


the series for log(exp x exp y). We will have no need of any of these in the
sequel.

1.12 EXTENSIONS OF MODULES

Let be a ring, and let be an abelian category of /^-modules. Given


modules A, B in an extension of .^4 by B we understand a module E of
if and morphisms a, such that we have the exact sequence

(1)
72 Lie Algebras

Given another extension

( 12) 0 B E' ^ A

oi A hy B we say that (2) is equivalent to (1) if there is a morphism y:


E E' so that the diagram

fi'
B E'
commutes.
It is elementary to see that y is an isomorphism and hence that equiva­
lence is an equivalence relation on the set of extensions of A by B. We
denote the set of equivalence classes by Ext(^, B).
It is customary to study Ext(^, B) in the context of homology theory
where Ext(^, B) is shown to be isomorphic to E x t^ ^ , B), However this
theory requires that the category has “enough projectives”; that is, for
every module ^ in ^ there is a projective module P in ^ and an epimor-
phism P ^ ^ ^ 0. Unfortunately, the category ^ of Section 2.6, for which
we need to understand extensions, does not have enough projectives, so this
theory is unavailable to us. Instead, we study Ext(A, B) completely in the
context in which it was defined, that is, in the theory of extensions. There is
in fact a complete theory of Ext in abelian categories [Str]. Our approach
here is to provide the definitions and basic facts needed to prove what we
need and leave the details to the exercises or to Strooker [Str] according to
preference. Much of this can be found in [HS] on which our exposition is
based.
Our first object is to make Ext( - , - ) into a bifunctor on if, covariant in
the second variable and contravariant in the first.
Let a: X ^ A , p. X ^ B, be maps in if. A commutative diagram

’1 i‘
B Y

with y in ^ is a push'Out (or fibred sum of A and B over AT) if it satisfies


the universal property that for any commutative diagram

B Y'
1.12 E xtensions o f M odules 73

there is a unique morphism A: y Y' with

AT-----

commutative. Of course the push-out is unique up to obvious isomorphism, if


it exists. Explicitly we can show that it exists by defining Y = (A e B)/K ,
where

K = {«( a:) - P(x)\x e X ) ,

and by defining

<p: a + K,
( 0 ,t ) + K .

Lemma 1
Let

1 . 4
B
be a push-out. Then

(i) (p induces an isomorphism coker a coker \jf {i.e. from A /im a


Y /im ij/X
(ii) if a is injective, so is i//.
Furthermore, if

X
/3

is a commutative diagram and <p' induces an isomorphism


coker a coker then Z in the diagram is a push-out. □
74 Lie Algebras

By reversing all the arrows in the definition of push-out we arrive at the


dual concept of pull-back. The analogous statement to Lemma 1 states that
given maps a\ A X, B X, the pull-back

X
/3

B
is characterized by the fact that <p induces an isomorphism

ker i/i ^ ker a.

Now let us consider the behaviour of Ext(^, B) in the second variable.


Suppose that cp: B ^ B' is a morphism in Let

E:

be any extension. We can form the push-out


P

<p
P'
B' E'
Since jS is injective, so is /3', and furthermore A — coker /3 = coker j8' by
Lemma 1. This allows us to uniquely define a': E' ^ A so that

E: 0 B -^ E -^ A ►
0

4
0 B' I ^ E '^ A ►
0

One now has to check that this procedure maps equivalent extensions into
equivalent extensions. Hence we have a map <p*

: Ext(/1, J5) E x t(^ , J3').


Furthermore

(i) if B ^ B ' ^ B", then (^V) *=<P'*<P*,


(ii) (idg)* = id

We will sometimes write E' = (p^E rather than the proper ( ¿ 0 = (p^{E\
where {E) and ( ¿ ') are the equivalence classes of E and
1.12 Extensions of Modules 75

In precisely the same way if il/: A' ^ A is a. morphism then starting from
the extension (¿ ), we can use pull-backs to get

0 ---- >B >0

1
0 ---- - ^ E ' ^ A ’ ---------^0

and hence derive

iff*:Ext{A,B) Ext{A',B)

with

Lemma 2
Let ¿’e ExtiA, B), and let ij/: A' ^ A and <p: B ^ B' be morphisms. Then

Following [Str] we prefer to write <p^ and (ftjt for and so that
Lemma 2 reads

Lemma 2 follows directly from the easily proved next lemma:

Lemma 3
Suppose that we have the commutative diagram

E: 0 B A' ---- >0

1 *1
E': 0 ---- » B' >E'

Then

< p ^ ( E ) = r { E ’).
76 Lie Algebras

Proof o f Lemma 2. Apply Lemma 3 to the top and bottom rows of

0 ---- » B E' *A’ ---- »0


I Pull-back I
1
E: 0 ---- > B E --------- *A ---- »0
Push-out I
1 1
0 ---- > 5 ' --------- > E" >A ---- >0

Using this we can now define an abelian group structure on Ext(A, B), Let

E: Q --> B -^E ^A ^Q
and

E^\ Q^ B jEj —^ A —>0

be representatives of two classes ^ and of Ext(yl, B). Then, with the


obvious maps,

E e E^: 0 ^ B e B ^ E ® E i - ^ A e A - ^ 0

is a representative of a class in Ext(^ 0^4, B 0 J?). Let


V: B ® B ^ B and A: ^4 0 ^ be the maps + ¿2
a (a, a), respectively. We define

^ = V (^ 0

(see notation following the statement of Lemma 2). We leave it to the reader
to prove that this is indeed an associative operation and that the split
extension

0-^B ^A ® B -^A -^0

is the identity element. The inverse of the class (E ) of

E: O ^B ^E ^A -^0

is (¿ ( —id)) obtained from the morphism —id: A ^ A, a ^ —a. This can be


1.12 Extensions of Modules 77

easily shown:

E + E ( - i d ) = V (£ © £ ( - id ) ) A

= V{E © £ )(id , -id )A = (V (£ © £ ) ) a ,

where K: A ^ A ® A given by a •-» (a, —d).


Now the construction of V(£ © £ ) is shown as the top two rows of

B ®B E e E A ®A
Push-out
A

V(£ © E): B E2 A ®A

(id,0) proj
V
0 B ®A A 0

In the lower part of this diagram, the part that needs explaining is the map
¡jl: a - ^ ^ 2* Given a ^ A , we let e ^ E be any lift of a, and define
fi(a) = X(ie, —e)). This is independent of the choice of e because V is
addition of components.
Now the bottom right-hand square is a pull-back because ()8, i n d u c e s
an isomorphism between ker(proj) —B and ker a — B (use the dual of
Lemma 1). But the pull-back of

E2 A ®A
Ta
A

defines (V(E 0 E))K, so this extension is split.


We now come to the main object of study. Consider an exact sequence of
modules in ^ :

B: 0 B ^ B' ^ B" ^ 0.

Let A be any module in Then we have the usual exact sequence [Ja 2,
Section 3.1].

0 ^ H o m (^ ,B ) ^ H o m ( ^ ,£ ') Ho m{A,B").

From the above we have a sequence of maps

E xt(yl,B ) - ! ^ E x t( ^ ,B ') ^ E x t { A , B " ) .


78 Lie Algebras

We construct a connecting homomorphism

5: H o m (^,B ") ^ Ext(y4,B)

by 5 (/) = (Bf). (The top two rows of the diagram below illustrate this.)

Proposition 4
The sequence

0 ^ H o m ( A , B ) ^ H o m (^ ,B ') ^ H om (/4,B")

- ^ E x t( ^ ,B ) - i^ E x t( /l,B ') - ^ E x t ( ^ , S " )

is exact.

Proof. We content ourselves with showing that the sequence of Proposition 4


is a null sequence. The rest appears in [Str] or may be taken as an exercise.
First observe that = (il/(p)^ = 0. Now let / g Hom(>l, B"), and con­
sider E = 8f. The definition of 8f is given by the top two rows of the
diagram.

B B' B"
. Pull-back
1 4 1^
y
E: 0 B E A

hB' (id,0)
Ufj) 1
B' ® A A
proj

The bottom two rows clearly form a commutative diagram. By Lemma 1,


since induces an isomorphism coker <p coker (id,0), we see that the
lower left square is a push-out. But this is the definition of and hence
£ splits; that is, {(p^E) = 0. Thus ^^<>5 = 0.
In the same way let

B' — > J?"


«A
1.12 Extensions of Modules 79

for some g, and let us see that 8f = 0. Looking at the diagram

>
l>
B' B" ^0

(»>.g)|
i'
0 ---- * B ----- *B e A
(id,0) proj

we see from the dual of Lemma 1 that the right-hand square is a pull-back.
But this is the definition of 8f, and hence 3 f = 0. □

Corollary 1
Let 0 ^ B ^ B' ^ B" 0 be exact. Then Ext(A, B') = 0 if both Ext(v4, B)
= 0 and Ext(^, B") = 0. □

Corollary 2
Let B = Bf^ Z) Bf^_i D • • z) Bi ^ B q = (0) be a filtration o f B. Then

E x t(^, = 0 / = ! , . . . , / : = > Ext(y4, B) = 0.

Proof Use the exact sequences

0 B j_ ^ Bj B j/ B j _ ^ -> 0

to prove inductively that Ext(A, B f) = 0, ; = 1 ,2 ,... .

Along similar lines given

B\ 0 ^ B ^ B '^ B " -^ 0

and A as above, we have the exact sequence

0 Hom(5", A) ^ H om (5', A ) ^ Hom(B, A)

and also

E x t(5 " ,^ ) ^ E x t ( B ', ^ ) - ! ^ E x t( B ,^ ) .

If the connecting homomorphism

5: H om (S, A ) ^ Ext(B", A )
80 Líe Algebras

is now defined by d( f) = (fB), then we have

Proposition 4'
The sequence

0 ^ Hom(B", A ) Hom(B', A ) Hom(B, A)


Ext(fi", A) ^ A) ^ Ext(B, A )
is exact.

Corollaries analogous to those following Proposition 4 can then be ob­


tained.

EXERCISES

1.1 Let F be a finite-dimensional vector space over a field and let


/3: F X F ^ IKbe a bilinear form.
(a) Prove that

9 = g(/3) — [a G End(V)\ß{ax,y) + ß ( x , a y )
= 0 for all jc, y e F}

is a Lie subalgebra of gI(F).


In the rest of this exercise we look at this important construction of Lie
algebras more carefully. An endomorphism a e g(jß) is said to be an
invariant transformation of ß. The Lie algebra g(j8) is the “linearized”
version of the group Giß) of isometries of F relative to ß; that is,
Giß) '= [a e GLiV)\ßiax, ay) = ßix, y) for all x , y ^ F}.
Henceforth assume that ß is nondegenerate; that is, for x g F,
ßix, F ) = (0) <=>X = 0 <=> ßiV, x) = (0).
(b) Let p: F ^ F* be the linear map defined by

{p(x),y) = ß (y ,x ) forall ac,y G F.

Show that p is an isomorphism.


(c) Show that for each a e End(F) there exists a unique element
e End(F) such that piax, y) = /3(x, a*y), and show that the
mapping a ^ a"^ of End(F) into itself is linear, bijective, and
satisfies iab)* = Similarly define by pi*ax, y) =
Pix, ay), and show that *ia*) = a = i*a)* for all a e End(F).
Exercises 81

(d) For each a e End(F) let a‘ e End(F*) be the transpose map


defined by
( a ‘f , x ) = ( f , a x } .
Show that
a* = p * ° a' o p

for some p e GL(V) and hence that

tr a = tr a* for all a e End( F ) .

(e) Show that for a e End(F),

a e g(/3) «»a + a* = 0<=>a +*a = 0.

Conclude that g(/3) c


(f) Let B be some basis of V, and relative to this basis write the
vectors of V as column vectors with entries in K. Let J be the
matrix of B relative to B so that

B(x, y) = x^Jy for all x , y e V.

Show that in matrix terms A* = and g(j8) = {A ^ qI„(IK)|


JA = -A^J).
(g) Show that the symmetric bilinear form

k: Q(B) X 9(/3) ^ K
k { a , b ) = tTy(ab)

is nondegenerate.
1.2 Let L be a finite-dimensional vector space over IK with basis B =
(Cj,. . . , e„).
(a) Define /3 to be the symmetric bilinear form on V whose matrbc
relative to B is I„ in X n identity matrix). The Lie algebra g(/3)
of Exercise 1.1 is called §o(n), or go(n,IK) in this case (special
orthogonal Lie algebra). Show that a e §o(n) iff a is skew-sym­
metric as a matrix. Determine dim §o(n).
(b) Suppose that n = 2m, and define /3 to be a skew-symmetric
bilinear form whose matrix B is
82 Líe Algebras

This time g()8) is denoted by ^p(m) (symplectic Lie algebra).


Determine dim §p(m).
(c) Let F be a vector space of finite dimension n over C, and let
<*, • >: F x F - ^ C b e a nondegenerate hermitian form. Prove
that the unitary Lie algebra

U{n) := [a e qX{V)\{ax,y) + { x, a y) = 0 for all x , y e V]

is a Lie algebra over R (but not over C!) and that the special
unitary Lie algebra

§U(n) := [a G Vi{n)\\x a = 0}

is an ideal of codimension 1 in U(«). Prove that relative to an


orthonormal basis of F, a g U(n) <=> ú: is skew-hermitian. Find
dim \X{n) and dim §U(n).
(d) Show that the Killing form of is negative definite. Show
that ^U(n) is a simple (real) Lie algebra and that its complexifi-
cation (Section 1.1) is §I(n,C).
1.3 (a) Let a and b be Lie algebras, and suppose that we are given a
homomorphism of Lie algebras b ^ from b into Der(a).
Define [ • , * ] : a e b x a 0 b - > a 0 b b y

[a + b, + b'] = [a, a'] + b¿,(a') - 5^.(iz) + [b, b']

for all a,a' ^ a and b, b' g b. Show that this gives the space
a 0 b a Lie algebra structure and that a is an ideal of this
algebra.
(b) In particular show that for a Lie algebra a and any subalgebra b
of Der a, a 0 b has the natural structure of a Lie algebra. The
holomorph of a is the Lie algebra a 0 Der(a) constructed in this
way.
1.4 Let ^ = 0 be an associative algebra which is graded by the
abelian group Q. Show that 1 g A^.
1.5 (a) Let A be an algebra over K, and suppose that IK is algebraically
closed. Let / g Der(.^4), and suppose that for each a g IK we
define

:= G y 4 l(/- a Y a = Ofor somes r g Z+).

Suppose that A = (this always happens if dim^^ ^ <


oo). Prove that this decomposition affords a grading of A by
(C, +). [Hint: ( / - (a + p)Xab) = ( / - aXd)b + a{f - /3Xb).]
(b) Find a similar statement for / e A ut(^).
(c) Let g be a finite-dimensional complex Lie algebra. Let 6 e Aut(g)
be of finite-order N. Set ^ and for each 0 < ; < iV
define
g-' = {ac e glflAc = ^^x).

Show that g = ®j=o 9^ constitutes a grading of g


by Z / N Z .
(d) Let 0 be the automorphism of gIsiC) defined by = £,+i,y+i
(indices mods) where is the ( i,;) matrix unit of gljfC).
Determine explicitly the spaces g 13(C)-', j g Z /3Z for 0.
1.6 Let F be a vector space over IK with basis { v / ^l k g Z}. Let a, /3 g K.
Show that by defining

= {k + a + I3{n + \))v„ +k J

we obtain a representation of the Witt algebra SB on V. We denote V


with this SB-module structure by F„^. Determine the conditions on a
and ¡3 under which ^ is irreducible.
1.7 Let g be a finite-dimensional simple Lie algebra over K. Let
g X g ^ IK be the Killing form of g. Let /B(g) be the IK-space of
invariant bilinear forms on g.
(a) Show that the elements of IB( q) are symmetric (use g = D(g))
and that they are either trivial or nondegenerate.
(b) Let Z := {A G End,^(g)|A ad(x) = ad(x)A for all x g g} (the
centroid of g). Show that Z is a field extension of IK and that g
has a natural Lie algebra structure over Z (henceforth denoted
9z)-
(c) Show that the following are equivalent:

CS 1: Z = J<.
CS 2: The IK-Lie algebra IK®||^g is simple (IK = algebraic closure
of IK).

If g satisfies either of the two equivalent conditions CS 1 or CS 2,


it is called central simple.
(d) For A G Z let k^(A): g X g IK be defined by k^(A)(x, y) =
K,^(x, A(y)). Show that / : A •-> K|,^(A) is a IK-linear isomorphism
of Z onto IB(q).
(e) Let g be a finite-dimensional simple Lie algebra over an alge­
braically closed field IK. Then
I B ( q) = K k .
84 Líe Algebras

1.8 Let 1^ c g be Lie algebras. Let


3 = 3g(l^) == {x e g |[j:,^] = 0},
It = iig(]^) ■■= [x ^ g|[jc, 1^] cl^}.

These are the centralizer and nonnalizer of in g, respectively.


(a) Prove that 3 an n are subalgebras and that 3 is an ideal of n.
(b) Show that if the representation of i) on g: p: f| ^ qK q) given by
X >-» adg(jc) is semisimple, then = 1^ ® 3g('^)-
1.9 Let F be a IK space and A e V*. Let A: S(F) IK denote the
extension of A to an algebra homomorphism. Show that

kerA = Z S ( V ) { v - k ( v ) l ) .
u^V

1.10 Let g be a Lie algebra and U(g) its universal enveloping algebra. Let
A^: U(g) -> U(g) be given by A^^: u ^ xu - ux. Let x, / e g c U(g).
For all e Z+ show that

L « ; ( A ^ ) r - '( a d / y ( x ) ,
j=i
where
N—j+\ f J.J. j
a,(AT) = ( - i y E
k=l \ ■'

1.11 Let g be a Lie algebra, g ¥= (0). Show that g admits modules that are
not semisimple.
1.12 Let G be a group and p: G -* GL(V) a representation of G on a
K-space V.
(a) Show that there exists a unique representation

p I: G G L (F ® • • • ® F ) (n-times), « > 1,
satisfying

p U x ) ( V i e> ■ ■ ■ <8> v „) = p ( x ) ( u i) ® • • • ® p (x )v „.

Let pqI G K be the trivial representation. Show that we obtain


a representation

pT:= Q pj;-. G G L ( r ( y ) = 0 r '* ( F ) ) .


n>0 ^ n>0 '
Exercises 85

(b) Show that determines a natural representation of G on


S(F) an also on each component S"(F) so that p^(xXf x • • ‘
= p(x)vi ■ p(x)v„.
• •

(c) Show that p^ determines a natural representation p^ of G on


A(F) and also on each component A"(F) so that

Л ■■■ AV„) = p ( x) { Vi ) A ■■■ A p{ x) ( v „) .

(d) Assume that G is finite and that dim^^ F = / is ^nite. Given


X ^ G, let WiCac), . . . , щ(х) be its I eigenvalues in K. Verify the
following identities in IK[[i]]:

П (1 - = L ( - 1 ) " Ч р „^(д:)) г".


i=l n=0

1.13 Let a be an ideal of the Lie algebra g. Show that U(g)a = all(g) is an
ideal of U(g) and U (g /a) = U(g)/U(g)a.
1.14 Let g be a finite-dimensional Lie algebra. Prove that U(g) is left
(right) noetherian. [Hint: Use the associated graded algebra gr(U) and
the noetherian property of 5(g).]
1.15 Let be a (not necessarily commutative) ring with no zero divisors,
(a) Prove that the following conditions are equivalent

ORE 1: Aa r\A b (0) for all a, b e A \ {0}.


ORE 2: There exists a division ring D and a ring embedding
A ^ D such that for all d e Z) \ {0} we have d = x~^y
for some x ,y ^ A.
ORE 3: For every left .,4-module M the set

= {m e M\am = 0 for some a e ^ \ {0}}

is a submodule of M.
A ring satisfying the above equivalent conditions is called an Ore
domain [see JA II].

(b) Let {b„)„ez ^ nonnegative real numbers such that


Iim sup„_Jb„)'/" < 1. If we define a„ := E"=ib„ show that
lim su p „^ ia„)‘/" < 1.
86 Lie Algebras

(c) Let A be an associative algebra with no zero divisors. Assume A


admits a filtration of subspaces A = \J7=o-^i

IK =. ^0 <=••••
AiAj c Ai+j for all i, j e N.
dimjj Ai < 00 for all i e N.

Defineb„ == d i m S h o w that if limsup„ < 1,


then A is an Ore domain. Let S3 and § l2 be the Virasoro
algebra and affine —§ l2 respectively. Show that U(9S) and
U(§l2(IK)) are Ore domains.
(d) Prove that if g is a finite-dimensional Lie algebra, then U(g) is
an Ore domain.
(e) Let g = ® “=ig" be a graded Lie algebra such that dim,^ g" < oo
so for all n e Z+. Suppose that limsup„_j[dim g")*/" < 1. Show
that U(g) is an Ore domain.
1.16 Let R be an Ore domain, and let S c i? be a subring.
(a) Show that if R is free as a right S-module, then S is an Ore
domain.
(b) Let a c g be Lie algebras. Show that if U(g) is an Ore domain so
is t/(a).
1.17 (See J. -P. Serre, Lie Algebras and Lie Groups, Benjamin, New York,
1965.) Let Y be a nonempty set. Let A ( X ) and ^(36) denote the free
associative and free Lie algebra on X. Recall the augmentation ideal
A+ (X ) = ®„^ qA(X)". Define a linear map

by

d( jii,. . . , [■^i> • • • >■^n]


:= a d ^ i ••• adx„_i(A:„)

for all eX
Let ad: § (X ) ^ gI(g(X )) be the adjoint representation. It extends
uniquely to an associative algebra homomorphism

ad: A { X ) ^ E n d g (Y ).

(a) Prove that for all u ^ A i X ) , v e ^ ( X ) + , d(,uv) = 2id(u)d(v).


(b) Prove that dlg(Ar) is the derivation of g (X ) for which d\^^xr is
scalar multiplication by n (Dynkin’s theorem).
Exercises 87

(c) Let X, y ^ X, X ¥=y, and let

z := Iog(exp X exp y)
m+1
( - 1)
= E m E —
m= l P+<J>0

and let z = Ez„, where z„ g g(AT)".


Prove the following explicit formula for the terms z„ of the
Campbell-Baker-HausdorflF formula:

~ ^ ^P,q’>
" p + q= n

where

(-!)"■ +' (a d x ,r(a d y i)« ' • • • (ad x)^"-(y)


Pl+ ••• +Pm=P m P i ' - - - q x '--“ P j
- +q^_^=q-\
Pi-^Qi>\

(ad X i/'(ad y,)^' • • • (ad y / ”- ( x )


P l + • • • + P m - l = P - ‘^
m P l'-Q l'- - -- Q m -l'-
^1+ ••• +<?m-l=<?
Pi +Qi>l

1.18 Let Z be a set, g(AT) the free Lie algebra on X over K, and let M be
an g(A!’)-module. Let d: X ^ M he. any map. Show that there exists a
unique extension of d to a linear mapping d: S(A') M satisfying

d{[a,b]) = ad (b ) - bd(a) for all a ,b ^^{X ).

In particular show that any mapping d: X g(A') extends uniquely to


a derivation of §(Ar).
1.19 Compute all automorphisms of the Witt and Virasoro algebras.
1.20 Let g be a perfect Lie algebra, and let § be its universal covering
algebra. Prove that every derivation (automorphism) of g lifts uniquely
to a derivation (automorphism) of §.
1.21 Prove all the results quoted without proof in Section 1.12. [Hint: See
Lang, Algebra, Exercise p. 175.]
1.22 The following exercise, based on Bourbaki, covers the fundamental
facts about solvable and nilpotent Lie algebras.
88 Lie Algebras

Let g be a Lie algebra over an arbitrary field IK. Recall that Dg := [g, g]
is an ideal of g. Define decreasing sequence of ideals D®(g) d
D^(g) D • • • and ^ T f\g ) 3 • • ♦ inductively as follows:

(derived series) D°( q) == g


D"^*(9) = [D"(9)>D"(g)] for all n > 0
(central series) ^ °(g ) = 9
^ " ^ ‘(9) = [9, ^"(9.)] for all n > 0.

(a) Show that all the terms in these series are ideals of g that are
stable under every derivation of g.
(b) Let /: g i) be a Lie algebra epimorphism. Show that
= 3% ii) and that /(-^"(g)) = Conclude that 3 g is the
smallest ideal a of g making g /a abelian.
Henceforth in this exercise we assume all Lie algebras to be
finite dimensional.
(c ) For a Lie algebra g show that the following conditions are
equivalent

N1: = (0) for some n ^


N2: There exists a decreasing sequence of ideals 9o ^ 9 ^ 9i ^
• ** ^ ^k = (0) such that [g, g j c g.^^ for all 0 < / < A:.
N3: There exists a decreasing sequence of ideals g^ 3 • • • d g^
as in N2 and such that dim,,^ 9i/9/+i = 1 for all 0 < f < A:.
N4: There exists N ^ N such that ad jCj ® ° ad = 0 for all
9.

(d) A Lie algebra satisfying the equivalent conditions of (c) is said to


be nilpotent. Show that subalgebras, quotients, and central exten­
sions of nilpotent Lie algebras are nilpotent.
(e) Let K be a (K-space, and let x e Endj^(F). Consider
ad x:gi(V) gl(V) by ad x : y [ x , y] — a : ° y — y © x. Show
that for all A/^ e Z .

a d : c ''( y ) - L ( ^ ) ( - 1)' X N —i oyoJC.i


/ =0 ^ ^
Conclude that if x is nilpotent so is ad x.
(f) Let V ¥= (0) be a K-space, and let g be a finite-dimensional
subalgebra of gI(L) consisting of nilpotent elements of Endj^(L).
Show that there exists i; e F, # 0 such that g • i; = 0. (Assume
that the results holds for dimension < n '= dim g. If is a
subalgebra of g of dimension m < n, use (e) and induction
Exercises 89

to conclude that some nonzero element of g/f) is annihilated


by all of adf). Conclude that g has an n - 1-dimensional
ideal ]^. If ;c e g \]^ , show that U •= {v ^ V\ii • u = 0} is a
nonzero subspace of V that is A:-stable.)
(g) (Engel’s theorem). Show that for g to be nilpotent it is necessary
and sufficient that ad ;c e gl(g) be nilpotent for all x e g. [Use
(f) to show that g has nontrivial center c. Now use induction on
dim(g/c) and (d).]
(h) For a Lie algebra g show that the following conditions are
equivalent:

SI: D"(g) = (0) for some n ^


S2: There exists a decreasing sequence of ideals g = go 3 g^ d
• • • D g^ = (0) such that g,/g,+i is commutative for all
0 < / < A:.
Such a Lie algebra is said to be solvable. Note that every nilpo­
tent Lie algebra is solvable.
(i) Show that subalgebras and quotients of solvable Lie algebras are
solvable.
(j) If a is an ideal of g and both a and g /a are solvable (as Lie
algebras), show that g is solvable.
(k) Let g ¥= (0) be solvable. Show that there exists an ideal of g of
codimension 1.
1.23 Let g be a finite-dimensional Lie algebra. Let

^ ( 9 ) == n
«=0
(See Exercise 1.22.) Show that for an ideal a of g, g /a is nilpotent if
and only if a D T^"(g).
Chapter Two

Lie Algebras Admitting


Triangular Decompositions
Agathon: But it was you who proved that death doesn’t exist.
A lle n : Hey, listen—Fve proved a lot of things. That’s how I pay my rent.
Theories and little observations. A puckish remark now and then. Occasional
maxims. It beats picking olives, but let’s not get carried away... So—what else
is new?
A g a t h o n : Oh, I ran into Isosceles. He has a great idea for a new triangle.
—Adapted from W. Allen’s “My Apology”

The Lie algebra gI„([K) can be decomposed in a rather obvious way into the
sum of three subalgebras: the strictly upper triangular matrices, the strictly
lower triangular matrices, and the diagonal matrices. This “triangular decom­
position” is a basic ansatz that appears in all the split semisimple Lie
algebras, the Kac-Moody algebras, and the Virasoro algebra. A considerable
amount of the basic theory of these various types of Lie algebras depends
upon nothing but this triangular decomposition, and hence is common.
In this chapter we develop the triangular decomposition theory. The
technical definition that we adopt here in Section 2.1 was inspired by the
work of Rocha-Caridi and Wallach [R-CW]. At the foundation of the
representation theory of Lie algebras with a triangular decomposition are the
highest weight modules, and particularly the universal highest weight mod­
ules or Verma modules of Sections 2.2 and 2.3.
Section 2.4 considers the special case of the three-dimensional simple Lie
algebra §I2(IK). It is hard to overemphasize the importance of the represen­
tation theory of this Lie algebra in the study of semisimple and Kac-Moody
Lie algebras (also in physics!). The material of this section is used extensively
in Chapters 4, 6, and 7.

90
2.1 Triangular and Weight Space Decompositions 91

A basic tool in representation theory is the notion of formal characters.


These characters are essentially formal power series that assemble into a
single expression all the information about the dimensions of the weight
spaces of a given representation. The fact that oftentimes characters can be
given by unexpected and remarkably beautiful formulas (see Section 6.4) is
one of their fascinations. In Section 2.5 we introduce characters, derive the
character formula for Verma modules, and relate it to the theory of parti­
tions.
In Section 2.6 we introduce the module category & of Bernstein-Gelfand-
Gelfand (BGG). This category provides a natural setting for highest weight
modules. All the modules in 0 have characters, and we can even make a ring
out of them. There are various series, particularly a weak form of composi­
tion series, that are available for modules in 0 , and these series play an
important technical role in the development.
Section 2.7 deals with various notions of radical ideals for Lie algebras
with triangular decompositions. Many of these ideals vanish in important
cases, although there are difficult, and even unanswered, aspects to this
which we will come in Section 4.5.
The structure of Verma modules has turned out to be a rich and subtle
subject, in both the Kac-Moody and Virasoro settings. One of the tools by
which we may understand this is the Shapovalov form and Shapovalov
determinant. We introduce these objects in Section 2.8 and then use them in
Section 2.9 to develop the Jantzen filtration, a device that allows one to study
the effect on the structure of Verma modules as the highest weight (the
single characterizing parameter) is allowed to vary.
The BGG duality of Section 2.10 introduces a new collection of categories
of modules, and in particular a class of projective modules that allow us to
see a remarkable duality between projective and Verma modules, on the one
hand, and Verma and irreducible modules, on the other. This result is used
in Section 2.11 to give a general result about the possible embeddings of
Verma modules into Verma modules. In Section 6.7, where we specialize to
the Kac-Moody case, we get definitive results on this.
The results of Sections 2.9, 2.10, and 2.11 are not used until Section 6.6
and not at all in Chapter 7. For this reason they might well be skipped on
first reading.

2.7 TmANGlJLAR AND WEIGHT SPACE DECOMPOSITIONS

Let ^ be an abelian Lie algebra, and let ^ qI(M ) be a representation


of in a IK-space M. Let g 1^* (the dual space of 1^). A nonzero vector
i; G M is called a weight vector of weight <f>(relative to and tt) if for all
A G

7r(h)v = <¡>{h)v.
92 Líe Algebras Admitting Triangular Decompositions

Recall from Section 1.8 that U(l^) may be identified with the symmetric
algebra 5(^) of 1& and that any representation 17 of lifts uniquely to a
representation tt of U(]^). Suppose that u is a weight vector for i) of weight
<t>. The linear mapping <^: 1^ ^ IK extends uniquely to a homomorphism

<t>: S(f|) ^ K.

If hi, /12, . . . , are any elements of then

••• *1*) = E «i<^(*i)‘'<^(*2)'" •••

where i = (/i, /2, . . . , ^ and a-^ e K, It follows that

( 1) tt( u) u = <t>{u)v for all w G 5(]^), i; G M.

For each </> e define the </>-weight space by

= {y G M \7r(h)v = (f>(h)u, for all /z e .

Clearly is an 1^-submodule of M.
Let 'f) be an abelian Lie algebra, and let ir: ij qI(M ) be a representa­
tion of We say that the 1^-module M admits a weight space decomposition
if M = This sum is necessarily direct (see Proposition 1). The (f)
for which ¥= (0) are called the weights of M relative to and tt. The set
of all weights of M will be denoted by P(M), For each e 1^*, dim is
called the multiplicity of in M and is denoted by mult^(<^). Evidently
mult^(</)) > 1 if and only if is a weight.

Proposition 1
Let ^ be an abelian Lie algebra, and let M be an f)-module admitting a weight
space decomposition. Then

(i) the sum is direct,


(ii) every ^-submodule N of M admits a weight space decomposition, and
moreover n N for 12// </> e f)*.

Proof Let be a submodule of M. We have n A/^ for all e ]^*


by definition. Let x ^ N and write x = where x"^ e and R is
some finite set of distinct weights of M . l f R = {</>}, then x e O N = AT^.
We show by induction on card R that each x"^ ^ Suppose that
card R > 1, and let and (f>2 ^ R, (l)^ ¥= 02- Choose /z g so that </>i(/z) #
<l)2(h). Then

h ■X - <l>2(h)x = 52 - <l>2Íh))x'‘‘ :N.


<I>s r \[4>2)
2.1 Triangular and Weight Space Decompositions 93

By the induction hypothesis we may assume that each component -


(f>2(h))x'^ e In particular so that ^
Applying the induction assumption again, it follows that each x"^ belongs
to N and hence that N admits a weight space decomposition. This estab­
lishes part (ii). Now part (i) follows from the same argument when x is taken
to be 0. □

Example 1 (^I„(IK)) Let be the set of all diagonal matrices in ^I„([K).


Then has a basis ..., where

0.
*0
hi. '= (1 in the ( k , A:) -position).
-1

*•0 /
Let M = K", and let tt be the natural representation of on M
(Example 1.6.4). Then

(1 in the A:th position)

lo
is a weight vector for ij with weight given by

( 0 if /: + 1,
< f > k { h j) = l 1 if k = j ,

[-1 if A: = ; + 1;
k = j = - 1.

Furthermore and M = showing that M admits a


weight space decomposition.

Example 2 (The Virasoro Algebra) Let 93 = E“ = _oolKL„ 0 (Kc be the


Virasoro algebra, and let = IKLq ® IKc. Think of S as an f)-module
through the adjoint representation restricted by Then
[L q, L„] =
94 Lie Algebras Admitting Triangular Decompositions

shows that L„ is a weight vector of weight A„, where A„ e 1^* is defined for
all Me Z by
An(^o) =«>
A„(i) = 0.

Since A„ = nAj, we have iB"'" = 93"'^' = KL„ if n # 0 and 93° = f). This
shows that 93 admits a weight space decomposition relative to ^

Proposition 2
Let Q be a Lie algebra, and let ^ Der %be a representation o f an abelian
Lie algebra as a Lie algebra o f derivations on g.

(i) Suppose that g admits a weight space decomposition

9 = 0 9“

with respect to under the representation tt. Then


(ia) for all a,l3 ^ f)*, [9“, 9^] c 9“"^^ (thus g is graded, as a Lie
algebra, by the Z-span of its weights)-,
(ib) U(g) admits a weight space decomposition {relative to the repre­
sentation o f on U(g) extending that o f b, on % {Proposition
1.8.3)), and for all a, /3 e

U (g )“U (g)'’ c U ( g ) “ ^ ^

Thus U(g) is graded, as an associative algebra, by the Z-span of


the weights o f g. Moreover this grading is the grading o f U(g)
inherited from that of g in part (i) {Section 1.8, Remark 2).
(ii) Suppose that g is generated as a Lie algebra {see Section 1.10) by a set
of weight vectors Xj, j e J, with Tr{h)xj = 4>j{h)xj for each j and for
all h e ]^. Then g admits a weight space decomposition g = ©g“ with
Xj e Q'I’J and each weight a e LZ<f>j. In particular the results o f part
{ia) apply to g and U(g).

Proof Identify g as a Lie subalgebra of U(g) in the usual way. We know that
every element of U(g) is a sum of products x^X2 . . . x ^ , where each x, e g.
In view of the assumptions on g (in either part (i) or (11)), we can assume that
each X; lies in a weight space, say, g“'. Then

TT{h){xiX2...Xk) = ix^...TT{h)xj...Xk

= («1 +ak){h)x^X2...Xk.
2.1 Triangular and Weight Space Decompositions 95

Part (ii) and part (ib) clearly follow from this (for (ii) invoke Proposition 1
with g taken as an l^-submodule of U(g). Since for all x, y e g, [x, y] =
- y;c we may deduce (ia) from (ib). □

Let g be a Lie algebra over a field K. A triangular decomposition of g


consists of an abelian subalgebra (0) and two subalgebras g+ and g_
such that

TDl: g = g_© f| 0 g^;


TD2: g+^ (0), [1^, g+] c g+, and g+ admits a weight space decomposi­
tion relative to f) (under the, adjoint representation) with weights
a =5^ 0 lying in a free additive semigroup 0 + ^ ^*5
TD3: there exists an anti-involution a (i.e., antiautomorphism of period
2) on g such that

<r(9+) = 9 -,
o-ift =

TD4: there exists a basis {ccj)j^j of 0 + consisting of linearly indepen­


dent elements of 1^*. In particular 0 + consists of all nonzero finite
sums of the form

rrijaj, mj

The triangular decomposition is said to be regular if in addition the weight


spaces of g+ in TD(2) are finite dimensional. Notice that the condition
0 ^ 0 + is equivalent to

( 2) C + n (- ! 2 + ) = 0
and also that

(3) o-(g_) = g+.

Remark 1 We will usually refer to a triangular decomposition of a Lie


algebra g as a 4-tuple (^, g+, 0+ , ir), where f),g+, 0+, and < t are as
above. It will be understood that g_== cr(g+) and that g = g_0 ^ 0 g+.
Sometimes we simply write g = (^, g+, 0+, c7)t o describe this situation.

Let g be a Lie algebra admitting a triangular decomposition, and assume


that the above notation is in effect. Let x e g+, a 0 0^_. Then [h,ax] =
[crh, ax] = a[x, h] = —a{h)ax for all A 0 from which we see that a x e
g l“. It follows that g_ admits a weight space decomposition relative to
96 Lie Algebras Admitting Triangular Decompositions

with weights in —0+;


g_“ = <TQ%,
(4)
dim g l" = dim g“ for all a e Q_^.;
and g admits a weight space decomposition relative to with
S+ if a e e + ,
if - a e
(5)
if a = 0,
otherwise.
Let Q be the group generated by We know that Q =
(which is a lattice in 1^* whenever J is finite). We call Q the root lattice
(regardless of whether or not J is finite). We call the a ^ Q for which
g" ¥= (0) the roots of g (relative to 1^) and the corresponding spaces g“ the
root spaces. We define the height function
h tiQ ^Z

by ht\ ^j^j] = H

This terminology, due originally to W. Killing in this work on finite­


dimensional simple Lie algebras, derives from the fact that for each h
the quantities a{h) are the characteristic roots of ad h (which is obviously a
diagonalizable transformation). There is some nonuniformity in whether or
not 0 is to be considered as a root. Classically in the case of semisingle Lie
algebras it is not, but in the more general context here it seems more natural
to do so (as we have just done). For each root a, dim g“ is called the
multiplicity of a and the set of all roots is called the root system of g rela­
tive to ^ ' = (t), Q^, a) and is denoted most commonly by A.
Define A+= A n 0 + and A_= - A+. Then TD2 and (4) and (5) give us that
A = A+u{0} U A_. We call A+ (resp. A_) the set of positive (resp. negative)
roots of A.
We call the diagonal subalgebra of the decomposition and g+ (resp. g_)
the upper (resp. lower) triangular subalgebra of the decomposition.
Remark 2 \i ^ = (1^, g+, 0+, a) is a triangular decomposition of g, then it
is immediate that g _ , - 0 + , cr) is another. We call this the
triangular decomposition opposite to S^.

Example 3 (g = ^I„(K), n > 1). Let g+ (resp. g_) be the subalgebra of


strictly upper (resp. strictly lower) triangular matrices. Let 1^" be the subalge­
bra of diagonal matrices in gI„(K), and let Ij = 1^" n §I„((K). B o th 't)" and
are abelian subalgebras of gI„(IK). The triangular decomposition that we are
looking for is
^I„(IK) = g_0 f) e g+.
2.1 Triangular and Weight Space Decompositions 97

Transposition X ^ serves as the anti-involution, pointwise fixing and


interchanging g_ and
To determine the roots, we introduce the matrix units 1 < /, j < n:
Eij is the matrix with a 1 in the (i,;)-entry and O’s elsewhere. Let Cj,. . . ,
e be defined by

and let a^,. . . , be defined by

Since + • • • +£„ vanishes on 1^, it is easy to see that

{ (E « ie,)|Ja,- s K, £ a ,. = o} =

and hence that a„_i is a basis of 1^*. For h e 1^*,

[h, = /t£,, - £,,£ = (s, - e ,)(;i)£ ,,.

For i < j this gives

= (a , + ••• +aj_i)(h)Eij.

Thus E^j e is a weight vector with weight a, + • • • +ay_i.


Set Q+-= {Em,a,|m, e l\j} \{0}. Then

9+= ® g: = 0 KE,j
i<;

is the required weight space decomposition. Of course Ej^ = Ejj has weight
-{ai + • • • -\-aj_i) for all i < j\ and

g_= © KE¿j
i>j

is the weight space decomposition of g_. Thus ^I„(IK) admits a regular


triangular decomposition. It follows that gI„(IK) also has a triangular decom­
position with the same g+ and g_ but with replaced by %. These two
examples are the origin of the terminology.

Example 4 (The Virasoro algebra) Referring to Example 2, set g+ =


^ ® Let Aj e 1^* be the weight
of Lj. Then with Q+ = we verify TD l, TD2, and TD4 of our definition.
Let a be the linear transformation defined by aLj = L_y, j ^ Z, = 0.
98 Lie Algebras Admitting Triangular Decompositions

This provides S3 with a regular triangular decomposition. Of course it is easy


to modify this for the Witt algebra.

Example 5 (The affine algebra ^ l2(IK) of type See the Appendix.

Example 6 (Heisenberg algebras extended by derivations) Let a = a_ 0


Kc 0 a+ be a Heisenberg algebra so that and a_ are abelian subalge­
bras that are nondegenerately paired by the associated skew-symmetric
bilinear from «/r. Suppose that there are bases {a^\ and {a}}, ; g J, of and
a_ such that [ay, a\] = 5y^c (this is always possible if a is finite dimensional).
Define cr by

cr{aj)=a'j, y e J,

cr(c) = c.

Then axioms TDl and TD3 of the definition of triangular decompositions


are fulfilled with = Kc. However, is the centre of a and hence cannot
produce the root spaces required by TD2. Just as in Example 5 this problem
is easily cured by enlarging 1^^. For each i e J and integer g Z+, we let d
be the linear operator on a defined by ¿(a,) = d¿a¿, d{a\) = -d^a\, d\i^^ = 0.
Then d is di derivation of a. We let if •= if^ ^ Kd = Kc ® Kd and make
f i : = a 0 l K d = a _ 0 ] ^ 0 a + into a Lie algebra containing a as a subalgebra
by defining (Exercise 1.3)

[d, x] = d{x) for all X G a ,


[d,d] = 0.

Then, if we define a g 1^* by a(c) = 0, a(d) = 1, and set = E^.=^IKa„


k a
we see that a has a weight space decomposition with a+=
i<—ka
« - = ®;t>0 a ij = a . After extending cr to a by a(d) = d,
(]^, a+, Z^_a, cr) is a triangular decomposition of a with root lattice Q = Za.
The problem that was raised in these last two examples, namely that of
extending the algebra in order to produce more eigenspaces, is one that we
will see again in our study of affine Lie algebras.

Remark 3 Later on in the text (Section 2.9) we will find it useful to have
the notion of triangular decomposition for Lie algebras over rings. Suppose
that ^ is a commutative ring of characteristic 0 (i.e., Z is a subring of AX Let
fj be an abelian Lie algebra over A that is free as an .^4-module (Remark 1 of
Section 1.1). Set

r = H o m ^ (^ ,^ ).
2.1 Triangular and Weight Space Decompositions 99

Suppose that M is a free ^-module that is also an 1^-module. We define


the weight space /x e 1^*, just as above:

:= {y e M\h • V = ii{h)v for all h ,

M has a weight space decomposition if M = The sum is necessar­


ily direct.
Let g be a Lie algebra over A that is free as an ^4-module. A triangular
decomposition of g consists of an abelian subalgebra f) and two subalgebras
g+ and g_ of g (all free as ^4-modules) for which TD1-TD4 hold.
Let g = (]^, g+, (2+, a) be a Lie algebra over A with triangular decomposi­
tion. Let jB be a commutative ring containing A, and let g == B 0^ g be the
Lie algebra over B obtained by extension of the base ring from A io B (see
Section 1.1). Let 'i) = B and g+= B 0^ g+. These are subalgebras of g
that are free as S-modules. Every ^ induces an ^-bilinear map

: i5 X
by
At: (6,/ï) biJi{h),

and hence an ^-linear map (also denoted ¡lI)

/1: B 0^ f) ^ 5 .

One verifies that ]I is jB-linear, and hence there exists an (injective) group
homomorphism

"ril* ^ 5 * .
Along similar lines one shows that there exists a unique antiinvolution â
of g satisfying
a{b 0jc) = b 0 ot( a:).

Then is a triangular decomposition of g which is said to be


obtained from (1^, g+, Q^, a) by extension of the base ring from A to B, The
most obvious examples are when g is defined over the field IK and IK' is some
extension field of IK.

Example 7 (Complexification of a real triangular decomposition) Let g be


a Lie algebra over the real numbers IR. Assume (1^, g+, Ô+, a ) is a triangular
decomposition of g. Let gc '= C 0ff^ g be the complexification of g. Then
evidently g^ = (gc)-® ® (9c)+ ^be obvious notation) and, with as
the C-linear extension of a to g^, (^ c?(9 c)+ j Ô+? a triangular decom­
position of g^ (see the preceding remark). However, it is convenient for
100 Lie Algebras Admitting Triangular Decompositions

Other purposes to extend a to a semilinear map so that cr(ojc) = acr{x)


for all e C and for all x e g. We denote this extension, which is unique,
simply by cr, and we call the 4-tuple the hermitian
complexifícation of(l^,g+,(2+>i^)- The real subalgebra 5 of g^ is recovered
as the set of fixed points of a- in Conversely, suppose that we are given a
complex Lie algebra m, an abelian subalgebra f, and a semilinear anti-in-
volution cr on m that stabilizes f. Let be the set of fixed points of a in f.
Then if is a real form of f since, for any jc e f, we have x = l/2 ( x + o-jc -
Kix + aix)). The 4-tuple (f, m+, Q^, a ) is called a (regular) hermitian trian­
gular decomposition if it is the complexification of a (regular) triangular
decomposition (^,g+,j2+,o-)of some real form g of m. Note that Q^czi^*
and ]^* can be thought of as the real subspace of f* consisting of those
functions that are real valued on i).

We will see in Chapter 4 that all the affine algebras, and indeed the far
wider class of contragredient Lie algebras, possess regular triangular decom­
positions that are fundamental to their study.

Proposition 3
Let q be a Lie algebra admitting a triangular decomposition (^, g+, Ô+, (r),
and suppose that
9= © 9"
aGA

is the corresponding root space decomposition. Let <t>be an automorphism o f g


stabilizing Then (¡) maps root spaces onto root spaces. More precisely
</>(f|) = Í), cind if (f>*: ]^* -> Í)* is defined by </>*(/) = f°(f>~^ for all f ^ 1^*,
then 0(g") = g"^*“ for all a ^ A. In particular <f>* determines by restriction a
bijectiue mapping o f A onto itself.

Proof Let CKe A, jc e g", and H Then

which shows that e g° = 1^. Thus Now

and hence </>(x) e Together these show that (f> maps root spaces into,
and hence onto, roots spaces in the desired way. It follows that <^*A = A.

The group of automorphisms of g that stabilize f) will be denoted by


Aut(g, ij). Let
Aut(A) := {<l> e AutK(i|*)l<^(A) = A}.
2.1 Triangular and Weight Space Decompositions 101

Proposition 2.3 gives rise to an exact sequence

\ K ^ A ut(g, ]^) Aut(A).

Let 6: Q ^ be a group homomorphism (from an additive into a multi­


plicative group). There exists a unique e Aut(g, 1^) such that

/elg“ = scalar multiplication by ^(a) .

We define
KQ ={f,\e^H om {Q ,K ^)}.

Clearly K q c K,

Proposition 4 [B el]
Let % be a Lie algebra admitting a triangular decomposition (i), Q+,Q+,(t ).
Suppose that dim < oo and dim < dim g+. Let D q c Der(g) be the deriva­
tions of g stabilizing each root space o f g. Then Der g = D q + ad g and
D q n ad g = ad 1^.

Proof Let 5 be any derivation of g mapping ^ into 'i). Then for any root a
and any e g“ we have for all e

a{h)8x'^ =

which we write as

(6) (ad/i - a ( h ) l ) 8 x ^ = ~ [ 8 h , x ^ ] e g“.

If 8x°" were to have some component in g^, /3 ^ a, then with /t e such


that p(h) a (h \ we have a contradiction to (6). Thus 6g“ c g“, and hence
8 e D q. It is obvious that D q n ad g = ad 1^.
Let dim ^ = n, and let ..., ^ A+ be arbitrary. We can find a basis
{hy, . . . , /i„) of f) such that a^ih^) = 1, i = l , . . . , n . In fact proceeding
inductively, if linearly independent h^,...,hf^_^ have already been chosen so
that afhi ) = l , i = 1 , k - 1, then set •= {h e ^ a f h ) = 0}. Since
cannot be written as the union of two proper subspaces, we can choose

X + ••• UH^}.

Then hi^ := x / a j f x ) is as desired.


Now let basis of g+ taken so that each Xj^ lies in some
root space g“^, e A+, A: e S. Let {h,^}f^^^ be a subset of such that

^k(^k) = 1 for k ^ S.
102 Lie Algebras Admitting Triangular Decompositions

By the above we can assume that spans 1^. Evidently {crAc^l^es is a


basis of g_.
Let d e Der g, and define bf^j e K and h\ e by

Since dim dO)) < «>, and b/^j are zero except for a finite set of values of j
(independent of k). In particular ajj and bjj are zero except for finitely
many j.
From
o = d[h^,h^] = [dh^,h^] + [h^,dh^]

we have

- b^jaj(h^) + b^¡aj(h^) = 0 for all p, q, j.

With p = j,

^ pp^ pV^p ) ^qp^


^pp^pi^q) ^qp’
Let
------ E«ppO-JCp + 'LbppXp e g.
P P

Then for all ^ S,

[^^>3^] “ p ^J^pp^p{j^q)^p
P P

= Hagpcrxp + T^bgpXp
P P

= d h ^ ~ h',,
so we have
(d + a d y ) ( / i J = A ' , e ^ .

Then d' •= d ad y: if if, since the {h^} span 1^, so d' e D q, providing the
proposition. □

As an example one may easily show that every derivation of the Virasoro
algebra is inner (Exercise 2.24).
2.2 Highest Weight Modules 103

2.2 HIGHEST WEIGHT MODULES

Let 9 be a Lie algebra over K admitting a triangular decomposition


(1^, 9+> Q+, and write g = g_0 0 g+. Let tt be a representation of g
on a space M. We know from Section 1.8 that ir lifts to a representation of
U(g) on M.
If y is a vector in M, then

7r{n(q))v {7t( u ) v \u

is a submodule of M [since it is closed by the action of 7r(g)], and indeed it is


the smallest submodule of M containing v. We call it the submodule of M
generated by v,
A nonzero vector z; e M is called a highest weight vector for g (relative to
the triangular decomposition) if

(i) z; is a weight vector relative to the action of 1^.


(ii) tt{ x ) v = 0 for all X g g+.

A g-module M is said to be a highest weight module if M contains a


highest weight vector u that generates it. The weight A g ]^* of the highest
weight vector v is called the highest weight of M. We call the pair (A, z;) a
highest weight pair of M. The representation of g on M afforded by M as a
g-module is called a highest weight representation.
Remark I We will see shortly (Remark 2) that highest weight vectors of a
highest weight module are unique up to nonzero scalar factors. In particular
all highest weight vectors have the same weight. In addition all the other
weights will be seen to be “lower” than A (Proposition 2.1). These two facts
justify the above terminology.

Highest weight modules occur very frequently in the sequel. Although in


detail there are many differences between highest weight modules for differ­
ent Lie algebras and many subtleties to be understood, the broad features
are quite uniform. In the physics literatures as well as in many mathematical
papers, highest weight vectors are called vacuum or null vectors.
Let M be a highest weight module affording the representation tt and
with highest weight pair (A, z;). Using the Poincaré-Birkhoff-Witt theorem, we
have U(g) = U(g_)U(]^)U(g+). Thus
M = 7r(U(g))z;
= 7T(U(g_))7r(U(^))7r(U(gJ)z;
= 7r(U(g_))7T(U(^))z;
= 7r(U(g_))z;,
104 Lie Algebras Admitting Triangular Decompositions

where the last equality is a consequence of the fact that is a simultaneous


eigenvector for 7t(U(]^)) = [see 2.1(1)].
According to Proposition 2.1.2 and 2.1(4), g and U(g) admit weight space
decompositions with weights in Q and are also graded by Q. All the weights
of g_ lie in -Q+, and those of ll(g_) in -(2 +^(0)-
Let P be a subgroup of 1^*. A (2 +-fan in P is any set of the form

A I 0 + - {A} U {A -/3li3 e 0+},

where A e P. We call A the source of the fan. Fans are useful objects for
keeping track of the weights of representation that we will study. Note that

Q-U{0} = 0 i Q ^

Proposition 1
Let q be a Lie algebra admitting a triangular decomposition (ij, Q+,Q+,cr),
and let Q denote its root lattice. Let M be a highest weight module with highest
weight pair (X,v) and ajfording the representation tt. Then

(i) M admits a weight space decomposition relative to where all the


weights lie in the fan A i Q^. Furthermore = '7r(U(g_)““)i;/or
all a ^ !2+^{0}. In particular = Ko.
(ii) the weight space decomposition o f part (i) defines a grading o f M by
ZA + Q and, for each a ^ ¡2? 9“ tind U(g)“ act as homogeneous
operators o f degree a on M.
(iii) if the decomposition o f q i s regular, then all o f the weight spaces o f M
are finite dimensional.

Proof By the preceding remarks, M is spanned by vectors of the form


v{u)v, where u e U(g_)“", a e (2+Li{0}. We have

(1) Tr{h)7r{u)v = \'Tr{h),'Tr{u)\v + 'Tr{u)7r{h)v

= —a { h ) 7 r { u ) v - \ - \ { h ) T r { u ) v

= (A —a){h)Tr{u)v,

which shows that M is spanned by weight vectors and hence has a weight
space decomposition. In addition it shows that = 7r(U (g_)"“)o and
that all weight spaces of M are of this form. (Since modules are by definition
graded only by groups, we use the group ZA + (2 for the grading even
though it is clear that all the weights lie in the coset A + Q.) That U(g)“ and
g“ act homogeneously is clear from (1).
2.2 Highest Weight Modules 105

Finally, suppose that the decomposition of g is regular. Now U(g_) has all
its weights in O iQ ^, and except for U(g_)® (= K.l), each weight space
U(g_)”^ is spanned by products x ^ .. . Xf^ of weight vectors with weights
-Pi, e - Q ^ and p = • -\-p¿. Since there are only finitely
many ways to decompose p ^ Q+ into parts that are also in Q+, and since
dim g l" = dim g““ < oo for all a e Q_^, we see that dim U(g_)“^ < oo. Thus
= 7r(U(g_)“^)¿; is also finite dimensional, and this proves part (iii).

Remark 2 The use of the terminology highest weight is now apparent. A is


a weight and all other weights are “lower” because they are of the form
A - a, a e (2_^. Proposition 1 shows that A and IK¿; are unique. Indeed, if w
is a highest weight vector for M of highest weight ¡jl, then \ —a =
(fjL — P) — a for some a, P ^ (2 + U{0}=>a+)S = 0 => a = p = 0 => fi =A.
Then dim = 1 => Kw = Ki;.

Example 1 (The Natural representation of §I„([K)). We use the notation


established in Examples 1 and 3 in Section 2.1 with M = K" and tt the
natural representation of ^I„(IK) on M. Then M = is a weight space
decomposition of M. Since E^je^ = 0 for all i < j, g+Ccj) = 0, so is a
highest weight vector. Since for / = 1 ,..., n, M is generated by
ej, and hence M is a highest weight module. The weights are </>2 = 0i -
Oil, <l>3 = <l>i - - a 2 , . . . , (f>n = <f>i - Oil - ' - All the weight
spaces are one dimensional.
Remark 3 It is a fact that if K is algebraically closed, all finite-dimensional
irreducible representations of ^I„(1K) are highest weight modules.

Example 2 (Heisenberg algebras) We return to the Heisenberg algebras of


operators described in Section 1.4. Thus a is the linear span of the operators
/(xy), d/dxj, and 1 acting on 5 = Let j j be any map of J into
Z+. Let a_= ElK/(jCy), a+= T.Kd/dxj. hct the degree operator D be defined
on S by

..., = («1/, + • • • + n . . . , x "Jnk


k

for all monomials in S. Thus Xj is given degree j. Let d •= —D. Then d is a


derivation of 5 (see Example 1.5.1), and

[d, K ^ j ) ] f = = - h j f + Xjdf - Xjdf = -¡l(xj)f

for / e 5
shows that [d,l(xj)]= —jl(xj). Similarly [d,d/dxj] = jd/dxj. Thus d be­
haves like the extension element of Example 6 of 2.1 and fi — Kd 0 a is a
Lie algebra of operators on S with triangular decomposition a _ 0 0 a+,
0 K l.
106 Líe Algebras Admitting Triangular Decompositions

Consider 1 ^ 5 . Evidently 1 = 0 and h • 1 = \(h)l for all h


where A e 1^* is defined by

A(l) = 1, X(d) = 0.

Since also <* - 1, we see that 5 is a highest


weight module. It is quite easy to see that it is irreducible, but we leave the
proof of this until Section 2.3. Let a e 1^* be defined by

a ( l) = 0, oi(d) = 1.

Then ..., is a weight vector of weight A - + ••• It


follows that the weights of 5 lie in A - Na. Except for = K ^ 1, the
weight spaces need not be finite dimensional unless J is finite.

Proposition 2
Highest weight modules are indecomposable.

Proof. Let M be a highest weight module with highest weight pair (A, y).
Suppose that M = ® M2, where and M2 are submodules of M. Then

= Mf e

forces either = (0) or M^ = (0). Say, M^ = (0). Then v : M^ so that


M = U(g) • V c Mj, and hence M2 = (0).

Remark 4 In general highest weight modules are far from being irre­
ducible. We will see this explicitly in Section 2.4.
Remark 5 Analogous to highest weight modules we may define lowest
weight modules M. In this case M is generated by a lowest weight vector
(G_* = 0, • i; c K v ), and the weights of M lie in upward fans of the form
At!2+- Lowest weight modules for g with triangular decomposition
Q+, 0-) are evidently highest weight modules for g with the opposite
triangular decomposition (see Section 2.1, Remark 2).

2.3 VERMA M O D U LES

In this section we show how to actually construct highest weight modules for
Lie algebras with triangular decompositions. In fact the modules that we
construct are universal in the sense that all other highest weight modules
occur as homomorphic images of them. They carry natural (contragredient)
bilinear forms (see Section 2.8), which are used later in the book to gain
insight into questions of submodules and irreducibility.
2.3 Yerma Modules 107

We begin with a Lie algebra g admitting a triangular decomposition


The root lattice will be denoted as usual by Q. Unless
explicitly assumed, the decomposition need not be regular.
Let M be a highest weight module with highest weight vector v and
highest weight A. In this section we will put more emphasis on modules than
representations. We will not usually specify the representation explicitly but
will simply write jc • w to denote the effect of the operator defined by jc e g
on an element w of M. Suppose that g_ has a basis of root vectors
indexed by some totally ordered set J. For each ; e J let e —Q^ denote
the root corresponding to Xj, that is, Xj e g^>. Then, by Proposition 2.2.1,
each of the vectors
Y^k V,
(1)
where A: > 0, ‘ ’ * > h, ^ Z+, is a weight vector of M of
weight A + + ••• and collectively they span M.
We are going to show how, starting with any A e ]^*, we can construct a
highest weight module M(A), as above, for which the vectors (1) form a basis
of M(A). We will also see that such a module is unique up to isomorphism.

Let A e A highest weight module M(A) for g with highest weight


pair (A, i;+) is called a Verma module if given any highest weight module M
with highest weight pair (A, v) (the same highest weight A), there exists a
g-module homomorphism r¡: M(A) ^ M such that v,

A Verma module is universal in the sense that every highest weight


module with the same highest weight is a homomorphic image of it. It is clear
that the homomorphism rj in the above definition is unique [for the pair
(A, i;)] and surjective: for if v, then for all u e U(g_), u * u • v;on
the other hand, M = U(g_) *v. Also M(A) is itself unique up to isomor­
phism by the standard type of argument (see Section 1.2).
The existence of M(A) will be shown in two different ways. The first is by
the method of inducing modules. The second, which is more prosaic, is much
more explicit and gives one a better feeling for what the module looks like. It
also takes more work.

Proposition 1
Let (]^, g+, ¡2+, a ) be a triangular decomposition o f g. Then, for all A e 1^*, g
has a Verma module M(X) o f highest weight A.

Proof 1. Let b be the subalgebra 0 g+ of g. Let Kv+ be a one-dimen­


sional vector space and make it into a b-module by defining
g+* v += 0,
h • Vj^= \{h)u^ for all /r e ]^.
108 Líe Algebras Admitting Triangular Decompositions

Now IKd+ is a left U(b)-module and U(b) is a subring of U(g)- Hence we


may form the induced module obtained from by extension of the base
ring from U(b) and U(s) (see Section 1.1, Remark 1)

M(A) == U(g) ®u(6)IKt;+.

Clearly, M(A) is a left U(g)-module, and hence a g-module. Consider


1® A/(A). For X e g+,

X • (1 ® Ü+) = 1 ® 0»

and for h

/i • (1 ® Ü+) = 1 ® /1 • v + = 1 ® A(/i)i;+.

Also U(gXl ® y+) = U(g) ®u(b)IKi’+= Af(A). Combining these, we see that
M (\) is a highest weight module with highest weight vector 1 ® of
weight A.
Let M' be any highest weight module with highest weight vector v' and
highest weight A. The mapping

/ : U(g) X
f ( u , av^) = au ■v',

is bilinear and U(b)-balanced: Namely for w s 11(b),

f(uw,av+) = au • w • v' = f { u , w ■av^).

Hence there exists an induced linear mapping

/: U ( 9 ) +

satisfying u (S> au • v'. This is a U(Q)-niodule map, and it satisfies


1 0 i; + L?' as desired (see [BA2] for more on tensor products). □

Before starting on the second proof we isolate the following fact which we
will use several times in the sequel.

Lemma 2
Let A e 1^*. Then

(2) U(g) = U (g _ ) e |u ( g ) g + + £ U(9)((* "


2.3 Verma Modules 109

The reader may wonder about the notation ~


which the sum is incredibly redundant. Evidently the sum may be replaced by
a sum over any basis of f).

Proof. The mapping A: IK extends uniquely to a homomorphism

A: 5 (^ ) IK

whose kernel is ~ A(/z)l) (Exercise 1.9). Thus

5 (i|) = K 0 L 5 (^)(/г - A (/i)l).

Let b_!= b ® 9_, and recall that (PBW theorem. Corollary 2), U(b_) is a
free U(g_)-module admitting any IK-basis of U(b) as a U(g_)-basis. Now the
triangular decomposition and the Poincaré-Birkhoff-Witt theorem gives us

U(g) = U (g _ ) U ( b )U ( g J
= U (g _ )U (b )(K l® U (g J g + )
= U (g_)U (b) ® U (g)g^

= U (g_)(K ® E U ( i ) ( A - A ( / i ) l ) ) ® U (g)g^

(because U(b) = 5 (b ))
= U (9_) ® E U (b_)(/i - A (/i)l) ® U (g)g^
Asf)
(because U(b_) is a free U (g_) module).

Now 9+(/i —A(/i)l) c U(g)g^ since g^. is an b-module under the adjoint
representation, and thus

U (g)(/i - A (/i)l) c U (b_)(/i - A (/i)l) + U (g)g^. □

We now give the second proof of Proposition 1.

Proof 2. Let

p :U (g ) ^ U ( g _ )

be the projection determined by the direct sum decomposition (2) of U(g).


lio Lie Algebras Admitting Triangular Decompositions

The kernel of p is clearly a left ideal of U(9>- Define an action of U(9)


U(g_) by
y-u= p(yu) for all Me U (g_), y e U (g).

This makes U(g_) into a U(g)-module: Let y', y e U(g), u e U(g_), and
write yu = piyu) + r, where p(.r) = 0. Then

y' • (y • m) =p{ y'p {yu )) = p {y '{ yu - r))


= p(y'yu) = (y 'y ) • u,

since y'r is in the kernel of p.


For u', u e U(g_),
u' ' u = p{u'u) = m' m,

and therefor the action of 11(9_) on itself is left multiplication. Thus U(9-)
is generated by 1 as a U(g)-module.
For all y G U(g+),
y • 1 = p ( y ) = 0,

and for all e ]^,

/z • 1 = p{h) = p(A (/i)l + h - A(/z)l) = A(/z)l.

This shows that U(g_) has been made into a highest weight module with
highest weight vector 1 and highest weight A.
If M' is any highest weight module with highest weight pair (A, v ' \ then
the kernel of p annihilates v'. Thus the mapping U(g_) ^ M', u•
a g-module map showing that U(g_) (as a g-module) is a Verma module.

Remark 1 It is clear from Proposition 1 (especially the second proof) that,


as U(g_)-modules, M(A) = U(g_) and M(A)^”“ = U(g_)““, where the ac­
tion on U(g_) is left multiplication. We see that as U(g_)-modules, all the
M(A)’s look the same, namely like U(g_). It is the actions of and U(g+)
that are different.

Since U(g_) is an domain (Proposition 1.8.5), we have the important fact


that a Verma module M = M(A) is a torsion-free U(g_)-module; that is, for
u e U(g_), m ^ M,

(3) w • m = 0 <=> M= 0 or m = 0.

Any homomorphism of M(A) that is not injective annihilates some nontrivial


2.3 Yerma Modules 111

element u • v, u ^ U(g_) \ {0}. Therefore we arrive at the following impor­


tant conclusion:
Proposition 3
Let M be a highest weight ^.-module. For M to be a Verma module it is
necessary and sufficient that M be torsion free as a \liQ_)-module [i.e., (3)
holds], O
Remark 2 If M is a Verma module, then as a U(g_)-module it is
monogenic ( = cyclic) and torsion free, and hence free.

Corollary
Any submodule of a Verma module that is itself a highest weight module is a
Verma module.

From the left U(g)-module mapping


U(g) ^ M ( A ) ,
u ^ u • u ,.
we conclude that
(4) M(A) = U ( g ) //( A ) ,
where /(A) is the annihilator of in U(g).
From (3) and the second proof, we see that

(5) /(A) =U(9)g^+ ElI(s)(^-A(/i)l).

Often (4) and (5) are used to define Verma modules.


The next result shows us that irreducible highest weight modules exist and
are unique.

Proposition 4
Let (^,g+,!2+jOr) ^ triangular decomposition o f g, and fei A e 1^*.

(i) I f M ( \ ) is the Verma module o f highest weight A, then


(ia) every proper submodule N o f M ( \ ) has a weight space decomposi­
tion relative to and the weights lie in A —0+j
(ib) M(A) has a unique maximal proper submodule MA);
(ic) L(A) := M(A)/MA) is irreducible, and up to isomorphism, L(A)
is the unique irreducible highest weight module o f g with highest
weight A.
(ii) Any highest weight module M with highest weight A contains a unique
maximal proper submodule N. Furthermore M /N = L(A).
112 Lie Algebras Admitting Triangular Decompositions

Proof, (ia) If A/" is a submodule of M then in particular N is an 1^-submodule


of M. By Proposition 2.1.1, N has a weight space decomposition with weight
spaces n M If A is a weight of N, then = «1;+ and
then N = M, since generates Af as a g-module. Thus, if N is proper, all
the weights of lie in A —
(ib) Using (ia), we see that the sum A^(A) of all the proper submodules of
M is still proper (since none of them can contribute to M^). This establishes
(ib).
(ic) L(A) = M(X)/N(X) is clearly a highest weight module with highest
weight A, and since A^(A) is maximal, it is irreducible.
Let L' be any irreducible highest weight module with highest weight A.
Then, since M(A) is a Verma module, there is a nonzero homomorphism

77: M(A)

with M{XY for all weights (j) of M(A). We have already seen that rj is
surjective, and therefore its kernel is a maximal proper submodule of M(A).
Thus ri(N(\)) = 0 by part (ii), and

M(A)/iV(A) = L '.

(ii) From the universal property of M(A), there is a surjective homomor­


phism M(A) M. This takes weight spaces to weight spaces from which it is
clear that N ( \) maps to a proper submodule N of M. Then N is maximal,
and M / N = M{k)/N{k) = L(A). □

Example 1 (Verma modules for Heisenberg algebras) Following Example


2.1.6, assume that a = a _ 0 lK c 0 a + is a Heisenberg algebra with an
anti-involution a and bases {fly}, (cr(fly)} (; e J) of and a_ such that
[fly, cr(fl^)] = 8jj^c j, A: e J. We extend a to & = a _ 0 0 a^_, where = IKc
0 Kd and where [ d , fly] = fly, [d,a(aj)] = —o -(fly ) so that fi acquires a
triangular decomposition with root lattice Q = Za(a e 1^* given by a(c) = 0
and a(d) = 1).
Let A G ]^*, and let M(A) be the Verma module for fi with highest weight
pair (A, ¿;+). Then M(A) = is the corresponding weight
space decomposition of M(A). We have M(A) = U(a_) ♦v^A = U(a) •
and the annihilator .^4 of v, in U(fi) is

U (a )a ^ + Vi{a){c - A(c)) + Vi{á)(d - k ( d ) ) .

This is the content of (5) for this example. Thus M(A) = Vi(&)/A.
2.3 Yerma Modules 113

If we restrict the representation to a, then the annihilator ^ of in


U(a) is quite easily seen to be

U (a)a^+U (a)(c-A (c)),

and M(A) = U(a)//4. Since we end up with the same representation space
[i.e., M(A)], whether or not we adjoin d to a, it is worthwhile to comment on
its role. The element c e is central in a (and in a) and consequently acts
as scalar multiplication by A(c) on all of M(A). Indeed for u e U(a),
c -( m - i;+) = w *c *i; += A(c)w • i;+. Thus c is unable to decompose M(A)
into its characteristic form as a one-dimensional “highest” weight space and
a sum of weight spaces of lower weights. The element d does not have this
defect, and as we saw already, leads to the decomposition 0M(A)^“'*'^. It is
primarily because we want these kinds of weight space decompositions that
we have formulated the definition of triangular decomposition as we did. Of
course û = û _ 0 lK c 0 û + is not a triangular decomposition [it violates
axiom TD2].

We return to our discussion of M(A) as an d-module. We propose now to


show the following:

(7) If A(c) # 0, then M(A) is irreducible.

We introduce some terminology: A multi-index is an element n e i^l'^of finite


support. If , «i j are the nonzero values of n. A: > 0, . . . , ^ J,
then we write

for the operator <2"^* • • * on M(A);


a" for the operator on M(A);
|n| for and n! for

In this notation a basis for M(A)^“ '^“ is the set of elements

+, |n| = r.

We have the following inductive calculations in U(a):

(8) [a,.,o-(a^)”] = V, n e Z+,


m
(9) =5,7« m ,n
r=l
114 Lie Algebras Admitting Triangular Decompositions

Thus, if m > « > 0,

flX fly )" • [a^,(r(aj)"]v^

= 8¡jncaf’-^a(ajY~^ • v+ by (9)
= • Ü+ by induction

_ j 8¿jn\c'^ • if m = «,
\0 in m > n .

It follows that for multi-indices m and n with |m| = |n|,

( 10)
0 if m n (then for at least one j, nij > rij)
\ n!c'"‘ • if m = n.

It is now easy to see that M(A) is irreducible. Let r e Z+, and let

X = Ze^a'!. ■ e„ e K .

Then for any multi-index m with |m| = /*,

ÍZ™ •X = •v ^ = m!e^A(c)*^ •v + G: N ( X ) ^ = (0).

Thus each ^„ = 0 [if A(c) # 0], and hence x = 0.


As for M(0), (10) shows that • u += 0 whenever |m| = |n| > 1, and
hence is a submodule of M(0). Obviously this submodule is
maximal, and hence is MO). We conclude that

(M(A) i f A ( c ) ^ 0,
L(A) =
([ if A(c) = 0.

Example 2 (The Virasoro algebra of Example 2.1.4). The diagonal subalge­


bra of the Virasoro algebra is two-dimensional with basis L q and 0. Thus
any linear functional A on is determined by the two values

/i :=A(L q) and c — A(0).

We denote the corresponding Verma module by M = M{h, c) and a highest


weight vector by Since 0 is in the centre of S3, we have 0 * x * i ; + = x * 0*
u^= cx ‘ for all X e U(S3). Thus 0 acts as scalar multiplication by c on
2.3 Yerma Modules 115

M. With Ai e ]^* defined by A^CLo) = 1 and A^Cc) = 0, we have M =

Using the Remark 1 the Poincare-Birkhoff-Witt theorem applied to the


ordered basis L _ i , L _ 2, . . . of SS_ we find that a basis of is the set

V^\n^ <ri2< and = n |.

The problem of determining the maximal submodule N{h, c) of M(/i, c) is


quite complex (see the references in Section 2.8).
Remark 3 There is obviously a lowest weight analogue of Verma modules.
Verma modules for g with the triangular decomposition are lowest
weight Verma modules for g with the triangular decomposition and vice
versa.

Proposition 5
Let g have triangular decomposition (f), g+, o-) and let =
be its opposite triangular decomposition {see Remark 2.1.2).
Let ( tt, M ) be any representation o f g. We define a pair ( tt^, M^) consisting of
a K-space and a mapping g ^ gl(M ^) as follows:

• as a vector space, = M,
• for X ^ V Ei 7t ' ^ ( x ) v = -7 r(a (x ))v .

Then

(i) (tt®", M^) is a representation o f g;


(ii) ((7r^)^(M -)-) = (7r,M);
(iii) = M~^ for all /1 e ]^* Furthermore M is an ^-weight module
if and only if is an ^-weight module’,
(iv) N (zM is a (¿-submodule relative to ir if and only if •= N <z is
a ¿-submodule relative to
(v) M is a highest weight module for (g, ^ ) with highest weight pair
(A, f+) if and only if is a highest weight module for (g, ^ ^ pp)
with highest weight pair ( —A, ¿;+);
(vi) L { \ y is the irreducible module o f highest weight —A for (g, ^ ^ pp).

Proof (i) and (ii) are trivial. For (iii) we have for all e 1^, for all e M

TT'^{h)v = (p,h)v,
<=> —Tr{o’h)v = ( p , h } v ,
<=> 7r(h)v = ( — p,h}v.
116 Lie Algebras Admitting Triangular Decompositions

Again (iv) is trivial, and for (v), (A, y+) is a highest weight pair for Af(A)

<=> U+ G M^, v + ¥= 0,
Tr(Q^)v + = 0,

M = ir(U (g))i;+ ,

« D+e 0
tt'^{Q-) v += 0,
M‘^ = 7T(U(g))£;^,
<=><—A,i; + > a highest weight pair for M.

Now (vi) follows from (iv) and (v). □

2.4 ^l2(K)-THEORY

We recall the Lie algebra ^I2(1K) consisting of 2 x 2 matrices of trace 0 with


entries in K. The representation theory of this Lie algebra is of the utmost
importance for the theory of semisimple Lie algebra.
A basis {E, H, F} of §I2(IK) satisfying the relations

[H ,E ]= 2E , [H ,F ]= -2F , [E,F]= H ,

is called an § 12-triplet.
We recall that ^I2(IK) = K f 0 Kh 0 Ke, where

/=(; 2 ). - ( i -M ' «=(2 D-


As we saw earlier, {e, h , f ) is an § 12-triplet. Evidently if we set

§_;= [K/, := Kh, § += Ke,

and define an anti-involution a on ^l2([K) by transposition, then we have a


regular triangular decomposition

^I2(1K) = ^_0 0 ^+

with Q+= Z+a, where a(h) = 2 (Section 2.1, Example 3).


Let A 0 f)*, and consider the ^I2(1K)-Verma module M(A) of highest
weight A. Fix a highest weight vector of M(A). Then by Propositions 2.2.1
2.4 ёКСЮ-ТЬеогу 117

and 2.3.3.

= K /" • +^ (0), for all n ^ N.

It follows from Proposition 2.1.1 that any submodule R of M(A) is either (0)
or has the form for some Hq > 0. In the latter case

(1) e -/"0 • y + = 0.

Conversely, if Hq is a positive integer for which (1) holds, then we see that
/'^0 • is a highest weight vector and U(g) *u += U(^_)/''o • +c M(A)
is a proper submodule of M(A). We conclude that the (unique) maximal
proper submodule MA) of Af(A) is generated by the highest weight vector
/"0 • ¿;+, where is the least positive integer satisfying (1).

Lemma 1
Let \(h) = a e K. Then for all n ^ N,

(i) /z • /" • u^= (a - 2n)f^ •


(ii) e • /" • z; + = n{a — n + 1 )/" “ ^ *z;+, where we make the convention
that f~^ ‘ v += 0.

Proof (i) follows from the fact that / " • +e M(A)'^“"". We prove (ii) by
induction on n, it being clear if n = 0. For n > 1 we have

e • ü+= [ e , / ] • / " ^ - v ^ + f - e - f n - l • V,
= (a — 2(n — 1) ) / " ^ • y + + (/i — 1)(л — n 2) / " ^
= n(a —n + 1) / ” ^ v +. □
Proposition 2
Let M(A) be the Verma module for §12(Ю with highest weight pair (A,z;+).
Then

(i) M(A) is irreducible if and only if \(h) ^ 141;


(ii) if A(A) e 141, then N ( \ ) is a Verma module with generator
and highest weight A —(A(/z) + l)a. The irreducible module L(A) is of
finite dimension X(h) H- 1 and is the only nontrivial quotient module of
M(A).

Proof Let a = A(/z). After the preparatory remarks and Lemma 1, we see
that M(A) is reducible if and only if

По(л —«0 + 1) = 0 for some n^ e Z^_.


118 Lie Algebras Admitting Triangular Decompositions

This is possible if and only if a e I4j and п^ = a This shows (i) and also
shows that equation (1) has precisely one solution whenever a e N. In that
case M(A) has only one nontrivial submodule, namely ЖА), and MA) is
generated as a highest weight module by ' v Thus MA) has highest
weight A —(a + l)a. Since (A —(я + \)aXh) = —л —2 e Z _ , it follows
from (i) that the Verma module M(A —(a + l)a ) is irreducible, and so MA)
is a Verma module, namely MA) = M(A - (a H- l)a) (Proposition 2.3.3). If
M(A) ^ L(A) is the natural homomorphism [with kernel MA)], then
{/^ • Ü+10 < 7 < й} is a basis for L(A). □

The finite-dimensional irreducible highest weight modules L(A) are pre­


cisely those for which A(A) = a e N. We sometimes denote L(A) by L(a) if
\(h) = a. In the physics literature these modules are often denoted simply
(a) or (a) [and also unfortunately by their dimensions (a + 1)]. If v+
denotes a highest weight vector of L(A), then we have the basis

(2) v^,f ■

and corresponding weights

A,A —a,A —2 a , . . . , A — aa = —A.


The values of the weights at h form the symmetric string

(3) a,a - 2,a - 4 , . , , , - a .

Trivial though it may seem, this symmetry will eventually lead us (in Chapters
4, 5, and 6) to a complete theory of symmetry in root and weight systems of
Kac-Moody algebras. If we set у _=/"* • i;+, then, up to scalar multiples, the
vectors in (2) can be written

e"* ' 'v v_,

which displays L(A) from the point of view of a lowest weight module.
Another way to look at the symmetry is to interchange the roles of e and /
in the triangular decomposition (and then to replace й = [e,/ ] by —h =
[/, e]). Then L(A) is a highest weight module with highest weight vector v_
and highest weight -A ( - A (- A ) = д).
One of the central results of the classical theory of semisimple Lie
algebras is Weyl’s theorem, which states that their finite-dimensional repre­
sentation are completely reducible. Eventually we will prove a version of this
theorem (set in a much wider context). However, we will find it useful to have
the result for §12(Ю much earlier. The proof that we present here is in
essence the same as the more general one that will appear in Chapter 6. We
will make use of the concept of primitive vectors, a subject that is more fully
developed in Section 2.6.
2.4 SljOlO-Theory 119

Let M be an §l2(lK)-module admitting a weight space decomposition


relative to IK/z. A weight vector v e is called a primitive vector if there is
an §l2(lK)-submodule N oi M such that u ^ N and such that e • v ^ N,

For instance, highest weight vectors are primitive [with N taken as (0)].

Proposition 3
Let K / 0 K/z 0 Ke be the triangular decomposition o f §I2(1K) described
above, and let M be a finite-dimensional ^l2(K)-module admitting a weight
space decomposition M = 0M ^ (p 0 1^*) relative to f). Then

M= © L„
¿=1

where each is an irreducible ^i2(K)-module isomorphic to L(A^) for some


= /¿a, I f e N. The decomposition is unique to the extent that the collection
(Aj,... , \jf) of highest weights is unique up to rearrangement.

To prove this result, we introduce the element

(4) C = |/z2 + e/ + /e

of U = U(§l2(IK)). The next lemma says that this element lies in the center
3(U) of U. In general, elements of 3(U) for semisimple g are called Casimir
operators. However, there is always a quadratic Casimir operator like (4)
which is essentially unique and is usually called the Casimir operator. We will
never have occasion to use any other. (In Section 4.5 we will introduce
quadratic Casimir operators in a far wider context.)

Lemma 4

C lies in the center of U.

Proof It suffices to show that [e, C] = [/, C] = 0, since e and / generate


§I2(IK) and hence U. We have

[e,^h^ + ef + fe] = \{ [e ,h ]h + h{e,h]} + e { e , f ] + [ e , f ] e


= —eh — he eh he = 0.
The case of / is similar □

Suppose that F is a highest weight module for §I2(1K) (relative to a given


triangular decomposition), and let be a highest vector of weight A. Let
120 Líe Algebras Admitting Triangular Decompositions

a = A(/z) e IK. Then

C •f f ++ e •/ • +
= + /e ) • f +

= { W + «)í'+-
Also for all MG U,

C • M• u ' C ' v^= [\a?’ + • v^.

Since i;+ generates F,

(5) C V= + a)v for all f g F.

Proof (of Proposition 3). For some algebraic extension IK' of IK all the
eigenvalues of C acting on M lie in IK' [if IK is algebraically closed, then
IK' = IK]. Let M' = IK' M. Then M' is a module for the Lie algebra
IK' ^ ^I2(1K) = §I2(IK'), and C can be interpreted as an element of 301'),
where U' == U(§l2(IK')). From M = we have M' = 0(IK' ^ M ^ \
Each space M'^ •= IK' 0^^ is a weight space for 1^' — IK' % 1^, from
which it follows that we have produced a weight space decomposition of M'
relative to f)'.
Let t be an eigenvalue of C with corresponding generalized eigenspace
M' := {i; G M '|(C - i)^ • = 0 for some A: > 0} (see Section 7.1). Then Af'
is an §l2(lK')-submodule for M' since C g 301')- In particular it is 1^'-
invariant and hence decomposes into weight spaces M[^ '= M[ n M'^. If
VG then e ' V E: (where we are identifying §I2(W inside §I2(IK')
via X 1 0 x). In view of the finite dimensionality of M' over IK', there can
only be finitely many nontrivial weight spaces, and hence there is a weight ¡i
and a weight vector u g \ {0} such that e • ¿; = 0.
Now is a highest weight vector, and in particular a primitive vector of
M'f. Let us show that for any primitive vector w of weight v in Af/, we have

( 6) K/t) and j v { h f + v(h) = t.

Let AT be a submodule of M'„ with w ^ N, e ■w ^ N. Then on w + N ^


M't/N, C acts as multiplication by \v(,hY + v{h), and hence

C • w = (\v{h)^ + v{h)^w mod N.

Since (C —i)* • *v = 0 for some A: > 0, we obtain \ v i h y + vih) = t. Since v


2.4 gl2(IK)-Theory 121

is a highest weight for the finite-dimensional quotient module M¡/N, v{h)


by Proposition 2. This establishes (6).
Now (6) shows that there is only one primitive weight in M/, namely
above, and that a — e N. In particular t e K. Thus all the eigenvalues
of C lie in K, and the extension of K to K' is unnecessary.
Let u \ , . . . , be a basis for M^. If e • 0, then + /:a is a highest
weight for some A: > 0, contradicting our observation above about primitive
weights. Thus for each i, L,(/¿) == U • is an irreducible ^l2(IK)-module
isomorphic to Liix). It follows that n (Ey^,U • = (0) (since
iv{) and that

R = ® L j ( f i ) CM,.
; =i

Consider the quotient module M^/R. It is a finite-dimensional §l2(W-mod-


ule with a weight space decomposition on which C acts as multiplication by t.
The argument above shows that if it is not (0), then it has a highest weight v.
Then V is 2i primitive weight of M ,, and hence it is equal to ¡ by (6). But
jl

{ M J R Y = M>t/R^ = (0). Thus R = M„ and

(V) © Lj(jx).
y=l

From M = (the sum taken over all the eigenvalues of C) and (7), we
have the decomposition of M into irreducible submodules, as required.
For the uniqueness property we consider the composition series

k k
M = © Ly D © Ly D э L ^ э (0)
y=l y= 2

(i.e., each submodule is a maximal submodule of the previous one). By the


Jordan-Holder theorem (which is a very general fact whose scope lies far
beyond Lie theory) the factors

j=i + \

are unique (including repetitions) up to the order in which they appear. [Ja2,
3.3]. □

Remark 1 It is quite impossible to expect any stronger uniqueness in the


decomposition of a module into irreducible submodules. For example, the
module M — L 0 L also decomposes nicely as L + 0 L_, where L = =
122 Lie Algebras Admitting Triangular Decompositions

{(i^, ± ^ L). However, one does have uniqueness of the isotypical com­
ponents (the Af, in the preceding proof) [Ja2, Thm. 3.11].

CoroUary 1
Under the hypotheses of Proposition 3, the Casimir operator C acts as a
semisimple transformation on M with eigenvalues of the form \a^ + a, a
The eigenspace of M is the sum o f all the irreducible submodules
isomorphic to L{a) [the L(a)-isotypical component].

Example 1 Consider ^I2(1K) c gl2(lK). The adjoint representation, adgi^^^^^,


restricted to ^I2(1K) determines a representation of ^Í2([K) on gl2(IK). We do
not need to invoke Proposition 3 to conclude that gÍ2(W is completely
reducible: clearly gl2(lK) = IKl e ^I2(1K).
As a preview of what is to follow below, we explicitly describe the
corresponding representation of 5L2(IK). Let A" g gl2(IK). Then ad =
L x — R x ’>where L x and R x are, respectively, left and right matrix multipli­
cation by X. If X is nilpotent, then exp ad A" = exp(L;^ —R x) = exp L;^(exp
— R x) = exp L ; ^ ( e x p s i n c e L x and R x commute [in End(gÍ2(IK))].
In other words, for all Y e gl2(IK), (exp ad X \ Y ) = exp(AOexp(y)exp(AT)“ ^
As we will see, this shows that the representation of SL2QÍ) on gl2(W is
p i - X ) : Y ^ x Y x - ^ for all x e SL2(K).

The Lie group associated with the Lie algebra ^l2(K) is the special linear
group

SL2ÍK) ad-bc = 1 .
-{l(:

Any of the finite-dimensional §l2(IK)-modules that we have been discussing


above carries a natural compatible 5L2(lK)-module structure. We begin with a
noteworthy fact about the representations of §I2(1K). Let denote the
space of 2 X 1 column vectors with coefficients in K, and let ir be the natural
representation of §I2(IK) on it (i.e., acting as matrix multiplication). Then in
fact K ^ s L(l) in our notation above. Let M= ^ j j , i ; = | ^ j i rm and let
5 = 5([K^) be the symmetric algebra on the space Then S is graded by
Z, 5 = following Section 1.2,

5'^ = © Ku^vL
i +j = n

Each X e §I2(IK) lifts to a derivation on S, say, X (Proposition 1.1.5), and it


is immediately clear that [X,Y] = [X,Y] [both are derivations extending
[AT,y] e §I2(IK)] showing that we obtain a representation on S. Since X(S")
c 5", we obtain a representation of §I2(IK) on each 5", n = 0,1,2,... .
2.4 § 12(1K) -Theory 123

Proposition 5
Let n ^ Then ( tt^, 5 ”) is an irreducible ^l2ÍK)-module isomorphic to L{n).

Proof. The case « = 0 being obvious, we can assume that n > 0. Let {e, h, /}
be the standard basis for §I2(IK). Then for all X e §I2(1K),

= nu^ ^tt{ X ) u .

Since

(s m^i i )
and

we have

7T„(e)w" = 0, 7T^{h)u’^ = nw".

Thus is a highest weight vector of weight n for §I2(1K), and the result
follows from part ii of Proposition 2 and the fact that dim 5" = « + 1.

Now we extend the action of SL2(K) to 5(IK^) in an analogous fashion.
Again, let 77 denote the natural representation of 5L2([K) on Each
X e SL2(K) determines a linear mapping on ® ••• 0 (« factors) by
i;j 0 • • • 0 7r(jc)i;i 0 • • • 0 v(x)u^, and hence on the tensor algebra
T(IK^) by linear extension [where on T([K^)° = K, the action is trivial].
Thus 5L2(IK) acts as a group of automorphisms of r([K^) (as an associative
algebra). Since it stabilizes the ideal J generated by the set of vectors
y 0 w - w 0 y , y,w e we have a group action 77 of SL2(K) on 5(1K^) =
T(K^)/J (see Section 1.2) with tt( x )( v ^ • • • v^) = irix)v^ • • * 77(jc)i;„. Let
^ Then 7T^ is our desired representation of SL2(K). Using the
isomorphism 5" —L(n), we have an action of 5L2(IK) on L(n). Next we
study in what sense t7„ is compatible with t7„ [see (8) below].
Let X e §I2(IK) be a nilpotent matrix; that is, X ^ = 0 for some /: > 0. In
fact = 0, fc > 0 => (K^ d ATIK^ d AT^IK^ = (0) the containments being
strict, so X ^ = 0. Relative to some basis {u^, U2) of IK^, A", as a linear
transformation, has the matrix form ( ^ ^ |, whence exp AT = 1 + X +
( : : ) '
ATV2! + • • • is similar to “j, which shows that exp X ' 5L,
124 Líe Algebras Admitting Triangular Decompositions

Lemma 6
For all nilpotent X e §I2([K) and for all n ^ N, is nilpotent on L(n),
and

( 8) expir„(A') = ■j7„exp(X).

Proof. We may take S" as our model of L(n). For Mj, . . . , m„ e and
ATe SI^OK),
n
^ n ( ^ ) ( “ i • • • « « ) = L «1 • • • ATwy • • • u„,
J=i

and one sees inductively that

7T„(Xf(u, ■■■ U„)= E . , •••


*,+ ••• +*„=* '^1---- '^n-

It follows at once that if X is nilpotent, so is tt„(X), and that

expv „(X )(u i • • • « „ ) = E — — ( “ i ••• u„)


k
_ _ X ’^^ X^-
k A:i+ • • +A:„=A: ^1*
= exp X{u f) • • • exp X(u^)
= 'n-„{expX)(ui ■■■ u„). □

We say that an §l2(K)-moduIe (p, F ) is integrable relative to the basis


{e,h, f ) if V is decomposable as a direct sum of IK/i-weight spaces and pe
and p f are locally nilpotent on V,

Proposition 7
An integrable ^l2(K)-module (p, F ) can be written as a direct sum o f finite-
dimensional irreducible ^l2(K)-submodules, In particular (p, F ) is completely
reducible. For every nilpotent element X e §I2(IK), p{X) is locally nilpotent.
The space V admits a unique SL2(K)-representation p with the compatibility
condition

(8) exp p { X ) = p(exp X ) for all nilpotent X e §I2([K).


2.4 êl2(K)-Theory 125

Proof. If f e K is a [K/z-weight vector, then the êl2(lK)-submodule generated


by V is

¿,;>0

(by the PBW theorem), and this is finite dimensional because of the local
nilpotence of pe and p /. Thus W is completely reducible by Proposition 3
and it follows that F is a sum, hence direct sum, of finite-dimensional
irreducible submodules of êl2(K) (Proposition 1.6.2). Using Lemma 6, every
nilpotent X e êl2(lK) is locally nilpotent in its action on V. Furthermore the
Lie algebra action êl2(IK) on each irreducible submodule W of F “integrates”
to an SL2(K) action, as we described above, and the compatibility condition
(8) holds on W. Piecing the representations together in the direct sum, we
have a representation p of 5L2(IK) on F which is compatible with p. The
uniqueness of p follows from the fact to be shown immediately below that

{(i o)L ““ ((! 2)},.K


form a set of generators of the group SL2(IK). □

We now gather together some useful facts about SL2(IK). Define

a b
B == \b ^ K
0 a-*
t 0
H -=
0
0
AT:=//U t e
- t -1

It is easy to see that each of these is a subgroup of 5L2(IK), that H <N, and
that N /H = Z/2Z.
The simple calculation

(o !)(-!-■ ?)(o ;)■ (-?-' o)

shows that the set

i(A #\/l n\
te K
- { ( i ;)■!(: Î)
126 Líe Algebras Admitting Triangular Decompositions

generates the element n{t) and hence also the matrices

( 10 ) h( t) ■■=n ( t ) n ( l ) ' = 1^ t^K ^.

Given an arbitrary matrix A = j


* e 5L2(IK), we want to reduce it to 1
by left multiplications by elements of U. Applying n(l) if necessary, we can
assume that a 0. After applying

1 0
—ca~^ 1

we can assume that

a b
A =
0 a-^

Then

/i(a-i)y4 = | j and = 1.

We have proved the first part of

Proposition 8

(i) U generates 5L2(IK).


(ii) SL2(K) is generated by B KJ N\ and H = B C\ N,
(iii) 5L2(1K) = ^ (Bruhat decomposition).

Proof. It remains to prove parts (ii) and (iii). From (9) we see that U lies in
the subgroup of 5L2(IK) generated by B and N, whence the first part of (ii).
Obviously B n N H. Set

=(-? i)-

Notice that s is a representative for the nontrivial coset of H in N. After


part (ii) we will prove part (iii) if we show that B U BsB is closed by right
2.4 SijiW -Theory 127

multiplications by elements of N. Since N = H U Hs, this reduces to showing


that sBs (Z B U BsB. In showing this, we may replace each occurrence of 5 by
any element that we wish from the coset Hs = sH of H in N. Using (9), we
see that

1 0
G BsB for all t e
-r-i 1

Thus

for all t e

Let a typical element of B be written as

[l /-)(J !)•
S i G K.

Then

^( o

which is what we wanted. □

Parts (ii) and (iii) of Proposition 8 express a basic ansatz, which we will see
fully developed in Section 6.3. An interesting application of Proposition 8 is
to use it to prove the following.

Proposition 9
The only proper normal subgroups o f SL2(K) are

( ( i ;)} - ( m : :)}■

The same proof {due to J. Tits) may be used to prove that many other groups
are ''essentially"’ simple [773].

Proof Suppose that K # 5L2(K) and that AT is a normal subgroup of


SL2(K). Consider the subgroup BK. Since it is a union of .B-double cosets,
BK = B, or BK = SL2{K). The latter implies that SL2(K)/K = B K / K =
B/B n K, However, is a solvable group; hence also B / B OK , whereas
5L2(IK) is its own derived group, and the same goes for SL2(K)/K. These are
128 Lie Algebras Admitting Triangular Decompositions

contradictory, so we have BK = B\ that is, B Z) K. However.

/ 0 l\la feW 0 -1 \ 0
1 0 /\0 0/ \ —b a

and hence K c . A trivial calculation with shows


( : ; - ) - ( : : )
that fl = +1. Thus K has the required form. □

Corollary
The action of 5L2(1K) on 5" is faithful if n is odd and has kernel {±1) if n is
even.

Proof By Proposition 9 the kernel is either trivial or is {± 1}. Since

the result is clear. □

2.5 CHARACTERS

Suppose that M is a vector space which is graded by an abelian group P and


for which the degree subspaces are finite dimensional. In many situations we
can, or would like to, make useful numerical statements about all of these
subspaces at once. For instance, we might like to know the dimensions
{dim or we might have a homogeneous linear operator T of degree
0, and we might be interested in the traces (tr(r|A/a)}^gp (of which the
dimensions are the special case for T = id^).
As a simple example consider 5(K), F = K x 0 lKy, asa Z-graded module
under the total grading. Then 5(K) == IK[x, y], and the degree subspace
S%V) of elements of degree n admits the monomials

as a basis. Thus dim S%V) = n + 1. This information is neatly stored in the


formal power series

(1) E (« +1)9"-
/1 = 0

This series goes by various names: the Poincare series of (the Z-graded
module) 5(K) or the generating function (for the dimensions of the degree
2.5 Characters 129

subspaces) of 5(K). We often call it the character of 5(K), for reasons to be


explained later.
The use of the word “function” in this context is not to be interpreted to
mean that (1) is some real or complex valued function of q. Of course such
an interpretation might be useful (in which case questions of convergence
appear), but our only intention at the moment is to consider (1) as an
element of the formal power series ring K[[^]]. Part of this section is devoted
to a basic introduction to power series. For the moment it suffices to say that
(1) is an invertible element in the ring K[[q]] and that its inverse is (1 - qY.
The ring structure of K[[q]] is manifestly useful when we go on to consider
the graded algebra K[x, y, z]. Regarding [K[x:, y] and K[z] as subalgebras of
K[x, y, z], the space K[x, y, z T of elements of degree n in K[x, y, z] is

E K [x ,y M zr"^
/ =0

and

(2) dim(IK[x, y, z ] ”) = ^ dim(lK[x, y]')dim(lK[z]"


i= 0

Clearly the generating function of K[z] is E7=o^' = (1 —^) \ and hence by


(1) and (2), the generating function of K[x, y, z] is

-3

n=0 n=0

Induction on dim V quickly shows that the generating function of SiV ) for
dim F = m is

(3) (1

It is not hard to expand this as

(!-«)■” - £ j" '

We now establish a suitable generalization of this for treating jP-graded


spaces. Let P be an additive abelian group. We let

e{P) = [e(a)\a e P}
130 Lie Algebras Admitting Triangular Decompositions

be a multiplicative copy of P; that is, we define

e(a)e(P) '■=e{a + P ) , a,P ^P .

The elements e(a) are called formal exponentials.


Given a commutative ring R, we define P[P] = Re(a) to be the
free module over R with basis e(P). Then R[P] becomes an associative
i?-algebra with 1 — e(0) as the identity element if we define a multiplica­
tion by

(4) H a„e(a) • E = E ( E
a^P /3e/> a +^ =y '

R[P] is called the group algebra of P with coefficients in R.


If M is a finite-dimensional P-graded space, then M has only finitely
many nontrivial degree spaces, and we can define the generating function or
character of M by

ch(M ) = i ; dim M“e (a ) e Z [P ].

Unfortunately, most of our graded spaces are not so simple, and we have to
extend Z[P] to express their characters. To this effect let us assume the
following.

(5) There exists a nonempty family {yy}y of elements of P indexed by a


set J such that the group Q •= Ey^jZyy is freely generated (as an abelian
group, by the {yy}yej*

Under this assumption we let denote the semigroup of Q generated


by the 7y’s. A typical example of this situation is Q+= \ {0} in Q = Z^.
Assuming (5) is in effect, we define sl Q +- fan in P to be a set of the form

U {A - ^ 1^ e (2 +},

where A e P (Section 2.2). We call A the source of the fan. We introduce a


partial ordering < on P by setting / x <A if ¡jl ^ ^ i Q+- Equivalently,
< A if = A - a for some a e (2+.
Let r .y
" -C\‘-
R [ p y ■■= { T , a , e { a ) \ a , ^ R , a ^ P } .

The elements of R[PT are formal infinite sums. For / = Y.a^e{a) e P[P]^,
2.5 Characters 131

we define the support of / by

su p p (/) == {a e Pla„ 0}.

Notice that R[P] = { / ^ R[PT\supp(f) is finite}.


We cannot supply R[PT with a composition law by simply extending (4)
because meaningless infinite sums of elements of R will arise on the
right-hand side. Consider, however,

R [ P i Q ^ ] ■■=
{ /= a ) e jR[P]"^|supp(/)lies in a finite number of ¡2+-fans|,

and

J.0+] == { / = £ a „ e ( a ) |s u p p ( / ) lies in A J. g + j ,

where A is an arbitrary element of P.


Both R[PIQ+] and i?[0i<2+] become associative algebras under (4).
Indeed let / = La^eia), and g = Lb^eip) lie in R[P j Q+] with

supp(/) c (J (a, i(2 + ), supp(g) c U


1=1 j=i

Let 5 ^ P, and consider writing 8 = a p, a e supp(/), p e supp(g).


Then 8 = pj - e ^ (a^ + Pj)i Q^, where s is the sum of two elements
of C+U{0}. By the assumption on <2+ we see that there are only finitely
many ways of writing 8 in such a manner. Thus

a +(B=8

is a well-defined sum and (4) makes sense. Clearly R[0i Q+l is a subalgebra
of R [ P i Q j .
For each A e F we define

^[■Pi!2 + ] ( A ) ~ { E « a ^ ( “ ) ^ R[P i Q +] \ ^a = Ofor all a > a).

Let { /J/ei ^ family of elements of R[P iQ+], and write

fi= £ « « ,« (« )•
a^P
132 Líe Algebras Admitting Triangular Decompositions

We say that the above family is summable if the following two conditions
hold.

(51) There exist , A„ in P such that supp(/,) c (J J=i(Ay i Q_^,) for


all i e I.
(52) For all A in F there exists only a finite number of i in I for which
/; i R[P i Q ^ \,y

Proposition 1
Let R[P I Q be as above.

(i) Two elements x and y o f R[P iQ ^ ] are equal if and only if x =


y mod R[P i !2+](a) ^
(ii) If (//)/ei ^ ^ summable family o f elements o f R[P IQ+] with f =
there exists a unique element f = of
>10 +] that a^ = ^i^ts sum has only finitely many
nonzero terms).

Proof (i) is straightforward.


(ii) Let a ^ P. Then =>f^ ^ R[P>IQ+](«). Axiom S2 then shows
that the sum E^ei^a, ¡s finite. Consider / == e R[PT • By
SI we see that supp(/) lies inside the union of a finite number of fans. Thus
f^R [P iQ ^l □

The element f ^ R[P i Q+] defined in Part (ii) of Proposition 1 is de­


noted by E, ei//-

Now suppose that M is a P-graded vector space with finite-dimensional


degree subspaces such that all nontrivial degrees [degrees a for which
M“ (0)] lie in a finite union of 0+-fans. Then we define the character of
M by

(6) ch(M) = £ dimM“e( a ) e Z [ P i 0 + ] .


a^P

All important case of this is the situation where M is a highest weight


module of highest weight A for a Lie algebra g admitting a regular triangular
decomposition (f), g+, Q+, a). Then M is graded by the group P = ZA + 0,
where 0 is the root lattice, so that (5) holds for this choice of 0 , and all
nontrivial weights of this weight space decomposition lie in the 0+-fan
A i 0+. Also by Proposition 2.2.1 all the weight spaces are finite dimensional
because of the regularity of the triangular decomposition. The character (6) is
2.5 Characters 133

thus defined. Likewise we may define

(7) ch(g_) = Y. dim g “e ( - a ) e Z[Oi g +]-

It is primarily because of this type of example that it is more convenient to


base fans and the definition of IK[0i (2+1 on - Q + rather than Q+.
The use of the word character in this connection has its origin in the
complex representation theory of compact Lie groups. In this situation is a
finite-dimensional abelian Lie algebra over R, and there is associated with
each h ^ if an operator on the complex space M whose action on the
weight space M" is multiplication by [whereas h acts as multiplication
by a(h)]. The operators form a compact abelian group T, If M has finite
dimension then tr^ = E d i m M “e^“^^\ The mapping tr^(e^) is a
character : T ^ C (in the sense of group theory). We have a formal version
of this with e(a) replacing the function e" : ->
We can try to define ch(M) as a function on i) through
cc(h)
h Y dim

This of course raises very interesting questions of convergence and the rate
of growth of the dimensions of the weight spaces, which we do not treat in
this book.
The ring R[0i 2+] is easily identified as a ring of formal power series. Let
the basis of Q+ given in (5). Each a ^ Q+ has a unique
coordinate representation in terms of the y/s:

( 8) ri:y:
JrJi
+ Jk'Jk'
The coordinates n = (itj) are elements of (mappings of J into [^) for
which ftj = 0 for all but a finite subset of J.
Let Xj •= e(-yj). Then for a with coordinates n as in (8)

e { - a ) =x;/i...

which we denote simply as x". Then

where the sums run over all coordinates of elements of 2 +u{0}. As we have
already mentioned, these are formal sums involving (in general) infinitely
many nonzero coefficients.
The most common situation is with J = {1,..., m} in which case each
monomial x" and /?[0J,2+] is the well-known ring of formal
power series in m commuting variables, R[[xi,. . . , x^]]. We will also have
134 Líe Algebras Admitting Triangular Decompositions

occasion to use the case J = In both cases the exponential and monomial
notations are useful, and we work with them interchangeably. Generally,
when we are using the monomial notation, we will denote R[0i Q+]hy i?[[jc]]
or
The first question that we settle is which elements of R[[x]] are units (i.e.,
have multiplicative inverses). For m = (rrij) ^ (2 +*-^{0} we define

k l :=

Recall the partial ordering < defined earlier. For m, n ^ ¡2 +^(0}?

m < n if and only if for all i.

Let R[[;c]]" be the R-span of the elements

X*" for which lm| = n.

In particular = R.\.

Proposition 2
For / = ^ ^[[^]] to be a unit, it is necessary and sufficient that
üfi a(0, . . . , 0) be a unit o f R, I f f is a unit, then its inverse g = Ebm*™ ^
recursively defined through

(9)
L aA-
r + s = m, s < m

Proof Writing fg = l leads directly to (9). If a^ is a unit, these equations


admit a unique solution which is found inductively on |m|. If a^ is not a unit,
then there is no solution to a^bQ = 1, and hence / is not invertible. □

A particularly important instance of this is the well-known

(10) (1 - x ) ~ ^ = l + x + x ^ + .

We have already pointed out the expansion

{i-x) - £ ( „ )* ’ .

We will have a number of occasions to write down product expansions

(11) n (i+ /,).


2.5 Characters 135

where /¿^R[[x]\ and / / = (i.e., = 0). These products


make sense if we assume that for each n e Q + there is a finite subset D(n)
of J such that

= 0 whenever j ^ D(n) and m < n.

Indeed in that case only finitely many /, can contribute to the coefficient of a
given monomial x". More precisely the coefficient of x" in

n (1 +//)

is independent of S provided that S d D(n). This coefficient is then the


coefficient of x" in (11).
Consider the rather natural looking product

( 12 ) n ( i + « 0 ^ »<[[«]]•'
y=l

This has an expansion

L Poin)q",
n= 0

and Po(n) has a simple and beautiful combinatorial interpretation. The


coefficient Po(n) of was already established by the time we expanded

fl(i +^0-
y=l

Since expanding this product amounts to forming simple products by choos­


ing either the 1 or the from each factor, we see that p^in) is the number
of sequences
1 < 7i < 72 < **• < Jr

such that E-=i 7/ = n, that is, the number of partitions of n into distinct
parts.
In the same way if m^, m2, . . . belong to N and we write

(13) n ( i + «0 ^= L r ( n ) g " ,
;=i
then r(n) can be interpreted as being the number of partitions of n into

^The use of q rather than x^, or simply x, is to conform with the standard notation used for
partition functions in the literature and elsewhere in the book.
136 Lie Algebras Admitting Triangular Decompositions

distinct parts where the number j may be considered available in


different colours. For example,

y=l

generates the partitions of n into distinct parts where the odd numbers are of
two colours but then even numbers of just one colour. Suppose that we
distinguish the two types of odd numbers by plain and bold fonts. Then, for
instance, the coefficient of is 6, with the actual partitions being

4,31,31,31,31,211.
The above partitions are restricted; this means that we can use each part only
a limited number of times. The product

(14) 0(1-90 = E p («)9"


; =i

enumerates the unrestricted partition function: p(n) is the number of se­


quences
j\ <J 2 < • • • <Jr,

with = n. To see why, expand each factor of (14) using (10) to get

{l + g + q^ + ■■■)[! + q^- + ( q ^ f + +q^ + { q ^ f +

Now interpret the /cth factor (1 + <?^ + (<?^)^ + • • • ) as referring to parts of


size k\
0 parts of size A: 1
1 part of size k
2
2 parts of size k <->

Then any given monomial

is interpreted as the partition

1 . . . 1 2 . . . 2 . ..A:...A: ofn = ,
Si $2 Sf^
2.5 Characters 137

and the entire expansion has the interpretation as the generating function of
the unrestricted partition function.
To interpret

(15) n ( i -« 0 > m,
; =i

we imagine that each integer j is available in rrij “colours.” This is the


generating function for the number of partitions of n into parts where we
may use each positive integer any number of times and also with any
combination of colours. If rrij = 0, then j cannot be used in any partition
(i.e., j is not available in any colour). For more on partitions, see [HW] and
[An].
We come now to the main results of this section.

Proposition 3
Let n be a Lie algebra graded by an abelian group P for which the following
conditions hold,

(i) All the nontrivial degrees lie in —Q+ where Q+ is a semigroup o f P


satisfying (5);
GO For all a e •= dim n ““ is finite.

Let U(n) be the universal enveloping algebra o f n with the inherited P-grading
{Section 1.8). Then

ch(U(n))= n (1

Proof. Let [yj}jej be the given basis for <2+- Let i;*., A: e K, be an ordered
basis of n consisting of homogeneous elements, say, g n “**, 8 ^ ^ Q+-
Then according to the Poincare-Birkhoff-Witt theorem, the products

(16) ki< ...< k„

form a basis for U(n). On the other hand, the expansion of

-1
(17) r i ( l - e ( - 8, ) )
A:eK

enumerates all the products

(18) e ( - 8 , y “' . . . e { - 8 , y \ k,< ...<k,.


138 Lie Algebras Admitting Triangular Decompositions

The coefficient of e ( - 8 ) in (17) is thus the number of products (18), with

which in turn is the number of products (16) that lie in U(n)“^. Since every a
appears times among the 5y’s, the proposition follows. □

Example 1 (Abelian Lie algebras) As a simple example we take a finite­


dimensional vector space a with a basis and consider it as an
abelian Lie algebra. By Proposition 1.8.1 U(a) is naturally isomorphic to the
symmetric algebra 5(a). If we set Q = Z, Q+= Z+, a~^ = a (all elements
have degree —1), and q = e( —1), then Proposition 3 gives

-N
ch (5 (a))= (l-« )

which is precisely what we had deduced before in (3). If instead we set


Q = Z", ¡2+= \ {0}, = (0,..., 1,.. ., 0) (1 in the ith place), and
= KXi, then we have a completely different grading, and

ch(5(a)) =¿=n(l-K-a,))“‘.
1

An immediate consequence of Proposition 3 and Remark 1 of Section 2.2


is the character formula for Verma modules:

Proposition 4
Let q be a Lie algebra admitting a regular triangular decomposition. Write
9= ^ ® 9+, and let Q denote the root lattice. Let

ch(g_) = L " !« « (-« ) ^ Z[04, e + ] .


a^Q^.

Then

(i) ch(U(g.)) = - K -a))-"*« e Z[01


(ii) for each A e 1^* the character o f the corresponding Verma module
M(A) is given by

ch(M(A)) = e(A) n ( l - e ( - a ) ) " ' ”“ e Z [ P i < 2 j

where P = Z \ Q,
2.5 Characters 139

Just as with (15) we can give the product in Part (i) of Proposition 4 a
combinatorial interpretation:

ch(U(g_))= E K{ p ) e { - I 3 ) ,
j8eQ+u{0}

where K(p) is the number of partitions of ¡3 into parts a ^ Q+ and

1. each a ^ Q+ occurs in different colours,


2. each a ^ Q+ of each colour may be used any number of times.

This gives rise to a function

K:Q

(defined to be 0 outside Q ^) which is called the Kostant partition function


[Kol].

Example 2 (§I2(K)) We have §I2(IK) = IK/ 0 Kh © Ke graded by Q =


l a c ]^* with / of degree —a. For A © 1^*, M(A) = 'L n > o ^ f * where
u+ is 3. highest weight vector, and by Proposition 4,

ch(M(A)) = e(A)(l - e ( —a))~^


= ^ ( A ) + ^ ( A — Of) + ^ ( A — 2o f ) + * * * .

Example 3 (The Virasoro algebra) We use the triangular decomposition of


S3 given in Example 2.2.4 with SS_= ^ ®
where Ai(L q) = 1 and A/c) = 0. Let q = e(-Ai). Then 1KL„ = and

ch(SS_) =
n= \

Let e ]^*. Then

(19) ch(M(/x)) = e(/i,) n (1 - /2 =1


\

and we see from (14) that

dimAf(/i,)'^ ' ' ^ = p { n ) ,

where p{n) is the partition function. This agrees with the explicit basis
constructed in Example 2.3.2.
140 Líe Algebras Admitting Triangular Decompositions

The final example is an application to the theory of free Lie algebras.


Suppose that card X = m < oo. According to the discussion in Section 1.10,
there are two natural gradings of S (^ ), one by Z in which every element of
X has degree 1 and one by ß = 2 '”- The next result gives formulas for the
dimensions of the degree subspaces with respect to each of these two
gradings. We do not need these formulas in the sequel, and they can safely
be omitted.

Proposition 5 [Witt]
Let X be a set o fm elements^ and let S(AT) be the free Lie algebra on X over K.
Let pi be the Mobius inversion function. Then

(i) for all n

^ r\n

(ii) for all m-tuples, n = (n^,. . . , n^) of nonnegative integers

where r|n means that there exists a positive integer k such that nj = krj for all
j = 1, . . . , m, and in that case n / r — k.

Proof Both parts are proved in essentially the same way using a result from
the exercises. We give the proof of the second one. We refer readers
unfamiliar with the Mobius function to [Jal] and [Bo2, Appendix].
For each m-tuple n let = dim^^ By Proposition 3 the character
of the universal enveloping algebra of ^ ( X ) is

n ( l - x " ) ~ ‘' " e K [ [ x ] ] ,


n

where x = {xj,. . . , x^}. (Do not confuse the commuting variables Xj,. . . , x„
with the elements of X ) On the other hand, we know that the free
associative algebra A ( X ) is the universal enveloping algebra of S(X ), and
its generating function is easily seen to be

(1 - (x i + ••• +x„,)) \
Thus

( 20) n ( l - x " ) " " = l - ( X i + --+xJ.


2.6 The Category 0 141

By Exercise 2.16, the dimensions d„ are given in terms of the set 5(r) of
partitions of r into other m-tuples and certain multinomials C(p) for each
partition p. Since the only coefficients on the right-hand side of (20) are
limited to terms of degree 0 or 1, the multinomials C(p) are 0 unless p is the
partition

r = r i ( l , 0 , . . . , 0 ) -I-r2(0, l , . . . , 0) 4- • •• -t-r„(0,0,...,l)

in which case it is

(/•i + ••• 1
\ 1• ***' m• ''i +
Formula (ii) follows immediately. □

Example 4 ( ^ ( ^ ) , card(Ji) = 2)

n dim
1 /x(l)2^ = 2
2 (1/2){ m(1)2" + m(2)2'} = 1/2(4 - 2) = 1
3 (1/3){ m(1)2^ + m(3)2'} = 1/3(8 - 2) = 2
4 (l/4){/x(l)2^ + m(2)22 + jtx(4)2'} = (1/4)(16 - 4) = 3
5 (l/5){/x(l)2^ + A^(5)2'} = (1/5X32 - 2) = 6

dim = (1 /6){m(1)6!/3!3!+ m(3)2!/1!1!} = 3

Note There is a procedure due to P. Hall [see Bo2] that produces an


explicit basis for each degree subspace of a free Lie algebra.

2.6 THE CATEGORY &

Throughout this section g will denote a Lie algebra admitting a triangular


decomposition (1^, 0+? which we fix once and for all. Our aim is
to introduce a category <^(g, [BGGl] whose objects might loosely be
described as consisting of those g-modules whose characters lie in Z[f)* iQ+].
In fact the last part of this section is devoted to putting a ring structure on
the isomorphism classes of modules in ^ to form the representation ring of
which is then related to Z[f)* J, ¡2+] via the character map.
Unfortunately, g itself, under the adjoint representation, does not appear
in & in general. With this exception ^ and its a opposite category
contain all the representations of g that we will need in this book, including
in particular highest weight representations. The first half of this section
introduces the (local) concept of composition series of modules in the
142 Lie Algebras Admitting Triangular Decompositions

notion of the multiplicity of an irreducible submodule of a module M in


category and more on primitive vectors which were briefly used in Sec­
tion 2.4.
The category is deflned as follows:

(i) Objects of 3^). Those g-modules M having the following prop­


erties
(ia) M admits a weight space decomposition

M= ©

with respect to the action of f);


(ib) the set P(M) of weights of M, that is,

is included inside a finite number of fans of ij* iQ+;


(ic) for all e P(M), dim^^iM^) < oo.
(ii) Morphisms of S^). All g-module homomorphisms between the
modules in Obj(<^).

We often write M e «^) to mean that M is an object of The


category of g-modules is the category <^(g,
We are already familiar with highest weight modules. Provided that its
weight spaces are finite dimensional, such a module lies in For instance, if
the triangular decomposition is regular, then for all A e 1^* the Verma
module M(A) and the irreducible module L(A) are in
Remark 1 (i) It is useful to note that if 0 + is countable, then every
module in & is of countable dimension.

(ii) We recall (from Section 2.5) that a partial ordering > was defined in
by declaring that for all A, /u, e 1^*,

A > / t < = > ^ t e A i ¡2 +*

We also introduce for each A e 1^* the depth function

defined for all /8 e ¡2 _^u{0} by

¿ , ( A - / 3 ) := ht(/3).
2.6 The Category 0 143

Lemma 1
Let F = U i 0+) union o f a finite number o f fans in 1^*. Let S be
a nonempty subset o f F, Then

(i) if p ^ S, the set {A e 5|A > p) is finite;


(ii) 5 admits maximal elements.

Proof (i) Let p ^ S and ; e {1, . . . , n}. By definition of <2+-fans, it is


immediate that if p ^ Ay i Q+, then there is no element ^ e Ay i with
^ > P, and if /1 e Ay I Q+, then there is only a finite number of elements
i e A; I (2 +, with ^ > p .
It is clear therefore that the set in question is finite. Part (ii) follows from
Part (i). □

Let M e ^ (g , and let M = be its weight space decompo­


sition. We recall the formal character of M

ch(M) = E dim M^ e( p) ^ Z [ V iQ+]^


/Ltei)*

If M = M', then the isomorphism must map weight spaces isomorphically


onto weight spaces so ch(M) = ch(M').

Proposition 2
Let M and N e ^ (g , ^ ) , and let P{M) and P{N) be their corresponding sets
of weights.

(i) I f is an exact sequence o f ^-modules, then


both A and B belong to Moreover, for a// e

0^ -> ^ 0

is an exact sequence o f K-spaces and

ch(M) = c h ( ^ ) + ch(B).

(ii) Every submodule and every homomorphic image o f M belongs to


^ (g , ^ ) .
(iii) The direct sum M 0 AT belongs to ^ (g , Moreover P{M ^ N ) =
P(M) U P( N) and ch(M ® N) = ch(M) + ch(A^).
(iv) The tensor product M ^ N belongs to ^ (g , Moreover P{M 0 N )
c P{M) + P{N) and ch(M ® N ) = ch(M)ch(iV).
144 Lie Algebras Admitting Triangular Decompositions

Proof. Let M = Identify A with a submodule of M to conclude


(Proposition 2.1.1) that A admits a weight space decomposition

A = e A'",

where = M'^ n A. It is clear then that A e 3^). Also

B = M/A= © M V 0 A -'s 0 M>^/A>^,


usi,* Ae6*

and hence

B= ® B>^,

where B'^ = M'^/A'^ for all fi e i)*. This shows that B , The rest
of (i) follows easily.
(ii) The proof is obvious from Part (i).
(iii) For every ju. e 1^*,

(M ® AT)'‘ = M'" ® JV**,


and thus

M®N= 0

Evidently P( M ® N ) = P(M) u P(N), and the weight spaces are finite


dimensional. Thus M ® AT e (^, and Part (iii) follows.
(iv) Let {y,}, g j and ek basis of M and N, respectively, consisting
of weight vectors, say, e M^ ‘ and e M*'* for all i e I and it e K. The
set {d, ® Wfcl/ei, *eK is a basis of the space M ® AT and these are weight
vectors: Indeed for h

/i • ( y,- ® H'fc) = h ■Vi ® + Vi (8) h •


= iJLi(h)Vi ® + Vi ® A * ( h ) w ^

= {Pi + ^k)(h)Vi ® w ^ . .

This shows that M ® N admits a basis consisting of weight vectors and hence
a weight space decomposition. Moreover, if y e 1^*, then (M ® AT)'*' has as a
basis the set {y, ® consisting of those i and k such that /r, + A;t = y. In
other words.

(1) ( M <8>N y = © (M>^(8i N^).


fl+\=y
2.6 The Categoiy 0 145

Thus F(M <S) c P(M ) + P(A/^), and this is contained inside the union of
finitely many fans. Finally, from (1) we get ch(M (S>N) = ch(M)ch(N).

Let M be a g-module. An element v ^ M is called a primitive vector if

1 y is a weight vector relative to 1^;


2. there exists a submodule N of M such that u ^ N and - i; c N.

If this situation arises and A the weight of v then we also say that A above is
a primitive weight of M {relative to N). Every highest weight vector of a
g-module is a primitive vector.

Lemma 3
L e t M ^ & {% ,Sr\

(i) For a weight vector v o f M o f weight /x, the following are equivalent:
(ia) V is primitive,
Gb) V ^ U(g) • V,
(ic) there exists a submodule N o f M such that N c U(g) *v and
U(g) • v/N = L{p),
(ii) For a primitive weight vector o f weight ¡jl to exist in M, it is necessary
and sufficient that L{ jjl) be a subquotient o f M [i.e., that there exist
submodules X and Y o f M such that X :d Y and X / Y = Lip)].

Proof (ia) => (ib) Let v g be primitive. Then there exists a submodule
M' of M such that v ^ M' but g^* i; c Af'. We have

v ^ M ' = Vi{%) - M ' D U ( g ) -g^-i;

proving Part (ib).


(ib) => (ic) Suppose that v ^ U(g) • g+* == M'. Then M' is a submodule
of M and g+* v c M'. This shows that v is primitive. Let M *= U( q) *v/M' ,
M is a highest weight module generated by the coset v + M'. According to
PartjGO of Proposition 2.3.4, M Ims a unique maximal submodule N and
M/ N = L{p). The preimage N of N in U(g) *v gives us U(g) • v / N = L{p),
proving Part (ic).
(ic) => (ia) The submodule N of Part (ic) serves to show that v is
primitive.
(ii) After Part (i) it remains only to show that if M has a subquotient
X / Y = L{p), then M has a primitive vector of weight p. But any preimage
¿;+ in of a generator D+e L { p Y of L{p) under the map X X/Y
Lip) is primitive: T, g+* + c y. LI
Líe Algebras Admitting Triangular Decompositions

Proposition 4
Let M and N be objects of ^ ( 9, ^ ) , and let f : M -^ N be an epimorphism.
Suppose that v ^ M is a highest weight vector o f weight A and that w ^ N is a
primitive vector of weight p. Then

(i) there exists a primitive vector x ^ M o f weight p such that f ( x ) = w;


00 if f(v) ^ 0, then f (v) is a highest weight vector o f N o f weight A.

Proof 0) Let N ' be a submodule of N with w ^ N ' but w c N', Since /


maps onto N^, we can find x ^ with f ( x ) = w. Set M' = f ~ \ N ' l
Then X ^ M' and g+- jc c A/'.
(ii) The proof is straightforward. □

Remark 2 It is generally not true that the homomorphic image of a


primitive vector is a primitive vector.

Proposition 5
Let M e ^ ( 9, be a nonzero module, and let P(M) be its set o f weights.
Then

(i) every maximal element o f P(M) (with respect to > ) is a highest


weight', in particular highest weight vectors, and hence primitive vec­
tors, for M exist.
(ii) The set of primitive vectors o f M generate M as a q_-module (and a
fortiori as a q-module).

Proof (i) Since M ¥= (0), we have P(M) # 0 , and hence P(M) admits
maximal elements by Lemma 1. Let p be such an element, and let v e M^,
V ¥= 0. Then 9+* V c ^ which shows that i; is a highest
weight vector.
(ii) To see the primitive vectors generate M as a 9 .-module, we consider
the space
v=
V
L
primitive
U(g_)i;cM .

Clearly V is an 1^-module and hence admits a weight space decomposition


(being included in M). Suppose M ^ V , and choose p e P(M) maximal
such that (tV .lix^ \ V, then x is not primitive, and hence
X e U(9) • 9+- ;c by Lemma 3.
Now U(9+) • 9^- X c c K by the maximality of p. But then
x e U ( 9 _ ) - U ( ^ ) *1 1 (9 ^ - 9 ^;c c U ( 9 . ) - U( f t ) - K c F

contrary to the choice of x. This shows that V = M. □


2.6 The Category 0 147

Our next two results characterize irreducible and completely reducible


objects of

Proposition 6
Let M ^ ^ ) , (M ^ (0)). Then the following conditions are equivalent

(i) M is irreducible,
(ii) M = L(A) for some A e 1^*.

Proof Assume that M is irreducible. Let v e be a highest weight vector


(Proposition 5). Then M = U(g)i;, so M is a highest weight module and
hence a homomorphic image of M(A). Since M is irreducible, the kernel of
this map is a maximal submodule of M(A) and it equals MA) (Proposition
2.3.4). Thus M = L(A) as desired. □

Proposition 7
Let M Ei ^ (g , ^ ) , (M =5^ (0)). The following conditions are equivalent,

(i) M is completely reducible [in ^ (g ,


(ii) M is generated by its highest weight vectors, and every highest weight
submodule of M is irreducible.

Proof, (i) =►(ii) Suppose that M is the direct sum of a family of irreducible
modules of <^(g, *^). By our last result

M = © L( A ,)
¿el

for some family {Aj^gj c Evidently M is generated by its highest weight


vectors.
Let be a highest weight vector of weight A, and write

t; = E i;,

where v¡ e L(A,) for all i e I. Then g^.- ü = 0 =» g^.- v¡ = 0 for all i => v¡
L(.X¡y‘ for each i. Now for all /t e we have

E a(/ i ) ü, = \ { h ) v = h • V =
/e l /e l
148 Lie Algebras Admitting Triangular Decompositions

which shows that

f = E
A, = A

It is easy to see then that the submodule U(g)i; generated by v is isomorphic


to L(A) and hence is irreducible.
(ii) => (i) By assumption M is a sum of irreducible submodules. Then M is
the direct sum of irreducible modules by Proposition 1.4.2. □

Corollary 1
Let M ^ be a module with the following property: I f A and p are
primitive weights o f M, then \ > p ^ k = p. Then M is completely reducible.

Proof Let be a highest weight vector of M, say, of weight A. If U(g)i; is


not irreducible, then any primitive weight /a of a proper submodule of U(g)i;
(such p exists by Proposition 5) will satisfy k > p, which is not possible. Thus
U(g)¿; is irreducible. Let

M' :=

with the sum being taken over all highest weight vectors of M. If M / M ' (0),
we can consider a primitive vectors of M, say, x e M^, such that x ^ M'
(Proposition 4 applied to the canonical map M M/M' ),
Now g+- a : # (0) (for otherwise x e M'), and hence the space U(g+)g+* x
contains some highest weight vector, say, of weight A'. Then X > p, contrary
to our hypothesis. □

We now introduce the concept of local, highest weight, and Verma


composition series. It can be shown that not every module in ^ (g ,
admits a composition series (see Exercise 2.13). We have, however, the
following local version of composition series.

Proposition 8
Let M e ^ (g , and A e if*. Then there exists a finite sequence M q, , ,.,Mf^
of modules in ^ (g , S^) and a subset I c {1,..., A:} with the following proper-
ties,

(LCSl) M = M¡^z> D • • • D Mo = (0).


(LCS2) I f i G I, then there exists A¿ e 1^* such that A¿ > A and

M,/ M, _i= L( A, )

(LCS3) I f i ^ I, then (M ,/M ,_i)^ = (0) for all p > k.


2.6 The Category 0 149

In words, all weight spaces o f M for weights equal or above A are involved in
irreducible quotients o f the series.

Proof Define á(M, A) = dim(M^). (This sum is finite by Lemma 1).


We reason by induction on d{M, A). If d{M, A) = 0, we simply set = M,
Mq = (0), and 1 = 0 . Assume that d(M, A) > 0 and that the result holds for
all modules N e <^(g, 5") for which d(N, A) < d(M, A). Let fi e P(M) be a
maximal element satisfying p > \ ( jjl exists by Lemma 1), and let v e
y ¥= 0. Then i; is a highest weight vector (part (i) of Proposition 5). The
highest weight module V = U(g)i; contains a (unique) maximal proper sub-
module U (Proposition 2.3.4) and V / U = Lip).
We observe that

d{M/V,X) <d(M,X), since dim(F^) = 1,

and that

d(U, A) < d{M, A), since dim([/^) = 0.

By induction hypothesis both M / V and U have sequences of submodules


with the desired properties. We now consider the inverse image in M of the
sequence of M / V under the canonical map M M / V , say, M = M^z> ♦
D Mq = F, and the sequence of U, say, U = U^z:> • • • ^ Uq = (0). Since
Mo/Up = Lifi) (see above), it then follows that the sequence

M = M _D ••• z^ M o ^ U ^ z> ••• D i / o = ( 0 )

is as desired.

A sequence of the type appearing in Proposition 8 is called a local


composition series of M at A. The factors for some /x > A
are called proper. The remaining ones are called extraneous.
Remark 3 We make the simple observation that if we insert a submodule
N between Mj and Mj_i, in the local composition series

M = Mf^^ • •• d Mo ={0}

relative to A, then the resulting series

M = M f,:D ••• M y D ••• d M o = {0 }

is also a local composition series at A and has the same proper factors,
including multiplicities, as the original series. Indeed there are only two
150 Líe Algebras Admitting Triangular Decompositions

cases:

1. is proper, in which case N = Mj or N = Mj_^, and one of


N/Mj_^ and Mj / N is proper and one extraneous [actually (0)].
2. is extraneous, in which case both Mj / N and N/Mj_-^ are
extraneous.

Two local composition series for M, say, M = d • • • =) Mq = (0) and


M = Ni ^ • • • z>N q = (0) are equivalent if k = I, and there is a permutation
i ^ i' of {1, . . . , k) such that for all i,

Proposition 9
Let M e (^(g, and fei A, /x e 1^*. Any two local composition seríes {M¿}
and {M/} o f M at A and /x, respectively, have common refinements that are
equivalent to a series {M") that is local at both A and jjl. Furthermore for
^ > A {resp. ^ > /x) the number proper factors of type L ( 0 in {M¿} and [M¡'}
(resp. {M/} and {M"}) are the same.

Proof Using the Schreier refinement theorem [Ja2] find a refinement {M-}
equivalent to refinements of both {M¿} and {M/} (as modules). According to
the last remark, this series will be local at both A and p and the number of
proper factors will not change. □

Let M e ^ (g , 5^) and p e 1^*. Let A g f)* be arbitrary with p > X, and
let {M¡\ be any local composition series of M at A. We claim that the number
of proper factors of type L(/x) in this series is independent of the choice of A
and the particular series {Mj that we have chosen. Indeed if {M/} is a local
composition series at some A' with p > A', then {MJ and {M/} have a
common refinement (M/'} as in Proposition 9. Then the number of proper
factors of type L(p) is the same in (MJ and (M"} and also in (M/} and (M/'},
hence in (M j and (M/}. □

Thus we can make the following definition.

For M e (^(g, and for /x e we define the multiplicity of L(p) in


M, denoted by [M : L(p)] to be the number of proper factors of type L(/x) in
any local composition series of M at /x.

We will see another interpretation of these numbers in Proposition 12.

Let M e ^ (g, A highest weight series (HWS) for M is an increasing


filtration

(0) = Mo c Ml c M2
2.6 The Category 0 151

of submodules of M so that

(HWSl). U M =M;
i=0

(HWS2). If M,_i M¡, then is a highest weight module, say,


with highest weight /i.,.

Let {A/,} be a HWS for M. If 9 e 1^* is any weight, then dim M ’’ < 00, and
the sequence {dim is increasing to dim M ’’. Thus

(HWS3). For each e f)* there is a A: e Z+ such that for all i > k,
M f = M ‘<‘.
If the highest weight series satisfies

••• •••>

we say that it is a finite highest weight series of length equal to n.

Proposition 10 [G L, DGK]
Suppose that the root lattice Q is finitely generated (.equivalently the index set J
of the triangular decomposition S ' is finite). Let M e Then M has a HWS,
(M,}7=o- Furthermore the series can be constructed so that

(HWS4). > fij => i < j, where is the highest weight of

Proof Since M e there exist e such that

P { M) c Ai i • • • U A^ i e +-

If A^ i (2+n Ay i (2 + =5^ 0 then there exists A with A i A¿ i <2+^


Ay i ¡2+- Thus we can assume that the union of the fans is disjoint. Then for
any e P ( M \ ¡JL = - p for some unique / e {1,..., A:} and /3 e (2 +’^{0}-
Let d()L¿) := = ht(p) (Remark 2.6.1(ii)). Clearly from the assumption
on QAP' ^ P(M)\d(/jL) = N} is finite for each N ^ N.
Let fi^ G P(M) be chosen with d(fji^) minimal. Then is a maximal
weight of M relative to > . Thus, if ¿;i e \ {0}, is a highest weight
vector, and generates a highest weight module M^. Now repeat the
argument with M/M^ to get a submodule M2 of M, M2 ^ Mj for which
M2/ M 1 is a highest weight module. In this way we obtain a filtration
(0) := M q (zMi d M2 c • • • . Since the value of d(fi¿) is increasing and can
only remain at one value for finitely many i, we see that for any weight
152 Lie Algebras Admitting Triangular Decompositions

<p e P(M) eventually = • • • . Thus UM, = M. The con­


struction gives (HWS4) automatically. □

For each M e O bji^iq, y ) ) we have its associated character ch(M) e


i Q A s we have pointed out before, M = M' => ch M = ch M'. Let
S = S( q, be the set whose elements are the isomorphism classes of
g-modules in We denote the equivalence class of M by M. Evidently ch
determines a mapping, also denoted ch

ch:

satisfying

ch(M),

where M ^ M.
We let R = R(q, denote the free abelian group on the set 5. A typical
element of R has the form

(finite sum). R can be made into a commutative ring by defining multiplica­


tion by bilinear extension of

M-N=

This is evidently well defined. The character map extends to a homomor­


phism ch : R ^ i Q+], We denote its kernel by K, and set R =
R(ij, ^ )8 } :R /K , We call R the representation ring of g relative to
Factoring ch through K, we have the character map

ch:R(q,^) iG+].
For M e O bj(^) we denote its image in R( q, y ) by 0^ A B ^
C 0 is an exact sequence of modules then ch(B —^ —C) = 0 by Proposi­
tion 2 and hence [B] = [A] + [C] in R. In particular for modules M and N
in
[M © AT] = [M] + [TV].

We also have

[A/® TV] = [TV/][TV]

by the definition of the multiplication in R.


2.6 The Category t 153

Remark 4 It is useful to note that any element Z oi R can be written in


the form [Z+] - [Z_], where Z+ and Z _ e For if Z = then
we simply set

Z^_== 0 (Af 0 • • * © Af) (^[Af] summands),

Z_:= 0 (M ® ®M) (-Crjv^j summands).

Proposition 11
Let {c;^}^ e 6* t>e a family o f elements o f Z.

(0 The family ch ¿(A));^eE)* o f Z[if* | <2+1 is summable if and only if


there exists a finite union F o f fans in Z[k*iQ^] such that [X e
0) c F .
(ii) For every element f if Z['i)* IQ+] there exists a unique family o f
integers {C;^}Aefi* such that the family {c^ o f zli)* J, 0 +]
is summable, and f = ch(L(A)).

Proof (i) For each ju. e h* let

= c^ch L { f i) = c J Y . - a)].

If the family (/^)^ ^ f,* is summable, then by definition (see (SI) of Section
2.5)
{supp/^lju e h*} c F ,

where F is a finite union of fans. Now since = 1, we see that ¡x e


supp 0, and hence

{p, e f)*|c^ 0} c F .

Conversely, assume that {/x e # 0} c F. Then clearly condition (SI)


holds, since
{supp 4 I m e c F.

Fix A e ]^*. Then for /x e 1^* we have

ffjL ^ Z[]^* i 2 + ] ( a) ^ ^ for some ¡jl — a > \


0 and jJL > \
=> jjL ^ F and fJL > \
154 Líe Algebras Admitting Triangular Decompositions

and this set is finite by Lemma 1. This establishes (S2) of Section 2.5 and
hence the summability of the family in question by Proposition 2.5.1.
(ii) Suppose that / = La^eifi) e Z[l^* i Q+], We wish to write / in the
form Ec^ ch(L(^t)) as in Part (i). We may assume, as in Proposition 10, that
all the weights in supp(/) lie inside a finite disjoint union of fans. Then there
is no loss in assuming that all the weights appearing in supp(/) lie in one fan
A i G + . For each ¡x ^ A i Q + let difx) == be the depth of fi (see
Remark 1). We compute the coefficients by induction on the depth.
Suppose that / has been written in the form

/= L c^ch(L(M)) + / .
dilxXd

for some d > 0, where contain nontrivial terms only for weights v with
d{v) > d (we begin at / = / q). Let denote the set of weights of depth
d in A i (2+, and for each i e I let denote the coefficient of c()ll^) in /¿.
By part (i) the set {c^. ch L(/x^)}^^/ is summable.
Setting /¿+1 ■■=fa - E,^,c^.ch(L(/u,,)), we have

/ = E d if j¿ )< d- ¥ 1 c c h L ( ^ ) +/ d+i.

where contains nontrivial terms only for weights v with d{v) > d + 1.
The entire family {c^ ch L(^t)} is easily seen to be summable, proving the
existence of part (ii).
Finally, if the uniqueness were to fail, then E^ g f ^a L(A) = 0 for some
union F of finitely many fans. Let 5 — {A e Flc;^ = 0}. If S 0, then it has
a maximal element Aq and

CA„chL(Ao)-----E C;^chL(A)
A # Aq

gives a contradiction. Indeed the coefficient of e(Ao) is on the left and 0


on the right. □

Remark 5 In Proposition 11 we have used the characters of the irreducible


modules in as a “Z-basis” for Z[l^* i 0+]. However, the only facts that we
have used about ch(L(A)) are

1. ch L(A) involves e(/¿) only if /a e A J,


2. the coefficient of e(A) is 1.

Thus Proposition 11 holds, for instance, if we replace each ch(L(A)) by the


character ch(M(A)) of the corresponding Verma module.
2.6 The Category & 155

Proposition 12
Let M e 5^). Then

chM= 12 [ M : L(iJi)]ch L(ix),


/xei)*

and this is the unique way o f writing ch M as an integral combination o f the


characters o f irreducible modules from

Proof The uniqueness follows from part (ii) of Proposition 11. We prove that
for each A e 1^*

chM = L [ M :L ( /i ) ] ch L ( /x ) mod Z [ ^ i !2 +](a)


fji>\

(see Proposition 2.5.1). For if {M¡)¡^o „ is a local composition series of M at


A, then for each /a > A, [M : L(/i)] is the number of proper factors of type
L(/i) in {M,}. We have

c h M = ¿ ch ^ [ M :L ( M ) ] c h L ( M ) m o d K [ n e + ] ( A ) - □
¿= 0 /11 ^ A

Let F e A Verma composition series (VCS) (of length r) for M is a


descending sequence of modules of V

F = M o = ) M i3 ••• DM, = (0)

so that M¿_i/M^ is a Verma module, / = 1,2,..., r.


Such an object need not exist for a given module V. However, if it does,
then

c h F = E chM¿_i/M¿,

and hence by Remark 5 the Verma factors M^_i/M^ are, up to permutation,


uniquely determined by V and are independent of the particular VCS
chosen. The number of times that M(/x) occurs in a VCS of M is denoted by
[M:M(p)l

Proposition 13
Let p be fixed. There exists a unique group homomorphism
156 Lie Algebras Admitting Triangular Decompositions

satisfying

Proof Let : R(g, -» Z be the unique group homomorphism satisfying

We show that this maps factors through the defining kernel K<zR oi R. It
X ^ K , then
N
x = E c ,M „
1= 1

where
N
E c,ch(M ,.) = 0.
¿=1
Thus in Z[]^* i 2+] we have by Proposition 12,
N

0= E [ M, : L ( n ) ] c h L ( f i )
¿=1

= E ÍE^,[M,:L(M)]]chL(M),
^ e f ) * \/ = l /

and hence for all e 1^* we obtain


N
£ c, [ M , : L ( m)] = 0
¿=1
by Proposition 11. Thus
N

¿= 1

= L c , [ M , : L ( m )] = 0 .
/=1
We have therefore an induced group homomorphism

as desired.
2.6 The Category ^ 157

Corollary 1

(i) If 0 ^ ^ M ^ M /N -» 0 is an exact sequence o f modules from


I, then for each ¡jl e 1^*,

[ M : L ( m )] = [AT:L( m )] + [ M / N : L { t i ) \ .

(ii) For M j, . . . , e i^(g,

[Ml Ф 0 M , : L ( m)] = Е [ М : Ц м ) ] .
k =\

Proposition 14
ch: i?(g, -» Z[if* 1 0+] ^ an isomorphism o f rings.

Proof ch is injective by the definition of i?. Let / e I[i>* j, g+l. By Proposi­


tion 11 we can write / = Ec^ ch(L(A)). Let

M+:= 0 c^L(A),
«A>0
M _ ~ © — c ^ L ( X ) .
c*<0

Both of these lie in and

ch([M j-[M _ ])= /. □

Let M e ^ (g , ^ ) . Then M has a weight space decomposition

M= © M^.

Recall that we have defined the restricted dual space M^es of M by

Mres == { / ^ M*K/, M^> = (0) for all but a finite number of e 1^*}.

Proposition 15
Let := M^es a K-space. The map

g X M^ -> M^
158 Lie Algebras Admitting Triangular Decompositions

denoted by
{x,f) ^ x - f ,
and given by

{x ' f , m ) = </, cr(x) • m> for all jc e g, / g m ^ M,

gives a Q-module structure.

Proof. The map in question is clearly bilinear. Moreover, if jc, y g g and


m ^ M ,^ e have

i [ x , y ] ■f , m } = < /, [o-(y),<r(A:)] ■m}


= ■a { x ) ■m) - ( f , a { x ) ■<r(y) • m>
= { x - y • f , m ) - ( y ■X ■f , m ) ,
so
[^>y] - f = x ■y ■f - y ■X ■f .
That AC• / e is clear. □

The g-module is called the restricted dual module of M.

Proposition 16
Let M G ^ (g , and let be its restricted dual. Then

(i) admits a weight space decomposition relative to

=e
with

(ii) e ^ (g , ^ y ,
(iii) ch = ch M;
(iv)
(v) M is an exact contravariant functor o f i^(g, into itself;
(vi) L(A)^ = L(A) for all A e 1^*.

Proof (i) Given / € we can define the map =/ ° ^ where


is the natural projection M that annihilates Af** for all ju. A. By the
2.6 The Category < 159

definition of M^, f ■= where 5 is some finite set (depending on /) .


Let us see that each is a weight vector of weight A. Indeed, given m ^ M
and writing

m = £ m ^, e M>^,

we have, for all /i g f),

{h ■f^ ,m ) = {f¡^,a{h) ■m) = {f ^ , h • m>

= < / a>

= k{h){f)^,m) = k{h){f)^,m).

This shows that f¡^ g Now part (i) follows.


Evidently - (M^)* as K-spaces, and hence dim,^ M'^ = dimi^ M^,
since all the spaces in question are finite dimensional. This establishes parts
(ii) and (iii).
(iv) Write = M for convenience. If m g M, let m g M be de­
fined by

ñi{f) = { f , m) ioxdWf^Mg.

The resulting linear map

~ :M ^M

is clearly injective. Moreover

c for all A G f)*

so that ~ is surjective because of the finite dimensionality of the weight


spaces and part (iii). Finally, ~ is a g-module map. Indeed, if a: g g,
w G M, and / G Mg, we have

X ■m ( f ) = { f , x ■m) = <cr(x) ■f , m )
= m{a{x) • /) = ( x - m ) { f ) .

as required.
SO X • m = X • m
(v) Let M and N be in , Given a g-module homomorphism
F: M N, define
F : AT* M*
160 Lie Algebras Admitting Triangular Decompositions

by

for all / e N* and m ^ M (thus is the transpose of F). It is clear that if


/ vanishes on N^, then F^f vanishes on M^. We therefore get an induced
map
F 'N M

We verify that this is a g-module homomorphism

F :N M .

In fact for all x ^ q,

{F„{x ■f ) , m) = ( x ■f , F{ m) ) = ( f , cr {x) ■F ( m ) )
= ( f , F { a { x ) ■m)>
= ( F^ f , a ( x ) ■m)
= {x • F„f , m).

If G : AT ^ P is another morphism in then

( G oF) ^ = F„ oG^,

showing that

is a contravariant functor of ^ (g , into itself. The exactness is left as an


exercise.
(vi) Let ¿;+ be a highest weight vector of L(A), and let v% be the linear
functional = 1, = 0 if ^ X. Then by part (i), u% e
(L(A)^)^, and for all x e g" , jc • y* e (L(A)^)'^'^“ = (0). This shows that i;*
generates a highest weight module with highest weight A. Since ch L(A)^ =
ch L(A), we conclude from Proposition 12 that L(A)^ —L(A). □

Caution is not the same thing as defined in Proposition 2.3.5.

2.7 THE RADICAL

In this section we introduce four different ideals of a Lie algebra g with a


triangular decomposition (i), g+, Q+, a). Each of them has a right to be
2.7 The Radical 161

called a radical of g. However, in general they are not all equal. They are

1. t y ( g ) == the set of all elements of g that annihilate every irreducible


module in
2. m ^(g) == the largest ideal of g intersecting trivially;
3. n(g) == the sum of all the residually nilpotent ideals of g;
4. rad(g) := n the intersection being over all triangular decom­
positions of g.

We discuss these in turn. We find that always r ^ = m ^ , and in the case of


primary interest to us, that is when E , e j 9~“' + h + E , e j 9“' generates g as
a Lie algebra, r^ ( g ) = m ^ (g ) = rad(g) and Z(g) ffi rad(g) = n(gX We
make use of these results in Chapter 4.
Throughout this section g will denote a Lie algebra over IK admitting a
triangular decomposition, and we will let ■^= (i),Q+,Q+,o’) denote one
such decomposition. To describe this situation, we will henceforth say that
(g, is a triangular pair.

Proposition 1
Let (g, y ) be as above, and write g = g _ ® ® g+. Let Z(g) and D(g)
denote the centre and derived algebra o f g respectively. Then

(0 Every ideal o f q is graded by Q;


(ii) Z(g) = {A e l^|a(/i) = 0 for all a e A};
(iii) D(g) = g_®([g, g] n fi) ® g+.

Proof (i) Let a c g be an ideal. Then [1^, a] c a and therefore a is a


submodule of g under the adjoint action of 1^. Now part (i) follows from
Proposition 2.1.1.
(ii) By part (i) and element belongs to the centre of g if and only if all of
its homogeneous components do. If x e g“, a e then a(,h) # 0 for
some h Thus [h, x] = a(h)x # 0, and hence x i Z(g). Similarly g_n
Z(g) = (0). It follows that Z(g) c fi, and part (ii) is then clear.
(iii) As in part (ii) we see that g_® g+c D(g) and hence that part (iii)
holds.

Let (m,), s i be any family of ideals of g such that m, n = (0). Then the
ideal Em, of g, being graded, satisfies (Em,) n = (0). We conclude that
there exists a unique largest ideal of g intersecting t) trivially. Thus m ^(g)
exists.
162 Lie Algebras Admitting Triangular Decompositions

Proposition 2

Let (g, be a triangular pair as above. Then r = m ^(g).

Proof. Let L be an irreducible module in ^ (g , ^ ) . Then L s L(A) for


some A e 1^* (Proposition 2.6.2). The annihilator of L is clearly an ideal of
g, and hence also is an ideal of g. Suppose e r ^ n ]&. Then for any A
for which A(/i) # 0, we have h • L(A) # (0) contradicting the fact that h e
r ^ . Thus r ^ c xn^. Conversely, let L(A) be an irreducible module of
(s, "^X and let i )+ g be a generator forit. Let V ~ v^. Since
m ^ n = (0) and since is graded, we have

V= ti+c g_- v^+ g+- v^= g_- v+<z Y, L(A)^.


/1 < A

Now

g+- V = 9+- Ü+C ( m ^ + m ^g + ) • v^= V.

Thus

U(g)-FcU(g_)U(^)-Fc E L(A)^
<A

which shows that U(g) • F is a proper submodule of L(A) and hence (0).
This proves that m^* u^= (0).
Let N '= {u Ei L(A)|m^* v = (0)}. Then N is a submodule of L(A) since

(9 • ^ g] • N + 9 • N = (0),

Since v +E N, it follows that N = L(A) and hence that m ^ annihilates L(A).


Since Awas arbitrary, this shows that m ^ c as desired. □

The ideal r^ ( g ) (or simply if g is understood from the context)


described in the last proposition is called the (Wedderburn-Jacobson) radical
of g with respect to We will see in Proposition 5 that for a large class of
Lie algebras admitting triangular decomposition, the radical does not
depend upon the choice of triangular decomposition
Define

r 5r : = r ^ n g + and r^:= r^ng_.

Then
2.7 The Radical 163

Since cr(r^r) is an ideal of g intersecting trivially it is easy to conclude that

= r^
and that

Example 1 (Extended Heisenberg algebra a) The notation is that of


Example 2.6, where we extended a Heisenberg algebra by a derivation in
order to obtain a triangular decomposition. Recall that S = IK[^y]ye j can be
made into a highest weight module of type (1, A) for any A e ]^* where
if = Kc ^ Kd. Moreover these modules are irreducible whenever A(c) 0.
(Example 2.3.2). Now Kxj^) ■■ ^ 0 so that r^ (fi)_ =
(0). It follows that r^(a)^.= cr(r^(a)_) = (Ó) and hence that a is radical
free with respect to the triangular decomposition in question.

Example 2 Let g = Ka © Kh © Kb be a three-dimensional Lie algebra


with bracket defined by [A, a] = —2a, [h, b] = 2b, and [a, b] = 0. It is easy
to see that g admits a triangular decomposition with root lattice Q = Za
where a e (K/z)* is defined by a(h) = 2. The two-dimensional space r =
Ka © Kb is an ideal of g intersecting Kh trivially, and hence r is the radical
of g relative to the triangular decomposition in question.

Proposition 3
Let g and ^ be as above, and suppose that r ^ ( g ) = 0. Then for g+ and
g_ we have

[ z+ ,s _ ] = (0) <»z+= 0,
[z_, g+] = (0) « z_= 0.

Proof Suppose that [z+, g_] = (0), g+. Without loss of generality we
may assume that z+ is homogeneous. By assumption U(g_) z+= Kz+,
where the action is given by restriction of the adjoint representation. Then

U(g) • z^ = U(g J U ( ^ ) U ( g _ ) • z , c g^,

which gives an ideal of g (i.e., the one generated by z+) intersecting


trivially. Thus z+= 0. The second assertion follows by applying a to the first.

Our next result deals with the problem of how Lie algebras with triangular
decomposition behave under extension of the base field. (For Lie algebras
obtained by extension of the base field see Remark 1.1.1).
164 Lie Algebras Admitting Triangular Decompositions

Proposition 4
Let q be a Lie algebra over IK admitting a triangular decomposition
(^5 9+> ¡2+? ^)* Let IK' be a field containing IK. Let g' = IK' the Lie
algebra over IK' obtained from g by extension o f the base field from IK to IK',
and let = (IK' IK' ®(k9+, Q+> I ® cr) be the corresponding triangular
decomposition of g' {see Remark 3, Section 2.1). Then

Proof. For convenience let us set i)' = IK' and g'^. = IK' <S>|,^g+. It is
immediate that IK' c To establish the reverse inclusion, it
suffices to show that r^ ,(g ')‘^c IK' (for we can then apply 1 ® a).
Let be a basis of IK' over IK. Let a e g+j and consider a nonzero
element z e r^ /(g ')“. Write

z = T, k^<8>x^,
A eA

where x “ e g“. Suppose that e g^',. . . , y, e g^' is a sequence of ele­


ments where /Sj • • • -1-/3, = —a. Then ad(l ® y , ) . . . ad(l ® y,Xz) e
r^Xg') n £)' = (0), and

0= ® adyi...ady,(ji:J').
AeA

We conclude that ad ... ad = 0.


It follows that for each A e A , the ideal generated by in g intersects
trivially and hence that each x “ er ^ ( g ) '^ . Thus z e IK'<8>j^r^(g)‘^ as
desired. □

Remark 1 Let g, y , r ^ , and be as above. Define

9 == 9 / r ^ ,

and let : g ^ g be the canonical map. Then

g = g_e fi e g+,

where g+— 9+ A 5r and §_== g _ A 5^ (we are identifying with a subalge­


bra of both g and g). It is clear that a induces an antiinvolution of g since
o-(r^) = r ^ , as we just saw. As long as 5+=?^= (0), we obtain an induced
2.7 The Radical 165

triangular decom position

of g with root spaces given by

9“ = g“ for all a e g .

Finally, observe that if a is an ideal of g such that d n = (0), then the


preimage a of d in g also intersects ij trivially (because the restriction of “
to if is injective). Thus a c r We conclude that

( 1) = (0).

We next consider the problem of the independence of r^r from We


begin with some general facts about Lie algebras. If A is an ideal of an
arbitrary Lie algebra I, we inductively define a decreasing series A =A^
A^ ^ • • • of ideals of I as follows:
A^ :=A,
A'^^^ := [ A, A^ ] .

The ideal A above is said to be residually nilpotent if


n ^ " = (o).
n>l

We define n(I) to be the sum of all residually nilpotent ideals of I. Observe


that n(I) itself need not be residually nilpotent.
Returning now to our Lie algebra g we define the (Wedderburn-Jacobson)
radical rad(g) of g by
rad(g) =

the intersection being taken over all triangular decompositions of g. We say


that g is radical free if rad(g) = (0). Notice that g/rad(g) is always radical
free (see Exercise 2.18).

Proposition 5

Let (g, ^ ) be a triangular pair as above. Assume that

(CTl) The subspace 5 ^ 9 generates % as a Lie


algebra.
166 Lie Algebras Admitting Triangular Decompositions

Then

(0 r ^ , r Jr, and are residually nilpotent ideals o f n


(ii) r ^ c for any other triangular decomposition o f g, in particular
= ^ 5^(9) = rad(g);
(iii) n(g) n Í) = Z(g);
(iv) ^ induces a natural triangular decomposition on g/Z(g), and we have

rad(g/Z(g)) = it (g/Z(g)),
Z(g) e rad(g) = it(g);

(v) n(g) is residually nilpotent]


(vi) I f S ci)* spans f)*, then rad(g) = fl a s 5 ann L(A).

Proof (i) Let U(g) • r j- be the ideal of g generated by r j- (here U(g) acts
by the adjoint action). We must show that

U (9) • r ; ^ c r ^ .

Suppose not. Then, since U(g) • r j- c r ^ , there exists some element z e


r ^ n g“, a e Q+, such that

[ x i [ x2,...)[-'"A;>-2] ' ’ ' ] ] ^ {0}

for some elements Xj,. . . , g g. From condition (CTl) we can assume that
Xi e g * “ >. for some g J. If we choose k above minimal, then X j g g “ “ A,

and

w ■■= [x2, [ x 3 , . . . , [ x * , z ] ] G

But w is an element of n = (0), contradicting the fact that [xj, w] # 0.


It follows that r jr is an ideal of g. Moreover

( 2) (rj.) c © g“,
ht(a)>n

which shows that r J- is residually nilpotent. The analogous results for


follow by applying a. To show that is residually nilpotent, first observe
that [t;^, rjr] c = (0) and that hence

r ^ c ( r » " + (rj-)".
2.7 The Radical 167

Combining this with (2), it follows that

r^ c e g“.
aeA
\ht{a)\ >n

Thus 0 = (0), and is residually nilpotent.


(ii) Let 3^' = be any triangular decomposition of g. We
claim that c that is, Pi 1^' = (0). For suppose not. Let h 0
belong to this intersection. If a'(h) = 0 for all roots a' of (g, then
h e Z(g), and hence ^ripart (i) of Proposition 1). Thus h e r ^ n =
(0), a contradiction. Therefore let a' be a root of (g, such that a'(h) 0,
and let X ^ X ¥= 0. Since

(ad h ^ x = a ' ( h y x ,

we conclude that x e for all contradicting the fact that is resid-


usually nilpotent. Thus x ^ c z x ^ r , and the rest of (ii) follows from Proposi­
tion 2 and the definition of rad(g).
(iii) Let a be a residually nilpotent ideal of g. If e a n and a(h) ¥= 0
for some a e A, then

(ad/l)"g- = g ^

and hence g" c a ” for all n, which contradicts the fact that a is residually
nilpotent. It follows that a n c Z(g). Let a^,. . . , be residually nilpotent
ideals of g.

(Uj + • • • + a^) n ]^ - (Ui + • • • +a^)^ = a? + • • • +a^


(3)
£ a , ni)cZ(g)
/=1

by the above. Now let h e n(g) Pi 1^. Then by definition h belongs to a finite
sum of residually nilpotent ideals of g, so e Z(g) by (3). This shows that
n(g) Pi c Z(g). The reverse inclusion is obvious.
(iv) We can assume that for each i e J, g“' ¥= ( 0 ) . Then from part (ii) of
Proposition 1 , o : ,lz (g ) = 0 for each i e J, so we have a natural interpretation
of j2 as a set of linear functions on íi/ Z ( q). Furthermore {«/l.ej remains
linearly independent in this setting. Let tt : g 9/Z (g) denote the natural
mapping. Then 'Tr(g) = '7r(g_) 0 7t(]^) 0 7r(g+) [note that 7r(g+) ¥= ( 0 ) ] ad­
mits a weight space decomposition relative to 7r(fi), and a factors through tt.
Thus we have an induced triangular decomposition Evidently Z(7r(g)) =
(0 ), and (CTl) holds for 7r(g).
168 Lie Algebras Admitting Triangular Decompositions

By Part (i), tyÍTTÍg)) c nCirCg)). To establish the reverse inclusion notice


that Tr(ii) n n(Tr(g)) = (0) by Part (iii). Thus rad(g/Z(g)) = n(g/Z(g)). Let
a be a residually nilpotent ideal of g. For notational convenience let
denote the map v . W contend that 5 is a residually nilpotent ideal of g.
(Note that this is not “obviously” the case.) Set 21 = Write 21 =
®2l“ relative to the induced triangular decomposition ^ of g/Z (g) (see
Proposition 5).
Suppose that ic g 21“ \ {0} for some a 9^ 0. For each n = 1,2,... there
exists x„ G (a")“ such that x„ = x. Since ~ restricted to g“ is injective,
x„= x„=- X ^ g“ for all m, n. Then x g = (0). Since any ideal of
g lying in 5 must be (0), 21“ = (0) and 21 = 21 n ^ c Z(g/Z(g)) = (0). This
finishes the proof of our claim. Thus, if a is any residually nilpotent ideal of
g, then 3 c n(g/Z(g)) = rad(g/Z(g)). Combining this with the fact that the
preimage of rad(g/Z(g)) in g is (clearly) rad(g) ® Z(g), it follows that
a c rad(g) ® Z(g). Thus n(g) c rad(g) ® Z(g). The reverse inclusion fol­
lows from (ii) and (iii). This fibnishes the proof of Part (iv).
Finally, n(g) is a residually nilpotent ideal. Indeed
n(g)" = ( r ( g ) ® Z ( g ) ) " = r(g)",
and r(g) is residually nilpotent.
(vi) ann(L(A) := [x G q\x • L(A)) = (0)}, so by definition rad g c
n as5 L(A) =' I. As for the reverse inclusion, note that / is an ideal of
g; hence if I <t rad g, then by Part (ii), I n i j ¥= (0). Now, if A e ( / n 1^) \ {0},
then A(/i) ¥= 0 for some A e 5. Hence h ^ ann L(A). This contradicts the
fact that h ^ I. □

Remark 3 The Lie algebras that are of most interest to us, namely the
contragredient Lie algebras to be introduced in Chapter 4, satisfy the
assumption (CTl) of Proposition 5.

2.8 THE SHAPOVALOV FORM

The problem of describing the submodules of a Verma module is in general


very difficult. For instance, it is not at all obvious whether M(A) is irreducible
and how irreducibility might depend on A. In this regard the construction of
the Shapovalov form is extremely useful. This is a so-called contragredient
bilinear form on U(g) with values in 5(1^) which is then transferred to Verma
modules M(A). In principle it provides information about the weight spaces
MA)'^"“ of the maximal submodule N( \ ) of M(A), although in practice a lot
more work is required before any specific information is forthcoming (see
Sections 6.6 and 6.7).
We assume that we have a Lie algebra g admitting a triangular decompo­
sition g_ 0 0 g+, root lattice Q, and anti-involution cr. This is the first time
the anti-involution is going to play a significant role.
2.8 The Shapovalov Form

Using the Poincare-Birkhoif-Witt theorem, we may write

(1) U (g) = U(^) e {g_U(9) + U(9)9+}.

Note that <r extends uniquely to an anti-involution (also called <r) on U(9)>
which interchanges 9_H(g) and U(g)9+ and pointwise fixes UO^). Both
components on the right-hand side of (1) are fi-modules under the adjoint
representation. We let q : U % ) ^ U(^) be the projection defined by (1) (so
that q annihilates the summand in braces). It is convenient to use Proposi­
tion 1.8.2 to identify U(fi) with the symmetric algebra S(^) so that q:
U(g) S(l^). We define the Shapovalov form

f : U ( g ) x U (9 )

by
F { x , y ) = q{<r{x)y).

We collect together some simple facts about F.

Proposition 1

(i) F is symmetric.
(ii) For all u , x , y ^ 11(9), F(ux, y) = Fix, a(u)y).
(iii) For a , P ^ Q and a ¥=p, U(g)“ ± U(9)^ relative to F.
(iv) H I, 1) = 1.

Proof, ii) Let X, y e U(g). Writing

<r(x)y = F i x , y ) + r,

where r e g_U(g) -l- U(9>9+, and applying <r, we have a iy ) x = Fix, y) +


crir) from which F(x,y) = F(y,x).
(ii) F(ux,y) = q(<r(ux)y) = q(cr(x)o-(u)y) = F(x,(r(u)y).
(iii) If X e U(g)“, y e B(g)^, then tr(x)y e U (g)^-“. Since the decompo­
sition (1) is a sum of fi-modules, we see that if ^ - a 0, then qiaix )y) = 0.
(iv) Fil, 1) = qia-iDl) = i d ) = 1 - □

Let g be a Lie algebra with an anti-involution a. A bilinear mapping

B: M X M V

of a g-module M into some space V is called contr^redient (relative to tr) if


170 Lie Algebras Admitting Triangular Decompositions

for all x , y Ei M and for all w e g

B{ux, y) = B{x,(T{u)y).

The Shapovalov form F is contragredient relative to the anti-involution of


the triangular decomposition.
Remark 1 More common in physical applications are the hermitian contra­
gredient forms. For these we need a complex Lie algebra g with a semilinear
anti-involution cr. Then, if M is a g-module, a hermitian contragredient form
on M is a mapping

which is linear in the second variable and antilinear in the first—that is,
{ax + y\z) = a{x\z) + <y|z> and {z\ax + y> = a{z\x) + <z|y>—such
that {u jc|y> = <jc|o-(m), y> for all x, y e M and for all w e g. If the
hermitian contragredient from < • | • > is positive definite, then we call it a
unitary contragredient form and say that the representation of g on M is
unitary (relative to < • I * » .

We will usually construct hermitian contragredient forms by beginning


with a real form f of g with an anti-involution a and a f-module M with a
contragredient form ( | •) on M. Then the complexification of g acts on
the complexification of M. Furthermore a extends uniquely to a semilin­
ear anti-involution cr on gc» (*!*) extends uniquely to a hermitian
contragredient form < • | • > on M^. We have

{x iy\u + iv) = {x\u) i{x\v) — i(y\u) + <y|i;> for x, y, m, y e M.

One checks immediately that {g • x\y) = <A:|cr(g) • y> for all g e f and
for all jc, y e M^. Then for g, e I and x , y ^ M^, we have

( ( g + ih) *jr|y> = <g *x|y> - i(h -xly) = {x\(Tg *y> - i{x\ah *y>
= {x\{(Tg - iah) *y) = { x \a { g + ih) *y),

which shows the contragredient. Also < • | • ) is unitary if and only if (*| •) is
positive definite.
Let us return to the discussion of the Shapovalov form. Any A e 1^*
extends uniquely to a homomorphism /- > /( A ) of S(ij) into K. Since in
particular the Shapovalov form takes values in we may define, for each
Ae a [K-valued symmetric bilinear form f (A) on U(g) by

F (A )(x ,y ) = ( F (x ,y ))(A ).

Now if y e U(g)g+, then for all x e U(g), F(x, y) = q((r(x)y) = 0 (since


2.8 The Shapovalov Form 171

cr(jc)y e 11(9)9+) and in particular F(A)(x, y) = 0 for all x e U(9)- Also if


y G Xli^Xh — A(]^)X then it is a straightforward exercise to show that F(x, y)
= f(h —A(/i)) for some / e 5(1^) and hence again F(A)(x, y) = 0. It follows
that the left ideal 1(A) [see (2.3(5))] of U(9) is in the radical of F(A). If we
recall that M(A) == U(g)/I(A) and use 1 e U(9) to determine the highest
weight vector ¿;+ of M(A), then we obtain an induced blinear form

F^: M(A) XM(A) K

satisfying

F;,(x • i; + ,y • u^) = F (A )(x ,y ) for all x ,y e U (g).

Remark 2 The symmetry between + and - in (1) is deceptive. The


definition of the Shapovalov form given here is appropriate for highest but
not lowest weight modules.

Proposition 2

(0 F^ is symmetric and contragredient.


(ii) Relative to F^, unless a = /3.
(iii) The maximal submodule is the radical o f F^.
(iv) Writing M(A) = lKi; ++ have for all x, y e U(9)
that F^(x • v^, y • v^) is the component of o-(x) *y • i^+ in

Proof Parts (i) and (ii) are straightforward. To move part (iii), let R be the
radical of F;^. F is a submodule by the contragredience of F and is proper
since F^(i^+, 1;+) = 1. Thus R c MA). Conversely, let w e MA). Then for all
jc e U(9),

Fj ,(x -v^ ,w ) =F^(v +,o-(x) -w ) = 0

by part (ii), since cr(x)w e N(Á) c This shows that MA) c


R. To move part (iv), write a-(x)y = F(x, y) + r as in the proof of Proposi­
tion 1, and apply it to i > □

We have established that M(A) is irreducible if and only if is nonde­


generate. However, we can be more precise than this. Let y ^ Q+ U{0}, and
define FL as the restriction of F to U(q_)"'’':

F |^ :U ( g _ ) " ^ x U ( g _ ) ' S (^).


172 Lie Algebras Admitting Triangular Decompositions

Proposition 3
F|y(A) is singular if and only if N ( \Y # (0).

Proof, We have by

u U ' V.

and

F|y(A)(Mi,«2) = F ( « i ,M2)(A) - v + , U 2 - v ^) .

For «2 e U (g_)'

( 2) «2 is in the radical of F|,y(A)

<=> U2 *v ^ ± relative to Fj^,

<=> «2 *^ + -*- ^ (A ) relative to

<=> «2 *^ + ^ □

Now suppose that g admits a regular triangular decomposition. Then for


any y G ¡2+ U{0} we may find a basis, say, « i,. . . , u„, of U(g_)~'^ (See the
proof of part Proposition 2.2.1 (iii)). Consider the n X n matrix

(3) Sh^(A) =

From (2) we have the following:

Corollary
I f g admits a regular triangular decomposition and Sh^ is defined by (3), then

(i) the dimension of N(XY~^ equals the dimension o f the null space of
Sh/A ),
(ii) M(A) is irreducible if and only if det(Sh^(A)) ^ 0 for all y ^ U{0}.

Amazingly det(Sh^) can be explicitly computed for some of the algebras


that are most interesting to us, for instance, affine algebras and Virasoro
algebras (see Section 6.6 for the invariant Kac-Moody case). The result is that
det(Sh^) factors into linear factors in 5(1^). The knowledge of these factors
obviously provides direct information about MA).
2.8 The Shapovalov Form 173

Example 1 (Heisenberg algebras) We continue directly from Example


2.3.1. In the notation established there, for |m| = |nl > 1 we have

(4) F ^{a 'lv^,a 'iv^) = F j^v^,a'la'Lv^)

= „m!A(c)'"’' [see equation 2.3(8)].

If A(c) ^ 0, we see that the basis of M(A) is an orthogo­


nal basis. If A(c) = 0, then every basis element |m| > 1 is in the
radical of Thus we see once again that M(A) is irreducible if and only if
A(c) ^ 0 and MA) = if A(c) = 0.
If is finite dimensional, then so is M(A)^“^“ for each r > 0 and

det(Sh,_,„(A)) = n m!A(c)^
|m| = r

If IK = R and A(c) > 0, then (4) shows that the form is positive
definite. If we complexify a and M(A), then according to Remark 1 above,
the resulting representation is unitary.

Example 2 (The Virasoro algebra) We use the notation of Example 2.3.2.


The diagonal algebra is spanned by L q and c, and M = M{h, c) denotes the
Verma module with highest weight A: A(Lq) = /z, A(c) = c. Let denote a
highest weight vector of M. The subspace M" has as a basis the set of
elements • K+, where < • • • < rij^ and «j + • • • = n.
We compute out the value of F^ on this basis for the values n = 0,1,2. This
requires some simple calculations:

L iL _i • V ( 2L q + L _ iL i) •
= —Ihv^

_\ F _j * “ 3 L jL _j F _1^ 2.^ —\ *^ +
= 6hu^
F^F^F_2 • 6hv^

L iL iL _ iL _ i • v += (8/z^ —4h)u +

L2L_2 ■u+= { - 4 h + (i)c)ü + .


174 Lie Algebras Admitting Triangular Decompositions

The resulting matrices for « = 0,1,2 with respect to the bases {v^},
{L_ • i; +}, and {L _2 *u+, • z;+} are

—Ah + (^)c 6h
(1), i-2h),
6h 8h^ - Ah

In the case that K = R, we may, as before, produce a hermitian form on the


complexification of M(A). The problem of deciding for which pairs h,c the
resulting representation M(A)c is unitary is much more involved and inter­
esting than first appears. (See [GO] for more on this.).

Example 3 ( ê l2(IK)) For A e and a highest weight vector of the


Verma module M(A), a basis for M(A)^“"“ is / " • The computation

^ a( / " • ■i'+) = «! n ((A + p)(h) - k ) ,


k= l

is left as an exercise (Exercise 2.19). We have written this using the function
p e p(h) = 1, because it appears this way in the generalization (Section
6.6). We observe that the formula shows us in another way that is singular
(hence M(A) is reducible) if and only if \(h) g I^.

Proposition 4 [Ka]
Let q be a Lie algebra over C with triangular decomposition ^ =
(^j 9+j Q+j cr). Suppose that the basis Yi o f Q is finite, and let a <z q be a
subalgebra of g satisfying

(i) cT-a = a;
(ii) there exists an element such that
• [/io,a] c a,
• there is a real valued linear functional f on C such that f({a, Aq)) > 0
for all a e n .

Then any unitary q-module M in category 0 is completely reducible as an


a-module.

Proof Let M be a unitary module. Define F: R>o by


♦= ^o)- For each c e C let (resp. g^) be the sum of the
a-eigenspaces of /iq on M (resp. ad /iq s) for which F(a) = c. Since the
weights of M lie in a finite number of fans and F takes a discrete set
of positive real values on ¡2+, it is clear that for all c e C, dim < oo. Note
that g^ = ^ and g^ • c for all c, d e C.
2.9 Jantzen Filtrations 175

Decompose a into ad/iQ-^ig^nspaces: a = Define (resp. a_,


resp. Oq) as Ea^^ summed over all c such that /(c ) > 0 (resp. /(c ) < 0, resp.
/(c) = 0). Let a := a + C/iq. Then
a = a_ + (ao + C/iq) + a+ and dg := Qq + C/iq c ]^.
Let < • I • > be the contragredient form relative to which M is unitary. We
begin by showing that M is completely reducible as an á-module. Indeed, if
N is an d-submodule, then N = where C\ N. Since < • | • > is
positive definite on M"" and dim < oo, has an orthogonal supplement
in M^: = N"" ± P"". Thus N M = N ® ^ , and N ^ is an
d-submodule because < ♦| • > is contragredient. This shows the complete
reducibility.
Now suppose that AT c M is an irreducible fi-module. We prove that it is
an irreducible a-module. For if not, then N ¥= (0), and we let x = Ex^,
be a vector chosen with the following properties:

1. x e N \ { 0 ) , U(a) - jc #iV,
2. cdLiá{c\x^ # 0} is minimal among all vectors x satisfying property 1.

Note that the weights of M lie in finitely many fans and hence the values
of /«A ,/zo» form a discrete set bounded above as A runs through the
weights of M. Thus among the vectors jc = E jc^' satisfying properties 1 and 2
we may choose one for which

3. m3x{ f(c)\x^ ¥= 0} is maximal among all vectors x satisfying properties 1


and 2.

Let X satisfy properties 1, 2, and 3, and let x = Ex^, with x^ satisfying the
maximizing condition of property 3.
Now from the choice of x, - x^ = 0 for all d > 0; that is, a+- = 0
and hence a +* = 0 for all c. Thus for each c with x^ ¥= 0, N = U(a) •
= U(a_) • U(ao) • U(a^) • c E ^ ^,.A^^, ^ 0. This can only be true
for one value of c, and hence x = for some c. Therefore x is a /zro-eigen-
vector, It follows that U(a) • x = U(a) X = N, contradicting the choice
of X . □

2.9 JAN TZEN FILTRATIONS

Jantzen filtrations provide a technical tool for deeper study of Verma


modules. The results that we establish here are not used until we discuss the
Shapovalov determinant formula in Section 6.5, and the reader can (and
probably should) omit this paragraph until it is needed. This section is a
natural continuation of Section 2.8, and we use the notation established there
without further comment.
176 Líe Algebras Admitting Triangular Decompositions

Let A e 1^*, and let M(A) be the corresponding Verma module. Let
ze and imagine that we make a “small perturbation” A^ = A + iz, i e K,
in A. Presumably M(A^) is closely related to M(A). Indeed, if we identify
M(A^) as a subspace of U(q_), these Verma modules can all be compared
directly. In doing this, we find that for w e U(g_), the Shapovalov bilinear
form F^(v,w) is a polynomial function in t. Loosely speaking is defined
to be the subset of vectors u of M(A) for which F^lu, M(A)) c t^K[t] where
t is now a variable. This gives rise to a filtration

M( A) = Mq ^ D M2 D *• •

of g-submodules of M(A), a Jantzen filtration [see [Jz], Chapter 5]. We use


this filtration to give us a detailed view of the way in which the singularities
of arise, the main result being Theorem 4. Preliminary to even writing
down the statement of Theorem 4, we require an algebraic apparatus to
formalize the heuristic discussion above.
Let g have a regular triangular decomposition (1^, g+2+,o-), and let
K[i] be the polynomial ring in one indeterminate over K. Set

9 == IK[i] ®k 9-

Then g is a free (K[i]-module of rank equal to dim^^g. There exists a


unique IK[i]-bilinear mapping

[*, •]: 9 X g ^ g

extending the Lie multiplication of g:

[ p ( i) ® x , q ( t ) ®y] = p ( t ) q ( t ) ® [j;,y ]
for all p ( t) ; q{t) e IK[r]; x, y e g.

This makes g into a Lie algebra over K[t], It is also a Lie algebra over IKwith
g as a subalgebra through x 1 ® jc. (See Section 1.1 for Lie algebras over
rings. We will use the full generality of the Poincaré-Birkhoff-Witt theorem
(see Remark 4 in Section 1.8) in the arguments to follow.)
Now g has a triangular decomposition as a lK[i]-algebra in an obvious way,
as we saw in Remark 2.1.3. We define

§ H[t] Í)

and

9+== ^ [t] 9,
2.9 Jantzen Filtrations 177

We identify 1^* with a subset of — Honi|,^[,j(§, K[t]) by linear extension of

A(p ( t ) (S>h) = p ( t ) \ { h ) for all p ( t ) e K[t]; /i e 1^; A e f)*.

We make this identification without comment hereafter. Extend cr to <7 on §


by linear extension:

a(p(t) = p ( i ) ® <r(x).

Then our triangular decomposition is

where Q+<^ 5*. For a e Q,

r == K [i] ® 9“.

We set about constructing a suitable “Verma module” for g. Let U(§) be the
universal enveloping algebra of g [viewed as IK[r]-module (Remark 4 Section
1.8)], and let A, z e 1^* be fixed in our discussion. Set

A = A + rz e 5*.

Define /(A) to be the left ideal of U(g) generated by g^. and all elements of
the form h — Ae

i { k ) = { U , h - \{h)\h i/(g )-le ff

Then

M = M(A) := U ( § ) / /

is a §-module and is our desired Verma module (see 2.3(5)).


For each /1 e let

M4 := ! ^ v ^ M \ h - v = p ,(h )v for all ft e §}.

Using Remark 3 of Section 2.1 and the arguments at the beginning of Section
2.2 we see that

M= ©
aeG+U{0}
178 Lie Algebras Admitting Triangular Decompositions

and that

= IK[í]í)^., where i+== 1 + /,

M‘' ““ = U (g_) “ via M• (1 + / ) <-^ M-

This uses the Poincaré-Birkhoíf-Witt theorem (over IK[f]). In addition we will


need the Verma modules (over g), A/(A + az), for various a e IK, which are
defined as U(g)//(A + az), where /(A + az) = (g+, /z — (A + azXh)
\h G ^>u(3)-ieft [see 2.3(5)].
Let a G K, and let

£^;K [t] ^ K,
p ( t ) *^p(«)>

be the evaluation map. Since U(§) = »<[^1 «k « (s )’ ^ e have a K-linear


homomorphism of associative algebras

e„:U (g) - ^ U (9 )

defined by

£a- P ( 0 ® ^

We evidently have
e j9 = idg (9 identified inside g ),

^0( 9 +) = 9±>

kei£a = (t - n)U (9)-

We recall the decomposition of 2.8(1):

(1) U(g) = U(^) ® (9 -fi(9 ) + ^ ( 9) 9+)

and the projection map


^ :U (g )-.U (^ )-W

determined by this decomposition. In the same way we have

(2) U(g) = U ( § ) ® (9-U (§)


2.9 Jantzen Filtrations 179

and a projection

i':U ( § ) ^ U ( § ) = 5 ( § ) .

These projections will enter the picture in a little while when we consider the
Shapavalov form. We next assemble the properties of that we need.
Given / e 5(1^) and A e 1^*, we sometimes use the notation /[A] to refer
to its value /(A) at A. We use similar notation for / e 5(§). This may avoid
some confusion in the thicket of brackets that occasionally appears.

Lemma 1

(i) ë ^ ° â = <7° s^.


(ii) For all Ü G U(§), = (qej^idxx + az],
(iii) éa(/(Â)) c /(A + az).

Proof.

(i) The proof is obvious.


(ii) Write M= «1 + «2, where ii, lies in the ith direct summand in (2).
Then q(ii) = Mj, qs^,(u) = ^(e^(iJi)). It is therefore enough to prove
the result for u = Ui, where has the simple form p(t) ® f,
pU) e IK[i], / G 5(f|). Then

e<,(«(“ i)[-^]) = + tz])


= p ( ^ ) f [ ^ + «z],

while

(i?M “ i))[A + = ( « ( P ( « ) / ) ) [ ^ + «2 ] = p ( a ) /[ A + az].

(iii) Since is a homomorphism, it suffices to check that

8+) ^ 9+j

which we already know, and that

e^{h - A(A)) = e^(h) - (A + az)(e^(A )).

It suffices to check the latter for h = p(t) ® h, p(t) e K[t], h


180 Lie Algebras Admitting Triangular Decompositions

Then

^a{pi0 i { p { t ) ® h))

= - e ^ (p (0 ((A + tz){h)))

= e^(h) - p { a ) { \ + az)(h)

= B^(h) - (A + az)s^{h)

as required. □

After part (iii) of Lemma 1 we have from a natural mapping

= U ( g ) //( A ) ^ U ( g ) / / ( A + az) =M (A + a z) .

In the sequel M(A) and M(A + az) are denoted by M and M, respectively.

Lemma 2

(i) i/i is a (U(§), U(g))-m<2/7 in the sense that for all v ^ M and for all
X e U (§ ),

• v) =

(ii) ker = it — a)M.

Proof, (i) This is clear since is a homomorphism.


(ii)

M = U (g )//(A ) - U (g_)

\<t>. i^a
M = U (g )//(A ) ^ « (9 -)

The result now follows from ker £„ - it a)U(§). □

Now recall that the Shapovalov form on M(A + az) is obtained by


defining F-. U(g) X U(9) ^ through F iu, v) = q i a iu ) y \ evaluating
2.9 Jantzen Filtrations 181

F(u, i;) at A + az and observing that the ideal /(A + az) is in the radical of
this form. Thus we have an induced form on U(g)//(A + az) — M.
Similarly we define F: U(g) X U(§) ^ S(§) through F(u, u) = q(a(u)v)
and, after evaluation at A, we induce a lK[i]-valued bilinear form on
M = U(§)//(A).

Lemma 3
For all a ^ K and all x, y ^ M,

Proof. Let u,v ^ U(g_) correspond to x and y. Then

£aFx{x,y) = e^(i((r(M )i;)[A])

= + az]

A+ a z

by parts (i) and (ii) of Lemma 1. □

We now make an important hypothesis

(NDG) F^ is nondegenerate i.e. M(A)) = 0 <=>jc = 0.

For instance, suppose that to e 1^* can be chosen so that F^ in M(o)) is


nondegenerate. Then for any A e f)* setting z = co - A, we have F^+z
nondegenerate. Now let A = A + tz, and suppose that F^ (x, M) = 0 for
some X Ei M. If X ¥= 0, then we can suppose that (t — 1) does not divide x
[otherwise, divide out (t - 1)]. But then il/^ix) ¥= 0, and this contradicts
Lemma 3. Thus, in this case (NDG) holds for A.
For each A: = 0 ,1 ,2 ... define

M,:= { i e M |( i - a ) * |F x ( i , M ) } .

In other words, i e <=> Fj^x, y) is a multiple of (t - a)* for all y ^ M.


Since is contragredient, it is obvious that is a §-submodule of M.
Thus we have the filtration

(JFl) M = Mq ^ z>M2 ^ ■■■

of M. Since F^ is nondegenerate, n “ =o^* = (0)- We define

M* = A: = 0,1,2, • • • .
182 Líe Algebras Admitting Triangular Decompositions

By Lemma 2, M* is a g-module, so we have the Jantzen filtration

(JF2) M(A + az) = Mo =) Ml D M2 3 • • •

of M(A + az). Furthermore This last result is not entirely


obvious. We prove it in Theorem 4. The case a = 0 gives us a filtration of
M(A). This completes the formalization of the heuristic discussion that
opened this paragraph.
If y e <2+ U{0}, then we have corresponding filtrations of weight spaces

M^

where

Mf-^ = M* nM^-^ = {jc eM^-T'|(i -a)*lFj;(jc,M^-^)}

which is mapped by onto

This last sequence stabilizes at (0) after a finite number of steps because
M^^az-y ¡5 a finite-dimensional K-space for each y.
Let y e <2+^{0}- We know that M^~'^ is a free IK[i]-module of finite rank
(since U(g_) is a free lK[i]-module by PBW and K[t] is a principal ideal
domain). Let {x^,. . . , be a basis for it. We are interested in the Shapo­
valov determinant

det(Sh,,) = det(fx(x,,ii),. .) e K [i].

This is unique up to nonzero factors from IK.

Let f i t ) e IK[i]\(0). The number o rd ,_ ^ /(0 is defined as the unique


nonnegative integer k so that (i - a)^\f(tX but (t - does not divide
fit).

Theorem 4 [Jan]
Suppose that Fj^ is nondegenerate. Then

(0 For all y ^ Q + n{0}, ord,_„(det Sh^(A)) = E^=i dim < oo;


(ii) ch M^. = E^eQ^n,o, ord,_^(detSh.^(A)c(A + az - y);
(iii) Mj is the maximal submodule N(X + az) o f M = M(A + az);
(iv) n “ =o^* = (0).
2.9 Jantzen Filtrations 183

The theorem is a direct consequence of the following general result that is


also due to Jantzen.

Proposition 5
Let A be a principal ideal domain and let p be a prime ideal o f A. Let (K be the
field A / p . Suppose that X is a free A-module o f finite type, and set X •= X / p X
with natural mapping ijj: X ^ X. Suppose that F: X X X A is a nondegen­
erate symmetric bilinear form, and let det(F) be the determinant o f the matrix
{F{Xi, Xj)) for some basis {Jcj of X over A (note that det(F) is unique up to a
unit of A). Let

X^ = { x ^ X \ F { x , X )
and set

Then

ordp(det(F)) = Y. X*
k= l

(where ordp has the obvious meaning).

Proof B. Let X ^ = Hom^(X, A) he the -.4-dual of X. Each x ^ X induces


an element F ( jc, • ) of Let Y (zX^ be the submodule of all functionals
induced by X in this way. Since the mapping x F(x, • ) is injective,
ranky = rank X = rank X^. Using the well-known basis theorem [Jal] for
modules over principal ideal domains, we can find a basis {e?,. . . , of X^
so that {a^e^,. . . , a^e^ is a basis for Y for some suitable elements a ^ ^ A \
{0}.
Choose fi e X with F(f^, • ) = a^e^, and let {e^,. . . , be the basis of X
dual to the basis {e^,. . . , We have

Since {/)), is a basis for X (because {a,e,}, is a basis of Y), there is a matrix
S = (Sij) e GL„(A) with /, = £y_iS,yey, i = Thus

d e tF = d e t ( F ( / ,,/,) )

= d e t|F (i;s ,.;C p /* jj

= ( a ^... a^)Aet S.
184 Lie Algebras Admitting Triangular Decompositions

Since det 5 is a unit of ^4, we conclude that


n
ordp(det F) = Y j ^rdp
k=l

Let X = T.bjfj e X, Then

Jc e <=> ordp F[ej, x) > k for ; = 1 ,..., n,


<=> ordp ajbj > k for y = 1 ,..., n,
<=> ordp bj > k - ordp aj.

Now feyCmod p) = 0 unless ordp bj = 0. Thus X¡^ = €


A / p ) = Eord„a )•/'(/;)> where J(,k) ■■={j e { 1 ,..., n}\k - ord„ a. <
0}. Thus '

diniK Xk = card(/ lordp aj > A:},

and hence

L dim^ = L ordp ÜJ = ordp(det F ).


k=l j=l

Proof o f Theorem 4.

(i) This follows from Proposition 5 with ^ = IK[r], X = M ^~f, p =


U - a)IK[r], and so on
(ii) This is an immediate consequence of part (i)
(iii) Let X e M, and let jc e M be any preimage of x. Then using
Proposition 2.8.2 and Lemma 3,

X e AT(A + az) <=i> M) = 0

<=>Jc e Ml <i=>jc e Mj.

(iv) From part i, fl = (0) for all y e Q+ U{0}. □

2.10 BERNSTEIN-GEUFAND-GELTAND DUALITY

The origin of this section is a paper by Bernstein, GelTand, and GelTand


[BGG2]. Their theory was worked out for finite-dimensional split semisimple
2.10 Bernstein-Gerfand-Gerfand Duality 185

Lie algebras. The generalization to the infinite-dimensional case was carried


out by Rocha-Caridi and Wallach [R-CW], and this section follows their
paper.
The motivation for [BGG2] was a certain duality principle that had
appeared in the category of finite-dimensional ^-modules for certain
finite-dimensional associative algebras A over a field IK. Suppose that
{Lj,. .. , L„} are the irreducible ^-modules (up to isomorphism). For each L,
there is a unique indecomposable projective module for which
Hom^(/^, L^) ^ 0. Let = [¡¿: Lj] be the number of occurrences of Lj in a
composition series of The matrix C = is called the Cartan
matrix of A (no relation to the Cartan matrices that appear in Chapter 3 and
play such a fundamental role in all of the rest of the book!). For certain
A,C = D^D for some integral matrix D (not necessarily square). In particu­
lar C is symmetric and positive semidefinite.
One natural way to interpret this would be to find a collection . . . , M„}
of “intermediate” modules that occurs as the factors in certain descending
series for each /¿. If we denote by [/^: M^] the number of occurrences of
in li in such a series and by [M^ : Lj] the number of occurrences Lj in
then c^y = [li : L y ] = : L y ] . Thus, if it were true that the
“duality principle” [/^: M^] = held, then C = D^D for D = (d^tyX
where == : L y ].
In this section we construct various subcategories of the category for
which just such a duality principle holds, the and Ly being the Verma
modules and irreducible modules in this subcategory, respectively. Our main
task is to construct the indecomposable projective modules for which the
Verma modules are to play the intermediate role. In Section 2.11 we will use
this duality principle to prove a basic result about embeddings of Verma
modules into Verma modules.
Let g be a Lie algebra over IK with regular triangular decomposition
(f), g+, Q+, cr). Fix A e ]^* once and for all.

We define if(A) to be the full subcategory of <^(g, T) whose objects are


those modules M of ^ for which

M= ©
tei)*

and = (0) if fjL ^ A i Q^. We define Fif(A) to be the full subcategory of


if(A) whose modules are finitely generated g-modules.

Note that if M is a module of F-^iA) with the finite set of generators


Xi = E x f, i = 1 ,..., AT, where x f g then {xf'l/ = 1 ,..., n, e 1^*} is
also a finite set of generators. Thus modules in F-^ have finite homogeneous
generating sets.
186 Lie Algebras Admitting Triangular Decompositions

Proposition 1

(i)
T^(A) is closed under taking submodules, forming finite direct sums,
and forming quotient modules. In particular i^(A) is an abelian
category.
(ii) Fif(A) is closed under forming finite direct sums and forming quotient
modules.
(iii) For all fjL E: M(fji) and L(fi) are objects in Fif{A) and -é'(A)
and the modules L(/i), At g A J, ¡2+? precisely, up to isomorphism,
the irreducible modules in if{A).

Proof. The only part that is not immediately obvious is the statement about
the irreducibles in A), and for that we use Proposition 2.6.6. □

Since A is fixed, we will feel free to abbreviate if(A) and F ^(A ) to and
F-é" respectively.

Lemma 2
Let M be a module in F€. Then any proper submodule N o f M in -é" lies in a
maximal proper submodule o f M in -é.

Proof Let jCj,. . . , be a set of generators of M, and let N be any proper


submodule of M in Choose k as large as possible so that the module
generated by {N, . . . , x¡^} is not all of M. Let ^ be the set of all
submodules K oí M ijm -é") such that Nj^<z:K but ^ K. Then ^ is
nonempty, partially ordered by inclusion, and inductive and hence has a
maximal element M' by Zorn’s lemma. Evidently M' is a maximal submod­
ule of M in if . □

We now define for each /jl ^ A i Q+ an induced module P Í jjl) that we will
eventually prove is projective in t^(A) and has M ( / jl) as a quotient. Fix
¡JL G A iQ+, and let IK^ be the one-dimensional (g+© l^)-module with basis
Vq and action h • Uq = ( ¡ jl, h}üQ for all e and x • Uq = 0 for all x g g+.
Set
Pi ■■= [a e Q, U{0}|ju, + a s A i <2+})
P2 == Q .\ P v

Lemma 3
For the fixed ¡ jl and P^, P2 as defined above,

(0 P^ is finite,
(ii) a G P , P^Q+\J{0}=>a+P
2 G P2 .
2.10 Bernstein-Gerfand-Gerfand Duality 187

Proof, (i) See Lemma 2.6.1.


(ii) If on the contrary a + /3 e then /x + o :H -)S e A l(2 + , and hence
also /i + a e A l Q ^ . This contradicts a e P2- ^

Consider the decomposition

1 1 ( 9 © U ( 9 j “ e © U(gj".
a^Pi ol^P2

Then from Lemma 3 we obtain

Lemma 4
is an ideal o f U(g+). □

For /1 e A 1 ¡2+ we provide U(g+) with the unique g+0 1^-moduIe struc­
ture satisfying

(jc + /z) • M= + a , h )u + XU
for all e /z e ]^, w e U (g+)“ ; a e (2+U{0}.

(This can be thought as the restriction to g+0 of the Verma module of g


with lowest weight ^t).
Let

IV= W ( pl) : = U ( g J / © U ( g j \
a^Pz

Now from the fact that 0^eP2^^^+^'' U(9+) we see that it is


stable under the action of g+0 which we have just described.

Lemma 5
W is a g+0 ^-module with the action x ' и = x и for all x e g+0 f) and
и e U(g+X ^here “ :ll(g+) ^ Wis the natural quotient map. □

P(ju) is defined to be the induced g-module

P ( p ) = P ( p ; A ) := и ( д ) 0 ц , ^ е ^ W( p ) .

Lemma 6
Let {iV/Xe / basis o f W{fx). Then P(/x) is a free left Vi(g_)-module with
basis {1 ® Furthermore, if for each i, w¿ is a weight vector o f weight %
188 Líe Algebras Admitting Triangular Decompositions

for 'i), then 7/ = M + ßi for some unique : P^, and in this case 1 w¿ is of
weight % in P(ix),

Proof By Corollary 2 of the PBW theorem U(g) is a free right U(g+0


module with any basis, say, {ujij^j of U(g_) as a basis. Thus P(/x) is the
linear span of the vectors {uj ® w^}j and these vectors are a IK-basis for
P(fi). Then the vectors {1 <S>w¡\¿ certainly generate Pifi) as a left U(g_)-
module. Moreover, if £¿^¿{1 w¡\ = 0 for some e U(g_), then writing
z, = LjCijUj, Cij e K, we obtain 0 = LijCijiuj ® w¿X whence each c^j = 0.
This proves the freeness. The statement about weights is obvious. □

Proposition 7
Pip) e F i f i M

Proof W - as a vector space. Since P^ is finite and, for all


a e ¡2^, dim U(g^_)" is finite, we see that dim W is finite. Thus, by Lemma 6,
Pip) is finitely generated as a g-module. Choosing a basis of W consisting of
weight vectors, we see from Lemma 6 that Pip) is a weight module and that
all of its weights lie in UaepjiM Л- a ) l Q ^ ( z К I Q ^ . Furthermore, since
U(g_) has finite-dimensional weight spaces, it follows from Lemma 6 that
Pip) does too. Thus Pip) ^ -^(A). □

Proposition 8
For every M e -^(Л), Hom^iPip), M) - Hom^(IK^, M). Explicitly the iso­
morphism is given by
Ф:Ф(Л)(г;о) =Л(1 0 1)
for all A e Нот ^iPip), M), and its inverse is given by linear extension of
4 ^ (/)(x ® y) =д: у -/(i^o)
for all f e Horn j(IK^, M).

Proof. Let us check that Ф(А) is an ^-module map from IK^ ^ M. For all
Ae
h ■Ф (^)(У о) = h A ( i ® T) = A { h ■(1 ® 1))
= A{h ® 1) =A{1 ® /г - I )
= A(1 ®
= Ф(А){({1,1г}оо) = Ф ( А ) ( к ■Vq),
where we have used the fact that 1 e U(g+)°.
Now consider First we should prove that it is well defined, namely that
X • у ■fivo) is independent of the choice of у e U(g+) with у y. But, if
2.10 B ern stein -G el’fand -G el’fand D u ality
1»9

also y ' ^ y , then y ' = y + u , where u e e„ei>2^(9+)“- However, V q ^


and / e Hom^iK^, M) - /(t;«) e U(g^)“ • fiv^) e = (0) f^r
all a Ei P2 since M ^ -^(A). Thus u ' fiv^) = 0. It is now clear that ^
exists and is linear.
Let z e g. Then

^ (/)(^ ■ ® ^)) = ■ ^ ( / ) ( ( ^ ) ® y)
= ( z x ) y - / ( V o ) = z - x - y -f{Vo)
=z • ®y ),

showing that ^ ( / ) ^ Homg(P(/i), M).


Finally, we have

< I > ( ^ ( / ) ) ( ^ o ) = 'P ( / ) ( l ® i ) = / ( ^ o )


and
^(<I>(^))(a;: ® y) = jc y • <I>(^)(i;o)
= X • y • A(1 <8> 1)
= .4 (;r y (l® I))
= A ( x ® y ),
showing that <I> and ^ are inverses of one another.

Proposition 9
P(li) is a projective module in the category ^(A).

Proof. Consider a diagram

Pitx)

N ^ M - 0
where all the modules and maps are in ^(A ). Then, in the notation of
Proposition 8, we have

IK.
/
f/ 4>(A)
/
N ^ M ^ 0
for which we can obviously find f e HomgOK , A/^), making the diagram
190 Lie Algebras Admitting Triangular Decompositions

commute. Then by Proposition 8, 'ir(f) ^ N). We want to


prove that g = A. But

g o ^ ir(/)(jc ® y ) = g { x - y -fiVo)) = x - y - gi f i vo) )

= X ■y • ^ { A) V q = 'P (< I> (^ ))(JC ® y)

= A { x ® y). □

Proposition 10
Every object of F-^(A) is the image o f a projective module of the form
0/L i P(A^) for some A, 0

Proof Let M be a module in Fif(A), and let be a finite generating


set consisting of weight vectors, say, m^ g From /¿: M^‘, 1 •->
as ]^-modules, we have ^(/¿) ^ Homg(F(A^), M). Then adding, we obtain a
surjective map

M.
i= l

But ef=iP(A^) is projective and lies in Fif(A). □

An important method for proving results in ^(A ) is the use of Verma


composition series (VCS). The definition and basic properties appear in
Section 2.6. The next few results deal with Verma composition series for
certain modules in i^(A). If M has a VCS, then /(M ) denotes the length of
one (and hence any) such series of M.

Proposition 11

(i) P(/x) has a Verma composition series.


(ii) For a l l V 0 A i (2 + , [ P ( ijl): M(v)] = dim =
dim H om g(P(^), M(v))
(iii) [Pip): M(v)] = 0 unless < v < A.
jjl

Proof Let {wJ/Li be a basis for W •= Wip), which is chosen so that


e for some /3^ e P^ and so that > Pj => i > j. Let Wf^ :=
k = 1,2,... ,n, and define + i == (0). Thus W = W^z>W2
D • • • => Z) + i = (0). Observe that g+IT; d since e
and Pi + a > p^ for all a e A^_. Since is also a weight module,
is a (g+0 ]^)-module.
2.10 Bernstein-Gel’fand-GeFfand Duality 191

Inside P iti) = U(g) define

M ,:= U (g_) • (1 ®
= U(g) -(1 ® IT,).

Then

P{li) = M i d Mz

Since the elements {1 ® w^} from a basis of P()ti) as a U(g_)-module, it is


obvious that is a free U(g_) module of rank 1. Since 1 ® has
weight ßk + f i , we see that = M(ß^ + ¡jl) (Proposition 2.3.3). This
proves part (i). Since the multiplicities [P{^í)■.M(.v)] are invariants of
the first equality of part (ii) follows from [P()u,):M(v)] = card{*lj3;t + ¡i = v)
= dim For the second part of (ii) we have for e A | Ö+.

dimHomg(F()u.), M (v)) = dimHom^(lK^, Af(v))

= dim M (v)^ = K(^v — ¡x) = dim U(g+)*^ ^ ,

where K is the Kostant partition function (Section 2.5). We get zero here
unless V — IX a ^ Q+ U{0}. Then ix + a = v < h. v — ¡x = a ^ P^.
Thus dim U(g+)’^-'‘ = dim
(iii) In the proof of part (i) the only Verma factors M{v) occurring have
V = !x + ßi^iox some jß^ e P^. □

Proposition 12
For M ,N ^ T^(A), M ® N has a VCS if and only if M and N have VCSs.
Furthermore, if M, N and M ® N have VCSs, then for all weights fx,
[M ® N, Miix)] = [M, Miix)] + [N, M(ix)l

We prove Proposition 12 using the following lemma:

Lemma 13
Suppose that M has a VCS, and suppose that fx ^ M is a maximal weight in M.
Let 0 V M ' ^ , and set M ' ■= U(g) • v+. Then M ' = Mip ) and M / M ' has
a VCS with l i M / M ’) = /(M ) - 1.

Proof M' is a highest weight module with highest weight pair (ix,v+). Let
M = M q Z) • • • D M; = (0) be a VCS (of length /) with M¡/M¡+l =
M(ju,,). We prove the result by induction on /. If / = 1, M ' = M q = M q/M^,
and we are done. Suppose that / > 1. If then by the induction
assumption applied to Afj we obtain M' - Mip), and A /j/M ' has a VCS
192 Líe Algebras Admitting Triangular Decompositions

with KM^/M') < Then M / M ' has a VCS by adjoining the top term
M /M j and KM /M') = 1 + KM^/M') = 1 + l(M^) - 1 = /(M) - 1.
We are left with the case / > 1, u +^ Let M' ^ M/M^ be the
canonical map. Then ¥= 0, and by the maximality of fi we
obtain fi = ijlq and M /M^ - M(/i) = U(g) • D+. From the universal property
of Verma modules we obtain a homomorphism (p : M ( f i ) M ' with v+.

Thus the exact sequence

0 ^ Ml ^ M ^ M /M i == M(fi) 0

is split, and

M Ml 0 M'.

Now M' == M(fi) and M /M ' has a VCS with /(Mi) = /(M ) - 1. □

Proof of Proposition 12. If M = Mq D Ml D • • • D M^ = (0) and N = N q D


Vi D • • • D = (0) are VCSs, then M e V = M o 0 A /^ D M i0 V D •••
D M^ 0 V = Vo D Vi D • • • D = (0) is a VCS for M 0 V.
Conversely, let / be the length of a VCS for M 0 V. If / = 1, then
evidently M = (0) or N = (0) since Verma modules are indecomposable
(Proposition 2.2.2). Thus we may assume that / > 1. Let be a maximal
weight of M 0 V, and for definiteness suppose that M^ ¥= (0). Let 0 # i;+e
M^, and let M' = U(g) • u^. By the last lemma, M' == and (M e
N ) / M ' — M / M ' 0 N has a VCS of shorter length than M. By induction on /
we may assume that M /M ' and N have VCSs, whence also M and N.
Now the statement on multiplicities follows from the first part of the
proof. □

The next sequence of results aims at proving that every projective module
in F ^(A ) has a VCS.

Proposition 14
Every module in F-^ is a finite sum o f indecomposable modules, which are also
in F-^.

Proof. Any M e Fi^ has a finite set of generators consisting of weight


vectors. Let G •= {¿;i,. . . , be such a set of generators, and let v^ 0 M^‘.
Define o-(G) == E dim M^. We define o-(M) — min^ o-(G) as G
runs over all such sets of generators of M. We prove this result by induction
on o-(M).
If cr(M) = 1, then M = U(g)i; for some v e M^ with dim M^ = 1. If
M = Ml 0 M2 for some submodules, then M^ = M f 0 M^ => M f = (0), or
2.10 Bernstein-GeFfand-GeFfand Duality 193

= (0). Say, = (0). Then v e M f and M = Thus M is indecom­


posable.
Suppose that <r(M) > 1. If M is indecomposable, there is nothing to
prove. Suppose that M = 0 M2, where and M2 are nonzero submod­
ules. Let G be a generating set of weight vectors of M with a{G) = cr(M).
For each g e G write g = gj + g2, g, e M,.. Then G^ := {gjg ^ G) ^ {0}
generates M^, i = 1,2. From = Mf 0 for each ¡jl we see that
< cr(Gi) < <j(G) = a(M), Thus and M2 have finite decomposi­
tions into indécomposables, hence also M.
If M = where each Ij is indecomposable then clearly M e =>
Ij e F if for each ;. □

Proposition 15

(i) For all p. Ei K iQ ^ , P(fi) is a finite sum of indecomposable modules.


(ii) A module M in Fi^(A) is projective if and only if it is a direct
summand of 0/LiF(A^) for some A^,. . . , 0 1^*.
(hi) Every projective module in Fi^(A) has a VCS.

Proof, (i) Use Propositions 7 and 14.


(ii) Let M e F # be projective. By Proposition 10 we have Aj,. . . , A^ e 1^*
so that there exists an exact sequence

© F(A ^) ^ M - ^ 0 .

Since M is projective the sequence splits, and hence M is a direct summand


of 0F(A¿). Conversely, any direct summand of a projective module is
projective.
(hi) Use Propositions 11 and 12. □

We now move toward the connection between indecomposable projectives


in F ^ and irreducible modules in F if.

Proposition 16
Each projective indecomposable object in Fif(A) has exactly one maximal
proper submodule {in i^(A)).

Proof. The existence of maximal submodules is given by Lemma 2. Now let I


be finitely generated, projective, and indecomposable, and suppose that Mj
and M2 are distinct maximal proper submodules of /. Then M^ + M2 = I.
194 Líe Algebras Admitting Triangular Decompositions

Form the direct sum of and M2. Then we have

Mj 0 M2 0,

where m ^) = + .m2 and i/i is known to exist so that cpij/ = id^ by


the projectivity of I. Let p^: M^ 0 M2 ^ M^ be the natural projection, and
define f¿ = / -^ c I, Then /1 + /2 = id/. Indeed for jc e 7, (/j +
f2)(x) = (Pi + P2^° = (p° il/(x) = jc. Thus in particular /1 ° /2 = /2 ° /1
in the ring Endg(7).
Let 7 be a finite sum of weight spaces 7^ so that J generates 7 (since 7 is
in Fif). Fix 1 = 1 or 2. Since is a g-module map, //(7^) c 7^ and hence
/.(7) c 7. Now we make a slight variation on Fitting’s lemma [Ja2]. Since
dim 7 < 00, the descending chain

/ d / , ( / ) d / 2 ( / ) d •••

stabilizes at some A/j, so we have f^>U) = = • • • . Since / = U(g) •


J, we have /■^'(/) = U(g)/^^'(/) = U(g)_^^''^*(/) = = • • • . Fix any
weight fjb, and consider ■= Then by Fitting’s lemma [Ja2]

where

s Z i n == n
k= l

and

« r ( 0 ) == U = 0}.
/t=i

Now by what we just proved = g^'C/**) = and hence


= /^^‘( /) which is a submodule.
We also see that if x = Ex^ e ®g~”(0), then for each nonzero x^ there
exists ^ with /^ '‘Xj^ = g^*‘x^ = 0, and hence for k ■= max{A:^}, /¿*x =
0. Thus, in the obvious notation, x e / ~ “(0). Conversely, x = Ex^ e
/ r “(0) x^ G g ; “(0) for all M. Thus / " “(O) = e g ; “(0).
In summary then,

/= e { g Z i n ® 8Z^(0))
At
= © /r(0 ).
2.10 Bernstein-GeFfand-Gerfand Duality 195

This decomposes I into two submodules. Since I is indecomposable, one of


them is (0). But /;.(/) c M, # /, and hence f ^ ‘i n * 1. Thus = (0).
Set N = max{iV,, Then finally, since /, and / , commute, we have the
contradiction

id, = id2^ = ( / i + / 2 f ” = 0.

This proves the uniqueness. □

Proposition 17
Let L be an irreducible object in -^(A). Then, up to isomorphism, there is a
unique finitely generated projective indecomposable module in i f that has L as a
quotient.

Proof (Existence): Since L is irreducible it lies in F-f. By Proposition 10, L


is the homomorphic image of a finitely generated projective module P of -f.
By Proposition 14, P = ®/Li/y where each /y is indecomposable. Since L is
irreducible, L is the homomorphic image of one of the Ifs.
(Uniqueness): Suppose that and I2 are finitely generated projective inde­
composable modules and that 7t¿: L are surjective homomorphisms.
From

// / TT2

A ^1 0

we obtain / , making 771° / = 772- Now ker 77 ^ is the unique maximal proper
submodule of (by Proposition 15). It follows that / is surjective, for if not,
by Lemma 2, / ( / 2) lies in a maximal proper submodule of I, and hence lies
in ker TTp But then 772 = ttj ° / = 0, which is not the case.
Since 0 ker / ^ /2 ^ /1 0 is exact and is projective, there exists a
homomorphism g: 12 splitting this sequence. Then 12 = k e r /e g /^ .
Since ¡2 is indecomposable, ker / = 0. Then I 2 - h- ^

Propositions 16 and 17 establish a 1-1 correspondence between irre­


ducible objects in -f(K) and finitely generated indecomposable projective
modules in -f(K). However, we already know (Proposition 1) the irreducible
objects in -f(hj), namely the modules L(p), e A 1 (2+.

Let / 1 < A . We define K / jl) to be the unique (up to isomorphism) finitely


generated indecomposable projective module in i f ( A ) having L(/i) as an
irreducible quotient.
196 Lie Algebras Admitting Triangular Decompositions

Every finitely generated indecomposable projective module in -^(A) has


the form /(/x), /jl < A,

Proposition 18
Let M G. t^(A). Then for all ¡jl ^ A iQ+,

[M:L{ii)] = dimHomg(/()Lx),M).

In particular dim Horng(/()Lx), M) = 0 if = (0).

Proof We begin with the simple observation that if I, M q, are modules in


a category, I is projective, and M q Z) M^, then

H om (/,M o/M i) = H o m (/,M o )/H o m (/,M i).

Indeed every map / : / ^ Mq induces another, namely / : / -> M q/M^ and


/ = 0 iff / ( / ) c Mj. Thus there is an injective map

A:Hom(/, M o)/H om (/, M^) H om (/, M q/M^).

Furthermore, if M q/M^, then since / is projective, there is a map g:


I -> M q for which g = g, thus proving that A is surjective.
It follows th at dim H o m (/, M q) = dim H o m (/, M q/ M ^ ) +
dim Hom(/, Mf),
Extending this by induction, we see that if M = Mq 3 z> • • d M^,
then

( 1) d im H o m (/,M )= E dim Hom (/, + dim Hom (/,M ;t)-


/=1

In our present case let M = Mq Z) Z) • • • d Mi^ = (0) be a local com­


position series at fi (Section 2.6). For each i either Mi_^/Mi —
¡Hi > f i o r M f ^ _ ^ / M i has no weights r > /x. In the former case

X VÍ = (0) if A AX,,
(2)
2.10 Berastein-Gerfand-Gerfand Duality 197

Since

1. I(fji) has only one maximal proper submodule, call it Rifi),


2. = Ufi),
3. Endg(L(/x)) - K.

In the latter case let us show that Н отд(/(д), = (0). Let


/ e Н отд(/(д), Since Li/i) is a quotient of I(fi) and
{M i-i/MiY = (0), we conclude that kei f Rifi), = I( ijl)/R(/ jl)). Now
Lemma 2 and Proposition 16 combine to show that / = 0.
Since [M: L(/x)] is the number of times L(/i) occurs as a quotient in a
local composition series at fi, Proposition 18 is proved. □

Proposition 19
Lei Д e Л 1 2+. Then

® (finite sum),

where
= dimHomg(P(jLi),L(i/)).

We have

m j^v) = 0 fi < v < A,

Proof. By Proposition 14, P(jLt) is a finite sum of indécomposables. Of course


each of these is finitely generated since P(/i) is. As we have seen, every
finitely generated projective indecomposable module is an I(v), i; < Л. Thus

for some finite set of p and integers m jiv). Furthermore

dim Horn g( P ( Д ), L(iü)) = dimHomg( ф mJ^p)I{p),L((o)^

= Lm^i/)dimHomg(/(i/),L(û>))

from 2. But from Proposition 8, dim Н о т д ( Р ( д ) , L((o)) =


dimHom^CK^, L((o)) = dim L{o)Y. Thus mjfii) = 1 and mjfo)) = 0 unless
(i < (o < A. □
198 Lie Algebras Admitting Triangular Decompositions

Theorem 20 (BGG duality)


For all ii,v Ei A i Q+,

[M{n.)-.L{v)] = [ / ( v) : M ( m)]

in the category i^(A).

Proof. From Proposition 18, it will suffice to show that

[ l( v ) : M { fi ) ] = dim Horn g(/(i/),A i(iu,)).

Let /Ji,v E A i Q^. Using Proposition 19, write

(3) P{v) = © m^{(o)I{(o),m^{a)) E I^.

We have

(4) [ P ( i') : M ( m)] = 'LrnXo>)[l((o) : M(v)]

by Proposition 12.
From (3),

Homg(P(v),M(M)) = Em,(a>)Hom3(/(a>),M(»^))

so from part (ii) of Proposition 11,

(5) [P(v): M {fi )] = E "i.(i^ )d im H o m g (/(a )),M (i/)).

As ¡jL,v vary we may consider (4) and (5) as systems of linear equations
with the same matrix of coefficients (m^(ci))). Since mj^v) = 1 for all v and
m^((o) = 0 unless i/ < co < A, we see that the matrix (m^(co)) is unipotent
relative to any ordering of A i Q ^ that respects < .
In particular the matrix is invertible, and we obtain at once that for all
(o,v E A i Q^,

dimHomg(/(io) = [l((o):M(v)]

as required. □

Corollary
For any VCS o f Kfi),

/ ( m) =/o=>/l=5 ••• =>/„-!=>/„ = (0),


2.10 Bernstein-Gerfand-Gel’fand Duality 199

we have

Io/I,^M (ß),

Ij/Ij+i = M(vj), vj> fi, j = 1 , 2 , . . . , n - 1.

Proof, [/(/x): Miß)] = [Miß) : Liß)] = 1. Thus in any VCS for I(ß) exactly
one subquotient is isomorphic to Miß). From

li ß )

Miß ) Liß ) 0

we have a map ilr. l i ß ) M iß ) and iJ/Hiß)'^) ¥= (0). Thus iffHiß)) contains


a generator of Miß), so if/ is surjective. Consider /j. Since it lies in the
unique maximal proper submodule of l i ß ) (i.e., /j c ker tt), we have if/Hi)
c Miß). Thus if/Ui) c N iß ) and I q/ I i -* M i ß ) / N i ß ) = Liß), which shows
that /o //i = Miß). □

We may interpret the duality theorem in the way suggested in the


introduction to this section. Set

== Liv)]
and

'■= [ ^ i f J - ) '■Liv)] for all V e A i Ö+.

Consider the matrices C = D = We have

c^„- =[l i ß) - H^) ]= E [/(cu):M (ir>)][M (a>):L(i^)]


/1, V<(0

= E [Mico):Liß)][Mi<o):Liv)]
fJi, V<(x>

fJL,V<0)

Thus C = and in particular C is symmetric and positive semidefinite.


200 Lie Algebras Admitting Triangular Decompositions

2.11 EMBEDDINGS OF VERMA MODULES

Throughout this section g is a Lie algebra with a fixed regular triangular


decomposition and & is the category Let
A e ]^* and consider the Verma module M(A). It is a very interesting
problem (going back to Verma’s original work on these modules) to decide
when another Verma module M(/¿) appears as a submodule of M(A). In the
case of an invariant Kac-Moody algebra g, we will obtain an exact answer in
Section 6.7 to this problem for a large range of values of A (all A when dim g
is finite).
If M ( ijl) is a submodule of M(A), then clearly Homg(M(/¿), M(A)) ¥=(0).
Conversely, if e Horn ^ ( M ( ijlX (p ^ 0, then ( p ( M ( f i ) ) - M(/x) by
Proposition 2.3.3, and hence M(/x) is in effect a submodule of M(A). Thus we
are looking for conditions under which Homg(M(/¿), M(A)) ¥= (0). The main
result of the section is

Theorem 1 (Ku [Kuly Neidhardt [Ne])


Let f)*. Then

[M(A): L(/x)] ^ 0 <=> Homg(M (/x),M (A)) ^ 0.

The hard direction is “ => .” Our proof is taken from Neidhardt [Ne] and
relies on the BGG duality theorem (Proposition 2.10.20).

Proposition 2
Let )LL e and let N be a ^-module in category & that has no weights A with
A > jLt. Then Extg(M(^i), N ) = (0).

Proof. Consider any extension

(1) 0^ ^ P ^ M( f i ) - ^ 0 .

Let P^ be chosen so that <pv^ is a highest weight vector of M{p).


Since neither M{¡i) nor N has weights Awith X > fjL, (pv^ is 2lhighest weight
vector of P. Since M(¡i) is a Verma module, there is a unique homomor­
phism (/í : M( jjl) -> P with {¡/((pu^) = u^. Clearly <p o = id^^^^, and hence
I d splits. □

Proposition 3
Extg(M(/x), L(A)) ^ (0) ^ [M(A): L(/i)] # 0.

Proof The hypothesis together with Proposition 2 shows that A > /x. Thus
M{iD and L(A) lie in the category ^(A). Consider the indecomposable
2.11 Embeddings of Yerma Modules 201

projective g-module /(/x) in ^(A). We are going to prove that [ K i j l) : M(A)]


> 0 and hence by BGG duality that [M(A): L(/x)] > 0.
Let

/ ( m) = / o ^ / i => ••• 3 4 _ i D/„ = (0)

be a VCS for /(/x). We have 4 / 4 +1 - for some weights /x^, where


/xq = fi. (See Corollary of the BGG theorem, Proposition 2.10.2). Let

0 -> L (A ) 0

be a nontrivial extension. From the natural map tt : I(fi) = 4 ^ M(/x) with


kernel 4 from the projectivity of /(/x), we have

/( m)
T_T/ /
/
0 ^ L(A) P ^ M (fi) 0

with (p o TT = 77.
Now (p o 77(4 ) = '^■(4 ) = (0), so ^ ( 4 ) ^ ker 9 = L(A). In fact 77(4 ) =
L(A), for otherwise ^ ( 4 ) = (0). Therefore 77 induces

¥: 7( m) / / i(= V A = M (n )) ^ P,

and then splits the sequence contrary to it being nontrivial.


Since L ( A ) = 7 7 ( 4 ) ^ D • ‘ • D 7 7 ( 4 - 1 ) ^ ^ ^ ( 4 ) ^ (OX there is a A:
such that ^ ( 4 ) = L ( \ \ 'ñr(4+i) = (0) and hence a surjective map M(jjLf^) ^
4 /4 + 1 ^ ^(>^)- This can only mean that /x^ = A, and hence [/(/x): M(A)] >
0, as we wanted. □

Proposition 4
Let M E: 0 be a highest weight module o f highest weight A. I f Extg(M(jLx), M )
(0), then Á. > jjb, and for some v e ij* with X > v > 11 we have
[M(X):L(p)] > 0 and [M(v): L(fji)] > 0.

Proof We have A > ju by Proposition 1. By assumption there is an exact


sequence

0 -^N M L(X)^0,

where N is some maximal submodule of M. Every weight cp of N satisfies


X > <p.
202 Lie Algebras Admitting Triangular Decompositions

Using Corollary 1 of Proposition 1.12.4 we have Extg(M(iU.), L(A)) # (0),


or Extg(M()u.), N ) ¥= (0). In the first case we obtain [M(A): L(^)] > 0 by
Proposition 3, and hence we obtain our desired result with v = k.
Suppose that Extg(M(/i), N ) # (0). Let

( 2) N = N ^ z> N ,^ ••• 3AT„_i DA^„ = (0)

be a local composition series for N dX yi. By Corollary 2 of Proposition 1.12.4,


Extg(M(^t), ¥= (0) for some ;. Using Proposition 2 = L{v)
for some v. Then by Proposition 3, [M{v) : L()it)] > 0. It remains to see that
[M(A): L(z/)] > 0. But is a subquotient of M(A) (since M is a quotient of
M(A)) and[A/^:L(i/)] > 0. The result follows from the Corollary 1 of Proposi­
tion 2.6.13. □

Proof o f Theorem 1, (<=) By assumption there is a nontrivial g-module


homomorphism <p: M ( ) jjl Af(A). By Proposition 2.3.3, (piMiyu)) —
so we can consider M(^t) as a submodule of M(A). Refining the series
M(A) D M (/i) D M m) to a local composition series at m gives [M(A):
L ( ijl)] > 0 at once.
(=>) The proof is for all pairs A, m,(A > m)? by induction on h t(\ — ¡i). If
h t{\ - m) = 0, then \ = ¡JL, and the result is trivial.
Suppose that A > m- Lot MA) be the maximal submodule of M(A). Since
[M(A): L( m)] > 0 and A > Mj we see by refining M(A) d MA) d (0) that

(3) [N (k) : >0.

Let

(0) = A^o c ATj c iVj c

be a HWS of MA) (Proposition 2.6.10). Each is a highest weight


module, say, of highest weight i = 1 ,2 ,... . Evidently

(4) A > )u., for all i.

From : L(/a,)] > 0 we obtain

(5) [M A ) : L (n ,)] > 0, [M(A) : L ( m,)] > 0.

By part (iii) of HWS of Section 2.6, there is a A: such that = N'^.


It follows at once that [M A )/M ^LC^u.)] = 0, and hence from (3), [M •
2.11 Embeddings of Yerma Modules 203

Lifi)] > 0. Thus

for some j < k , [Nj/N j _ i : L(ju,)] > 0,


(6)
and [M(/ty) : L(/ii)] > 0 My ^

Case 1. fjL Ф fjLj. Then к > ¡Xj> ji. Using the induction assumption,

Ф (0), n o m ^ {M ifij),M (X )) Ф (0),

and so combining, we obtain Homg(M(/x), M(A)) Ф (0) as required.


Case 2. 11 = fij. Then Nj/Nj_^ is a nonzero homomorphic image of
Mill), and clearly Нотд(М (д), ^ (0). Let 0 < / < ; be
chosen minimal so that Нот J^Mifi), N ik )/N¿) Ф (0). If i = 0, then
Homg(M(/i), MA)) Ф (0). Consequently Нот^iM ifi), M(A)) Ф (0), and
we are done. If i > 0, then from the exact sequence

0^N,/N,_, ^Nik)/N,_, -^Nik)/N,-^0

we have the exact sequence (Proposition 1.12.4),

Since the first term is (0) and the second is not, neither is the third; that
is, Extg(Ai(/i), Ni/N i_i) ^ (0). Applying Proposition 4, there is a weight
V e ]^* with X > ( i j > V > f x s o that

: L (v)] > 0 and [ M { p ) : L (/i)] > 0.

The induction assumption gives us Homg(A/(v), ¥= (0) and


Homg(Ai(ju,), M{v)) ¥= (0). Furthermore from (5) and the induction
assumption Homg(Af(ju,,), Af(A))) # (0). Combining all these, we obtain
the desired result. □
2(M Lie Algebras Admitting Triangular Decompositions

2 .12 D EC O M PO SITIO N OF M O D U LES IN CATEGORY Û

According to Section 2.11 the embedding of Verma modules Mijx) ^ M(A)


is possible if and only if [M(A): Lifi)] > 0. A central issue is then when is
[M iX ):U fi)] > 0? This question can only be answered adequately by spe­
cializing the class of Lie algebras g under discussion. This is precisely what
we do in Chapter 6, where g is assumed to be an invariant Kac-Moody
algebra. However, the relation on 1^* defined by A /i if and only if
[M (A):L( m>] > 0 does lead to a useful equivalence relation on certain
subsets (characteristic subsets) of 1^*. The main result of this section shows
how this notion leads to a natural decomposition of any module M lying in Û
into submodules, each of which lies in a subcategory of ^ governed by the
equivalence relation.
The material is derived from [DGK] with some simplifications from
Section 2.11. These results are not used until Section 6.8 and are not
necessary for any other part of this book.
We will assume throughout that g is a Lie algebra with fixed regular
triangular decomposition <5^= (^, g+, Ô+? ^ind that Q is finitely gener­
ated. A nonempty subset 5 of 1^* will be called characteristic if it has the
following property

Ae 5 and [M(A) : L(/¿)] > 0 => e 5,

or equivalently,

A e5 = > ch M (A )= Y. c^ch L{/jl)


fJL^S

for some family in N. The following are examples of characteristic


sets:

1. Í)*.
2. P := {A eri< A ,a,-^ > Z for all i G J}.
3. A 1 (2+ for A e

For A, we write Á > ¡jl if and only if [M(A): Li/i)] > 0. Let 5 be a
characteristic subset of 1^*. We define an equivalence relation = '^s ^
as the transitive closure of in 5; that is, A /x <=►, there exists
Aq, Aj, . . . , A„ e 5 such that Aq = A, A^ = ¡jl, and either A¿_j -< A¿ or A¿
A,_i for all i < i < n. For y e 5 we will let f denote its equivalence class in
S. Each equivalence class of S is itself a characteristic set.
Let 5 be a characteristic set. A module M in <^(g, 5^) is said to be of
type 5 if
[M :L (A )] > 0 => A e 5.
2.12 Decom position o f M odules in Category & 205

Equivalently M is of type A if

c h M = £ c ;,c h L ( A ) .
Ae5

Let S c ]^* be a characteristic set. We define a category = ^ 5(9,


as follows:

1. The objects of 0^ are the objects M of ^ of type 5


2. The morphisms of 0^ are all the g-module homomorphisms between
the objects of 0^,

It is clear that direct sums, submodules, and quotient modules (see


Proposition 2.6.13) of modules in 0^ are still in 0^.
In the sequel we will generally have a single characteristic set 5 in mind
(this will he K ' in Section 6.8). The primary interest will be in the equiva­
lence classes of • For A e 5 its equivalence class A defines a subcategory
which for simplicity we sometimes denote by 0^ (so ^ if A /a).
The following is obvious:

Proposition 1

Let 5 be a characteristic set, and let A e 5. Then

(i) M(A), L(A) e


(ii) every highest weight module with highest weight A is in 0^\
(iii) E ^ 0 x ^ 0^.

Proposition 2
Let S be a characteristic set, and let S2 ^ S be subsets that are unions of
equivalence classes. Suppose that 5^ Pi 52 = 0 . Then for E ^ 0^ , F ^ 0^^ we
have

(i) E n F = (0X
(ii) H om g(£,F) = (0).

Proof (0 Evidently [E n F : L(p)] > 0 => [E : L(fi)] > 0 and [F : L(p)] > 0.
Hence fi ^ Si n S2, contradiction.
(ii) If e Homg(E, F), then <p(E) e 0^^. Hence cp(E) = (0) by part (i).

In the sequel the functor is Extg will be denoted by Ext. The reader is
referred to Section 1.11 for a discussion of Ext.
206 Lie Algebras Admitting Triangular Decompositions

Proposition 3
Let S be a characteristic set, and let A and fx be in S. Suppose that E and F are
highest weight modules in &(,%, o f highest weights A and fi, respectively. If
A and ¡X are not equivalent, then

E x t(F ,£ ) = (0).

Proof. Consider first the case F = M(fx). By Proposition 2.11.4, if


Ext{M(.fx), E) ¥= (0), then \ > fx and there is a ^ e such that
[M(A): L{v)] > 0 and [M{v): Lip,)] > 0. Thus Á ~ v ~ p , contradiction.
Thus ExtiM ip), E) = (0).
Now in the general case we have an exact sequence of g-modules

0 K ^ M (p) - ^ F ^ O

and hence an exact sequence (Proposition 1.12.4')

H om g(/¡:,£) ^ E x t ( F , £ ) ^ Ext(M (jit),£ ) .

But K is of type ¡X, and E is of type Á. Thus by Proposition 2, Homg(£, E)


= (0). Since ExtiM ip), E) = (0) by what we have just proved, we obtain
Ext(F, E) = (0). □

Corollary
Suppose that Á , p ^ S , are not equivalent. Assume E g F g are
modules each of which has a finite highest weight series iSection 2.6). Then
ExtiF, E) = (0).

Proof. Let He ) and /(F) denote the lengths of the given HWS of E and F.
If HE) + /(F) < 2, then the result is immediate from Proposition 3. In
general we reason by induction on /(F ) + /(F). For example, if

(0) c F i c c £ = F

and n > 2, then from

0 ^ F „ _ i - ^ F ^ £ /£ „ _ ! - > 0

we have the exact sequence

E x t(F ,F „_ i) ^ E x t(F ,F ) ^ E x t(F ,F /F „ _ i),


2.12 Decomposition of Modules in Category 0 207

and the outside terms are (0) by the induction hypothesis. Thus also
Ext(F, = (0). The case for /(F) > 2 is similar. □

Theorem 4
Let S be a characteristic set, and let S denote its set o f equivalence classes
under . Let M be a %-module in 0^, Then M decomposes as a sum

M=

where Furthermore is the unique maximal submodule o fM that


lies in In particular this decomposition o f M is unique.

Proof Let

(0) = Mo c Ml c M2 c • • ‘

be a HWS for M (Proposition 2.6.10). Thus each nontrivial module M„/M„_i


is a highest weight module, say, of highest weight A„, n = 1 ,2 ,... . We are
going to construct for each equivalence class ft e 5 a module
having a HWS

(0) = M o o C M i o C

such that

(1) ® M „n,

if e ft,
( 2) if A„ ^ ft.
i(0)

We will then have M = © ^ ^ required.


We construct all the M„ ft g 5, at once by induction on n. For n = 1
we set Ml = Ml, Mi ^ = (0) if ft Ai.
Suppose that n > 1 and by induction we have constructed M„_i
ft G 5, satisfying (1) and (2) for n — 1. We have M„/M„_i ~\ . Thus
^ we
define

(3) if ft ^ A„.

To define M„ we consider the exact sequence

(4) 0 ^ M„_i/M„_i,;^^ 0.
208 Lie Algebras Admitting Triangular Decompositions

From the induction assumption

SO

Ext(M„/M„_i, = e Ext(M „/M „_„ = (0)

by the Corollary to Proposition 3. Thus (4) splits, and hence there is a


submodule E of such that

and

Define M„ ■= E. Since Af„_i ^ and are in , so also


£ G . We" check (1).

M„ = £ + M„_,

=£ + ® 0
^ n#A„ >

=£+ 0 M„ by(3).
n#A„

But £ n 0 ^ ^ A „ ^ n , ii = (0) since £ g Thus

M„ = £ ® 0 M„_n= 0

Condition (2) is obvious. This completes the induction step.


Define — U “ =iAf„ Clearly e and we have our required
decomposition of M.
As for the second part of the Theorem, let be the sum of all
submodules K oi M such that K e Then e and z) M^, But
^ n ^ n ' "" (0) Proposition 2, and hence = M^, □
Exercises 209

EXERCISES

2.1 Let H be an abelian group and tt : H G LiV) a representation of


//. Let e Hom(//, K^). We define

= {u ^ M \7 r(h ){u ) = (f>{h)v for all h ^ H } .

Show that is a submodule of V and prove the “group version’’ of


Proposition 2.1. Describe also the “group version” of Example 1 of 2.1
for the group H = {A ^ SL(n, K)|>1} is diagonal.
2.2 (a) Let (^ ,g + ,2 + ,< j)b e a triangular decomposition of g. Show
that any subalgebra § of g satisfying ^ and c § also
possesses a triangular decomposition.
(b) Let g¿, i = 1 ,..., r, be Lie algebras with triangular decomposi­
tions ^ g^+, o}). Prove that the direct product gj
X • • • X g^ of the Lie algebras g¿ also has a triangular decompo­
sition.
(c) Let (]^, g+, (2+, a) be a triangular decomposition of g and
let a be an ideal of g satisfying a n = (0), ga = a, g+<i a.
Prove that g /a possesses a triangular decomposition.
2.3 Let g = §I„(IK) be considered with the triangular decomposition of
Example 3 of 2.1, and let M = IK" be the g-module with the weight
space decomposition given in Example 1 of 2.1. Show by direct compu­
tation that these two gradings are compatible.
2.4 Classify (up to isomorphism) all three-dimensional Lie algebras admit­
ting a triangular decomposition.
2.5 [Be2] Let g be a Lie algebra with triangular decomposition
(Q+j G+j 9 ^ 9 be any automorphism of order 2 such
that i/f(g“) = g“" for all a e g .
The objective of this exercise is to obtain some information about
the Lie subalgebra § of all fixed points of í/í. Of course § = {jc +
il/(x)\x G. g}. Let / : 0 -» Z be a group homomorphism satisfying f(a )
e Z+ for all a e 0+. Set

9" = L 9“,
f ( a) = n

thereby obtaining a Z-grading on g for which g+= E„>o9", ft = 9^*


Let {e^li e K}, where each e g"', n- e Z+, be a set of genera­
tors of g+ as a Lie algebra. Define == ^ i E: K.
210 Lie Algebras Admitting Triangular Decompositions

Also define the subspaces and n = - 1 ,0 ,1 ,... inductively


by

9-1 = (0), 9 o = 5, 9„= E 9^


0<j : ^n

§_i = (0), = [x + il/(x)\x e 1^} = i),

§„ = linear spcin of and all products ... ,e,^],


where M; + • •• +n¡ < n .

Evidently 9 = U "=_i9„, § =>


(a) Prove by induction on N that

{x + tlf{x)\x G g^jm od §^_i

and

[«,V • • • ’ • • •’ = h v ••• ^N-V

(b) Prove that § = U and that {e¡\i g K} generates §.


(c) Let

gr(§)+=«e= 1§„/§„_!
be the graded Lie algebra associated with the filtration ^
c • • • of Prove that 9+— under the mapping x ^ x
+ il/(x)mod for all jc e g" , « = 1 ,2 ,... .
2.6 Show that every highest weight submodule of a Verma module is a
Verma module.
2.7 Let (a ) and <6>, a, ft e N, denote the ^^(W-modules of dimensions
a 1 and & + 1, respectively. Prove that

(a ) ® (b ) - (a b) ^ (a + b - 2)
e ••• e <k - fc|>.

2.8 Let B c 5L2(IK) be defined as in Proposition 2.4.8. Let = {X^\X


e B}. Prove that 5L2(IK) = B^B U J (Birkhoff decomposi­
tion) and that this union is disjoint.
Exercises 211

2.9 Let M be any finite-dimensional § l2(lK)-module admitting a weight


space decomposition M = where M"* = {x e M\h • x = ¿zx}.
Prove that for all a ^ Z,
(a) dim = dim M “"*,
(b) the sequence {dim is monotonically increasing until a = 0
and monotonically decreasing after 0.
2.10 Let Af(A) be a Verma modules of the category T). Show that for
M(A) to be isomorphic to M(A)^, it is necessary and sufficient that
M(A) be irreducible.
The results of Exercises 2.11 and 2.12 are due to D.-N. Verma.
2.11 Let n be a finite dimensional Lie algebra graded by a free abelian
group Q — Z^ whose degrees lie in —Q+= —(N^\(0}). Let

ch U (n )= n (l-e (-a ))' L K{<p)e{-<p)


cp^Q+UiO]

be the character of U(n) (see Proposition 2.5.3).


(a) Prove that the coefficients K(<p) of ch U(n) are bounded above by
the values f(cp) of a polynomial function / on with coefficients
in Z.
(b) Show that for any real number c > 1 and any <pQe Z^, there
exists (p ^ Z^ with K((Pq + <p) < cK{<p).
2.12 Let g be a finite-dimensional Lie algebra with triangular decomposi­
tion Q+, (t ), and let M(A) be a Verma module for g.
(a) Show that any two Verma submodules M ( / jl) and M ( p ) of M(A)
must have a nontrivial intersection. (There is a counting argu­
ment using Exercise 2.11.)
(b) Prove that M(A) is Noetherian (see Exercise 1.14).
In the remaining exercises we will assume that every Verma
module has a (finite) composition series. This is not in general
true (see Exercise 2.13). However, it is true for split semisimple
finite-dimensional Lie algebras (see Exercise 4.7).
(c) Prove that every Verma module M(A) has a unique irreducible
highest weight submodule M (v) (v depending on A) and that it is
a submodule of every highest weight submodule of M(A).
(d) Prove that for all A, /x e 1^*, dimHomu(g)(M(ju), M(A)) < 1.
(Hint: Let M and M' be the image of two distinct embeddings of
M ( ijl) in M(A). Then M(v) of (c) is in both. Write the generators
of M(v), M, M' in terms of the generator í;+ of M(A), and use
the fact that U(g_) is a domain.)
212 Lie Algebras Admitting Triangular Decompositions

2.13 Give an example of a triangular pair (g, (with dim g < oo) and a
module M e ^ ( g , for which M does not admit a (finite) composi­
tion series.
2.14 Let be an abelian Lie algebra and V an l^-module admitting a weight
space decomposition. Let V* be the dual space of V viewed as a
]^-module.
(a) If / e for some A e ]^*, show that f(V ^ ) = (0) whenever
JJL

(b) (K^)* —(K*)"'^ as K-spaces for all A e 1^*.


2.15 Let M e <^(g, ^ ) . Let N+ and N_ be l^-modules admitting weight
space decompositions:

© N i , N _ = © N Z \

where all weight spaces are finite dimensional. Let

if/: N _

be a vector space isomorphism satisfying

{¡/(h • n) = —h • il/(n)

for a\\ h and n e N_. Show that there exists a unique monomor­
phism

satisfying

f{ m ® n) = f(>lf{n))(m )

for all / G Uom^(N+, M “^), n e N_, m ^ M. Moreover for all fi


the above linear map composed with the isomorphism of Exercise
2.14(b) determines an isomorphism

~ :Hom(A^+,M‘^)'' s

2.16 [BM] Let Q = ®/=iZa, be a free abelian group. Set Q+'= (Ei=iNa,)
\ {0}, and consider the power series ring Q[0T Q+\ which is the set of
all formal sums e Q. Suppose that (s^, ^2,. - is
a sequence (possibly finite) of distinct elements of Q+ arranged in
order of increasing height, and suppose that for each 5^ there is
associated a sign e(5,) = ± 1. For each sequence (n) = (n^, « 2?• •) of
Exercises 213

nonnegative integers, all but a finite number of the n, being 0 and for
all A e (2+ U{0} define

X(n) = 2 ( ( n ) ) = 2 « , ,

(X n,)!
B {n) = B {(n )) -

sgn(n) = sgn((«)) = r ie ( s ,) " ',


S(A) = {(n)|Xn,i,- = A}.

For A, a e write X\a if a = rk for some r g Z+. In this case


A /a := 1/r.
Let m: ¡2+ ^ N, a be any function, and suppose that for
some choice of s = ^2, . . . ) as above:

n ( 1 - € ( « ) ) '" “ = 1 - £ £ ( 5 ,)e(s,).
i

(This situation occurs for several types of graded Lie algebras n = ®n “,


where = dim n “ (see Proposition 2.5.5 and the Corollary to Theo­
rem 6.4.1).) The objective is to give a closed formula for the integers
m „.
(a) Let X := 1 - E,e(s,)e(s,). Show that

e{ka )
-log X = £ m,
a, k

(b) Apply the operator

I d
E:= £ e ( a , )
i= l
de{ai)

to obtain

£(X )
= XC(A)e(A),

where C(A) — E„|AW„/ii(a).


(c) By computing -£ (X )/X in a different way, show that

C(A) = X e(i,)/ii(5,) £ sgn(n)B (n).


( n ) s 5 ( A - i ,)
214 Lie Algebras Admitting Triangular Decompositions

(d) Show that the right-hand side in (c) can be rewritten to give

C (A )=Ä i(A ) E sgn(n) •


(«)e5(A )

(e) Use Möbius inversion (see Bourbaki, Ch. 1-3, App. Ch. 2) to
prove that

"ior E (ri£i(i/) ) pr ,
Ala ^ “ (n)e5(A ) 1 1 '^/-

2.17 (a) Show that the homomorphic image of a residually nilpotent Lie
algebra need not be residually nilpotent.
(b) Show that the sum of a finite number residually nilpotent ideals
of a Lie algebra is residually nilpotent.
(c) Show that the sum of nilpotent ideals of a Lie algebra need not
be residually nilpotent.
2.18 Let (g, be a Lie algebra with triangular decomposition. Show that
g/rad(g) is radical free.
2.19 Prove that the Shapovolov form for §I2(1K) is given by

^ a( / ” * *^+) = n (('^ + P )(^) “ for n e N.


k= l

(See Section 2.8 Example 3).


2.20 Determine all the characteristic sets 5 c 1^* and their equivalence
classes in the case of §I2(1K).
2.21 Prove that for the Lie algebra §I2(IK) the decomposition

M = ® Mj,
Ae5
of Theorem 2.12.4 is a decomposition into eigenspaces of the Casimir
operator.
2.22 Show that for t e C , a b one has M(a) ^ M(b) as ^12(C)
modules if and only if i) e N and a = —b — 2. (See Section 2.4 for
notation.)
2.23 Let I = ]^ 0 h be a Lie algebra consisting of an abelian subalgebra
and an ideal b that decomposes as a l^-module under the adjoint
representation into weight spaces b = where G +c 1^* is a
free additive semigroup.
Exercises 215

Let t)' be a copy of b as a Lie algebra, and let o-: b b' be the
anti-isomorphism between them that associates to each i; e b, the
element - u , as seen in b'. We make b' an 1^-module by defining
h • cr(v) = —o-[h, u],
(a) Prove that f' == f) 0 b' has the structure of a Lie algebra compat­
ible with this action of on b'.
(b) Let g := b' 0 0 b, and extend a to an endomorphism of g by
defining (rli) = id ij,c7 ^ = l. Prove that g may be given the struc­
ture of a Lie algebra with subalgebras b and b' by defining

[b,b'] = [b',b] = (0),

and show that (1^, b, ¡2+? is a triangular decomposition of


9.
(c) Let g be any Lie algebra with triangular decomposition
0), ^+,Q+,o-)- Set 5g = g_0 0 g+ as a vector space with
multiplication defined as in g except that now [g+, g_] = (0) =
[ 9 - , 9+ ].
Prove that ^g is a Lie algebra with triangular decomposition
s ^ = (ij, Q+,Q+,(t ), We call sq the suspension of g. A Lie
algebra g, with triangular decomposition is
called suspended if [g+, g_] = 0.
(d) Let (g, be a suspended Lie algebra. Prove that

^ t (Q) = i^ r(9 ) = n (g ) = rad(g) = g++ g_.

(e) Let (g, <5^) be a suspended Lie algebra, and let M(A) be a Verma
module. Show that every subspace fx e A J, (2^., is a
submodule of M(A). Show that all the irreducibles modules in
<^(g, D are one dimensional.
2.24 Show that every derivation of the Vivasoro algebra is inner.
Chapter Three

Lattices and Finite Root


Systems
Symbols are to mathematics as wooden pieces are to chess
—L. Wittgenstein

Any Lie algebra g with triangular decomposition У has associated with it a


root system Д c 1^* so that g = There is no a priori reason to
expect Д to have a particularly rich structure of its own. But, as Example
1.1.8 shows, it can happen. In fact Д is remarkably interesting for all the
finite-dimensional semisimple Lie algebras over C. The root systems that
arise from such semisimple Lie algebras are simply called finite (reduced)
root systems. Over time it has been realized that finite root systems appear in
other contexts that often have no apparent Lie theoretical connection. Thus
there has developed an axiomatic approach to root systems (exemplified by
Bourbaki [B03]), free from Lie algebras, which allows them to stand on their
own. We adopt this approach here, prefacing it by a section on lattices and
concluding with sections that show how to construct Lie algebras out of
lattices, in certain cases.
The investigation of root systems was begun at the end of the 19th century
by W. Killing and E. Cartan in their study of complex semisimple Lie
algebras g. Both realized that to each nonzero root there is an associated
involution that stabilizes Д. Later on, influenced by the work of H. Weyl
[Wy], the group W generated by all these involutions (now called the Weyl
group) assumed a major role in understanding and describing the representa­
tion theory of g. These Weyl groups are finite groups generated by reflec­
tions (in euclidean space) and as such fall into H. S. M. Coxeter’s classifica­
tion of reflection groups [Cxi], A simplified version of the classification of
simple and semisimple Lie algebras making use of root systems was achieved
by B. L. van der Waerden [vdW].

216
3.1 Lattices 217

3.7 LATTICES

A geometric lattice is a pair (L ,( | *)) consisting of a free Z-module L of


finite rank together with a symmetric bilinear form ( I ) : L X L ^ Z.

Normally the bilinear form is understood from the context, and we simply
speak of geometric lattices L. By assumption, L has a basis Cj,. . . , over Z:
L = Z^i 0 • • • e Ze„.
If we extend the scalars to IKby forming V - = K ^ i L and extend the bilinear
form to a K-valued bilinear form on V in the only way possible, then V
becomes a quadratic space by considering the quadratic form associated to
( I ) (see below). It is useful to be able to view a geometric lattice as a
subgroup of a real or rational space (take IK = R or IK = Q above). The
spaces R Q ®jL are called the realization and rationalization of
L, respectively. Notice that L is a discrete subgroup of R
The principal example of a geometric lattice is
0 • • • 0 Zs„ in R" = R^i 0 • • • 0 Re„,
where {e^,...,e„} is the standard orthonormal basis of R" as a euclidean
space. This lattice is denoted by . The simplicity of this example belies the
subtlety of the subject.
Let L be a geometric lattice, and consider the mapping
q:x
of L into |Z . Then q is the quadratic form on L associated with ( I ). The
quantity (x|jc) is called the norm or square norm of x. Of course (-I*) can be
recovered from its quadratic form in the usual way:
(jcly) = q { x + y ) - q(,x) - q{y) .
The concept of isomorphism between geometric lattices is defined in the
obvious way: We say that (L,(-10) and ( L ,(•!•)') are isomorphic if there is
an isomorphism f : L - ^ L ' of groups such that (x|y) = (/(x )|/(y ))' for all
x,y ^ L. If q and q' are the corresponding quadratic forms, then this last
condition can be replaced by q(x) = q'(f(x)) for all x e L. The concept of
automorphism of a geometric lattice is defined in the obvious fashion. If L is
a geometric lattice, then its group of automorphisms is denoted by Aut(L).
We list below a series of invariants for geometric lattices—that is, quanti­
ties attached to a lattice that depend only on its isomorphism class.

1. Rank, Two free Z-modules are isomorphic if and only if they have the
same rank. In particular the rank, rank(L), of a geometric lattice L is an
invariant.
2. Determinant. Let e = be a Z-basis of the geometric lat­
tice (L ,(1 )), and consider the n X n integral matrix = (B^ ), where
218 Lattices and Finite Root Systems

Bij '= If e' = {e\,. . . , is another basis for L and A is the matrix
corresponding to the change of basis; e\ = Ae^ = then we have

(< l< ) = E E Aue,


a =l 1=1

Ij
k,l

where A^ is the transpose of A. Thus


B^,=A^B^A,
Since det(y4) e Z and A has an integral inverse, det(^4) = ±1, from which
d et(5 ,0 = d e t(B J.
The common value of this determinant, which is independent of the choice of
basis, is called the determinant of L and is denoted by det(L). It is an
invariant of L. The bilinear form (*|*) is nondegenerate or nonsingular if and
only if det(L) 0. In the case that det(L) 0, we say that L is nondegener­
ate or nonsingular. Otherwise, it is said to be singular.
3. Signature, Suppose that L is a nonsingular geometric lattice and that
IK c R. Then by extending (•!•) to V •= we may apply Sylvester’s
theorem (Jacobson [Ja2, Ch. 6] and write V as an orthogonal sum 1. V~ of
a positive definite space and negative definite space V~. The dimensions
and m " of these subspaces are invariants of (L,(*|*)) and are indepen­
dent of the choice of IK. Their difference sgn(L) = m'^ — m~ is called the
signature of L. L is positive definite (resp. negative definite) if sgn(L) =
rank(L) [resp. sgn(L) = -rank(L)]. It is indefinite if it is neither positive
definite nor negative definite.
4. Type, We say that the geometric lattice L is even or of
type II if the square norm of x is even for all x. Otherwise, L is said to be
odd or of type I.
If L is a free Z-module of rank n and L is a subgroup (Z-submodule) of
L, then L is a free Z-module of rank n' < n, [Ja2, Ch. 3]. It follows that if L
has a geometric structure (-I ), then L together with the restriction of ( I )
to L X L is itself a geometric lattice. We describe this situation by saying
that L is a (geometric) sublattice of L, We say that L is a fiill sublattice of
L in case rank(L') = rank(L).
If L is a full sublattice of L, then the index [L : L] of L in L; that is, the
number of cosets of L in L is finite. There is a simple way to compute this
index: Let ..., be a Z-basis for L and / i , . . . , a Z-basis for L , Then
the two bases are related by the relations matrix («¿y) defined through

fj 52 j = l,,,,,n.
¿=1
3.1 Lattices 219

It is a standard fact [Ja2, Ch. 3] that by suitably replacing the bases


{«1, and { /,,...,/„ } , we can arrange to have a relations matrix (a'y),

/ / = E a'ije'i,
¿=1
where (a' ,) is of the form

d'y

and the d, e Z^. satisfy ¿¡\dj if i < j. Then


L / L = Z /d ^ l e • • • ® Z/d„Z,

and therefore [L : L'] = d, • • • d„ = det(a'y). Since det(a'y) = |det(a,y)|,

(1) [ L :L '] = |det(a,y)|.

The reader will observe that in obtaining (1) we use only the facts that L', L
are free abelian groups of the same rank n and that L' c L. Now using the
bilinear form,

det(L') = dct{{f!\fj)) = det(d,(e;.|e;.)dy)

= dl ■■■ dl det((e'|c}))
= dl • • • det(L ).
Thus

(2) det(L') = [L :L ']^ d e t(L ).

5. Dual lattices. Let L be a free Z-module of rank n. We denote by L°


the dual module of L. By definition, L° is the Z-module of all homomor-
phisms of L into Z. Let e = [e^,. . . , be a basis of L, and define ef e L°
for all 1 < i < n by

(Cj, e f ) = Sij for all 1 < j < n.

Clearly e* — { e f , . . e f } is a basis of L°, which shows that L° is free of the


same rank as L. The basis e* is said to be dual to e.
Suppose now that L is a geometric lattice. For each jc e L we define
G L° by A:®(y) = (3:|y) for all y e L. The mapping x ^ of L into L°
220 Lattices and Finite Root Systems

is a homomorphism of modules and is injective if and only if L is nonsingu­


lar. If this is the case, we can identify L with a submodule of L°. Since L
and L° have the same rank, [L®: L] is finite. This index is called the index of
duality of L and can be computed as follows: By the definition of e* we have

y=i
and hence from (1)

(3) [L»:L] = |det(L)|.

Especially important is the case = L, which happens if and only if


|det(L)| = 1.

A geometric lattice L is said to be unimodular if det(L) = ± 1.

From the preceding remarks it is clear that

Proposition 1
The following conditions are equivalent:

(i) L is unimodular,
(ii) L° = L {under the above identification).
(iii) The mapping x •-> jc° is an isomorphism o f L onto V .

Unimodular or not, the bilinear form ( I ) of a nonsingular geometric


lattice L extends to a nonsingular bilinear form on the extension V =
of L. One has a canonical identification of V and V* through x x®, where
x°(y) = (x|y) just as before. Then L® can be identified with the following
subgroup of V:

= [v ^ V\{v\x) ^ Z f o r a ll x e L } .

6. Direct sums. If {L¿,{'\‘)¿), / = 1 ,..., r, are geometric lattices, then we


may form a new geometric lattice by considering the Z-module

L = © L,
¿=1
with symmetric bilinear form (• | ♦) defined by

E y, = E (^ily,), for all X,, y, e L,-.


¿= 1 i=l I /=1
3.1 Lattices 221

Evidently

rank(L) = Erank(L,),
sgn(L) = E sgn(L^),
det(L) = ndet(L^),
L is even if and only if each is even,
L is positive definite if and only if each is positive definite.

It is easy to identify L° with One identifies with the Z-valued


linear functionals of L that vanish on Then the linear mapping
x° of L L° takes into L® for each i. We have

[L “ :L ] = n m - L , ] .
¿= 1

The lattice (L,(*|*)) thus obtained is called the orthogonal sum of the
family (L^,(-I*)/), i = 1 ,..., r. To describe this situation, we write

L = JL L,.
/=1

Example 1 (Matrix representation of lattices)


Let B = be any n X n symmetric integral matrix. Let L be a free
Z-module with basis ..., and define (*1*) on L X L by {e^\ej) = B^j.
This makes L into geometric lattice that is nonsingular if and only if
det(5) # 0 and even if and only if B^^ is even for all i. The isomorphism
problem amounts to determining when symmetric integral matrices B' and B
are related by invertible integral matrices A so that B' = BA. This prob­
lem is far more difficult (and interesting) than first appears, and in one form
or another has a long history in number theory. For more on the arithmetic
and geometric properties of lattices see [Se2] and [CS].

Example 2 (Some unimodular type I lattices).


Let m, e and consider with a standard basis and
the bilinear form defined by the matrix

-1

-1
222 Lattices and Finite Root Systems

We denote the lattice together with this bilinear form by


This is an odd geometric lattice of rank m + «, signature m - n, and
determinant ( - D". For convenience we denote by and by I".
It turns out that every indefinite odd unimodular geometric lattice is of the
form ^ 0
If L is any geometric lattice and we let
L q == {x e L\(x\x) = 0m od2},
then it is obvious that L q is a subgroup of L and that 2L — {2x\x e L} c Lq.
Thus L q is an even full sublattice of L. If L is even, then of course L q = L.
If, however, L is odd, then [L: L q] = 2, and detiLg) = 4det(L). To see this,
let X q ^ L with (jcqIjco) odd. Then, if y e L with (y|y) odd, y - X q E: L q. So
we see that L = L q \ J ( L q + X q). The value of det(Lo) follows immediately
from (2).
The simplest example of this procedure for constructing even lattices is
the construction of the lattices.

Example 3 (Lattices of type D/)


Beginning with I/", we construct
L := {x: 0 I/‘l(jc|jc) = 0mod2}.
Here L is an even lattice of determinant 4. L is called the lattice of type Di.
The terminology comes from Lie theory and E. Cartan’s classification of
simple Lie groups and Lie algebras. In Section 3.7 we will construct the Lie
algebra of type Di directly out of this lattice.
If £ i,. . . , £/ is the standard orthonormal basis of I'l' and / > 2, then the /
elements

•“ ^/-1
form a basis of L. Each of these Z elements has norm 2. Norm 2 elements
play an important part in the theory of lattices, particularly as they appear in
relation to Lie theory. An element x = e l/" has norm 2 if and only if
X is of the form ±e- ± Sj with i ¥=j . There are 2(/^ - /) of these elements.
By definition they all lie in our lattice L.
The matrix ((«¿1«^)) determined by the basis above is
2 -1
-1 2 -1 0
0 -1 2 -1
B =
-1 2 -1 0
-1 2 -1 -1
0 -1 2 0
0 -1 0 2)
3.1 Lattices 223

Finally, let L° be the dual of L. Recall that [L°: L] = |det(L)| = 4. Thus


/L is isomorphic either to Z/4Z or to Z/2Z X Z/2Z. Let co = e
Q Then (ajo)) e Z for all / so that o> e L°. If / is odd, it is clear that
po) ^ L =>p = 0 (mod 4). Thus

lP ^ 1
77"" 4Z if / is odd.

On the other hand, if / is even, then lo) e L. It follows that if L °/L - Z/4Z,
then there exists y e L° such that 2y = o) mod L. But this is impossible
since ^ L°. Thus

L °/L - Z/2Z X Z/2Z if / is even.

Example 4 (Lattices of type Ai)


Fix / e Z+, and form the lattice I/+i with the standard orthonormal basis
£i,..., £/+i. Define

//+1
L := I X; n,e,- E « i = 0 •

L is a subgroup of and it is easy to see that defined by

a: := 6: - 6:

form a basis of L as a Z-module. Thus L is a sublattice of rank /. The


definition of L guarantees that L is even and, since it is a subgroup of
it is positive definite and is called the lattice of type Ai. Relative to the basis
«1, .. ., the matrix of L is

-1
-1 2 -1
B =
-1 2 -1
-1 2

Let z •= + *• • ^ I /+1? and define


224 Lattices and Finite Root Systems

One checks by direct calculation that (cojay) = from which we have


0 ••• 0

From the matrix of one easily sees that

det(y4/) = 2det(y4/_i) - det(y4/_2),

with d et(^ i) = 2, d e t(^ 2^ = Thus [L°: L] = det(^/) = / + 1 and the in­


dex of duality of is / + 1. Since (o^ has order / + 1 modulo L, we have

lV l -

Example 5 (The hyperbolic plane)


The simplest example of an indefinite even unimodular lattice is the
so-called (by E. Artin) hyperbolic plane U, Here U = Ze^ Ф Ze2, and the
matrix defining ( I ) is

(? :) ■
This lattice has rank 2, determinant - 1 , signature 0, and is even and
indefinite. It is a useful building block in the classification of indefinite
unimodular lattices [Se2].
Up to now we have not seen any examples of even unimodular positive
definite lattices. In fact such objects must have rank divisible by 8 (not
obvious). The next example gives a method for constructing one for each
rank of the form Sk.

Example 6 (The lattices


Fix a positive integer k, and set n = Ak. Let L = , and view this as a
subgroup of Q " with the standard inner product. Let L q = {x e L |( x lx ) =
0 mod 2}, let v ^ Q”, and let
r = := ZV + L q.

From Example 3 we see that is a group halfway between the lattice D,Ak
and its dual D®*.
Let r ' be the subgroup of Q comprised of elements
satisfying for all 1 < i, j < n.

(4) 2?,- s Z,
(5) Qi - Qj e Z,
n
( 6) ^ = 0 mod 2.
¿= 1
3.1 Lattices 225

We show that T = F :
(r c F): Clearly y e F (since n = 4A:). Suppose now that x = E7=i«,e, ^ L.
Then X e L q <=» T,nf = 0m od2 <=> = 0mod2. Thus

(7) L q = I E «¿e,; n, G ZiEn, = 0mod2


ki = 1

In particular L q c F , and hence F = l u + L q F'.


(F c F): Suppose that x = ^ F', Qi e Q. If some e Z, then
e Z for all i [because of (5)] and hence x ^ L q [because of (7) and (6)]. If
no Qi e Z, then there exist integers ..., such that

X = $21(1 + 2m^)Si = i; +

[because of (4)]. On the other hand, because of (3.6)

n
$2 i ( l + 2 '” /) = 2A: + $2'^/ = 0mod2.
/=1

Hence e L q by (7), and therefore x e F.

This finishes the proof that F = F'. Moreover it shows that every element of
F is congruent either to 0 or to i; modulo L q, Thus [F: L q] = 2.
Let us show that F is a geometric lattice. We have

Ak
{v\v) = — = f c e Z .
4

If X e L q, X = En^e^, n, e Z, then En, = 0mod2, and hence

(ylx) = i E ^
i =\

Thus ( I ) takes integral values on F x F. Notice also that for x g L q we


have {v -\r x\v + x ) = (¿;|i;)mod2 and hence F is even if and only if k is
even, or equivalently, if and only if n = 0 mod 8. We know that F is positive
definite (being included in Q ® /^). Finally, recall that [L: L q] = 2 = [F: L q].
Thus 1 = det(L) = 4det(Lo) = det(F). This proves

Proposition 2
For each positive integer k the geometric lattice is positive definite^ even^
and unimodular. □
226 Lattices and Finite Root Systems

Example 7 (The lattices Ei, I = 6 ,1 ,8,9,10)


Let be the free Z-module of rank 10 with basis
a_i, a^, ag. Consider the following diagram (this is a Coxeter-Dynkin
diagram; see Section 3.4).
8
?
O -o -o -o - 0 - 0 - 0 -o -o^lo) (
“ 1 0 1 2 3 4 5 6 7
We give £’io ^ geometric lattice structure by defining (-I*) E^q x E^10 Z as
follows: For all - 1 < i, j < 8,

(aja^) = 2,
(a,lay) = - 1 if / ^ y and the /th and ;th node of the diagram above are
joined by an edge,
(ajay) = 0 otherwise.

We define geometric sublattices E¿, i = 6, 1, 8,9 of E^^ as follows:


8 8
£ 5 = 0 Za,-, ^7 = 0 Za^,
¿=3 /=2
8 8
F^g = 0 Za^, £ 9 = 0 Za^.
i=\ i =0
We could (and later will) attach to each of these four lattices a diagram that
carries the information of the corresponding bilinear form. This is done by
removing an appropriate number of nodes from the original £^0 diagram in
the obvious way.
We will later see that the lattices £g ,. . . , E^q are intimately related to Lie
theory. Corresponding to each of E^, £ 7 , and £ g , there is an exceptional
finite-dimensional simple Lie algebra. £9 corresponds to the affine Kac-
Moody Lie algebra E\. The lattice £^0 corresponds to a hyperbolic Kac-
Moody Lie algebra that is of much interest to physicists.

We study these lattices in a number of steps ( - 1 . . . 3).


[1] It is convenient to start with £g. Let . . . , eg be an orthonormal
basis of I^, and define ..., ^ by
Pi = ^i+i - ^/+2. 1 < / < 6,
^ 7 = i ( £ l + eg) - K «2 + • • • +£7).

^8 = e? + eg.
It is trivial to verify that • Pj = (ajay) for all 1 < /,; < 8 (here • is the
usual inner product of Q Ig), and hence £g = ©f^iZjS,. Using the j8/s,
3.1 Lattices 227

we are now going to show that ^8 —Tg (see the previous example for
notation). It is clear that eZ/S^cTg. Conversely, suppose that x =
(jCj,... , Xg) belongs to Fg. By using Pj if necessary, we can eliminate jCj. The
resulting vector lies in L q [see (4)] and has = 0. By means of j8g we can
assume that the sum of its coordinates Ef=2^/ = 0*We then have
(0,X2, . . . , Jig) =X2(s2 - £3)
+ (^2 + ^3)(^3 - ^4) + • + (^2 + **• +Xj){Sj £g)
= X2)3i + (X2+X^)P2 + • • • + {x2 + ••• +Xj)l3^ e eZ/3,-.
We conclude that is an even unimodular positive definite lattice.
[0] To determine the structure of Eg, we consider the element a = qiq +
2ai + 3a2 + 4a 3 H- 50:4 + 6a^ + 4a^ + 2aj + 3ag. The coefficients of a
have considerable significance as we will have reason to see later in Section
3.5. For now we note that they have the remarkable property that each n¿ is
half the sum of the n/s for the nodes j that are adjacent to i in the diagram
—for instance, «5 = 6 = |(5 + 4 + 3) = ^(«4 + «5 + n^). Writing this as

2n,. = (a ,|a ,> ,. = - 52 («,1«;)«;

leads to

(a,la) =Ia,[52 j =0, i =0,...,8.


In particular (a\a) = (Lw/aja) = 0, so a is isotropic. Since {a, a^} is
a basis of Eg it follows that Eg is singular; that is, det(£9) = 0. We also
observe that since E^ is positive definite. Eg is positive semidefinite.
[-1 ] Consider the basis { - a _ i - a, a, a^,. . . , ag} of E^q. The matrix of
Eiq with respect to this basis is given by
/M« 0
0 1
1 0

where M,g is the matrix for Eg


Thus E jo = E^ ± U, where [/ is a hyperbolic plane and detCEjo) =
det(£g)det(t/) = - 1 , showing that E^q is unimodular. E^q is even, and its
signature is sgn(Eg) + sgn(t/) = 8. An example of an element of E^q of
negative norm is given by a _ j + 2a. Summarizing, E^q is a rank 10 (indefi­
nite) even unimodular lattice of signature 8.
[2] Ej, being a sublattice of Eg, is positive definite and even. To compute
its determinant, we write a above in the form
a = «0 +
228 Lattices and Finite Root Systems

Let L '= Z2aj 0 Z«2 0 * * * 0 Zag. Then {)3, «2» • • • ?^8^ ^ basis of L and
[^gi L] = 2. Writing p = a - aQ immediately gives (a ,1)8) = 0, for 2 < / < 8,
and ()S|)8) = 2.
The matrix of L in this basis is therefore
M, o\

0
y0 •••0 2/
where M7 is the matrix for E-^ and hence det(L) = 2 det(£7). On the
other hand, det(L) = [£^gi L p det(jEg). Thus det(L) = 4, and therefore
det(^7> = 2.
[3] As above, is positive definite and even. Let )8 be as in [2] and write
)3 = 2«! + )8'.
Set L := Z3q'2 0 Zo:3 0 • • • 0 Zag. Then {)8', « 3, . . . , ag} is a basis of L',
(a ,1)8') = 0 for 3 < i < 8, and ()8'|)8') = 6. Thus det(L') = 6 det(£g). On the
other hand, [E^: L ] = 3 and det(L') = [E^: L Y so that det(L') = 18.
Combining these two, we conclude that deti^^) = 3-
[4,5,6,7 ,8] The reader who is finding this game amusing may continue by
removing nodes 3 ,4 ,5 ,6,7 in succession, obtaining lattices that might be
called ^ 5, J&4, JE3, E2, El with determinants 4,5,6,4,2,. In fact it is easy to
identify these with the lattices of type D5, .^4,^42 ± A i , A i ±A^, and A^,
respectively. (These games, 24 in all, can be found in Exercise 3.6.)

If L is a geometric lattice, we define for all n e Z,


L(n ) : = { a e L | ( q :| q : ) = n) .

Proposition 3
Suppose that L is a definite geometric lattice. Then

(i) each set L{n) is finite,


(ii) Aut(L) is a finite group.

Proof. Replacing (*| ) by - ( I*), if necessary, we may assume that (*1) is


positive definite. Then L(n) = 0 if n < 0, so we consider L(n) for n > 0.
There are two ways to see that L(n) is finite, each with its own merits

1. Let £ = be the realization of L. Then E is a euclidean


space, L is a discrete subgroup of E ^ d for each n > 0, S{n) •=
{x e E\(x\x) = /1} is a sphere of radius yfn . Since S(n) is compact and
L is closed discrete, L(n) = S(n)C\ L is finite.
2. (*1*) extends to a positive definite bilinear form in V = Q Lti
e = {ci,. . . , e^} be an orthogonal basis of V, and let ^ Q>o be such
that (eje^) = q^. Each element of L can be written as a rational linear
3.2 Finite Root Systems 229

combination of the e/s. Since L is finitely generated, we conclude that


there exists an integer N such that all elements of L can be written as
integral linear combinations of e\ , e ', where e- — N for all
1 < i < m. Define

L © Ze'.
i= l

If x' = e L', then {x'\x') = N~^Lnjq^. Hence for each n e Z


there exist only a finite number of x' e L' such that (x'\x') < n. Since
L c L, we conclude that each of the sets L{n) is finite.

To prove that Aut(L) is finite, let {ej,. . . , be a basis of L. Define


e Z >0 by Qi '= and set 5 == U Then 5 is a finite set that
generates L and is stable under Aut(L). Two automorphisms of L are equal
if their actions on S coincide (because S spans L). But there are only finitely
many ways in which Aut(L) can act on S (because S is finite). Thus Aut(L) is
finite. □

3,2 FINITE ROOT SYSTEMS

In this section we introduce the beautiful combinatorial objects called finite


root systems. Historically these first appeared in the analysis of finite-dimen­
sional simple and semisimple Lie algebras. Since then they have been used
also in the theory of lattices, the representation theory of associative alge­
bras, and the theory of singularities, just to mention a few instances. The
simplest example of a finite root system is the set of norm 2 vectors of a
positive definite geometric lattice L.
Usually root systems appear in the theory of Lie algebras during the
classification of split simple and semisimple Lie algebras. These algebras are
decomposed into what amounts to a triangular decomposition, and the
corresponding root systems are finite root systems in the special sense
defined below. The classification problem reduces to classifying finite root
systems.
Our approach is the reverse. We begin by defining finite root systems,
determining their basic properties in Sections 3.2 and 3.3, and using them to
construct simple and semisimple Lie algebras in Section 3.7. The notion of a
base for a finite root system, which is the main subject in Section 3.3, plays an
important part in the theory. In the first place, bases allow us to classify finite
root systems. Equally important for our purposes, they allow us to write down
a canonical presentation for the corresponding Lie algebras, which in turn is
the starting point for generalizing the finite-dimensional semisimple Lie
230 Lattices and Finite Root Systems

algebras to infinite-dimensional counterparts. The abstract theory of root


systems was initiated by Witt [Wil].
We begin with a few generalities about reflections. Let L be a vector
space over a field IK of characteristic 0. An endomorphism 5 of F is called a
reflection (of K) if

1. = 1 and 5 ^ 1 ,
2. s pointwise fixes a hyperplane (= subspace of codimension 1) H of V.

If V is finite dimensional, it follows from this definition that the minimal


polynomial of 5 is — 1 and that s admits - 1 as an eigenvalue of
multiplicity 1 and 1 as an eigenvalue of multiplicity dim F - 1. In particular
the minimal polynomial of s has no multiple roots, and all of its eigenvalues
lie in IK. As a consequence 5 is a diagonalizable automorphism of F (see
Section 7.1). Furthermore dct(s) = - 1 .
If a e F is an eigenvector of s with eigenvalue - 1 (i.e., if sa = - a \
then

(1) V=Ka®H,

where H = {x ^ V\s(x) = x}. We call H the hyperplane of fixed points of s.


Given an element a e F, a ¥= 0, we say that an endomorphism 5 of F is a
reflection in a if 5 is a reflection of F and sa = —a. Notice that there are as
many reflections in a as there are hyperplanes H of V supplementary to a
[i.e., satisfying (1)]. In particular this number is infinite whenever
dimtt,(F) > 1.
Let 5 be a reflection in a, and decompose F as in (1). Let F* be the dual
space of F, and let F X F* -> IK be the natural pairing. Define
«V ^ by

(a, a ^ ) = 2,
= ( 0),

and consider the endomorphism of F given by

( 2) > X - (x,a^)a.

Then it is evident that s = Conversely any decomposition (1) of F into


a hyperplane H, and a one-dimensional supplement IKa gives rise to a
reflection in a by using (2).
Note When using this notation, it is inevitable that we have to make a
choice whether to define (., .> from F X F* to IK or from F* X F to K. We
have chosen the former, but later, when F is finite dimensional and we talk
about reflections on F and F* at the same time, we will identify F** with V
3.2 Finite Root Systems 231

in the usual way and write reflection on V* in the form x* x* —


(a, x*)a^.

Proposition 1
Let V be finite dimensional, and let and S2 be reflections of V in the same a.
If Si S2, then S1S2 is not a semisimple transformation o f V.

Proof Let Hi and H 2 be the fixed point hyperplanes of and ^2, respec­
tively, and set Í = Clearly

1. det(0 = 1,
2. t pointwise fixes Ka 0 (//1 fl //2)-

If t is semisimple then t becomes diagonalizable by a suitable extension of


the base field K (Section 7.1). It is clear then that conditions 1 and 2 imply
that t = 1 and hence that Si = S2- □

Proposition 2
Let ^ be a subset of V, and let Aut(A) be the group o f automorphisms of V
stabilizing A. Suppose that A is finite and that A spans V. Then

(i) Aut(A) is a finite group,


(ii) For each a e A there exists at most one reflection s o f V in a such that
s e Aut(A).

Proof (i) If two automorphisms of V coincide on A, then they are identical


(since A spans V). Thus Aut(A) is finite whenever A is.
(ii) Let a e A, and suppose that Si and ^2 are distinct reflections in a
that stabilize A. Then t = S1S2 ^ Aut(A) is an automorphism of V of finite
order [by (i)] and hence semisimple by Maschke’s theorem [Ja2]. Now (ii)
follows from our previous proposition. □

Suppose now that V carries a nondegenerate symmetric bilinear form ( 1 )


and that 5 is a reflection of F in a which is also an isometry of (*|*). Then, if
we decompose F = Ka 0 / / as in (3.8), we have for all x e / /

(a\x) = ( 5( a)U (x )) = ( - a l x ) ,

so (a|x) = 0. Thus Ka = H ^ , and we have

(3) V=Ka±H,

It follows that the restriction of ( I*) to both Ka and H is nondegenerate. In


232 Lattices and Finite Root Systems

particular (a\a) ¥= 0. Conversely, if a e F is nonisotropic, then the restric­


tion of ( I ) to / / := IKa-^ is nondegenerate and then the reflection in a
given by the orthogonal decomposition (3) is an isometry. This is the unique
isometry of V that is also a reflection in a. It is called the orthogonal
reflection in a and is denoted by
If one uses (• | •) above to identify V with F* (so that x e F < ^ ( | A : ) e F * ) ,
then it is clear that
la
(4)
(a|a)

In other words,

2(x\a)a
(5) s^(x) = x
(ala)

Let F be a flnite dimensional vector space over a field K of characteristic


0. Assume that dimB^(F) > 1. A subset A of F is called a finite root system
in F if the following conditions hold:

RSI: A is finite, 0 ^ A, and A spans F.


RS2: For each a e A there exists a '' e F* such that ( a , a ' ' ) = 2 and
such that the reflection of F defined by

^ ^ - (v,a-}a

stabilizes A.
RS3: <j8, a " ) ^ Z for all a and p in A.

The reflection whose existence is assumed in RS2 is necessarily


unique (by part (ii) of Proposition 2). In particular a" is uniquely defined for
each a, and this makes RS3 meaningful. In the sequel we will write
instead of
The dimension / of the space F spanned by A is called the rank of the
root system. We refer to F as the ambient space of A and sometimes write
(A, F ) to emphasize this. The elements of A are called roots. (This terminol­
ogy will be made clear when root systems are linked to Lie algebras.) For
a e A,a'' is called the coroot associated with a. The set of coroots is
denoted by A".
The elements of GL(F) that stabilize A are called automorphisms of the
root system A. The group of all these automorphisms is denoted by Aut(A).
Axiom RSI together with part (i) of Proposition 2 imply that Aut(A) is a
finite group. By assumption, if a e A, then e Aut(A). Thus {r^;|a e A}
generates a subgroup W = IF(A) of Aut(A). This group is called the Weyl
group of the root system A and is of fundamental importance in Lie theory.
3.2 Finite Root Systems 233

Since rj^a) = —a for all a e A, it follows that A = —A and that —1 e


Aut(A). Whether or not - 1 is in depends on the root system (see the
^3,^3 Examples). Notice that if Aa e A for some A e (K^, then is a
reflection in a stabilizing A. By the above observation In particular
Ta ^ ' ' - a a e A.
Let V and V' be IK spaces and A and A' root systems in V and V',
respectively. We say that A and A are isomorphic if there exists a linear
isomorphism F ^ F ' such that </>(A) = A. We write A = A to describe
this situation. Note that if o- e Aut(A), then e Aut(A) and that the
map or <l>cr(f>~^ is a group isomorphism from Aut(A) onto Aut(A). More­
over, if a e A, then <f>r^(f>~^ is a reflection of F ' in <^(a) stabilizing A, and
therefore

(6) 4>rA ‘ = r.<Koc) for all a e A.

This shows that under the above isomorphism IF(A) is mapped onto 1F(A).

Proposition 3
Let is he a root system in F. Then IF(A) is a normal subgroup o f Aut(A).

Proof If cr G Aut(A), then cr g GL(F), and we can view a as an isomor­


phism of A into itself. Our present result then follows from (6). □

As our first examples of root systems we have the following:

Proposition 4
Let L be a positive definite geometric lattice o f rank I with rationalization F'.
Suppose that the set L(2) of norm 2 vectors in L is nonempty and spans a
subspace V of V'. Then L(2) is a finite root system in V and its reflections are
the orthogonal reflections r^ o f V in the elements o f L(2).

Proof By Proposition 3.1.3 L(2) is finite, and hence RSI holds. For each
a e L(2) the orthogonal reflection r^ in a is

2(x\ a)a
r ^ - . x - ^ x ----- , I , = - (jc|a)a.
(a|a)

Evidently r^ is an automorphism of L and hence stabilizes L(2); therefore


RS2 holds. Identifying F* and F by ( 1 ) , we see from (4) that = a. Thus
for G L(2), = (j8|a) ^ Z, and RS3 holds. □

Example 1 Figure 3.1 illustrates the root system A formed by the


vectors L(2) of the lattice of type A^. The roots are the vectors from the
centre 0 to the vertices of the cube-octahedron. The edges join vectors a, p
234 Lattices and Finite Root Systems

A, 0-0-0 B3 o — o = ^ o
Figure 1.

for which a./3 = 1. The planes through which the reflections take place pass
through the diagonals of opposite faces of the surrounding cube. The Weyl
group is isomorphic to the symmetric group 54. In fact, if T denotes the set
of four pairs of opposite triangular faces of the cube-octahedron then each of
the six reflections generated by root pairs ± a acts as a distinct transposition
on T, thereby giving all of the transpositions. If we adjoin to A the vectors
from 0 to the centres of the square faces, we obtain another root system A'
(later we will see it is of type B^) with 12 + 6 = 18 roots. The three
reflections determined by these new roots evidently lie in Aut(A). However,
their product is the mapping - 1 (central inversion), which acts trivially as a
permutation on T and is not in W(A). On the other hand, the product tt of
any two of them is a 180° rotation about the line L that is the intersection of
their reflecting hyperplanes. The same line L is the intersection of two
orthogonal reflecting hyperplanes of A so that tt e IF(A). Thus

W{A) = {±1} X IT(A),


[PT(A'):IT(A)] = 2.

In fact we have

W{A) = Aut(A') = Aut(A).

The pair of graphs labelled and ^3 are explained in Section 3.4. The
3.2 Finite Root Systems 235

reasons for changing the labelling of the roots in passing from -^3 to will
become clearer when we discuss bases in Section 3.3

Example 2 {Ai) (See Example 3.1.4) The set of roots L(2) is precisely the
set of vectors of the form - Sj with i j, and the orthogonal reflection r
in Si - Sj is the linear automorphism of V defined by

In other words, r is the linear automorphism defined by the transposition


Si <-> Sj of the basis {e^,. . . , ei+J. It follows that W = 5^+^.

Although we have defined root systems for spaces over arbitrary fields of
characteristic 0, it is more convenient to develop the theory for spaces over
the rational numbers Q. At the end of the next section we will see that it is
easy to reduce the general situation to the rational case. However, until
further notice, V denotes a rational vector space of dimension / and A c F a
finite root system.
Proposition 5
Let tx he a root system on V. Then there exists a positive definite symmetric
bilinear form (*|*) on V that is invariant under Aut(A). I f V and F* are
identified via such a form, then a" = 2a/ {a\ a) for all a e A.

Proof Let (.,.) be any positive definite symmetric bilinear form on F.


Define (•!•): F X F ^ Q by
(^ly ) = L forall e F.
o-e Aut(A)

It is clear that this form has the required properties. The statement concern­
ing a = 2a/{a\ a) now follows from (4). □

Proposition 6
Let ^ C.V be a root system. Then

(0 A" is a root system in F*;


(ii) PF(A) = W(A^) and Aut(A) = AutCA""), the isomorphisms being de­
fined by inverse transposition (see below);
(iii) q; = Qjv V all a ^ A and A = ^ (under the canonical identifica­
tion o f V and F**).

Proof Let (*1*) be a positive definite form on F that is invariant under


Aut(A). Using (*|*) to identify F and F*, we have (as above) a" = 2a/(a\a).
This shows that A" c F* satisfies RSI.
236 Lattices and Finite Root Systems

For each a e Aut(A) define <r* e GlXV*) by

(7) <£;,(r*(i;*)> = for all y e F, r* e F*.

(<r* is the inverse transpose of a.) For j8 e A and y e F we have

2P \
^ (y ),i3 '') = |(7 ‘(y)!
{v ,(T * {l3 '')) = ( a
W ) I
2(vla(/3)) 2(v\o-(/3))
(^1^) ( ( r(^ )k (/3 ))

= (v,{aiP)y},

and hence

( 8) a*{p'^) = ( a { ^ ) y .

It follows that O'* stabilizes A"; in particular r* stabilizes A'" for each
a e A. Now for i; e F and v* e F*,

( v , r : { v * ) ) = (r^(v), v*y
= (v - ( v , a ^ } a , v * y
= (u,v* - ( a , v * ) a - ) ,
which shows that

r * : i;* i;* - < a ,i;* > a''.

Thus r* is a reflection in a" and = a (see the note on pairings in the


previous section). This proves RS2 and part (iii). Finally, for all a ' ' , ^ A'",

which proves RS3.


The mapping o- >-> o-* is evidently a homomorphism of Aut(A) into
Aut(A") [resp. W(A) into W(A" )]. It is seen to be injective from o-** = a
and surjective by reversing the roles of A and A". □

The root system A" is called the root system dual to A. One should be
aware that the mapping A ^ A" defined by a a '' is not the restriction of
a linear map in general. In fact it is obtained by inversion with respect to a
suitable (/ - l)-sphere.
3.2 Finite Root Systems 237

Example 3 The root systems dual to ^3 and are illus­


trated in Figure 3.1. Notice that for a e A(y43), a" = 2a/ ( a\ a) = a, whereas
for a e A(^3) \ a" = 2a/ (a\ a) = 2a. The root system AiB^Y is of
type C3. We leave it as an exercise to verify that AiB^) and A(^3)" are not
isomorphic.

The reader may wonder about the role of ( •| •), which seems to be quite
arbitrary and yet quite significant. In fact it is essentially unique. We leave
the precise sense of the uniqueness to Propositions 3.4.7 and 3.4.8. Some
indication of this already appears in the following discussion.
Let (*|*) be a symmetric bilinear form on V as constructed in Proposition
5. Then

(9) 2iß\a)
( a la )
for a,/3 e A.

Recall, on the other hand, that in the euclidean space U the angle 6
between a and is given by the expression

(i3|a)
cos(6) = 1/2
[(a |a )(/3 lj8 )]

Hence

( 10 ) ( ß , a ^ ) ( a , ß" ) = 4cos^(ö).

The left-hand side of (10) is the product of two integers, whereas the
right-hand side is in absolute value < 4. Thus the possible values of 6 are
severely restricted. We collect below all information that follows from the
above.
There are two main cases to consider:

Case 1. a and ß are proportional. Suppose that ß = Aa, A e Q. Since


both ( ß , a " ) and ( a , ß " ) are integers, it follows from (9) that
both 2 A and 2A“ ^ are integers. We conclude that
A e {± ± 1, ± 2}. These are the only possible multiples of a
root. [In the sequel we will say that a root a g A is indivisible if
^ A. If a root system consists only of indivisible roots we will
say that it is reduced. Otherwise, we say that it is nonreduced. We
denote the set of reduced roots of A by
Case 2. a and ß not proportional. The only possible solutions of (10) (up
to interchanging a and ß) are given by Table 3.1.
238 Lattices and Finite Root Systems

Table 3.1
0 (a|a)/(j8li3) Order of
0 0 IT/ 2 Undefined 2
1 1 7t/3 1 3
1 2 7t/4 2 4
1 3 ir/b 3 6
-1 -1 27t/3 1 3
-1 -2 3-77/4 2 4
-1 -3 57t/6 3 6

We determine the order of as follows: Both and act on the


space = Qa © Qp, and on this space the transformation has the
prescribed order. There are two simple ways, one geometric and one alge­
braic, to see this. Embedding ^ into real 2-space U one can use
the fact that the product of two reflections is the rotation through twice the
angle between their reflecting hyperplanes. A completely algebraic way is to
observe that the matrix for relative to the basis [a, of ^ is

\ -1 )

and its characteristic polynomial + (2 - 4cos^0)A + 1 has roots


Now let and be the hyperplanes of V consisting of fixed points of
and respectively. Let V' = Then L = K' ® Since
acts like the identity on K', it has the prescribed order as a transformation on
V, The rest of the information contained in the table is clear.
Here are a few more facts about root systems.

Proposition 7
is a reduced root system.

Proposition 8
Let a, p ^ A be nonproportional roots. I f (a|j8) > 0, then a — p is a root. If
(a\p) ^ Oj theft ot “1“ ^ is Cl foot.

Proof. If (a|)3) > 0, then <j8, a '' > > 0 and > 0. Since their product
is less than 4, at least one of these two quantities equals 1. If </3, a ' ' ) = 1,
then r„/3 = P - </3, a" > a = )3 - a e A, and hence a - /3 e A. If
<a, jS" > = 1, then r^a = a - ¡3 ^ A. The second assertion follows from the
first by replacing )3 by —/3. □
3.3 Bases for Finite Root Systems 239

Let i= , r, be K spaces, and for each i let be a root


system in Let V = and identify each L; with a subspace of V.
Each can then be thought as a subset of V. We let A == U [=iA^. Let us
show that A is a root system in V. It is clear that A spans V. Let
a e A^ c L; c V, and let a" e K* be defined by means of the canonical
identification F* = Then is a reflection on V stabilizing A (it
acts like in Vi and like the identity on Vj, j i). Also <Ay, a" > = (0), and
(A^,a ") (z I . Thus RS2 and RS3 hold for A. We denote the reflection in a
by both for a e A^ and for a e A.
The root system A in F constructed above is called the direct sum of the
root systems A^. To describe this situation, we write
A = Ai V V A.

Let W d e ^ te the Weyl group of A and WK the Weyl group of A^. The
subgroup of W generated by the r^, a e A^, is evidently isomorphic to
(the restriction map w ^ w\y. being the isomorphism). S i n c e a n d
commute if a g A^, j8 e Ay with i ^ j \ we conclude tjiat W = W j X • • • X
Wj , - Wi X • • • XW^. There is no harm in identifying and and we do
so freely in the sequel.

A root system A is called decomposable if there are root systems Aj, A2


such that A = Ai V A2. Otherwise, A is called indecomposable.

Any root system can be decomposed uniquely into indecomposable root


systems (see Section 3.3).

Proposition 9
A root system is decomposable if and only if there exist nonempty subsets A^, A2
of A such that

(i) A = Ai UA 2,
(ii) (aljS) = 0 for all a e A^ and j8 e A2.

Proof (=>) The proof is clear from the definition.


(«=) Let Fi and F2 be the span of A^ and A2, respectively. It is trivial to
verify that Aj and A2 are root systems in F^ and F2, respectively, and that
ASA1VA2. □

3.3 BASES FOR FINITE ROOT SYSTEMS

Let A be a finite root system with ambient space F. It is evident that we can
choose a basis of F from A. However, it is possible to do this in a very
special way that naturally divides A into positive and negative subsets A+
240 Lattices and Finite Root Systems

and A_ with A_= -A+. The key result (Proposition 5) is that any two of
these special bases are translates of each other by the Weyl group W. Thus
the structure of a base is an invariant of W. This leads immediately to the
definition of the Cartan matruc of a finite root system and the introduction of
the Coxeter-Dynkin diagrams.

Let A be a root system in a I -space V. A subset n = [a^,. . . , a:/} is a base


for A if

Bl: n is a basis for K as a vector space,


B2: every root can be expressed (necessarily uniquely) as an integral
linear combination of the a^, where all the coefficients are either
nonnegative or all are nonpositive (loosely, all the coefficients have the
same sign).

We now return to the case K = Q. Throughout this section (-I*) denotes


an Aut(A)-invariant positive definite symmetric bilinear form on V (see
Proposition 3.2.5).

Proposition 1
Every finite root system has a base.

Proof We start by putting a total ordering > on the rational space V that
makes V into a totally ordered vector space. By definition this means that >
is a partial ordering on V and that

1. for all v,w ^ V either u > w,w > v, or w = u, the possibilities being
mutually exclusive;
2. if V > w and X e F, then v + x > w + x;
3. if V > w and c G Q+, then cu > cw.

For instance, if we choose a basis . . . , Ui), we may define the lexico­


graphical ordering by declaring that for two unequal elements, 'La¿v¿ > Lb^Vi
if and only if at the smallest i for which a^ b^ we have a, >b, . ^
In any case let V+= {z; e V\v > 0} so that V = U{0}U V- (U denotes
disjoint union). Let A+:=Ari]^+ and A_— Afll^-. A root a e A+ is
called simple if a cannot be written as ¡ 3 y, where j8, y e A + . Let
n = {«1, . . . , a^} c A+ be the set of simple roots. We reason in steps:

1. I f a, P ^ Yl, a ¥=p, then (a|j8) < 0 and (a, ) < 0: We have ex­
cluded the possibility that a and jS are proportional since then, say,
P = 2a = a a and p is not simple. If (a\p) > 0, then a — p and
p - a dire roots (Proposition 3.2.8). One of these, say, p - a, belongs
to A+, and then p = (p — a) a shows that p is not simple.
33 Bases for Finite Root Systems 241

2. The elements o f O are linearly independent : Suppose that = 0,


where not all the c, are 0. Let / ‘^= {i|c, > 0} and / “ = {i\c^ < 0}. Then
both and I~ are proper subsets of k} (otherwise, 0 would
belong to either or V~). Let x = Then x ^ 0, and
X= c^a^. Thus 0 < (x|x) = ^
step 1, which is a contradiction.
3. Yl is a base o f A: If y e A+ is not simple, then y = a + )3, where
a, p ^ A^. Then y —a = )3 > 0, so y > a, and similarly y > p. It
follows by induction on > that every element of A^. is a sum of
simple roots. Since A_= —A+, B2 holds; and since A spans F, O is a
basis of V, □

Example 1 Let L be a positive definite geometric lattice for which L(2)


0. Consider the root system L(2) (Proposition 3.2.4). Suppose that 11 c L(2)
has the properties

1. The Z-span of Yl contains L(2);


2. (a|j8) < 0 for all a, P ^ Yl, a

Then n is a base of L(2). For suppose that y e L(2), y = E ^ e n '^ a ^ ’


e Z. If the are of mixed signs, we may write
y = y++ y_,
where y+— y_:= y - y+. Then (y|y) = (y+|y+) + 2(y+|y_)
+ (y_ly_). Since (y J y +) e 2Z+ and the assumption (2) gives (y+|y_) > 0,
we have (y|y) > 4, which is impossible. The elements of II are linearly
independent by the argument of step 2 of the last proposition and span the
same rational space as L(2).
Referring back to Section 3.1, this shows that the a-bases given there for
the lattices L of types Ai, j g are actually bases for the corresponding
root systems L(2).

Proposition 2
Let ^ d V be a root system, and let S cz A be any nonempty subset. Let V' be
the subspace of V spanned by the elements of S, and set A = A fl F'. Then

(i) A is a root system in F',


(ii) any base of A can be extended to a base o f A.

Proof If a e A, then a" can naturally be viewed as an element of F'* by


restriction. Then for all a, e A,
r , ) 8 = ) 8 - < ) 8 , a - > a e A n F ' = A,
and it is clear then that A is a root system in F'.
242 Lattices and Finite Root Systems

Now let {«1, . . . , be a base of A', and consider a basis for V of the form
{v^, q:^}. Then the base of A obtained from the lexicographic
ordering of V determined by this basis contains « i , . . . , □

A subset 2 of a root system A is called a positive system of roots if


S = A n 1^+ for V+'-= {v ^ V\v > 0}, where > is some total ordering on V.

We have seen that the simple roots of a positive system of roots 2 form a
base n in X. In fact FI is the only base of A lying in X. To see this, we
introduce the height function

(11) ht := h t^ : A
by

htl £ m „a\ = E
aen
Notice that /3 e A is in if and only if ht(jS) > 0. Let O ' be another base
of A inside X. Then ht(a') > 1 for all a' e O'. Each a e O is expressible as
a = where the m^, e Z are all of the same sign. In fact the
m^> e N for otherwise a e A_. Now 1 = ht(a) = Xm^.ht(o:') =>« = «' for
some a' e O'. Thus O c O ' and, in view of card(n) = dim(F) = card(n'),
we have Yl = O'.
Every base Yl can be used to determine a lexicographical ordering of V
and then a positive system in which it lies as the simple roots. We call
A+= A+(n) the system of positive roots corresponding to O.
Clearly

A ,(n) = e A /3= E rn^a,m ^ > o}.


Summarizing, we obtain Proposition 3.

Proposition 3

(i) Every positive system o f roots contains exactly one base. This estab­
lishes a 1-1 correspondence between positive systems o f roots and
bases.
(ii) The elements of any base are indivisible.
(iii) I f {«1, is a base, then (a^, aj" > < 0 for all i ¥=j. □

Proposition 4
Let Y\ be a base o f A. Then

(i) { r j a G n} generates W as a group-,


(ii) A^^^ = WY\ == {wa\w ^ W , a ^ O}.
3.3 Bases for Finite Root Systems 243

Proof, Let n = {aj,. . . , aj}. Let Wq be the subgroup of W generated by


Tj,..., A*/ (for convenience, we denote by r^). We begin by proving

( 12 ) = W ^n.

We use induction on the height function ht = htp^ relative to Yl to prove


that := A^®^ n A+c W^Yl, Let /3 e If ht(iS) = 1, then j3 e n .
Suppose that ht(j8) > 1. If (jSla,) < 0 for all i, then writing )3 =
e N, we have

0 < ( ^ 1^ ) = Lnii{l3\ai) < 0,

which is impossible. Thus (p\ a , ) > 0 for some i. Then

(13) r,.(/3) = 13 - {p, a,'' >a, = Y, + (m, - </3, a," >)a, e A"®'*.
j * i

If rrij = 0 for all i i then ^ = m,a,. Since /3 is indivisible, /3 = a„ contrary


to ht(/3) > 1. Thus some Wy ¥= 0, and hence nij > 0. It follows that r,(/3) e
A+'’. Since = 2()3|a,)/(o:,|Q:,) > 0, ht(r,(/3)) < ht(/3), and we may
assume by induction that /-¿(jS) e WqTI. Write r,(^) = for some w'^ in
Wq. Then /3 = WqU^. with Wq = /-¿Wq e Wq. In addition by (6),

-1

The conclusion of all this is that c IFoIl and that e Wq for all
j3 G A™**. Since whenever /3 and Aj8 e A, we see that g for
all -y G A. Thus Wq = IT, proving part (i). Finally, r,(a,) = —a,-, showing that
- n c ITon, so IT o n -----ITon and A^f“ = -A'®“ c W q U . This proves that
^r®d (- ]YqTI. The reverse inclusion is obvious. This finishes part (ii). □

If n is a base for A+ and a e O, then Q a fl A+ is one of {a} or {a, 2a}.


In either case we have the following simple facts:

1. stabilizes A + \(Q a fl A+)


(14)
2. r ,A ^ = - ( Q a n A ^ ) U ( A A ( Q a n A ^ ) .

The point is that for any root )3, /3 and rj^p) differ only by a multiple of a. It
follows that if )8 e A+ and j8 ^ Q a f l A+, then the expression for in
terms of elements of II still contains nontrivial (hence positive) contributions
from elements 7 ^ n \ { a } [see (13)]. Thus r^(j3) e A +\ Q a fl A+ also.
Simple as these remarks are, they play an important role in understanding
the structure of W (see Chapter 5).
244 Lattices and Finite Root Systems

We have the following key result:

Propositions. (Conjugacy o f bases)

Let (A, V) be a root system.


(i) I f li is a base o f A and w then iv(n) is also a base.
(ii) Any two bases o f A are related in this way.

Proof (i) The proof is obvious. To prove part (ii), let and II2 be two
bases of A with corresponding positive systems A+ and A+. Let A_ := -A+
for / = 1,2. It suffices to prove that there is w ^ W with w A \ = A+ (see
Proposition 3).
If A + n A^_ = 0 , then A+ c A+, whereupon they are equal and hence
III = Il2 by (i) of Proposition 3. Suppose that card(A+ n A^_) = n > 0.
Then III ^ ^ since, otherwise, 111 c A+ and Hi = II2. Choose a e
III n A^_, and consider r^A \. It differs from A+ only in that —a (and
possibly —2a) replaces a (and 2a). But a e A^_, so - a e A+. Thus
card(r^A^^ n A^_) < n. Since r^A \ is the system of positive roots for r^IIi,
we are done by induction on card(A+ n A^_). □

This proposition says that the Weyl group is transitive on the set of bases
of A. In fact it is simply transitive on the set of bases; we prove this in
Chapter 5.

Proposition 6
Let tsbe a reduced root system in V and let A'' c F* be its dual root system.
Let Yl be a subset of A. For U to be a base o f A, it is necessary and sufficient
that ••= [a"\a ^ A) be a base o f A". Moreover, if A^ and A \ denote the
positive systems o f roots defined by II and II'', respectively, then a e A+ if
and only if ^ A+. In other words, {A ^Y = A+.

Proof Let identify V and F* via (• ). Fix an ordered basis of F*, and use
its lexicographic order to define a system of positive roots A+ and a base
n '' := {aj',. . . , a I } of A ''. The above choice of ordered basis of F* = F
defines a positive system of roots A+ of A. Since = 2 a / ( a , a), we
conclude that for all a e A,

(15) a e A+<=> a '' e A \.

We claim that II — {ai, .. .,ajf is a base of A. Since II c A+, we must


show that a^ = p + y, with and 7 in A+, is not possible. Now if a^ = + 7
33 Bases for Finite Root Systems 245

with P and y in A+, then

= ''a,«/ =

SOthat either r^,p e A_ or r^y e A_. Assume that e A_ (the case of


y is similar). Then {r^ p Y e A l, and hence by (8)

= r%p^ = ^ a: .

By (14) (applied to A" and the base II"), it follows that P" = a ," , whence
P= Then y = 0, which is absurd. This shows that II is a base of A. The
rest of the proof is easy. □

Up to now most of our results have been proved about root systems under
the restriction IK = Q. In what follows we show how the general situation is
dealt with.
Let F be a vector space over a field IK of characteristic 0. By restriction of
the base field to Q c IK, we can think of F as a rational space, which we
denote by F(Q). Notice that dimQ(F(Q)) = dim„^(F)dimQ(IK). A (rational)
subspace U of F(Q) is called a rational form or a rational structure of F if
the canonical K-linear mapping
K 0Q i/ F

is an isomorphism. (This map is the unique linear map satisfying k ^ u ^ ku


for all A: e IK and m e i/). Equivalently C/ is a rational form of F if 1/ has a
(rational) basis that is also a basis for the K-space F. In any case notice that
dimQ(i/) = dim„^(F).

Example 2 Let ^ ^ be a basis of F over IK. Then

i / := © QU^
AeA

is a rational form of F.

Next we show how the concept of rational form allows us to understand


the relationship between root systems (A, F ) for IK-spaces and rational root
systems.

Proposition 7
Let (A, V) be a finite root system where the ambient space Vis a K-space. Then
the rational span Vq o f L in V is a rational form o f F, and (A, Fq) is a finite
root system. Furthermore for each a ^ ^ the reflection r^ o f (A, Vq ) is the
246 Lattices and Finite Root Systems

restriction to Vq o f the corresponding reflection o f (A, F), and the coroot of


Vq is the restriction to Vq of the corresponding coroot of

Proof Let Vq be the rational span of A in V. Let a e A. Clearly a" (F q) c Q


and hence a" induces an element also denoted by a " , of Fq . It is now easy
to see that (A, Vq ) is a root system.
It remains to show that Vq is a rational form of F. Let II = {a^,
be a base of (A, Vq ), hence a Q-basis of Vq . Then II spans F (over K) since
A does. Suppose, if possible, that

Y, CiUi = 0, c,- e K,
/=1
where some c^ ¥= 0. Then for each jS e II,

Ec,<a,,i3"> = 0,
¿=1
which shows that the matrix A = « a ,, a p ) defined by II is singular (where
A is viewed as a matrix with coefficients in IK). However, we know that A is
nonsingular when viewed as a matrix with coefficients in Q. Since nonsingu­
larity remains unchanged by extension of the base field, we conclude that the
elements of II are linearly independent over IK and hence that II is a IK-basis
of F as desired. □

Remark 1 The above proposition makes it clear that all our results about
the structure of rational root systems generalize to arbitrary root systems. For
example, any base of (A, Fq) is a base of (A, F ) (showing the existence of
bases). Similarly any base of (A, F ) is a base of (A, F q ) , and hence any two
such bases are conjugate under W.

Let n = {«1, . . . , O'/} be a base of A. The Cartan matrix of A (with respect


to n ) is the / X / integral matrk

A =
where

Aij ■■= <a,-, a j' > =


(«yl“ /) ’

Interchanging the ordering of and changes A by simultaneous inter­


change of rows i and j and columns i and j. Combining this with the
IF-conjugacy of bases, we see that A is uniquely determined by A up to
simultaneous row/column permutations.
3.4 Graphs and Coxeter-Dynkin Diagrams 247

We note the following properties of A:

CMl A¡¿ = 2 for all i.


CM2 Aij G - N if / ¥=j.
CM3
CM4 There is a positive diagonal matrix D (D = diagidj,. . . , d j,
> 0) such that AD is symmetric.
CMS: AD in CM4 is positive definite.

For CM4 and CM5 simply observe that

, (“ ii«i) («/I«/)
A diag = ((a,l«y)).

In Section 3.4 we use CM1-CM3 as the definition of a “generalized” Cartan


matrix. We will also see that CMl-CMS characterize the Cartan matrices of
finite root systems. Notice that CMS implies that A is nonsingular.

3.4 GRAPHS AND COXETER-DYNKIN DIAGRAMS

We have many occasions to use square matrices of integers in the sequel,


notably matrices like the matrices A satisfying the conditions of CM in
Section 3.3. Representing these by graphs is an extremely economical and
useful idea. In this brief section we show how we wish to attach a graph to a
matrix and review the basic definitions and facts about graphs that we will
need. There are no proofs in this section. The reader may refer to [Bg] or
provide them as an (easy) exercise.
Let N* = U {oo}. Let J be a nonempty set. A matrix M = is
said to be combinatorially symmetric if for all i and j in J,

CSl: Mij G N*,


CS2:
CS3: M,, = 0.

Example 1 Let F be a IK-space, and let ^ family of automor­


phisms of V, For all i and j in J, define M^j as follows:

^ I order of s^Sj if s^Sj is of finite order,


if i ^ ;,
otherwise,
M,- = 0.

Then M = is combinatorially symmetric.


248 Lattices and Finite Root Systems

Let M = be a combinatorially symmetric matrix. We may


represent M by a labeled graph

T = (V,E,M)

with vertices (or nodes) V = {vj\j ^ J] and directed edges (or arrows) E =
e J X J, i ¥=j, M^j 0}, which are assumed to be marked by the
quantities (M^y). Note that if M^j = 0, then there is no edge e¿J. By definition,
then also will be missing. M is called the incidence matrix of F. We often
think of as being M^j arrows from i to j [or of type (i, j)] and refer to
them as such.

Example 2

(0 2 0
M= 2 0 00 J = {1,2,3}
lo 1 0
r; o V2 0 ^ 0 .

There are several simplifying conventions that we will use for our graphs:

1. If M,y = 1 and Mji > 1, we omit the edge e,y.


2. If Mij = Mji # 0, we replace the edges and eJ¿ by a single undi­
rected edge labeled by their common value.
3. If M,, is small (< 4 usually), then we may elect to replace a labeled
edge by a corresponding multiple edge. We also use this convention
in 2.
2 2
Example 3 0 ^ 0 becomes O O by convention 1, and then O => 0
by convention 3. O ^ O becomes 0 — 0 by convention 2, and then
O — O by convention 3. M of Example 2 now appears as

0 = 0 ^ o.

Example 4

0
1
0— 0 o
3.4 Graphs and Coxeter-Dynkin Diagrams 249

corresponds to

fo 1 0 o'
1 0 1 1
0 3 0 0
io 1 0 oj

In the sequel all graphs are assumed to be reduced by these conventions.

We assume that the reader is familiar with the standard notions of graph
theory. For definiteness we repeat the definitions that we will adopt for this
book together with a few well-known facts.
Let r = (V, E, M), F = (K', M'), where M and M' are indexed by J
and J', respectively.

1. r and F' are isomorphic if there exists a bijection a: J ^ J' so that


^a(i),«(;) = ^ i j for all (j, y) e J X J.
From this we obtain the concept of an automorphism of F.
2. F is a subgraph of F' if there is an injection a: J J' so that
K(0,a(;) = for all i , j e J.
3. F is disconnected if J is the disjoint union of two nonempty sets Jj and
J2 so that = 0= for all Jz) ^ Ji X Jz-
4. If F is not disconnected it is connected.
5. A path (of length n) in F is a sequence (/ q, ^1, • • •, of elements of J
so that for each k = ¥= 0. The path is said to connect
^0 to V ih
V:

6. F is pathwise connected if every two vertices are connected by some


path.
7. Pathwise connected and connected are equivalent.
8. A cycle or loop in a graph is a path (/ q, ¿1, . . . , of length n > 3 for
which ¿Q, are all distinct and
9. A forest is a graph with no cycles. A tree is a connected forest.
10. Two vertices (or nodes) are adjacent if there is an edge between them.
11. The valence of a vertex is the number (possibly infinite) of vertices
adjacent to it.
12. Any finite tree with more than one vertex has at least two vertices of
valence 1.
13. A connected graph F is bipartite if J can be partitioned into two
nonempty subsets Ji and J2 so that M^j = 0 whenever /, j e or

14. Any finite tree with at least two nodes is bipartite.


250 Lattices and Finite Root Systems

Let J be a nonempty set, and let A = be a matrix with integer


coefficients. We say that ^ is a Cartan matrix (some people call it a
generalized Cartan matrix) if

CMl: A,.= 2 for all / g J,


CM2: A¿j< 0 for all i ¥= j in J,
CM3: A¿j= 0 <=>Aj¿ = 0 for all i ^ j E: J.

Example 5 Let II = {a^,. . . , a j be a base of a finite root system A. Define


A¿j for all 1 < i, j < I by

A¡j = <a,, a / >.

Then ^ is a Cartan matrix.

Let A he a Cartan matrix. Then

M = 2I-A

(where / is the J X J identity matrix) is a combinatorially symmetric matrix


and defines a graph

r = r(A) ^ (V,E,M).

r is called the Coxeter-Dynkin diagram of A.

Example 6

2 -1 0 O'
-1 2 -2 0
A =
0 -1 2 -1
0 0 -1 2.
l: O — o =?• O — O

Example 7 Consider the finite root systems L(2) of types D,, Ai, E^, E^, Eg.
(Example 3.3.1). Referring back to Section 3.1, we see the Cartan matrices of
D, and A i displayed. The corresponding Coxeter-Dynkin diagrams are

T{Ai): 0 -0 - ------0 - 0
O
r(A ): 0 —0 — —0 —0 ^
O

For the E series we have already used graphs to define the matrices. The
3.4 Graphs and Coxeter-Dynkin Diagrams 251

Coxeter-Dynkin diagrams are

0
1
0 —0 —0 —0 —0
0
1
^ £ 7 ): 0 —0 —0 —0 —0 —0
0
1
r(£ 8 ) : 0 —0 —0 —0 —0 —0 - 0

The Cartan matrix A is indecomposable if T (^ ) is connected, and it is


decomposable otherwise.
To say A is decomposable is equivalent to saying that A can be reindexed
(by simultaneous row/column permutation) so that it assumes the form

0
0 ^2

where A^ and A 2 are (necessarily) Cartan matrices.


A Cartan matrix A = i.Aij)i is symmetrizable if there exist nonzero
rational numbers e„ i e J, so that

for all i , j e J

or equivalently

e,- '-<4,7 = ej for all i , j e J.

This concept is of considerable importance in the sequel so it is good to have


an easy way to recognize a symmetrizable matrix.

Proposition 1
The Cartan matrix A is symmetrizable if for each cycle (/ q, • • •, in r(>l),

A: i = A: : A: : .

Since \Aij\ = is the number of arrows (edges) joining i to this


condition can be restated as the products of the numbers of arrows in one
direction around any cycle equals the product of the number of arrows in the
other direction.
252 Lattices and Finite Root Systems

We now return to the case of finite root systems to illustrate and apply
some of the concepts just introduced.

Example S Bases for the root systems of types A 2, and along with their
Coxeter-Dynkin diagrams are shown in Figure 3.1. We will see shortly that to
each diagram there is (up to isomorphism) only one reduced root system.

Let us clarify the relationship between root systems and Coxeter-Dynkin


diagrams. Given a root system A and a base II of A, once we index the
elements of II, we can write down a Cartan matrix. From the matrix we
derive an incidence matrix and corresponding diagram. Reindexing the
elements of II leads to the “same” diagram with the nodes similarly
reindexed. Choosing a different base II' of A, we find a w e IF such that
wU = n '. Since (w(a)\w(p)) = (a|j8), II and vvll lead to the same matrix A
and then to the same diagram. Now we show that the diagram r(A)
determines A to within isomorphism.

Proposition 2
I f A and A' are two reduced root systems such that F(A) —F(A'), then A ~ A'.

Proof We may assume that F(A) = F(A'). Index the / nodes of the diagram
with the numbers 1,..., / in any way. This gives rise to Cartan matrices A
and A' for A and A. Let II = { a ^ , a n d II' = {a[ , . . . , a]} be the
corresponding bases. Let V and V' be the ambient spaces of A and A'. There
is a unique linear isomorphism </>: F ^ F ' such that <f>(a¿) = a', i = 1,. . ., /.
Let r¿ and r¡ be the reflections in a¿ and a', respectively. Then

Since Aji = A'ji, it follows immediately that

r[ = <l>r¿(l>-1

Hence we have an isomorphism W ^ W' with <^(w) = <f)W(})~^, and this


leads to an isomorphism (<^, (f>) in the category of group representations; that
is, the diagram

W X V ---- >

♦1 [*
W' X V ---- V
commutes.
3.4 Graphs and Coxeter-Dynkin Diagrams 253

In particular, <^(A) = <^(W^II) = ^(W)(l>(Ji) = W'U' = A' (Use Proposi­


tion 3.3.4) which shows that (f>is an isomorphism. □

Remark 1 It would be more accurate to require that r(A) ^ r(A) (in the
sense of graph theory) rather that F(A) = F(A') in the last proposition. It is
easy to make this minor formal modification if needed.

Example 9 (Rank 1 and rank 2 root systems)


Suppose that A is a root system of rank 1. Then A has a base II = {a},
and the Coxeter-Dynkin diagram is a single node O and the ambient space is
V= Ka. The Weyl group W is generated by and rj^a) = - a . Thus
^red = w H = {a, —a}. If A then ± 2 a are also roots (Section 3.2,
case 1) and A = {a,2a, —a, —2a) (said to be of type If A =
then A is said to be of type
Suppose next that A is a root system of rank 2 and that II = {a^, «2) is a
base. Then A ^2 '= («i, «2" ) ^21 '= ( ^ 2^^ 1" ) ^^e nonpositive integers.
Reindexing, if necessary, so that |^2il ^ fhe possibilities listed
in Table 3.2 (see Section 3.2, case 2)
Reduced root systems for each of these diagrams are easily constructed.
They are shown below in real euclidean 2-space (as far as a sheet of paper
can represent euclidean 2-space!) The reader can easily construct these: For
instance, for A 2 one has the Cartan matrix | ^ ^j and its Weyl group W
is generated by and ^2- where

^1( ^ 1) = '*1(^2) = «1 + 0^2.


''2(^1) = Oi^+ a2, ''2(«2) = “ «2-
Thus

. -« 1 <— ^ «1

- («1 + «2) (aj + a2)


—ao <— > an
determines IF • II =

Table 3.2
^12 '■ 21 Coxeter-Dynkin diagram Name
0 0 0 0 Ai VA
-1 -1 0 —- 0 Ai
-1 -2 0 <= 0 Cl
-1 -3 0 - ^ 0 Gi
254 Lattices and Finite Root Systems

vAi

+ 3«2

Figure 3.2. Reduced finite root systems of rank 2

If A is not reduced, then 2 a is a root for some indivisible root a. Using


W, we obtain 2a¿ as a root for / = 1 or 2. Then (2a^)'" = and we must
have <j8, \af ) e Z for all )8 g A. This happens only in and in C2
with / = 1. Up to isomorphism all the possibilities are shown in Figures 3.2
and 3.3.

Proposition 3
Let s be the reflection in a vector a given by a decomposition IKa ® H of a
vector space V. I f M is a subspace o f V invariant by s, then either IKa a M or
M(zH,

Proof. If there exists x g M \ / / , then sx —x = ca ^ Ka \ {0}, so IKa c M.


Proposition 4
The following are equivalent for a root system A in a K-space V:

(i) A is indecomposable.
(ii) r(A) is connected.
(hi) V is an irreducible W-space.
3.4 Graphs and Coxeter-Dynkin Diagrams 255

2ao
«2
, «2
«1 2«!
«1 2(

A, V (BC\ iB C \ V (,BC\

(B0 2

Figure 3.3. Nonreduced finite root systems of rank 2

Proof, (ii) => (i) Suppose that A is decomposable: A = Aj V A2. Any bases
III and II2 of Ai and A2 combine to form a base II = Hi U Il2 of A. Since
for a e III, jS e ^2? = <)3, a '' > = 0, the diagram F(A) is composed
of the two disconnected subdiagrams r(Ai) and F(A2).
(iii) => (ii) If r(A) decomposes into two disconnected subdiagrams Fi
and F2, then any base II of A decomposes nontrivially as IIi U II2 with
= 0, whenever a e 111, p e II2. Then the linear spans
Fi and F2 of III II2 are both W invariant, so V is reducible.
(i) => (iii) Suppose that M is a nontrivial IT-invariant subspace of V.
Using Proposition 3.4.3 we partition A as Ai U A2, with

Ai = ( a e Ala e M},

A2= ( a e A|<M,a-> = (0)).

If A2 = 0 , then A c M, so M = F. If Ai = 0 , then {M\a^ > = (0) for all


a e A, so M = (0). By assumption neither of these happens. Thus Ai # 0 ,
A2 ^ 0 . By definition (p, a '' > = 0 for jS e Ai, a e A2. Thus A = Ai V A2.

256 Lattices and Finite Root Systems

Proposition 5
Let V be a K-space and A a root system in V, Define a bilinear form К on Vby

K{x,y)= Y,
аеД

Then K is nondegenerate, symmetric, and invariant under Aut(A). Moreover


the restriction of K to the rational span Vq of A is positive definite.

Proof K is clearly symmetric, and its restriction to Vq is positive definite


since <A,o:''> c Z . It follows that this restriction is nondegenerate and hence
that K itself is nondegenerate (since A spans V). The invariance of K
follows from the fact that <cr"^(x), a" > = = (;c,(c7-(a))" > for
all X El Vq and for all a Aut(A) [see Proposition 3.2.6 and equation
(3.2.8). □

The bilinear form K constructed above is called the natural bilinear form
of A. It is deeply connected with the Killing form of certain semisimple Lie
algebras (see Exercise 6.4).
In the previous paragraph we indicated that invariant bilinear forms are
essentially unique. We now set about giving precise meaning to this.

Proposition 6
Let Vbe a K-space, and A be an indecomposable root system in V. Let B be a
bilinear form on V that is invariant under the Weyl group W o f A. Suppose that
B Q. Then

(i) B is nondegenerate, symmetric, and invariant under Aut(A).


(ii) I f B' is any bilinear form on V that is W-invariant, then B' = kB for
some fc G IK.

Proof Consider the left radical of B, namely {x e V\ B(x, y) = 0 for all


y e F}. Since B is IT-invariant, this is a IT-stable subspace of V, which is also
proper (since B ^ 0). By Proposition 4 the left radical is (0). Similarly the
right radical of B is trivial, and hence B is nondegenerate.
Let B' be as in part (ii) of Proposition 6. Since V is finite dimensional,
there exists / e Endj^(F) such that

B\x,y)=B{f{x),y).

Let us show that / commutes with IT. Indeed, if vv e IT, then using the fact
3.4 Graphs and Coxeter-Dynkin Diagrams 257

that B and B' are PT-invariant, we obtain

B{fw{x),y) =B'{w{x),y) = B ' { x , w - \ y ) )


= B{f{x),w-\y))=B{wf{x),y).
Thus Bdf w - wf)(x), y) = 0 for all x and y in V, and hence Jw = wf since
B is nondegenerate.
Now let a e A, and consider e W. Write V = Ka O H, where H is the
hyperplane of V consisting of fixed points of r^. Since fr^ = r^f, we have
f { K a ) = / ( l - r J ( F ) = (1 - r J / ( F ) c Ka,
and therefore there exists A: e K such that /( a ) = ka. Consider now V' —
ker(/-Al). Then V' is IT-invariant (because / commutes with W) and
V ^ (0) (because K a c K'). By Proposition 4, V' = V, and therefore B' =
kB. This establishes part (ii). Now part (i) follows from part (ii) together with
Proposition 6. □

Proposition 7
Suppose that A is decomposable and that B is a Wdnvariant bilinear form on V,
Let V ”^iA^ be the decomposition of A into indecomposable factors. Let
Bi denote the restriction o f B to where is the ambient space o f A T h e n
B(Vi,Vj) = (0) for i ¥=j and B = Furthermore, if is the natural
bilinear form of A^, then B = for unique c^s in K.

Proposition 8
Let A be an indecomposable root system in V. Let (*|*) be a non-degenerate
W-invariant symmetric bilinear form on V. Then for a and P in A we have
(ala) = (P\p) ^ W a = Wp.
Proof is obvious since ( I ) is PT-invariant by definition.
=» We may assume that K = Q and (• | •) is positive definite (Proposition
3.3.7 and Proposition 6). Since Wa spans V, we may replace a by a suitable
wa and assume that (a\p) 0. Then
2(a\p)

and by replacing p by - p if necessary, we may assume that (p, a " ) > 0.


From (a, p" ) ( p , a ^ ) < 4 we are left with two possibilities:

1. {a, p^ ) = 2 in which case a = p (see Section 3.2),


2. (a, p " ) = (p, a - > = 1 in which case r^r^r^p = r^r^ip - a) = r j , - p
—(a —p)) = a. □
258 Lattices and Finite Root Systems

3 .S CLASSIFICATIO N OF CARTAN M ATRICES A N D FINITE


R O O T SYSTEM S

Let A = (Ajj) be an (/ + 1) X (/ + 1) Cartan matrix. (For convenience we


will number the rows and columns of A from 0 to /). We think of A as an
endomorphism of via

A \ \ ^ \A
for all X = ( xq, Xj, . . . , Xi) e

An element n e is called a null root of if n is a (left) eigenvector


of A of eigenvalue 0.

In other words n = ( wq, satisfies

(1) n A = 0,
(2) n ^ 0,
(3) Hi > 0 for all 0 < / < /.

A Cartan matrix is said to be affine if it is indecomposable and admits a


null root. If A is affine then det(^) = 0.

Example 1 Let ^ = h t sl 2 X 2 affine matrix. Then ab = 4. The


possibilities are

' ^ - ' ^ ■’ - ( - 2 i ) ” -< )

(we may assume that a < b). Examples of null roots are given by (1,1) for
A^l^ and (1, 2) for A^^^.

Our next objective is to classify all affine matrices. As a consequence of


this classification we will obtain a classification for all finite root systems. (In
this section we establish only that finite indecomposable reduced root sys­
tems lie in a certain list. In Section 5.3 we will complete the classification by
showing that all these root systems actually exist.) Our approach follows
[BMW].
Let A = (^¿y), 0 < /, j < /, be an affine matrix, and let n = (riQ,
be a null root. Then

=0 for all 0 < y < /,


/=0
3.5 259
C lassification o f Cartan M atrices and Finite Root System s

and hence for all 0 < 7 < /,

(4) Y , \ ^i j \ ni = 2tij.
1=0
i*j

Proposition 1
Let A be an (/ + 1) X (/ + 1) affine matrix, and let n be a null root o f A. Then

(i) all the coefficients o f n are positive and the 0-eigenspace o f A is Un. In
particular, if n' is a null root o f A, then n' = qn for some q ^
(ii) There exists a unique null root { n ^ , n f ) o f A satisfying

(5) ni ^ for all 0 < i < I,

(6) g.c.d. {«0, •••,«;} = 1( 6)

Proof. Write {0, = I U J, where

I = {i|n, > 0},


J = = 0).

Because of (4) we see that Aij = 0 whenever / e l and j e J. Since A is


indecomposable, either I or J is empty. Hence J = 0 , since n 0. Thus
> 0 for all /.
Let m = (mo,. . . , m^) be a left 0-eigenvector of A. We can choose A e R
such that

n^ — Am, > 0 for all 0 < / < /,


n, - Am, = 0 for some /.

Then n - Am is a null root. By the above argument n, —Am, = 0 for all /.


Thus m = An. This proves part (i). As for part (ii), since A has rational
coefficients and 0 e Q is an eigenvalue of ^ has a rational 0-eigenvector.
By part (i) it is a multiple of n. Thus we may assume first that n e then
by clearing denominators that n e and finally, removing common
factors, that n is as in part (ii). Uniqueness follows from part (i). □

The null root n = (« q, . . . , n^) of A described in part (ii) of last proposi­


tion will be called the null root of A, We call n, the mark of the / node of
the Coxeter-Dynkin diagram of A.
260 Lattices and Finite Root Systems

Example 2 Let

2 -1 0
-1 2 -3
0 -1 2

The Coxeter-Dynkin diagram of A is

00 0^0
1
-
2

and n = (1,2,3) is the null root, since = 0. The Coxeter-Dynkin diagram


with its marksisgiven by
1 2 3
0 0^0-
0 1 2

Next weclassify allCoxeter-Dynkin diagrams arising from affine matrices


A, Let r be the Coxeter-Dynkin diagram of A. The nodes are indexed from
0 to /, and we let be the mark of the i node.
To each (/,;)-arrow (an arrow from the i to the j node with i j) we
attach a weight n^/rij. Since there are ;)-arrows, and we have from
(4) that for all 0 < j < I,

n:
(7) L \Au\- = 2
¿= 0
i^j
We conclude that

1. The sum of all the weights o f the arrows arriving at any node is 2. If there
is a (y, /)-arrow there is also an (/, ;)-arrow and its weight is n^/n^ (the
reciprocal of nj nj ) . Together with property 1 this implies property 2.
2. The weight w o f any arrow satisfies \ < w < 2, and hence we have
property 3,
3. There are at most four arrows arriving at any node. Using properties 1
and 2, it follows that property 4 holds.
4. There exists no weight w satisfying f < w < 2, and hence by taking
reciprocals, we have property 5.
5. There exists no weight w satisfying \ < w < \ . Because of property 5 it is
impossible to complete a sum of weights lying between f and f to 2.
Thus we have property 6.
6. There exists no weight w satisfying j < vr < f , and hence taking recipro­
cals results in property 7.
7. There exists no weight w satisfying \ < w < \ .
3.5 Classification of Cartan Matrices and Finite Root Systems 261

By using the seven properties above, we investigate the diagrams that may
arise according to the number of arrows arriving to a node.

Four Arrows Arrive at a Node. Suppose that 4 arrows arrive at a node i. Each
of these arrows must have weight If one of these arrows is of type (;, /),
then there is at least one arrow of type (/, j) and each of these is of weight 2.
Hence there is a unique arrow of type (/, j) (of weight 2), and hence no other
arrow may arrive at j. The only possibilities are (each diagram is given with
its marks).

1 1
o 0
1
0 0 -0- 0 ^ 0
1 |2 1 |2 1

0 0
1 1

1 2 1 1 2 1

0 =>0 ^ 1 G® 0 0^0 -

1 2
Bcf> o » 0 -

Each of these diagrams is complete in the sense that it satisfies property 1.


(See below for some remarks on the labeling convention.)

Three Arrows Arrive at a Node. Let the weights of these arrows he w^ <
W2 <Wy Recall that w^ > ^. Suppose that w^ = ^. Then W2 w^ = f - Thus
W2 < I, and hence either W2 = or We see that the only possibilities
are

(w^,W2,w^) = {^ , ^ , 1) , or or

Suppose that vwj = f . Then W2 + w^ = j. Thus W2 < f , and hence w^


W2 = W3 = f . We have

{W^,W2,W3) =

Suppose w^ > f . Then W2 + w^ < j , and hence vi^2 < f which violates w^ <
W2. Having determined all possible weights we study each case separately
262 Lattices and Finite Root Systems

Case (j, 1). We start with

‘Q\ 2 2 1 2 2
p - O or 0 ^ 0 - 0
O

In either case all nodes except the extreme right-hand node are full (sum of
arriving weights = 2).
These diagrams can be extended to include an arbitrary number, say, /, of
nodes of the form

Q\ 2 2 2 1 2 2 2
/ 0 - 0 ---------------O " 0 ^ 0 - 0 - -o
0
1

and then completed to one of the following diagrams (with 1 + 1 nodes)

. a /O
0-0---- o
0
1
o 1

2 2 2 2

0 - 0 ----------- o « o
0
1

1 22 2 2
BCP>: 0 ^ 0 - 0 ----------0 ^ 0
1

1 2 2
O
0^0-0---- o'\
o
1 2 2 2 1
cp>: 0 ^ 0 - 0 ---------------- 0 ^ 0
3.5 Classification of Caitan Matrices and Finite Root Systems 263

Case ( |, f , f ). We start with

0—0—0
3 4 3
o—o<=o
In either case the node with a mark 2 is full. In the first diagram the nodes
with mark 3 have one arrow of weight | arriving at them. There must be one
more arrow of weight f arriving at them. This leads to

o-o-o-o-o
2 3 4 3 2

and, by a similar argument, completes to

0 0 0 6 0 0-0
1
-
2
-
3
-
4
-
3
-
2 1

The same type of reasoning shows that our other diagram leads to

O -O ^ O -O -O
2 4 3 2 1

Case (j, f , f ). We start with

o-o-o
5 6 4

The node with a 3 is full, and the diagram can only be completed to
3

0 0 0 0 0 6 0-0
1
-
2
-
3
-
4
-
5
-
6
-
4 2

Case ( f , f , f ). We start with


2 2 3 2 3 2
Q 0 - 0<=0 0^ 0
0 0-0 -
2 3 2
264 Lattices and Finite Root Systems

and these complete in unique ways to

?
£<i>:

0 0 6 0 -0
- - -

O -O -O ^ O -O

6 6-0
^

A t Most Two Arrows Arrive at Every Node. We start with one of the following

1 .1 1
A^P:
o-3!o
1 1 1

^ o ^ o = ^ o
0-0 ----- 0-0
1 1 1 1
or

0^0-0----
1 1 1
o1
The first two diagrams are complete, while the other two complete to one of

.o

0-0 ------- 0-0


1 1 1 1
or

o « = o - o -------------- 0 ^ 0 -
1 1 1 1 1

A t Most One Arrow Arrives at Every Node. Since the diagram is connected,
the only possibility is Q —Q , which is not affine.

No Arrows Arrive at Every Node. Since the diagram is connected, the only
possibility is Q , which is not affine.
3.5 Classification of Caitan Matrices and Finite Root Systems 265

Proposition 2
Let A be an (/ + 1) X (/ + 1) affine Cartan matrix. Then its Coxeter-Dynkin
diagram T is of the form where is in the following list {all diagrams
have / + 1 nodes and are given with their marks): In particular affine matrices
are symmetrizable.

1
,o /> 2

0-0 ---------- - 0-0


1 1 1 1

Q
\,
0 0 - — --------- 0 ^ 0 ^^ 3
2 2
o^^ ^
1

CP: 0 ^ 0 - 0 - - ------ 0 - 0 < ^ 0 ^^ 2


1 2 2 2 2 1

1 1

/O
O i” : ^ 0 - 0 ------- --------- 0 - 0 , ' i “

1 1

E<‘>;

o-o-O-o-o
1 2 3 2 1

E y '- - 0 0 0 6 0 0-0
1
-
2
-
3
-
4
-
3
-
2 1

^ 5 01 02 03 04 05 66 04 - 02
" ^ - - - - - -

01 0
2
03 0
-
4
-02 - = ^
266 Lattices and Finite Root Systems

G<»:
r: 0 - 0 ^ 0
1 2 3

O 0
1

1
Q
^ 0 - 0 ---------0 ^ 0 ^^ 3
2 2 2 1

0^ 0-0 ------------------------------------------------------------------ 0 - 0=^0 ^^2


1 1 1 1 1 1

fiC p : 0 ^ 0 - 0 - ---------0 ^ 0 ' ^ 2


1 2 2 2 2

o-o-o^o-o
1 2 3 2

B C P>=^P:

G p:
0 0^0-
1 2 1

Remark 1 The meaning of the number a e {1,2,3} appearing in can


be found in [Mo3]. An alternate nomenclature is used in [Ka5].
Remark 2 Suppose that A is an affine Cartan matrix. Then is clearly a
Cartan matrix. The Coxeter-Dynkin diagram of A ^ is obtained by reversing
the arrows on the diagram of A (since we are interchanging the roles of
Aij and Aji), By inspection in Proposition 2 it follows that A ^ is also
affine (or see Proposition 3.6.5). For example, has matrix
2-1 O]
^ = ^-3 2 - i j and diagram Q ^ Q _ Q , while A'^ has diagram
Q __Q and is affine of type G^^\

We now begin the classification of the Coxeter-Dynkin diagrams of all


finite root systems. Let A = (A^jX 1 < i, j < I, be the Cartan matrix of a
finite indecomposable reduced root system A arising from a base II =
( a j,. . . , of A. Thus

^ ij { }•
3.5 Classification of Cartan Matrices and Finite Root Systems 267

Let e A be chosen so that the height of 4> is maximal. That is,

ht(o;) < for all Of e A.

(We will see that <f) is unique.) Then for all e n we have ht < ht (f>,
and hence

( (f), a / ) > 0 for all 1 < i < 1.

If we choose a positive definite PT-invariant bilinear form as in Proposition


3.4.6, we have

> 0, 1<i <L

Define for all 1 < i < I,

^ 0/ = - I.. ^ = - ( « ; )
(a ,la ,)

and
2 (g ,l< /> )
^io —

We can then define an (/ + 1) X (/ + 1) Cartan matrix A by A^j = {A^j) for


all 0 < i, j < I where A qq == 2.

Lemma 3
A is an affine Cartan matrix.

Proof. That ^ is a Cartan matrix follows from construction. We show that A


is indecomposable. Indeed the 0 node of the Coxeter-Dynkin diagram f of
is connected to some other node for otherwise (<^|<^) = 0. With the 0 node
removed, the diagram left is connected (since A is indecomposable). Thus f
is connected, and hence A is indecomposable. To establish the lemma, we
must show that A admits a null root. Write

/
</) = ^ n^a^ and define -(f>.
¿=1

Clearly n^ > 0. Let Wq == 1, and let n == (nQ, n^,..., n^). Then for all 0 < ; < /
268 Lattices and Finite Root Systems

we have

= A qj + ^
1=0 /=1

= 0.
(« > ;) ( “ yl“ ;)
This shows that n A = 0 and hence that A is affine. □

Theorem 4
(i) Up to isomorphism there exists a 1-1 correspondence between finite
indecomposable reduced root systems and the diagrams listed below.
This correspondence is given by attaching to a root system its
Coxeter-Dynkin diagram with respect to any base,

0-0-0----------O
1 2 2 2
Br. 0 - 0 — 0 ^ 0
2 2 2 1
Q: 0 - 0 — 0<=0 ^^2
1

1 2 2/ o
Or- 0 - 0 ---------------- /> 4

0
1

1 2 ^ 2 1
o-o-o-o-o
z.

1 2 3 ?|4. 3 2
£7: o-o-o-o-o-o
2 3 4 5 s |6 4 2
o-o-o-o-o-o-o
2 3 4 2
f.- o-o=o-o
2 3
^2- 0 ^ 0
3.5 Classification of Cartan Matrices and Finite Root Systems 269

(ii) If A is an indecomposable reduced finite root system, H = • • • aj)


a base o f A, and T is its Coxeter-Dynkin diagram with respect to H,
then A has a unique root (f>whose height with respect to H is maximal.
Moreover, if

4> = H riiOLi,
i= l

then the n^ are the marks o f Y as they appear in the graphs o f part (/)
of the Theorem, and for all a E: ^ we have a < <f).

At this point we will prove all of Theorem 4 except the existence of the
root systems. There are two approaches to the existence problem. The most
straightforward is to construct each finite root system by brute force. This
method is outlined in the exercises and is highly recommended for getting a
feel of what these root systems look like. The second approach, which is
uniform, is to use the abstract theory of root systems developed in Chapter 5
together with some facts from the Perron-Frobenius theory of Section 3.6,
and some Lie algebra theory from Chapter 4. This is our approach (see
5.2.12).
We begin with some definitions and preliminary steps. Let A be an
indecomposable reduced finite root system, and let n = {a^,...,« /} be a
base of A. Let ( | • ) be a positive definite IT-invariant form as in Proposition
3.2.5. Then {a, a ^ ) = 2(a|a^)/(aja^) for all a e A, / = 1 ,..., /.
Define a E A to be dominant if a > r^a for all 1 < f < /. That is,
r^a = a - ka^, where A: > 0, or equivalently (a, a ^ } > 0 for all 1 < / < /, or
equivalently (a|a^) > 0 for all 1 < i < 1.

1. If a = is dominant, then m^ > 0 for all i:


First, 0 < (a |a ) = Urnfala^) together with (a|a^) > 0 for all i im­
plies that some > 0, and hence a e A+. Write {1, ...,/} = I U J,
where I = {i\m^ > 0} and J = {i\m^ = 0}. Since (a, a ^ ) > 0, we see
that <a,, a /> = 0 for all i e I and j e J. Since A is indecomposable,
we conclude that J is empty.
2. If a = and p = are both dominant then with respect
to the Q^-partial ordering (see Section 2.5) either a > ¡3 or p > a:
From argument 1 we have (a\p) > 0. Thus {a, =
2(alj8)/()3|/3) > 0, and hence a - p e A (Proposition 3.2.8). Suppose
> Pi for some i. Then a — p e A forces a - p e A+ c (2+. Thus
mi > Pi for all i.
3. Let a E A be of maximal height with respect to H. Then a is dominant.
Moreover (a |a ) > ip\p) for all roots P:
That a is dominant is clear. Suppose that A has roots of unequal
lengths. If e A is of length different than that of a, we must show
270 Lattices and Finite Root Systems

that (a\a) > (p\p). By replacing p by wp for some vv e we may


assume that p is dominant. Indeed, if p is not dominant, then
(p,a^^) < 0 for some i and r^p > p. Repeating this argument if
necessary, we obtain at last a dominant root. The argument of Step 2
shows that a - j8 e A^. Thus

(«1«) - (P\P) = ( a + P\a - P)

= {a\a —)8) + (P\a — p) > 0

(the last inequality since a and p are dominant and a — p ^ Q+),


Finally, (a\a) > (p\p), since a and p are supposed to have different
lengths.
4. There exists a unique root (f> = of maximal height. Moreover
i<i>\(t>) > {p\p) and 4> > P for all e A:
The first two assertions follow from arguments 2 and 3. Let j8 e A.
Using the same argument as in 3, we conclude that there is a dominant
root y with y > p. Now (j) > y > p hy 2.

The root (f> above is called the highest root of A (with respect to II). We
reserve the notation ..., for its coefficients. A root is called long if it
has the same square length as (f>.

5. If n^ = 1 for some 0 < f < / , then is long:


Note that Uq = —(!> is long by definition. If f > 0 and if is not a
long root, let i = Iq, ¿1, . . . , t)e a minimal set of distinct indices in
{ !,...,/} so that = («/J«/,) = ^
each 1 < j < k, A: , ¥= 0. Thus from the Table 3.1,

0 - 0 -----------0 - ^ 0
^0 ^k-l

with m > 1.

Then r,^0 r,^k-la:^k = ma:^0 + ma, + H-m^ + a, < S contradicts


n, = 1.

Proof of Theorem (except for existence). Let A be an indecomposable finite


root system, and let II = {a^,...,« /} be a base of A. Suppose that the Cartan
matrix of n (hence of A) is A. Let (f> = be the highest root of A
3.5 Classification of Caitan Matrices and Finite Root Systems 271

with respect to II. Using Lemma 3, form the affine matrix A, The Coxeter-
Dynkin diagram t of A is the Coxeter-Dynkin diagram of A extended by
one node, the extended node being determined by —(f>. The marks are
riQ= 1, /Ip , rii and —</) is a long root. We observe:

1. r is a connected subdiagram of f with one less node that f.


2. The extended node of f has mark = 1 (since Hq = 1).
3. The extended node of f has no multiple arrows into it (because — is
long).
4. r and f have the same number of different node square lengths (since
-(f> has the same square length as some i = 1, . . . , /).
5. f has marks = 1 only over nodes corresponding to long roots (by
argument 5 above).

Regarding claim 3, it is useful to observe that whenever we have the


appearance of two nodes like Q s ^ Q , Q ’ O
Coxeter-Dynkin diagram F or F, then, naming the corresponding roots a, ¡3,
we have 2(a\p)/(p\p) = —4, —3, or —2 and 2(p\a)/(a\a) = —1, whence
(a\a)/(p\p) = 4, 3, or 2, respectively. Thus the node has multiple arrows
into it only if there are roots of longer square length than a¿.
Now looking at the classification of affine matrices in Proposition 2, we
observe that claims 1-5 eliminate all the diagrams except those of the form
Moreover, up to a diagram automorphism these diagrams have a unique
node with mark = 1. The diagram Xi is obtained by removing such a node.

Remark 3 The number of nodes whose mark = 1 in is actually the
index of connection [P: Q], where P and Q are the weight and root lattices
of the root system of type Afp
Simply by staring at the diagrams in Theorem 4, we obtain

Corollary
Let A be an indecomposable reduced finite root system with Coxeter-Dynkin
diagram F. Then

(i) there are at most two different root lengths in A;


(ii) there is at most one multiple arrow in F;
(iii) there is at most one node that is connected to more than two nodes,
and such a node is connected to exactly three nodes;
(iv) the corresponding Cartan matrix is symmetrizable.
272 Lattices and Finite Root Systems

3.6 THE PERRON-FROBENIUS THEOREM


AND ITS CONSEQUENCES

On the set oi m X n real matrices we define two partial orderings > and ^
as follows: Given m X n matrices C = (c,,) and D =

C >D Cij > dij for all i <i < m and 1 < j < n,

C^ D either C = D or > d^j for all

1 <i <m and 1 < ; < n.

As usual we write C > D (respectively C > D) io denote that C ^ D and


C > D (respectively C D), We say that C is positive (resp. nonnegative) if
C > 0 (resp. C > 0, where 0 denotes the zero matrix of appropriate size.
If C in a square matrix, we say that C is primitive if 0 for some
A: e Z+ C is called semiprimitive if e l C is primitive for all e e
Throughout this section will denote the set of all column vectors
X = (jCi,. . . , such that x > 0.
The main result of this section is the following:

Theorem 1
Let C be a semiprimitive nonnegative n X n matrix. Then C has a real
eigenvalue r satisfying

(i) r > 0. Furthermore r > Q if C is primitive or if n > 1;


(ii) there exist positive left and right eigenvectors for r, each unique up to
positive scalar factors;
(iii) r > 1^1 for all other eigenvalues s of C and the multiplicity o f r is 1;
(iv) if C > D > 0 and d is an eigenvalue of D, then \d\ < r and \d\ = r if
and only if D = C.

The eigenvalue r is called the Perron-Frobenius eigenvalue of A.

Example 1 Let T be a finite labeled graph, and let C = C(T) be the


incidence matrix of T. Thus the vertices of T are labeled ! , . . . , « in some
way and is defined as the number of arrows from vertex i to vertex j for
i # j and c¿¿ = 0 for all i. Certainly C > 0. It is straightforward to see that C
is semiprimitive if and only if T is connected. Indeed, if « = 1, this is trivial.
Suppose that n > 1, C is semiprimitive and (e / + C)^ >- 0. Then for all
i, j, Ec;.,., • c;.Ill2 > 0, where the sum runs over all . . . , j ^[i,n]
3.6 The Perron-Frobenius Theorem and Its Consequences 273

and dpq '= (s i + C)p^. Observe that the inequality itself is independent of s
(> 0). Suppose that i ¥=j. Take any summand > 0, and
omit all factors that are diagonal entries (Cp^X The remaining product is still
positive, and by definition of C defines a path from vertex i to vertex j.
Conversely, suppose that we have a path of length m = m(i,j) from i to
Then we obtain a product ‘ j > 0, and we see that (C'”),y > 0.
Then, for all e Z+ and e > 0, ( ( e / + includes the summand
• ** c, fS — e > 0 (k factors of eX and hence we see > 0.
Now if r is connected taking M == max{m(/,y)}, it is easy to see that
(el + C )^ 0. Thus C is semiprimitive. In particular, if ^ is a Cartan
matrix, then C '= 2 —A is its incidence graph, and C is semiprimitive if and
only if the graph is connected (i.e., A is indecomposable). Later on we will
prove the remarkable fact that the Perron-Frobenius eigenvalue of C(T) is 2
if and only if r is the Coxeter-Dynkin diagram of an affine Cartan matrix.
Our proof of Theorem 1 follows [Sn]. We begin with a sequence of
lemmas. We assume first that C is primitive and that m has been chosen so
that > 0.

Lemma 2
If a> Q and Cx = ax for some x e q, then either x ^ Q or x = Q.

Proof If X ¥= 0, then a’^x = C ^x 0, so jc 0. □

Lemma 3
Let C j,. .. , C„ be the rows o f C, and define

by

I \ T
R ( x ) = m i n i -----> forallx = ( x ^ , . . . , x ^ ) .
I i
(IfXj = 0, CjX/Xj is understood to be +oo.) Then R assumes a maximum value
r at some point x e q\ {0}.

Proof Let M = max^ y{c,y}. Then for x e R > o \

R(x)Xj < CjX = ^ for each j = 1, . . . , n.


k

Adding, we obtain

R{ x) Y,Xj < nM(l,Xi^),


274 Lattices and Finite Root Systems

whence < nM. Thus R is bounded above. Define

r '= sup

Evidently 00 > r > 0.


Since each of the functions CjX/Xj is continuous where it is defined, R is
continuous. Furthermore R(x) = R{ax) for all a e and hence

r = sup
x^S
where

S ’= [R>o ^ [x\x^x = 1}.

The latter being compact, R assumes a maximum on S, hence on R>o\№}.


This proves the existence of x as required. □

Henceforth r will be the quantity defined in Lemma 3.

Lemma 4
If X e \ {0} if rx < Cc, then rx = Cx and x 0.

Proof By assumption (C - r)x > 0. If this is not zero, then

C(C'”x) - r{C^x) = C^{C - r ) ( x ) > 0.

Setting x' := C"*x, we have rx' < Cx' and hence rx'j < Cjx' for all ;, contrary
to the definition of r. Thus (C - r)x = 0 and x 0 by Lemma 2.

Proof (of Theorem 1). We begin by assuming that C is primitive. By the


definition of X (Lemma 3), rx < Cx and hence by Lemma 3.4,

rx = Cx, X 0.

Suppose that Cx = sx for some x e C" \ {0}, ^ e C. Then for each j =

( 1) \sxj\ = |C^-x| < Y,Cji\Xi\

(= +ooii Xj = 0).
3.6 The Perron-Frobenius Theorem and Its Consequences 275

Set

Then

\s\ < R { x ) < r.

Now if \s\ = r, then from (1)

rx < Cx
so by Lemma 4

rx = Cx, jc >- 0.

Set Then

Since the Xy e C \ {0} and the > 0, we see from this equality that
jCj,..., all have the same argument, say, 6. Thus Xj = e'^\xj\, j = 1, . . . ,
and X = e^^x. Hence Cx = sx => Cx = sx, which proves that s is real and
finally that s = r.
Suppose that x, x' are nonzero eigenvectors for r. We can assume from
the argument just given that x, x' > 0 and choose a e R minimal so that
some component of x — ax' equals 0. Then x — ax' > 0, and by Lemma 2,
X - ax' = 0. This proves that x and x' are linearly dependent (hence
multiples of Jc) and the multiplicity of r is one.
Replacing C by C^, we obtain a positive left eigenvector for C with
Perron-Frobenius eigenvalue r'. Thus all eigenvalues s of satisfy \s\ < r'.
Since C and have the same eigenvalues, r = r'. This concludes the proof
of parts (0, (ii), and (iii) in the primitive case.
For part (iii) suppose that D < C and that D > 0. Suppose that y is a
nonzero eigenvector for D with eigenvalue d. Then using (1), we get

Idly < D y < C y ,

where y = ( ly j,. . . , ly„l)^.


Let 3c^ be a positive left eigenvector for C with eigenvalue r. Then

|dl;c^y < x^Dy < x^Cy = rx^y.


276 Lattices and Finite Root Systems

SO \d\ < r. If \d\ = r, then ry < Cy, and by Lemma 4,

ry = Cy, y > 0.
Thus
ry < Dy < Cy = ry.
It follows that Dy = Cy, and since all entires of y are positive, D = C.
Now suppose that C is semiprimitive. Then for all e > 0, C(e) ■= el + C
is primitive. Now C(e) has the same eigenvectors as C and has all its
eigenvalues shifted by e. Thus, if x is the right Perron-Frobenius eigenvector
of C(e) for some, hence all, e > 0 and the corresponding eigenvalue is r^, we
have
C ( s ) x = ex + Cx = r^x
so
Cx = rx.
where r ‘= r^ - e is independent of e. Obviously r > 0, since > 0. If
n > 1, then some row, say, Q, of C is not 0, and 'LJ^iC^Xj = rxj together
with X > 0 gives r > 0.
The remaining statements, parts (ii)-(iv), are now obvious. For part (iv) we
use C(e) > D(e) > 0 for all e > 0. □

Let A be an indecomposable Cartan matrix, and write A = 21 — P. Then,


as we saw in Example 1, P > 0 and P is semiprimitive. We let r = r{A) be
defined as the Perron-Frobenius eigenvalue of P.

A is called of finite (resp. affine, indefinite) type according as

r{ A) < 2 (resp.= 2, > 2).

The apparent conflict in the use of the word affine is resolved in the
following proposition.

Proposition 5 [Vi]
Let A be indecomposable Cartan matrix indexed by J = { ! ,..., /}. One of the
three mutually exclusive cases occurs:
(i) Finite. A is o f finite type

<=> e \ {0}, jc > 0 with Ax < 0,


<=> e IR^\ {0}, X > 0 with Ax > 0, Ax 0,
<=>3 x ^ with Ax > 0.
3.6 The Perron-Frobenius Theorem and Its Consequences 111

(ii) Affine. A is o f affine type

<=> \ {0}, a: > 0 with Ax = 0,


<=> G -+ with Ax = 0.

(iii) Indefinite. A is o f indefinite type

<=» 3 j : e R ^ \ {0}, a: > 0 with Ax < 0, Ax 0,


«=> G \ {0}, jc > 0 with Ax > 0,
<=> 3 x G Z+ with Ax < 0.

Moreover, A and A ^ fall under the same case.

Proof Write A = 21 - P, and let r > 0 be the Perron-Frobenius eigenvalue


of P. Let u and u be associated left and right Perron-Frobenius eigenvectors
for P. Thus w 0, i; 0 and uP = ru, Pv = ru.
If r = 2, then uP = 2u => uA = 0 A is affine. Conversely, if A is affine
with null root n, then nP = 2n and 2n • i; = nPi; = rn ' v. Since n • i; > 0,
r = 2. Thus

A affine r{ A ) = 2.

The rest of the affine case is already done in Proposition 3.5.1.


Henceforth assume that A is not affine so that r # 2. Suppose that
AG R^ \ {0}, a : > 0, and Ax < 0. Then A x < 0 = > P x > 2 x = > uPx > 2wc =>
rwc > 2wc => r > 2. Thus A is indefinite. In precisely the same way,
A G R^ \ {0}, A > 0, and Ax > Q =>A is finite. This proves all the implica­
tions “ < = 3 a ...
Suppose that A is indefinite. Then the set

F := {a G WjPx > 2 a }

is a nonempty open cone (note that u ^ V) and hence there exists a point
AG K n Q +. Scaling A , we can assume that a g Z +. Then ^ 0. In
precisely the same way, A is finite => 3 a g Z+ with Ax > 0. This proves all
the implications “ => 3 a ... The remaining implications now follow auto­
matically. □

Remark 1 Proposition 5 is a version of Vinberg’s lemma [Vi]. We have


phrased it in terms of Cartan matrices in order to make it fit most easily into
the material to follow. Vinberg uses some properties of dual convex cone
theory to establish his lemma, although he points out that the proof can be
carried out by the Perron-Frobenius theory, as we have done.
278 Lattices and Finite Root Systems

We have already classified the Cartan matrices of affine type. We now


show that the Cartan matrices of finite type are precisely those whose
diagrams appear in Theorem 3.5.4.

Let A = (Aij) and A' = (A]j) be Cartan matrices of dimensions / X /


and r X /', respectively. We say that A is inferior to A if there is an
injective map

such that

IA'a \ < I I for 1 < i, j < V.

Thus A is inferior to A if the Coxeter-Dynkin diagram of A can be


obtained from that of A by removing nodes and arrows.

Proposition 6
Let A be an indecomposable Cartan matrix with Coxeter-Dynkin diagram T.

(i) I f r appears in the list o f Theorem 3.5.4, then A is inferior to some


affine matrix,
(ii) I f r does not appear in the list o f Theorem 3.5.4, then there is an affine
matrix inferior to A,

Proof (i) The proof is clear by the way in which we obtained the list of
Theorem 3.5.4.
(ii) Suppose that T is not in the list of Theorem 3.5.4. If T contains a cycle
then some A^l^, A: > 2, is inferior to F. If F contains Q Q with ab > 4,
then either Q^ O O^ O inferior to F. If F contains a node of
valence > 4 (i.e., a node with 4 or more adjacent nodes) or two nodes of
valence 3, then is inferior for some k. If there are two multiple edges
(i.e., O # 0> !>’ then O = 0 - 0 --- 0 - 0 = O
some orientation of the arrows is inferior. If there is a single multiple edge
and a node of valence 3, then

,o
o =o-o-o-o;
'O

(again with some orientation) is inferior. If there is a single multiple edge and
every node is of valence < 3, then Q — Q — 0 ^ 0 — inferior
since Bi, Cl, and are excluded by hypothesis. Finally, we are reduced to
3.6 The Perron-Frobenius Theorem and Its Consequences 279

the case of only single edges, and the hypothesis leads us to conclude that
one of or is inferior. □

Let A' = 21 — P \ A = 21 — P be two indecomposable Cartan matrices


of dimensions /' X /' and / X /, respectively. Then A' is inferior to A if and
only if there is an injective map *: /'} ---- > such that F/y <
for all 1 < i, j < /'. Then /' < /, and we may assume, if we wish, that
/* = i for each i. If I' < I , then extend P' to an / X / matrix P' by filling the
new entries with zeros. Then 0 < P' < P, and the spectrum (= the set of
eigenvalues) of P' differs from that of P' only by the possible addition of
zero. By part iii of Theorem 1 all eigenvalues d of P' satisfy d < r(A)
(:= Perron-Frobenius eigenvalue of P), and equality occurs if and only if
?' = P. It follows that KA'X the Perron-Frobenius eigenvalue of P \ satisfies

r(A') < r ( A )

with equality if and only if ^ ' = ^4. Thus

Proposition 7
Let A and A be indecomposable Cartan matrices. Then A inferior to A =>
r(A) < r(A) with equality if and only if A = A.

Putting Propositions 6 and 7 together, we have

Proposition 8
The Cartan matrices o f finite type are precisely those whose diagrams are listed
in Theorem 3.5.4.

Proposition 9
Let A be a symmetrizable indecomposable Cartan matrix, and suppose that A is
symmetrized by e = diagie^,. . . , e^}, where each > 0. Then

(i) Ae is positive definite <=>A is o f finite type,


(ii) A s is positive semidefinite A is o f affine type,
(iii) Ae is indefinite A is o f indefinite type.

Furthermore the eigenvalues o f A are all real.

Proof Let sY'^ be a square root of s^, and let s^^'^ -= diag{sj/^,. . . , sY^}.
Then the equations A^Sj = A -^s^ show that is symmetric, and of
course A is similar to Furthermore for y e U^, y^(As)y =
so A s is positive definite, positive semidefinite,
or indefinite as s'^'^^Ae^^^ is. But the latter is determined by the eigenvalues
280 Lattices and Finite Root Systems

of being all positive, all nonnegative, or not all nonnegative,


respectively. However, the spectrum S p e c i e o f is real
and is equal to S pecif) = Spec(2 —P). Since the eigenvalues of P are
maximized by the Perron-Frobenius eigenvalue r, Spec(2 —P) is minimized
by 2 —r. The result follows from the definition of finite type, affine type,
indefinite type. □

The approach to Cartan matrices through the use of the Perron-Frobenius


theory came to our attention through the work of N. Iwahori [Iw]. The theory
of the spectra of graphs is quite extensive; see, for instance, [CDS].

3 .7 CONSTRUCTING L IE ALG E BR A S F RO M LATTICES

In this section we show how one can construct a Lie algebra from any
positive definite, even geometric, lattice L by using the elements of norm 2 in
L. The construction produces a free Z-module of finite rank (a finitely
generated torsion-free abelian group) of the form

( 1) 9^ = L ® ©
e L (2 )

together with a multiplication

[•, •]: 9 l X 9^^ ^ 9z.

that satisfies the two Lie identities: skew-symmetry and the Jacobi identity. In
as much as our scalars are integers rather than elements of some field, is
a Lie algebra over Z, sometimes called a Lie ring. This Lie ring (see Section
1.1) will be graded by L with as the elements of degree a and L itself
as the elements of degree 0. Of course should it happen that L(2) = 0 , none
of this would be of any interest. Henceforth we will assume that L(2) 0.
One easily obtains L-graded Lie algebras Qo< over a field IK(even if IKis of
characteristic 0) by extension of the base ring (see Section 1.1)

9iK == = ( 11^ ® ©
e L (2 )

The Lie algebras obtained in this way from the lattices of types ^4/, Z)/, £5,
Ej, and £’g are the Lie algebras of the same name, and they provide us with
about one-half of the so-called finite-dimensional split simple Lie algebras.
Our approach follows [FLM].
We make the following observation: Suppose that ( L , ( | • )) is a positive
definite geometric lattice. Then for a, /3 e L(2) precisely one of the following
3.7 C onstructing Lie Algebras from Lattices 281

holds:

(2) a + /3 ^ U {0} and (a|/3) > 0,


a p ^ and («1/3) = - 1 ,
a p = 0 and (a|/3) = - 2 .

by the Schwarz inequality.

|( a |^ ) |< ( ( a k ) ( / 3 |/ 3 ) ) ‘/^ = 2

with equality if and only if a and p are linearly dependent, and (2) follows at
once.
From the point of view of defining the multiplication on the main task
is to define the products [X^, X^]. Of these the most interesting is the case
when (a\^) = —1. Then

[ X ^ ,X , ] =

The choice of sign is a very interesting problem which ultimately is related to


making a suitable central extension of L by Z/2Z. The details of this are left
to Section 3.8, where we study such central extensions more carefully. For
now it suffices to assume that there is a biadditive mapping

s: L X L -
satisfying

(3) e (a , a ) = | ( a |a ) (mod2) fo ra lla e L .

and hence also

(4) e(a, p) + e(/3, a) = (a|j8) (m od2) for all a, p ^ L

We then define [X^, X^] = ( — for (a\p) = —1. Notice how


skew-symmetry follows form (4).
The last part of the section is devoted to determining the basic properties
of like showing that admits a triangular decomposition with an
anti-involution o) interchanging X^ and - X _ ^ for all a e L(2). As we know
L(2) is a finite root system in its Q-span. Once a base {a^,. . . , a/} of this root
system is chosen, we use it to determine a set of Lie algebra generators of g^^
and a set of relations, which, as we will see in Chapter 4, actually gives a
presentation of g^.
Since L(2) is a root system, we call the elements of L(2) roots. Define g^
by (1).
282 Lattices and Finite Root Systems

Define [•, • ] on as follows:

[L,L] = (0),
[a, Xp] = = - [Xp, a] for all a e L, j8 e L(2);
(5) 0 if(«l/3)>0,
Xjg] = I if ( a |/ 3 ) = - l , for all a,/3 e L(2),
-a if( a |/3 ) = - 2 .

Proposition 1
with the bracket defined by bilinear extension o f (5), is a Lie ring.

Proof. First we look at the skew-symmetry. The only nontrivial case is when
(a\fi) = —1. Then, as we noted above,

Note that (a|j8) = - 2 happens precisely when j8 = - a in view of the fact


that (• I ■) is positive definite.
The Jacobi identity involves looking at a number of cases. The hard part is
checking that

(6 ) [x.,[x,.x,]\ + -0

when a, j8, y e L(2) and a + /3 + y ^ f^(2). There are three subcases:


(1) two pairs of roots add to 0, (2) exactly one pair of roots adds to 0, (3) no
pair of roots adds to 0.

1. Suppose that a + /3 = 0 = ^ + 'y- Then a = y and (6) reads

which trivially holds.


2. Suppose that a + /3 = 0. Then /3 + y = - a + y, and certainly at most
one of a + y and - a + y may be a root [we need (±a:|y) = -1 ]. If neither
is a root, then by (2) either (a\y) = 0 or a = +y. Since we are in case 2, the
3.7 Constructing Lie Algebras from Lattices 283

latter is impossible, and hence (a\y) = 0. Also (p\y) = so clearly (6) reads

[X ^,[X ^,X .J]= 0.

We may assume then that exactly one of + a + y is a root, say, a + y (the


other case is similar, since then j8 + y is a root). Now the left-hand side of
(6) reads

[ x _ ,,[ x ^ ,x ,]] + [X ^,[X ,,X _J]

= 0,

since e(a,a) = ^(ala) mod 2 = 1 mod 2 and (a|y) = ~ 1-


3. If none of a + ^ , ^ + y , y + a are roots then (6) is trivially true.
Suppose then that /3 -I- y is a root. Then (a|/3) -t- (a|y ) = («1^ + y) = —1,
since a -t- /3 -I- y e L(2). Because we are in case 3, (a\P) > —1, (a|y ) > —1.
The only possible solution is for («1)3) = 0 (resp. —1) (“ ly) = ~ 1 (resp. 0).
Assume the first case. We have then

(a|/3) = 0 , ( a |y ) = - 1 , and (^ ly ) = “ 1-

The left-hand side of (6) reads

(7) {( - ^

Adding e(a, ^ + y) -I- e(/3, y) = e(a, /3) -f e(a, y) + y +


+ e(y, a) = sip, y) -I- e(/3, a ) + e(y, a), we get

e ( a , p ) + £(j8, a ) -I- e ( a ,y ) -I- c(y>“ )


= ((ali3) -I- (a ly )) mod2 = lm od2,

which shows that (7) is 0.


The other case is similar.

Define a bijective linear mapping

9l 0L
284 Lattices and Finite Root Systems

by
= for all a g L (2) ,
« li = id.

It is easy to see that a> is an anti-involution of g^: Obviously it has order 2,


and we have

[<u(Z^),a»(a)] = = [a,X_^]

= - ( a \ P ) X _ p = (a\l3)<a{Xp)

= (o{[a,X^])

and

= [X _,,x_„]

if(a|/3 ) > 0 ,
if(a|/3 ) = - 1 ,
if(a |j8) = - 2.

= <o {[X^ ,X,]).

Next we extend the bilinear form (• | • ) on L to all of 9^ by defining

(/3|Z^) = { X JP ) = 0 for all a e L(2), /3 g L,


—1 if a + ¡3 = 0,
W x ,) =
0 otherwise.

Proposition 2
The symmetric bilinear form on 9^ is even, nondegenerate, and invariant.
It admits co as an isometry. Relative to {'V) we have the orthogonal decomposi­
tion

g^ = L ± ± (ZAT, e Z 2 f _ J .
{a, - a )

Each submodule 0 Z X _^ is a hyperbolic plane. The determinant of (-I*)


on is ( - 1)^ det(L), where k = \ card(L(2)X

Proof. The orthogonal decomposition is obvious, and it is clear that 0


ZZ_^ is a hyperbolic plane (Section 3.1). The nondegeneracy follows imme­
diately, and the statement about determinants is left as an exercise. The
3.7 Constructing Lie Algebras from Lattices 285

invariance condition has to be checked on a basis made up of the and a


basis of L. There are four cases. Case (;), j = 0,1,2,3, accordingly as there
are 0, 1, 2, or 3 basis elements of L in the expression

( [ a , 6 ]|c) + ( è |[ a ,c ] ) .

We treat only the case (0), which is also the most involved. We suppose that
we are looking at

(8)

We have to show that it is 0. If a + j8 + y 0, this is automatic. Suppose


that a + /3 + y = 0. Then - y = a + /3 e L(2), and (8) becomes

(9) - ( - 1)
eioc, /3)
( - 1)
e(a,y)

However,

£(oc,p) + e ( a ,y ) = e ( a ,j8) + e{a, - a ) + s (a , -/3 )


= e(a, —a) = s(a, a ) = 1 (mod 2),

which shows that (9) reduces to 0. Finally,

= (x jx ,),

from which we can see that co is an isometry. □

Using the results of Section 3.3, we choose a system of positive roots


L(2)+ and a base c L(2)+. Then every element of L(2)+ is a
nonnegative integral linear combination of a ¡ and

L(2) = L(2) + U - L(2) ^ , L(2) + n - L(2) + = 0 .

Let 9 - '“ ®aeL(2)+ ^ ^ -a “ ^^9+)- Then and g_


are obviously Z-subalgebras of and

= g_0 L 0 g+.

The spaces ZX^ are root spaces in the usual sense that they are eigenspaces
for ad L (strictly speaking we should be talking about Z-modules rather than
spaces). The corresponding eigenfunctions are the functionals á: h ^ {a\h).
We usually identify á and a. With this convention we have a triangular
decomposition (g+, L, (2+, co) of g^ with root lattice Q = ©/^jZa¿. (Here
286 Lattices and Finite Root Systems

we are working with a Lie algebras over Z rather than over some field; see
Section 2.1.) Note that

[K 9^ = IK 0^ 9_© K L ®K

is a triangular decomposition of K 0^ 9 l *
In the case where L is generated (as an abelian group) by L(2), we have
Q = L. This can easily be arranged by replacing L by the Z-span of its norm
2 vectors from the outset.

Proposition 3
Suppose that L is an even positive-definite geometric lattice. Let {a^,... ,a¡\ be
a base of the finite root system L(2). Let

be the corresponding triangular decomposition o f the associated Lie algebra


Then

(i) = 1, . . . , /} generates g+ and = 1, . . . , /} generates g_


as Lie algebras over Z;
(ii) ifL(2) generates L (as an abelian group), then {X^,, X _ ^fi = 1 ,..., /}
generate as a Lie algebra over Z;
(hi) if L = L^ L L2 (see Section 3.1), then 9^ = 9^^ X 9^^^.

Proof (i) It suffices to prove the first statement. Let A+ be the set of positive
roots of L(2) relative to the base {a^,. . . , and let m be the subalgebra of
9^ generated by X ^^,. . . , X^^. Let a = Xc^a^ e A+ be a root of height k > 1
(see Section 3.3 for the definition of height). By a standard argument
(Proposition 3.3.4) we have that (alay) > 1 for some 1 < j < I and hence that
(alay) = 1 by the Schwarz inequality. It follows that

a — Oj = rjU ^ L (2).

Since ht(a - af) = k - 1 > 0, a - ay e A+, we may assume inductively that


^ Then

shows that g m. It follows that m = 9+.


(ii) Use part (i) and the relations [X^,, X_^] = -a^.
(iii) If L = Lj _L L 2, then L(2) = L^d) V ¿>2(2). The subalgebra 9 ^; i =
1,2 of 9^ spanned by L^ and [X^\a g L f l ) ) is evidently isomorphic to g^.,
and it is clear that 9^ = 9 1 © 92 and [9^, 92] = (0). □
3.7 Constructing Lie Algebras from Lattices 287

Theorem 4
Let L be an even positive definite geometric lattice. Let L(2) be the set of
elements of norm 2 in L, and let L be the sublattice o f L spanned by L(2). Let
K be a field o f characteristic 0, and consider the Lie algebras

g= K
§ = (K 02
where and are the Z~Lie algebras o f the lattices L and L . Let 5 be the
centre of g. Then

(i) g = 3 X
(ii) 3 is the subspace o f K L orthogonal to K ^ respect to
the invariant bilinear form o f given by Proposition 2. Moreover
dim(3) = rank(L) - rank(L'), and [g ,g ] =

(iii) ^ is semisimple. I f L(2) = Aj V • • • V is the decomposition of


L(2) into indecomposable root systems and L^ is the sublattice o f L
generated by A^, then = IK 9/, is a simple Lie algebra and
^ X *** X

In particular g is simple if and only if the root system L(2) is indecomposable


and spans L.

Proof Let I) = K L. Then

g = i) © © KX^
aeL(2)
is a triangular decomposition of g. Let a: e 3, and write

X=a ^ x “,
aeL(2)
where a and a:“ e By the way the multiplication in g is defined it
is clear that all the ;c“ = 0 and that (ala) = 0 for all a e L(2). In other
words, 3 c L' , where

L'^ := (jc © f)l(A:|y) = 0 for all y © L'}.


Conversely, it is clear that L' c 3, so we have 3 = L' and dim(3) = dim()^)
- rank(L') = rank(L) —rank(L'). Moreover

g = 3 ® (K 02 L') © = 3®

and since [3, §] = (0), we have g = 3 X


288 Lattices and Finite Root Systems

From Proposition 3, ^ X ••• X Suppose that we know that each


is simple. Then, since 3 = (0) ^ L = L', g is simple if and only if L = L'
and L(2) is indecomposable.
To finish the proof of the theorem, we prove that each is simple. For
this we may assume that L is generated by L(2), that L(2) is indecompos­
able, and that g = § =
Let a be a nonzero ideal in We begin by showing that a n L # 0. As
above, let ]^ = IK 02 Then a is an 1^-module by the adjoint representation
and hence is a sum of 1^-weight spaces (Proposition 2.1.1). The weight spaces
of § are f) and the one-dimensional spaces KX^, a ^ L(2). If e a, then
[X_^, X^] = a e a. Thus in all cases a n L ¥= 0.
Let {«1, . . . , a/} be a base of L(2), and let 0 9^= g a n L. Then for at
least one ¿, (j8|a^) 0, and we have [X^,,p] = -(a^\IB)X^, e a. Thus suc­
cessively X^,, X^], and X_^_ = a-]/2 lie in a. The argu­
ment can be repeated to deduce that X^,, aj, X_^. are in a if (aj\a^) 4^ 0,
that is, if «y and are joined by an edge in the Coxeter-Dynkin diagram.
Since L(2) is indecomposable, the Coxeter-Dynkin diagram is connected
(Proposition 3.4.4) and hence all the X lie in a. By Proposition 3, a = ^.

Remark 1 We have shown that the Lie algebras of type Ai, Di, F7,
and £g over fields of characteristic 0 are simple. It is not hard to consider­
ably relax the restriction on the characteristic.

In the light of Proposition 3, we have a set of Lie algebra generators of


which arise from a choice of a base for the underlying root system L(2).
There is obviously something rather special about this since we know that
bases of L(2) are all conjugate by the Weyl group W, In the next section,
where we look at the way in which the sign problem is solved, we will also see
that W lifts (in a slightly subtle way) to a group of automorphisms of g^.
These automorphisms take a given set of generators into some other
sets {±X^^^J, w ^ W, showing that all the generating sets obtained in
choosing a base of L(2) are essentially alike. This means that any relations
that we find amongst the are essentially invariants of the algebra
(relative to L).
Let L be an even positive definite geometric lattice which is generated by
L(2). Let {«1, . . . , a/} be a base of L(2), and let A = (A^j) be the Cartan
matrix of L(2): A^j = (ajay). Set
( 10 ) fi a/’ ^ 1, . . . ,/.
Then we easily have
(11) [hi,ej] =Aji€j,
[ h i , f j ] = - A j i f j ,

[ci, fj] = for all I, y e ( 1, . . . , .


3.8 Central Extensions of Lattices 289

Now consider i * j. If = (ay|a,) = 0, then a, + i A and


= 0. If = - 1, then a^ + aj e A, but 2a, + aj i A, so [e,,[c,, e^]]
= 0. Since these are the only possibilities, we obtain

( 12 ) (ade,.) = 0 for all i j,

and similarly

(a d /,)-^ " " V , = 0 for all i j.

The relations (11) and (12) provide the correct starting point for constructing
a far more general class of Lie algebras. This is the subject of Chapter 4. The
apparently unnatural ordering of the indices which appears in (11) and (12) is
chosen to conform with this more general setting. Of course it is immaterial
here, since A^j = Aj^.

3.8 CENTRAL EXTENSIONS OF LATTICES

Let L be an abelian group, and let be a central extension of L by a


(necessarily abelian) group A. By definition this means that we have an exact
sequence of groups

( 1) 0 0

and that i{A) lies in the center of .


It is convenient to adopt additive notation for L and A and multiplicative
notation for (which will not be abelian in general). Let (f>: L ^ be a
section of L in L " , that is, an arbitrary mapping satisfying tt • (¡>= id^.
For convenience we assume that <^(0) = 1. Then every element of L" is
uniquely expressible as for some a ^ L and a E: A. When the maps
(j>and i are understood, we will sometimes, for convenience, write (a, a)
instead of
Given a central extension L" of L by yl, as above, and a section of (f) of
L in L ^, we define a mapping

e: L X L >A

by

[This definition makes sense, since i is injective and <f>(a + p)~^(f>(a)(f>(p) e


ker(7r) = i{A).] If our central extension is fixed, then the map s depends on
the choice of section (f>. Later in this section we will discuss this dependence
290 Lattices and Finite Root Systems

in more detail. Notice that

( a , a ) ( ^ , b ) = <i>(a)i(a)<l>(p)i(b)
= <f>(a)4>(P)i(a)i(b)
= <f){a + I3)i{e{a, ¡3) + a + b).
Thus

(2) ( a , a ) ( ^ , b ) = {a + p , e { a , p ) + a + b),
and therefore s allows us to understand the structure of L ^ in terms of that
of A and L.
It is a straightforward consequence of the associative law that s satisfies
the condition

(3)
e(a -h p , y ) + e (a , = e ( a ,)8 + y) s(P,y) for all a, p , y G L.
A map e: L X L ^ A satisfying condition (3) is called a 2-cocycle on L
with values in A. One knows (see [Ja2], Chapter 6 [Ro]) that any such map
can be used to make the set L X ^ into a group by using (2) as multiplica­
tion; moreover the resulting group L " is then a central extension of L by A.
A simple way to obtain a 2-cocycle is to take any biadditive mapping s:
L X L -> A. Then e{a p , y ) = e(a, y) + y) and e(a, /3 + y) =
s(a, p) + s(a, y) so that (3) trivially holds, and one can easily see how the
above construction makes into a central extension of L by
We are interested in the case A = Z/2Z. For our purposes it is enough to
assume that e is bilinear (this is in fact not a restriction). Then e(2a, j8) = 0
= e(a, 20) for all a, p ^ L, so s determines a bilinear form
e: L / 2 L X L/2L —
Conversely, any bilinear form s on L / 2 L lifts uniquely to a Z-bilinear
mapping on L with values in Z/2Z.
At this point we introduce quadratic forms over the field F2 == Z/2Z. This
theory was originally worked out by Chevalley [Chi]. The principal thing to
be aware of here is that the usual relationship between quadratic forms and
symmetric bilinear forms is no longer valid. Instead we have a relationship of
quadratic forms to alternating forms and arbitrary bilinear forms to quadratic
forms.
Let K be a vector space over F2. A mapping
q:V F.
is called a quadratic form on V if the associated mapping
b : V x V ---- > F,
3.8 Central Extensions of Lattices 291

defined by

(^) + — —^ (y )

is bilinear. It is evidently symmetric. Furthermore

0 = ¿ ( 0, 0) = q(0) + g (0) + ^ (0) = > q{0) = 0,

whence b(x, x) = qix) + qix) = 0 for all x, and hence b is alternating.


The set Q(V) of all quadratic forms on V is also a vector space over F2 if
we use pointwise addition of functions. Let B(V) denote the vector space of
all bilinear maps on V, and let denote the subspace of alternating
bilinear forms b on V. Obviously the bilinear forms in B%V) are symmetric.
By definition we have a linear map

II : 0 ( F ) ~ ^ B \ V ) ,

which attaches to a quadratic form its associated bilinear form. Notice that
II (q) = 0 if and only if ^ is a linear map (linear maps are quadratic forms).
A quadratic form is entirely determined by knowing the values q(xi) and
b{Xi, Xj) on a basis {x,}, s j of F. In fact this follows simply from the identity

q( x + y ) = q ( x ) + q{y) + b { x , y ) .

Given any bilinear form s on F, we may define a quadratic form q^ by

= e{x,x)

thus establishing a linear map □: B ( V ) ---- > 0 ( F ) with kernel precisely


B%V). The associated bilinear form of q^ is b^, where

(5) be(x, y) = s{x, y) + e{y, x).

In the next result we put this backwards: Given a quadratic form q, find a
bilinear form e such that q = q^-

Proposition I

(i) 0 ^ B%V) B(V) 0 (F ) 0 is an exact sequence-,


(ii) if dim(F) = «<<», then dim B%V) |

Proof Let q G 0 (F ). Let {x,|i e J} be a basis for F over F2, where J is some
well-ordered set. We define s by establishing values for £(x„ X j ) for all
292 Lattices and Finite Root Systems

/, j e J. Choose six^, Xj), i < j\ arbitrarily and define

e(xj, Xi) = b(Xi, Xj) + s(Xi, Xj) for / <j,

where b is the bilinear form associated with q. This is forced by (5). Finally,
define
e(x,,jc,) =q(Xi).

Consider the quadratic form defined by e. It agrees with q on the x^ and


its associated bilinear form, given by (5), agrees with b on the basis {;cJ. Thus
q^ = q, and
q ( x) = s{x, x) for all x ^ V,

Obviously, if card(J) = n, then there are 2( 2) choices for the bilinear form s.

We now turn to the problem that was left unresolved in Section 3.7: to
determine a cocycle for a geometric lattice (L,(*| • )) that satisfies equations
(3.7.3) and (3.7.4).
Proposition 2
Let L be an even geometric lattice. Then there exists a central extension L" of L
by Z/2Z with cocycle s : L X L ^ Z/2Z satisfying

(i) e(a, j8) + s(p, a) = (a|j8) mod 2


(ii) s(a,a) = I (a |a ) mod 2

for all a, p ^ L.
Proof Let L = L /2 L , and let : L L be the natural homomorphism. L is
a vector space over F2 = Z/2Z. Define

q :L F.
by
q{a) = ^(a\a) mod2.

This is a quadratic form with associated bilinear form

b ( a , p ) = (a\p) mod2.

By the last proposition there is a bilinear form e: L X L F2 such that

q(a ) = s(a, a) for all a ^ L,


3.8 Central Extensions of Lattices 293

and we know that

(6) e (a , P) + e()8, a ) = ¿ ( a , jS) = (a|/3) mod 2, and


(7) e ( a , a ) = j( a \ a ) mod2.
Finally define e: L X L ^ Z/2Z by s(a, p) = e(a, $), □

Remark 1 In order to avoid a lot of cuml^rsome notation we will often


omit the overbars in symbols like q{a), b(a, p) writing simply q(a), b(a, p),
and so on instead.
Remark 2 The proof of Proposition 2 can be used, with only minor
modification, to show that for every quadratic form q on L / 2 L there is a
2-cocycle e on L with values in Z/2Z such that e(a, a) = q{a) for all a. The
second cohomology group H \ L / 2 L , Z / 2 Z ) of L / 2 L and the space Q of
quadratic forms in L /2L are isomorphic as groups.
The construction of L" leaves open the question of uniqueness, since
there are many choices for a cocycle satisfying condition (7). To establish the
fact that L" is unique up to isomorphism, we need Proposition 3.
Proposition 3
Let V be a vector space over F2, and let b be any alternating bilinear form on V,
Let B = {x^\i E: J } be a basis of V , and for each i e J let c, e F 2 be arbitrary.
Then there is a unique quadratic form q on V such that

(i) the bilinear form associated to q is b


(ii) q(xi) = Ci for all i e J.

In particular we have

0 -------- 'B(V) Q(y)

where the kernel of I1 is F* as noted above.

Proof We can assume that J is totally ordered. If q exists, it must satisfy


(8) q{x + y ) = q{x) + q{y) + b { x , y ) .
As a consequence, if {w^,. . . , is any subset of the basis B, then we must
have

(9) «(«iMi + ••• +a„M„) = “y)


i<j

for all a-^,,,,,a^ e F2.


294 Lattices and Finite Root Systems

Use (9) and the conditions qix^) = c, to define q on V. It is straightfor­


ward to show that it is a quadratic form satisfying (8). □

Now let us suppose that s and e' are two cocycles on L satisfying the
hypotheses set down in Proposition 2. Let the bilinear mapping 6 on L be
defined by

6( a , )8) = e (a , jS) + e '(a , )8) for all a , )8 e L.

Then

& (a ,a ) = e(a ,a ) e'(a,a)

= |( a |a ) + ^(ci\a) = 0 mod2.

Thus b is alternating and by Proposition 3 there is a quadratic form u on L


satisfying

u(a + jS) + u(a) + u(p ) = b(a, p) for all a, ¡3 ^ L.

Write this in the form

( 10) s'{a,P) + u(a + )8) = s ( a , ^ ) + u{a) + w(jS).

Use s to define the group structure on L ^ = L X (Z/2Z); that is

( a , a ) ( p , b ) = {a + p , e { a p) + a + ¿).

Define

(!>': L

by

(f)'(a) = ( a , u { a ) ) .

This section gives rise to another cocycle s" of L. Assuming that the
extension group Z/2Z is identified as a subgroup of L "" in the obvious way,
then s" is defined through
3.8 Central Extensions of Lattices 295

But

= ( a + ^ , s { a , p ) + u{a) + u(l3))
= ( a + ^ , s '( a , )8) + w(a + /3))
= <f>'(a + p ) s ' ( a , p ) .

Thus e" = s'. In other words, e' is a cocycle on L arising from some other
section of L in L ^. Alternatively, the group extensions and L'^deter­
mined by e and s' are isomorphic: We use the set L X (Z/2Z) and the
corresponding group laws given by (2). The mapping 0 : L" L'^defined by
(a, a) ^ (a, u(a) + a) is an isomorphism, and we have the commutative
diagram

( 11)

Proposition 4
Let L be an even geometric lattice. Up to isomorphism o f the type given by (11);
the extension o f L by Z/2Z given in Proposition 2 is unique. □

Let L = ® j^ j Z a j be a free abelian group, and let

0 ---- > Z/2Z L" L ---- > 0

be a central extension of L by Z/2Z. Fix a section of L in L ", and let


e: L X L ^ Z/2Z be the corresponding cocycle. Define z ^ L " by z *•= /(1)
where 1 •= l(mod2) g Z/2Z. Then z^ = 1, and if we adopt the notation e"
for (f>{a\ every element of L'" can uniquely be written in the form e“z",
where a ^ L and n g Z/2Z. The multiplication of is now given by [see
(2)]

( 12 ) _ ^oc+fi^eia,P)+n+m

Let J? be a commutative ring, and let /?[L"] be the group algebra of L"
over R. Thus /?[L^] is a free i?-module admitting the set {e“z"|a g L,
n G Z/2Z} as a basis and has multiplication given by bilinear extension of
296 Lattices and Finite Root Systems

(12). Consider next the free submodule j of i?[L"] defined by

;■= © + e“z)
a^L

(the sum is direct since the terms e" + are linearly independent over R
by assumption). By (12) we have

( e ^ + e ^ z ) e ^ z ^ = _|_ ^ a + / 3 ^ e ( a , / 3 ) + H - n

= e"^^(l + z )
= + e^-^^z e j,

showing that j is a right ideal of A similar argument shows that j is


also a left ideal (recall that z is in the center of L ") and hence that j is two
sided. Because 1 + z = + e^z, it is immediate that j is generated by the
single central element 1 + z.
Consider the .R-module homomorphism

f:R [L ^]^R [L ]

of i?[L"] into the group algebra /?[L] of L (Section 2.5) defined by

/ : e“z" e^(-iy.

This map is surjective, and its kernel is precisely j. Hence R[L^]/j == R[L]
as i?-modules. The multiplication of the quotient algebra R[ L^]/j induces,
via this isomorphism, a new multiplication on R[L]. We denote the resulting
algebra by R[LY , and call it the twisted group algebra of L (by the cocycle
e; see Borcherds [Bel]). Explicitly R[LY is a free /^-module admitting the
family {e“|a e L} as a basis and has multiplication satisfying the identity

(13)

Evidently f?[L] (with its usual algebra structure) is not in general isomor­
phic to since R[L]' need not be commutative.
Let L" be the central extension of L with a cocycle e determined by
Proposition 2. Let <f>e Aut(L). We will show that there is an automorphism
<f>" of L " (as a group) that stabilizes the central subgroup Z/2Z in L^ and
that induces <f>on L.
We begin by considering the cocycle on L defined by

£*^(«,/3) = e(<t>(a),<f>{P)) for all a,f3 ^ L .


3.8 Central Extensions of Lattices 297

Since

e * ( a , ^ ) + e'^{p,a) = e{<p{a),<j>(^)) +
= (<f>{a)\<p{P)) mod2
= (al/3) mod2,
and similarly
e'^{a, a) = |( a |a ) mod 2,

we may define a symmetric bilinear mapping i> on L by

b (a,/3 ) = s { a ,P ) + £"^(«,/3),

just as we did above, with b(a, a) = 0 mod 2 for all a. Let u be a quadratic
form satisfying

(14) u{a + /3) + « (a ) + m( j8) = b ( a , p ) .

Define <f>'' [in the notation of (12)] by

(15) ^>"(£“7") =
We have

<f>'{e“z ‘’e'^z‘’) =

= ^<f>(a) + <i>{p)^uia) + u ( ^ ) + b ( a , ^ ) + e ( a , P ) + a + b

^ g<^(a)+a>(P)^«(a) +u(£) +£*<“•


_ ^ 4 > ( a ) ^ u ( a ) + a^<i>((3)^u(P) + b

= <i.^(e“z ‘')<i.'(e^z")
and
4 > ^{z )= z.

Thus <l>^ is an automorphism of L" fixing z. Evidently <f>^ lifts </> on L.

Lemma 5

(i) Every lifting o f an automorphism <f) to Aut(L is given by (15) for


some quadratic form u satisfying (14) where b = e
(ii) If (f>^ and iff^ are liftings o f (j> and if/ with quadratic forms u and v,
then is a lifting o f il/(f> with quadratic form u + v<f>.
298 Lattices and Finite Root Systems

Proof. Begin with the case ф = 1. Let ф" be any lifting of ф. Then

for some Л: L F2. From (12) we see that Л is Z-linear. Now recall that
linear maps are quadratic forms, and with и replaced by Л and b replaced by
e H- equation (14) is satisfied. This proves part (i) when ф = 1.
Now suppose that ф ^ and ф^ are as in part (ii). Then

ф^ф"{е^г^) =
_ ^фф(а)^иф(а) + и(а) + а ^

which is what we wanted to show.


Finally to finish part (i) in general, let ф" be any lifting of </> e Aut(L).
Let ф^ be 2i lifting of ф given by a quadratic form u. Then (o^ = ф^
lifts 1 and so is given through some linear map Л, as above. Then ф" = ф"о)^
is given through the quadratic form Л + w by part (ii). □

Define Auto(L") to be the group of all automorphisms of L" that fix z.


Each ф^ ^ Auto(L) determines an automorphism of L, and we have an
exact sequence

(16) 1 ---- --------- > A uto(L^) ---- > A ut(L) ---- ^ 1,

where К is the subgroup of elements of Auto(L^) inducing 1 on L.

Proposition 6
к = (L/2L)* {the dual space of L /2 L ) via the map Л •-> Л" for all Z-linear
maps Л: L ---- > F2, where A^(c“z^) =

Proof The mappings A" are indeed automorphisms of fixing z, and the
proof of Lemma 5 shows that every lifting of 1 e Aut(L) is of this form.
From part ii of the lemma, A^ = (/i + A)" for all /x, A e L. □

To obtain results about the automorphisms of the corresponding Lie


algebra 9^, we apply the above results on automorphisms to the special case
when L is a positive-definite geometric lattice. We will assume that L " is the
extension of L given by a cocycle e satisfying the thesis of Proposition 2. We
return to the ordered pair notation (a, a) for elements of L '".
3.8 Central Extensions of Lattices 299

Let (l>^ ^ Auto(L^), and let <f>denote the automorphism of L induced by


(¡>^, We know that for some quadratic form u,

(f>"(a,a) = {(l>(a),u{a) + a)

where, for & = e + u satisfies (14). Let = L 0 ®aeL(2)^'^a the


Z-Lie algebra of the lattice L, as constructed above, and define an endomor­
phism of g^^ through

= a e L (2 ).

Let us verify that is an automorphism of g^. Indeed, if a e L(2) and


^ G L, then

= ( - ! ) “<“>(

while if a, ^ G L(2) then

fO i f ( a |^ ) > 0 ,

-<^(o) if («1/3) - - 2 ,

and

iO if{<f>(a)\<l>(p))>0,

if(<^(a)|«^(/3)) = - 2 .

We see that <f>"([X^, X^]) = [<f>"(Xj,<f>"iX^)] by taking into consideration


the following facts:

. ( a |/3 ) = (< ^(a)l< f>(i3)) and <f>(a + P) = (f>(a) +


300 Lattices and Finite Root Systems

• From (14),

u(a + /3) + e(a,IB)


= u (a ) + m( j8) + b ( a ,P ) + e(a,/3)
= u{a) + m(/3) + e(a,l3) + + e(a,/3)
= uia) + u(p) + ei<j>(a),HP))-
♦ if (a |^ ) = —2 then j8 = —a

and hence

(16) m( “ ) + m()S) = 0.

Suppose next that <(>'' and e Auto(L^), and write

4>"(a,a) = { ^ ( a ) , u { a ) + a),
i/r"(a,a) = (tl /(a),v(a) + a).

Then by Lemma 5

il>"4>"(a,a) = , {u + v<ff)(a) + a).

Therefore for all a e A,

(u+v4>Xoi) ^
</>(«)•

On the other hand,

(u+v<i>Xa)
= (-l) X.

Thus we have established the existence of a natural homomorphism 4> of


Auto(L^) into the group Aut(g^, L) of automorphisms of g stabilizing L
(see Section 2.1).

Proposition 7
If L(2) generates L, then AutoCL"") - Aut(g^, L) under the mapping i>.

Proof If e Auto(L") determines the identity automorphism of g^, then


(f> induces 1 on L, and hence the quadratic form determined by is a linear
3.8 Central Extensions of Lattices 301

mapping A by Proposition 6 and (16). Since A van­


ishes on L(2) and hence on L. From (15) we see that (f>" = 1 and hence that
is injective.
Now let e Aut(g^, L). Then by restriction r¡^ determines a (group)
automorphism <f>of L. By IK-linear extension of ry" we obtain an automor­
phism of IK ® stabilizing IK <8>L. Extend (*| * ) to a IK-bilinear form on
K 0 L. Recall that K 0 g^ has a triangular decomposition with root spaces
K0 corresponding to the roots a: h ^ (a\h) (see Section 3.7). By
Proposition 2.1.3, 7]^ determines an endomorhism <f>* of (IK 0 L)*, which
permutes the roots.
Using ( 1 ) to identify IK 0 L and (IK 0 L)*, we may view <f}* as an
endomorphism of IK 0 L with the property that

= ^a^<f>*(a) foT 3.11 a e L(2),

where e IK^ (actually e {1, —1}). Now we have

(Q^ I^ )^ /3 ^ < /> *( /3 ) “ '*7^ ([<^5 ^ / 3 ] ) ~ ['*7^ ( ^ ) j

= a^(<A(a)l</>*(^))AT^*(^),
from which

(17) (a\p) = (<f>(a)\<l>*{p)) for all L(2).

On the other hand, the equation

V - { [ X , , X p ] ) = [r,-X^,v-Xp]

together with the definition of multiplication on g^ leads directly to

(18) (a\p) = (<f>^(a)\cl>*(p)) fo ra lla ,i8 g L (2).

Since L(2) generates L, (17) and (18) give

= 0 and (i> eA u t(L ).

Let denote both a lifting of 4> to Auto(L") and the corresponding


automorphism of g stabilizing L. Then ¡1" :=</>" “ ^77" e Aut(g^, L) and
pointwise fixes L. Let { aj,. . . , a„} be a base for L(2). Then = c¿X^,,
where c¿= ±1. Let A: L F2 be the unique linear mapping defined by
Ci = ( - 1)'^^"'^ Using Proposition 6, we may lift A to A" ^ K and replace <f) "
with <^"A" to get = X_^., i = Then from [X_^,, X^] = a¿
we find that ii"X_^. = X_^.. Since g^ is generated by L, and { X ^ ^ \ i =
1,..., n}, /X" = id. This proves that 17^ is determined by an element of
Auto(L " ) and hence that i> is surjective. □
302 Lattices and Finite Root Systems

Remark 3 As an important application of these last two propositions we


consider the Weyl group W of L. We assume that L is positive definite and
that L(2) generates L. Then, since W c Aut(L), by (16) and Proposition 6, W
has a preimage W" in Auto(T"), which is a 2*^^"*^^-fold cover of W.
According to Proposition 7, W" may be identified with a group of automor­
phisms of stabilizing L. If w" e and w" ^ w then

(19) W^\l = W,
and for all roots a,

( 20) = ±X ^^.

The group W is sometimes called the Demazure-Tits group. For more on its
structure, see [MPS].

EXERCISES

3.1 Let 6i, be the standard basis for /-dimensional euclidean space

(a) Show that {±e, ± Sj\i j} U { ± e j is a root system of type


and that {±e^ ± Sj\i # /} U {±2eJ is a root system of type Q
(see Sections 3.1,3.3,3.4).
(b) Show that the Weyl groups of these root systems are isomorphic
to Si tx (Z/2Zy. (Interpret this as all permutations and sign
changes of the e/s.)
(c) Consider the root system {±Si ± 6j\l < i, j <1, i ^ j) of type D/
(Sections 3.1,3.3).
( i) Prove that the long roots of Bi and the short roots of C / form
root systems of type
(ii) Prove that the Weyl group of Di is isomorphic to 5/ K
(Z/2Zy~^, (Interpret this as all the permutations and even
number of sign changes of the e/s.)
(d) Show that the set

{±£, ± £y|l ^ i, j < 4, i ¥=j} U {^{ ±£i ± £2 ± ^3 ± «4}


u (± £ ,|l < i < 4}}

is a root system of type F^. Show that both the sets of short and
long roots of form a root systems of type D^.
(e) Let A be a root system of type D^, Show, by looking at the
Coxeter Dynkin diagram, that [Aut(A): W(A)] = 6. Show that
Aut(A) is isomorphic to the Weyl group of type F4.
Exercises 303

3.2 Compute the determinants of the root lattices of types Bi, Ci,
and G2.
3.3 Let A be a finite indecomposable reduced root system. Let II =
{ai,. . . , a/} be a basis of A. Show that there exists a unique short root
in A of maximal height (with respect to II). Show that if =
then p is dominant, and hence all c, > 0. p is called the
highest short root of A. Compute this root for A of type G2, F4, Bi,
and C/.
3.4 Let A be a finite indecomposable reduced root system of rank / on a
K-space V. Let II = { aj,. . . , a j be a base of A. Let A^ be the dual
root system and II [a^ , . . . , a /} the base of A^ dual to II. Recall
the root and coroot lattices

Ô == E 2 a = 0 Za, c V,
ae A ^= 1
I
Q ''= E 2 a © Za^cz V * .
aeA ^'=1

(a) Let

P^:= [y e F*|<jt:,y> e Z for all x ^ Q )

and

P := [ x ^ K|<x,y> e Z for all y e c K.

Show that P (resp. is a lattice in F* (resp. V) (called the


weight and coweight lattice respectively).
(b) Show that Q P and that the group P /Q is finite (the order of
this group is called the index of connection of A. The cosets
(o Q, (0 ^ P, are called the congruence classes).
(c) Let (o^,. . . , (Oi ^ P be defined by a ^ ) = 8^j. (The co/s are
called fundamental weights. They form a basis of P dual to the
basis of (2^.) Given a, e II, show that

Oii

where A = (-^¿P is the Cartan matrix of A.


(d) Show that the index of connection of A equals det(^).
(e) Let W be the Weyl group of A. Show that W stabilizes P,
thereby acting as automorphisms of P. Show that W acts trivially
on P/Q,
304 Lattices and Finite Root Systems

(f) Compute P / Q and the congruence classes for all indecomposable


reduced root systems A.
3.5 Let A c K be a reduced finite root system of rank /, and let II =
be a basis of A. Let Q and P be the root and weight
lattices. (See Exercise 3.4.)
(a) Let Z[P] be the integral group algebra of P. Show that Z[F] has
no zero divisors.
(b) Let W be the Weyl group of A. Then W acts as a group of
automorphisms of Z[P] via

w' • E n^e{ii) = ^ n^e{wii.).


fX^P /i, sP

Let Z[P]^ := {jc e Z[P]\wx = x for all w e W}. Let K' and K
denote the field of fractions of Z[P] and Z[P]^, respectively.
Show that Z[P] is the integral closure of Z[P]^ in K' and that
K' is a Galois extension of K with Galois group canonically
isomorphic to W.
(c) Fix iV e Z+, and let ^ e C. Fix z* e 0 and define
ring homomorphisms

by

and by restriction

f:Z {P T

Let p' = k e r/', p = k e r/, k' = field of fractions of Z[P]/p', and


k = field of fractions of Z [F ]^/p. Show that k' is a Galois
extension of k with Galois group canonically isomorphic to
Wd/ W j , where

Wo = { w ^ W \ w V ' = X>'),
Wj = [w e W\wx = X mod p' for all x e Z [F ]} .

Let i := {/i e z*) = 0 mod N). Show that

IV^ = {w ^ W\w\ = i},


W[ = [w ^ W\wii. = fi mod i for all /j, e P }.
Exercises 305

3.6 Consider a connected affine Dynkin diagram F with nodes / e { 0 ,...,


and marks rii. We play a “game” on F as follows:

Step 0: Pick a node (starting node).


Step 1: Remove the last node j you picked, and let Fy be the diagram
left over (y = if you are starting).
Step 2: Compute the determinant of the root lattice with diagram
Tj. If Dj = rij, you may continue playing the game. Otherwise,
stop: You’ve lost the game.
Step 3: Pick a node connected to the last node you removed. If you
cannot do this, you have finished playing a winning game. Otherwise,
return to Step 1 and keep playing.
Show that in the only diagram all of whose games (24 of them)
are winning games.

3.7 (J. Milnor and D. Husemoller, Symmetric Bilinear Forms, Springer-


Verlag, 1973). In this exercise we construct the famous Leech lattice:
the unique positive-definite even unimodular lattice of dimension 24
with no lattice points of square length 2. Let F2 = {0,1} be the field
with two elements. For

define

§0 := (1 < i < N\Si = 0}

= {1 < i < N \S i = 1}.

Let V = (i^i)o^i^io = (1, h 1,0,1,1,0,1,0,0,0) e Define an 11 X


11 square matrix A = (a^j) with entries in F2 by

aij := i^(/+y-2)modll for all 1 < j < 11.

The 11 rows of this matrix we will denote by a^ ,. . . , Let |5| denote


the cardinality of an arbitrary set S,
(a) Show that \a] n a]\ = 3 for all 1 < i < j < 11.
(b) Fix S c { 1 ,..., 11}. Let s(5) = s = ^ F2^ where by con­
vention s = 0 if 5 = 0 . Show that |s^| = 0 mod 2 and that for all
j € S, |s^ n aj\ - |s^ n aj\ = 0, ± 4 if |5| is odd and = ±2, ± 6
if |5| is even.
306 Lattices and Finite Root Systems

(c) Let B = (bn) be the 12 X 12 square matrix defined by


0 1 1 • • 1
1
B ■■= 1 A

1
Show that any two distinct rows of B are orthogonal and that
B^ = I.
(d) Let C be 12 X 24 matrix obtained by gluing together I and B
/1 0
0 1 0
C ■■= B
,0 1 t

and let 7 1 , , 712 ^ be the rows of C. Show that I7/I = 8,12


for all 1 < I < 12. Show that any two distinct rows of C are
orthogonal.
(e) Let S c F|‘* be the span of the rows of C. Show that
(i) (1 ,1 ,...,1 ) g 5,
(ii) Is^l = 0 mod 4 for all s G S,
(iii) |s^| > 8 for all s G 5 \ {0},
(iv) dim,:^ 5 = 12.
(f) Let L q c be a lattice with an orthogonal basis (b,)i^is24
satisfying b, • b, = i- Define two subsets E and I of L q as
follows.
24
^ {L such that for all 1 < / < 24,
U= i
• i/ ^ 2Z,
24
• = 0 mods,
i=\
mod 2 , .. ., 1^24 niod 2) e S.

/ = I 51 tib}j such that for all 1 < / < 24,

• ^ 2Z + 1,
24
• 51 ^ - 4 mods,
/=1
• { j( t i + 1) m od2,. . . , ^(^24 + 1) mod2) e 5.
Exercises 307

We define the Leech lattice L by

L = £ U / c L q.

(g) Show that if jc = ^ L, then 'Ltf = 0 mod 16. Use this to


show that L is an even integral lattice.
(h) Show that [L qI L] = 2^^.
(i) Let

M = {(^¿) e = 0 mod 2 for all 1 < / < 24},

and let

0= e = OmodS^.

Show that [Mo: E] = 2^\


(j) Use (h) and (i) to show that L is unimodular.
(k) Show that L has no vectors of square length 2.
3.8 Follow all the constructions of Section 3.7 to construct explicitly the
Lie algebra of the root lattice of type Follow Section 3.8 to
explicitly determine the cocycle needed for your construction.
3.9 Consider the root system of type Di

A = [ ± s ^ ± sj\l < i, j < l , i ^ ;}

and its base


n {a^,. . . ,

where —^2? • • • ?^ /- i ^ ^/-i ~ ^ ^i-i


(a) Show that the automorphism cr: -> W which fixes 1<i <
I - 1 and maps -5 / induces an automorphism ": a a of
A which is not in W, Furthermore show that if a e A and a a
then (a |a ) = 0.
(b) For each a e A, let a == (a + d)/2. Prove that 2 == {d\a e A}
is a root system of type with base [d^,. . . , a/_2?
corresponding positive roots S + — {d\a e A+}.
(c) Let L be the lattice in spanned by A, and let 9^^ be the
corresponding Lie algebra over Z with generators a e A,
and L determined by some cocycle e on L (Section 3.7). Show
that e can be chosen so that s(a, j8) = s(d, 0) for all a, e A,
and then a extends to an automorphism, also denoted by cr, of
in which aX^ = X^^ = X^ for all a e A.
308 Lattices and Finite Root Systems

(d) Show that the fixed points in of o- form a subalgebra I and


that f is linearly spanned by the elements of S, the elements
a = d, and the elements + X^, a ¥=a.
(e) For each define 6 ^ = 2 5 /(5 |5 ). Let a , /3,y e A , and
suppose that a ^ a, ^ y = y.
(0 Show that {X^ + X^, a d, -X _ ^ - X_^} is an ^l2-triplet.
(ii) Show that
[a + d,X^] = ( fla ^ )Z ^ ,
[y,\x,+x^] =

[a + d , X p + X^] = + x^).
(iii) Let
fi ^ 1, . . . , / 2,

^ / —1 ^ a i_ i f l —l a / _ i —a/ J

^ /-1 = ^ /-1 +

Prove that e^,. . . , ei_^, h^,. . . , hi_^, f ^ , . . . , fi_^ satisfy the relations
(11) and (12) of Section 3.7 for the Cartan matrix

of the base [d^,. . . , a/_i) of 2.


(f) Determine a triangular decomposition of I.
(g) Let K be a field, and let = IK Show that the elements
10 1 0 h^, 1 0 /^ derived from (e) (iii) generate as a Lie
algebra.
3.10 (This exercise will be used in the exercises of Ch. 6). Let T be a graph
that is a forest with vertices /}. Suppose that G is a group and
that

is a mapping with the property that whenever i and j are not joined,
/(/) and / ( ; ) commute. We wish to show that for all permutations tt of
{ ! ,..., /}, := /(7 t(1)) *• • f(7r(l)) lies in the same conjugacy class.
(a) Consider / points (also labeled 1 ,..., /) placed uniformly in some
order on a circle. Allowing only the rule that adjacent points i
and j on the circle may be interchanged if / and j are not joined
by an edge of F, show that (up to cyclic permutation) any
ordering of the points 1 ,..., / may be obtained from any other.
(b) Use (a) to prove that the conjugacy class of is independent
of 77.
3.11 Prove Proposition 3.4.1
Chapter Four

Contragredient Lie Algebras


T h e o p in io n s e e m s to h a v e g o t a b r o a d t h a t in a f e w y e a r s a l l th e g r e a t p h y s ic a l
c o n s t a n t s w i l l h a v e b e e n a p p r o x i m a t e l y e s t im a t e d , a n d th a t th e o n ly o c c u p a t io n
w h ic h w il l t h e n b e l e f t t o m e n o f s c i e n c e w i l l b e t o c a r r y o n t h e s e m e a s u r e m e n t s to
a n o th e r p la c e of d e c im a ls . . . . But we have no r ig h t to t h in k th u s of th e
u n s e a r c h a b le r ic h e s o f c r e a t io n , o r th e u n t r ie d f e r t ilit y o f th o s e f r e s h m in d s in t o
w h ic h t h e s e r i c h e s w i l l c o n t i n u e t o b e p o u r e d .
—James Clerk Maxwell, 1871, quoted from N ie ls B o h r 's T im e s ,
Abraham Pais (Oxford: Clarendon, 1991).

A contragredient Lie Algebra is, roughly speaking, a Lie algebra g, with a


triangular decomposition S^, that is generated by the diagonal algebra if and
subalgebras ^ l 2\ j ^ J, each isomorphic to êl2(!K). The remaining chapters,
beginning with this one, are devoted to the general structure of contragredi­
ent Lie Algebras and constitute the core of the book.
Section 4.1 introduces the contragredient algebras and concentrates largely
on the necessary and sufficient conditions for the subalgebras t>e
“integrable” to actions of 5L2([K) as an automorphism of g. The concept of
the Weyl group IV of a contragredient algebra arises naturally in this context,
although a full discussion on the structure of W is the subject of Chapter 5.
Every contragredient algebra has an associated structure matrix A =
{A^jX i, j, e J that plays a similar role to the Cartan matrix of Chapter 3.
Using a relaxed version of the conditions CM of Section 3.3, namely
CM1-CM3, we prove the basic result that a radical free contragredient
algebra is integrable if and only if its structure matrix is a Cartan matrix.
Section 4.2 is devoted to the existence of contragredient algebras with a
given structure matrix A. The construction proceeds by first constructing a
suitable diagonal subalgebra i) (a realization of and then by constructing
a universal contragredient algebra u around which has A as its structure
matrix. If y4 is a Cartan matrix, then u/rad(u) is integrable, thereby proving
the existence of integrable contragredient algebras. There are a number of
things about embeddings, field extensions, radicals, and decomposability

309
310 Contragredient Lie Algebras

which are rather obvious but nonetheless need to be written down. These
make up Section 4.3.
A very important class of contragredient algebras consists of the invariant
(or symmetrizable) contragredient algebras, those carrying a (proper) invari­
ant bilinear form. Such a form can exist only if the structure matrix of g is
symmetrizable (a generalization of symmetric). After establishing this, the
remainder of Section 4.4 is devoted to showing the converse of this result.
This is a crucial fact since most of the subsequent development depends
significantly on the existence of such forms. For instance, in this section we
have the construction of Kac’s generalization of the (quadratic) Casimir
operator. In reality this is a whole family of operators one for each
module in the category ^ (g , ^ ) , with the property that and inter­
twine g-module maps from M to N, The construction of these operators is
possible (as far as is known) only if g is radical free and invariant. The
Casimir-Kac operator is used at critical points throughout the entire theory,
mainly as a tool to keep track of the type of subquotient modules that can
occur in modules from ^ (g , *^).
Our first application of these operators appears in Section 4.6, where we
prove the Gabber-Kac theorem. This gives information about the generators
of rad(u), where u is the universal contragredient algebra of a symmetrizable
structure matrix. An important consequence is that one obtains an explicit
presentation for each integrable invariant contragredient algebra g (always
assuming symmetrizability) and in fact concludes that integrable is equivalent
to radical free. A slight modification of the Casimir-Kac operator gives us an
operator F acting on g itself. In Section 4.7 we define F and use it to prove
the remarkable fact that there exists a contragredient bilinear form on g,
obtainable by a slight twist of the invariant bilinear form, that is positive
definite on the nonzero root spaces of an integrable invariant contragredient
algebra g over R. This leads to the existence of a hermitian contragredient
form on g^, which is likewise positive definite on nonzero root spaces.

4.1 CONTRAGREDIENT L IE ALGEBRAS

Let g be a Lie algebra over a field K, We say that g is contragredient


triangular or simply contragredient if there exists a triangular decomposition
* ^ = (g + ,^ ,í2 + ,cr)o f g satisfying the following:

CTl: If {«ylyej ^ is the given basis of according to the axiom of


triangular decomposition TD4 of Section 2.1, then the Lie algebra 5/^^^
generated by g“^ and g““^ is isomorphic to ^12(1K).
CT2: The elements of g“>, g"“^ j e J, and generate g (as a Lie
algebra).
CT3: The sum Eye 9“^] is direct.
4.1 Contragredíent Lie Algebras 311

If we wish to indicate the triangular decomposition relative to which g is


contragredient, we refer to the contragredient Lie algebra or contragredient
pair (g, S^).
Let j G J. Using CTl and cr it is clear that dim g“> = dim g”“^ = 1. Let
Ej G g“>, Fj G g~“/ be any nonzero elements. Then Hj ~ [Ej, Fj] ^ 0 (by
CTl) and [Hj, Ej] = kjEj for some kj # 0. With Cj ■■=2 k J % , a/== I kJ^Hj,
fj •■=Fj, we have an il2-triple for §1^^:

( 1) [«; 2 e y , [tty , / y ] 2 / y , [^ y ,/ y ] C ij •

There is no reason why aej should be /y, but we can easily alter a to have
this happen. Indeed for each j e J, let crej = tjfj, tj e Then there exists
an unique automorphism 0 of g such that

Oej = iyey,

= t-%
e\if = id^.

(see Section 2.1, Remark 2). Of course 0g“ = g" for all roots a. Set r = da.
Then T is an antiautomorphism of g satisfying r\t) = id^. Furthermore
T6j = 0iy/y = /y, r/y = = 6j. Thus by modifying the triangular decom­
position ^ by replacing a by r, we can assume at the outset that cr(ey) =
fjJ e J. Note that = id from CT2.
Henceforth we assume that for each j e J,

Cy e g«^/y e g-->, a /e i|

satisfy (1) and


( X l C j ^ f j .

We call such a choice of Cp a j, and /y a display of (g,

The display {cy, a j , /y}ye j is finite if J is finite.


Remark 1 Let {cy, a / , /y}ye j and {e'p , //lye j two displays of (g, S^).
Then there exists a family {^ylyej such that

e '= t e f- =

Since a{e'j) = / / we obtain iy = ±1 for all j. In particular a / is unique,


since it is the only element in [g“>, g““^] that has g“^ as a 2-eigenspace. We
leave it as a simple exercise to show that if every element of IK has a square
root then displays can always be found without altering a.
312 Contragredíent Lie Algebras

We define

yej

Remark 2 The assumption CT3 is extremely mild. In fact it is not used


here at all. We retain it to avoid a certain awkwardness in the presentation,
but it is straightforward to adjust the results appropriately in the case that
the set {a/ly e J} is linearly dependent. In Chapter 5, when we consider
general root systems, dependence relations are an important consideration.

Proposition 1
Let (g, T) be contragredient, and let [ej, a j , be a display.
Then

(i) for all i e J, dim g“' = 1;


(ii) g+ is generated by {ej\j e J} and g_ is generated by {fj\j e J} as Lie
algebras;
(iii) for all a ^ L \ {0}, dim g" is finite.

Proof, (ii) Let g'+ denote the subalgebra of g generated by the e,, i e J, and
g'_ the subalgebra of g generated by the /,, i e J. It suffices to show that
9' == g'_+ ^ + g'+ is a subalgebra of g. First, ad /i(g') c g' for all g is
clear. Next it is easy to see that

ad , e ,J e ]^ + g'^. for all A: > 1,

and hence ad /y(g'+) ^ il + 9'+- the same way ad ey(g'_) c + g'_. This
proves part (ii).
Since [e,|,. . . , e,^] lies in gt^“o and g“' is spanned by e„ it is clear that
dimgt^“' is finite. Using cr we complete part (iii). □

We next attach to (g, a matrix A = (/1,^), by defining

An = ( a ,- ,« / ) .

The matrix A is called the structure matrix corresponding to (g,T). The


condition eJ = 2ei shows that

An = 2 for all/.

Example 1 Let g be one of the Lie algebras of type A , £>, and E given in
Section 3.7. Each of these is based on an even lattice L in which L(2)
generates L and forms a finite root system. For each we have a base
4.1 Contragredient Lie Algebras 313

n = {«1,. . . , relative to which L(2) = L(2)+U L(2)_. We have

g= © 0 g“ = 0 g"“ © Í) © 0 g“,
asU2) asi,(2)+ a^U2)^
the latter being a triangular decomposition of g. The spaces

9"“' + [9"''^ 9“^] + 9"^ j = 1, • • •, ^


are isomorphic to ^ Í 2ÍK) and collectively generate g as a Lie algebra. It
follows that g is contragredient. The elements i^y>^y»/y}y=i,.defined in
(3.7.10) form a display and the corresponding structure matrbc is the Cartan
matrix A of L(2).

Example 2 (§I2(IK)) (see the Appendix

iT¡(lK) = < /o,/i> e $ ® <eo,ei>

determines a triangular decomposition of ^l2(IK). The subalgebras

= Kf¡ + Kh¡ + Ke¡, j = 0,1,

are isomorphic to §I2(IK), from which we see that §I2(IK) is contragredient.


The corresponding matrix is

(-^ 1 )'
We now define a very important group W (called the Weyl group) of linear
automorphisms of As we will see, the Weyl group is closely related to the
group Aut(g, ]^) of elementary automorphisms of g that stabilize if. Initially
we use W to get a better idea of what the root system A looks like.
Let j e J. Define
0: r
by
rji p ^ P - <j8, a ^ ) a j for all P
and

by
rjy : h ^ h — (aj, h ) a j for all /2 e

It is easy to see that r,(ay) = - a ^ and that Vj pointwise fixes the hyperplane
= {j8 e ]^*|<j8, = 0} of Thus rj is a reflection of f)* in aj
314 Contragredíent Líe Algebras

(Section 3.2). Similarly ry is a reflection of ^ in a / . In particular

r / = id^*, r / ^ = id^,

rj^G L ir), r/e G L (^ ).

Notice that

0( ^ u ’

Thus

rj stabilizes Q A^j ^ Z for all / g J,

stabilizes A e Z for all / g J.

The group W c GL(ij*) generated by those r, for which ad Cj is a locally


nilpotent transformation of g is called the Weyl group of (g, T):

W = (rj\j G J, ad Cj is locally nilpotent).

Similarly we define the dual Weyl group

GL{^), W ^= {r^ I ; G J, ad is locally nilpotent).

If none of the ad ej are locally nilpotent, then by convention W = {1} and


W ^= {!}, where 1 denotes the identity map of the appropriate space.
Remark 3 Both W and depend on (g, but not on the chosen
display. The reason for asking that ad ej be nilpotent will become apparent
as we move along. In the end we are going to be interested in the case when
ad 6j is locally nilpotent for every j g J. Readers seeing this for the first time
might like to assume this from the outset.

Proposition 2

(i) W ^ and W are isomorphic groups, an explicit isomorphism being given


by r,, j G J.
(ii) I f these two groups are identified by the isomorphism in part (/), then

(a , h) = { w{ a ),w { h) ) for all a G 1^*, G 1^, and w E:W.


4.1 Contragredient Lie Algebras 315

Proof. Let ^ e and /i e Recall the inverse transpose (Proposition


3.2.6). We have

= { ^ ,h ~ { a j,h )a f)

= {l3 ~ { l 3 ,a ^ ) a j , h )

= { r j( ^ ) ,h ) fo r a ll/3 G f)*, G f).

This shows that (ry^)* = rj and also shows that

(2) { ^ , r / ( h ) ) = { r j ( p ) ,h ) fo ra ll^

Now it follows that

fr/
V Jk J
= r,j \ . .. r ,Jk

and hence that

(3) (■ )* :W ''^W

is a (surjective) group homomorphism. Now consider the inverse transpose r*


of rj,
y ^^

Then for all t e 1^** and for all /3 e 1^*,

(4) { P , r * i t ) ) ={ rj(P),t)

={l3,t-{aj,t)a^).

Here we have written the pairing of 1^* and 1^** as

< • , • > : f|* X ^ IK

and have identified a f g 1^ as an element of 1^** via the canonical identifi­


cation of 1^ in 1^**. Thus (4) reads

rr If. = r / ,

and we have a homomorphism

(5) W ^ W '^
316 Contragredíent Líe Algebras

given by
VT w* 1^
which is evidently the inverse of the homomorphism (3).
(ii) This follows immediately from part (i) and equation (2).

Proposition 3
Let t ^ K^, and let j e J. Assume that ad ej is a locally nilpotent endomor­
phism of g. Then

(i) n^U) '= expad(iey)expad( —i~ypexpad(iey) is an {elementary) auto­


morphism of g stabilizing f);
(ii) ny(i)li, = r /.

Proof Since ad e^ is locally nilpotent so is ad aiej) = ad fj. Thus nj(t) is an


automorphism of g (Proposition 1.5.3).
Let
Hj = { h ^ f i \ ( a j , h y = 0}.

Since (oij, a j ) = 2,

and since
[h, ej] = 0 = [/i, fj] for all h e Hj,

we have
\h ¡ = 1-

On the other hand,

0 (« /) = ad(tey)exp ad( - 1~‘/ y ) ( a / - 2ie^)


= exp ad(ie^) ( - « / - 2iey)
= —a¡

This shows that n^it) stabilizes 1^. Moreover let e 1^, and write h = ka^ + h!
with W e H:. Then

nj{t){h) = k a j h ! = h — { a j , h ) a j
= r]'{h).
4.1 Contragredíent Lie Algebras 317

Proposition 4
Let w and write w = rj . . . r j . Let ..., e be arbitrary. Then

(0 n = n{w) •= is an elementary automorphism o f g


such that ng“ = g^“ for all a e A. Moreover n\\^ = w under the
identification o f W with W ^\
(ii) ivA = A and dim g“ = dim g"^" for all o: e A;
(iii) for j ^ J with ad Cj locally nilpotent and for all w El W, the elements
of g"^“^ are diddocally nilpotent.

Proof (i) We see that n stabilizes by Proposition 3, and hence by


Propositions 2.1.3 and 2 above we obtain ng“ = g'^/i o*« = with n |^ = w,
as desired.
(ii) This follows from part (i).
(iii) «(g“0 = g'^“^ and n is an automorphism. Since g“>= Ke^, g"^"> =
IK/ie , and part (iii) is immediate. □

Of course the n{w) of Proposition 4 is far from unique for a given w. We


come back to this point in detail in Chapter 6. Notice, however, that if n{w)
and ri{w) are as above, then n{w\n'{w))~^ e K, where K c Aut(g, 1^) is as
in the discussion following Proposition 2.1.3.

Let (g, ^ ) be contragredient. A root « e A is said to be real if a = wa^ for


some w EiW and some ; e J for which ad e^ is locally nilpotent. If a root is
not real, then it is said to be imaginary.
Set

'’^A := (a e A I a is real},

'^A := A V"A,

'■"A^:= A+n""A, '■"A_:= A_n''^A,

^■'”A^:= A+n^'”A, "'”A_:= A.n'^^A,

n =={«;!;• e j } , n'^== { a / l / e J},

:= {«y I y G J, ad €j is locally nilpotent},

‘11'^:= I y G J, ad Cj is locally nilpotent}.


318 Contragredient Lie Algebras

Thus "A = w r W . We have

g= © 9“ 0 9“ © 0 g"" 0 0 0 g"" ® 0 9"*


ae A ae"A_ ae"^A+ aG''”A+

The root spaces g", a are relatively easy to understand. This is not
so for the imaginary root spaces.

Proposition 5
Let (g, r ) be contragredient, and let Wbe its Weyl group. Then with the above
notation we have

(i) Q a n A = {a, 0, —a) for all a e'^^A;


(ii) dim g“ = 1 for all a e'^^A;
(hi) ^"A = -""A ««¿"'"A = -"'"A;
(iv) Wstabilizes '^^A/'”A^_, and
(v) ""A^= -""A_ and"'”A^= -'^A _;
(vi) ^"n = nn^"A.

Proof (i) Let a e^'^A, and suppose that na ^ A for some n e Q^.. By
assumption there exists aj and w ^ W such that

w{na) = naj.

Since W{A) = A, it follows that naj e A. Assume that nuj e A+c Q_^. Then
by TD4 of Section 2.1, n e Z+. The space g"“> is generated by all expres­
sions of the form

[ej^,...,ej^]

with 0 J, A: 0 Z^_, and «y. + • • • +ay^ = noj (Proposition 1). It follows


that k = n and that j) = j for all 1 < / < A:. Thus g"“>= (0) unless n = 1. If
naj e A_ or if n e Q_, we reason in the same fashion. This proves (i).
(ii) We know that dim g“>= 1 for all ; e J (Proposition 1). Since
dim g“ = dim g"^“ for all a 0 A and w ^ W (Proposition 4) (ii) follows.
(hi) Let a e'^^A. Choose w ^ W and aj e'^^II such that a = waj. Then
- a = wrjOj e'^^A, showing that '^^A = -'^^A and hence that "'"A = - " ”A
(since A = -A ).
4.1 Contragredient Lie Algebras 319

(iv) We have W C ^ ) and hence that ='"’A, since M A ) = A.


Let us show that

(6) if aj ^ n ,th e n rX A A { a;} ) = AA{«y}.


Indeed, since for all n e Q^, naj ^ A+\{ay}, we can establish (6) by reason­
ing as in (3.3.14). Now we want to show that +) Let a
and let aj e '’^n. Then rja e'^'A and by (4.6) Vja e A+. Thus Vja e A^_n^'”A.
It follows that ITa c (A +n^'”A) c^'^A^ which is what we want. Finally,
WCA_) = 1^(-^‘'”A+) = -1TC'”A+) = -^■'”A^ = ''”A_.
(v) The proof is obvious.
(vi) Clear from part (iii) of Proposition 4. □

Example 3 {A, D, E) This continues Example 1. We have A = L(2) U {0}.


Since A is finite, the set
jo: + kaj I A: e Zj n A

is finite for all a e A, ; e J. Thus ad and ad /y are nilpotent. Thus the


Weyl group W (in the current context) is

and from the definition of each rj and Proposition 3.3.4, it is clear that W is
precisely the Weyl group of the finite root system L(2). The same proposition
shows that L(2) = ITIl ='^^A, and hence that "'"A = {0}. The action of W on
= LK a^ = (]^*)* is precisely the transpose action discussed in Proposition
3.2.6. The role of the dual root system A^= {Wa^ | ; = 1 ,...,/} is made
apparent below.

Fix ay Then ad ey and ad/y are locally nilpotent, and we have


already had occasion to see, exp(ad te^) and exp(ad i/y), i e IK are well-
defined automorphisms of g. By Proposition 2.4.7 the representation
== ^I2(IK) Ql(g) given by

:X ad X
gives rise to a representation

^<^>:5L2(IK) ^ A ut(g),
satisfying

(o ^ exp(adtó^),

(] ?)
320 Contragredient Lie Algebras

We can interpret this as “integrating” the representation of to a group


representation. It is interesting to see that under 7t^^\

0
- t -1 " ( ( ; :i(; :il
= nj(t) fo ra lli e K ''

(see Proposition 2.4.7). We denote the group 7r^^^5L2(lK)) by or SL^{\


that is,
:= (exp ad tej,exp ad tfj U e 1K>.

According to the way in which g decomposes into irreducible §I^2^^‘niodules


under ad, S L f - 5L2(K) or - SL2(K)/{t ± 1} (see Proposition 2.4.9).
The group generated by those exp ad te^ and exp ad tf¿, i e K, / e J for
which ad and ad are locally nilpotent is called the adjoint group of g
and is denoted by ^ad- There is much more to come on <Jad in Chapters 6
and 7.

Proposition 6
Let (g, ^ ) be contragredient as above, and let a Let w and
aj be such U
that a = waj. Then

(i) the space


:= g“ 0 [g“,g "] 0 g “

is a three-dimensional subalgebra o f g isomorphic to ^I2(1K).


(ii) The elements of and g~“ are locally ad-nilpotent, and the group
SL^2 ^ := (exp ad e, exp ad / | e e g", / e g~“>

is a subgroup o f isomorphic to SL2(K) or 5L2(IK)/ {± 1}.


(iii) Either of the following properties characterizes the same unique ele­
ment ]^:
(iiia) belongs to an 2-triplet of in other words, there exist
^a’fa ^
[ a ^ ,e „ ] = 2 e „ , [a'",/« ] = - 2 / ^ , [e^,fj= a'^;

(iiib) a ^ = wa^ {under the identification o f W with W ^).

Proof By Proposition 4.1.4 there exists n = n{w) such that

«9^ = 9*^^ for all /3 e A.


4.1 Contragredient Lie Algebras 321

Thus
0 [g«y, g-«y] 0 g"«y)

which establishes part (i). Moreover, since n(g“>) = g", it follows that g"
(and hence g““) consists of locally ad-nilpotent elements. This shows that
is a well-defined subgroup of Aut(g). It is easy to verify that

n(expad = expad(njc)
wherever n e Aut(g) and x e g is locally ad-nilpotent. Thus

= SL^^\
whence we have part (ii).
From part (i) it follows that {ncj, n a j , n/y} is an ^l2-triplet for Since
n\^ = w, the existence of having properties (iiia) and (iiib) follows by
setting a^:= As for the uniqueness of we recall that g", g““, Ka ^
are the 1^-eigenspaces of 5/^"^ and that the space [g“, g““] has a unique
element that can be extended to an ^ 12-triplet of ^
Remark 4 Let a and let w e IF and be such that a = waj.
Let be the (unique) element defined in part (iii) of Proposition 6.
Then a ^ = wa^, and this identity holds independently of the choice of w and
aj above. We call the coroot of a. The set of coroots is denoted
by Notice that from the definition it follows immediately that for all
a e ""A and w e IF,

{waY=w{a^)

and that '’^A^c 2 ^


From [a^, e^] = 2e^ we clearly have <a, a^> = 2.

Proposition 7
Let a e '^^A, and let be its coroot. Then

(i) the reflection r^ E GL(i)*) deflned by

belongs to W:
(ii) I f w ^ W and aj are such that a = waj, then

r^ = wrjW~^.

In particular r . = r,;
322 Contragredient Líe Algebras

(iii) under the identification o f W with we have = r^v, where for all
h
r^w :h ^ h — {a, h )a ^ .

Proof Let w ^ W and Uj be such that a = waj. By Proposition 6,


way. If e ]^*, we have

wrjW~^p = w(^w~^fi - (w~^P,ay)aj^

= )8 - ( p , w a ^ ) a (by part (ii) of Proposition 2)

= fi - <)8,a^>a

which shows that r^ = wrjw~^. This establishes part (ii), and also part (i),
since wrjw~^ e W. Finally, part (iii) follows from Proposition 2 (ii). □

We continue to assume that g is a contragredient Lie algebra.

Proposition 8
Let A be the structure matrix o f g. Fix j e J.

(i) ad Cy is locally nilpotent if and only if for all i e J, / ^ ;, we have


(ia) ^ , y e Z < o ,
(ib) dfj := (ad Cy)"^'^^e¿ = 0.
(ii) I f ad Cy is locally nilpotent, then for all i ^ j,

Aij = 0<=^ [e„ey] = 0,


^ , y = 0 = > ^ y , = 0.

(O f course, if ad e^ is also locally nilpotent, we obtain A^j = 0


- ^ y , = 0.)
(iii) Properties (i) and (ii) hold if ej and c, are replaced by f andfi.

Proof (i) Suppose that ad e^ is locally nilpotent. Since for each i e J,


rja^ = —A^jUj e A, and since no root can have ‘'mixed signs,” ^¿y ^
whenever i ^ j. Furthermore

+ (1 “ ^ i j ) ^ j ) = «<• - «y Í A
which shows that
4.1 Contragredient Lie Algebras 323

Conversely suppose that for all i ¥=j, e Z < q and (ad = 0.


Note that (ad = 0 for all i and (ad ejYh = 0 for all /z e 1^. Now set

a = |jc G g I(ad = 0 for some rij{x) e l\l|.

It is straightforward to prove that a is a subalgebra of g. Since a contains a


set of generators of g (see CT2) we have a = g.
(ii) Suppose that ad ej is locally nilpotent and that A¿j = 0. Then (ad ej)e¿
= 0 by part (ib), and we have

^ji^J = [«."'>«;■] = = 0

by the Jacobi identity. Thus Aj¿ = 0. This last argument works equally well to
show that [e¿, ej] = 0 =>A¿j = 0, thus finishing the proof of part (ii). □

A contragredient Lie algebra (g, y ) is called integrable (relative to if


for some (hence any) display {ej, a^, ad 6j (hence, using the involution
(7, also ad fj) is locally nilpotentfor every j g J.
Example 4 The Lie algebras A, D, E and §I2(IK) are integrable.For an
integrable contragredient Lie algebra the Weyl group W is generated by the
set of all rj, j G J.

Proposition 9
The structure matrix o f an integrable contragredient Lie algebra g satisfies three
of the conditions CM o f Section 3.3:

CMl: All = 2 for all i g J;


CM2: Aij G —N for all z, j g J, z ;;
CM3: A^j = 0 <=>Aji = 0.

Using the notation established above, we have the relations

R l: i[ar,f,] = -AjJj,
for all i, j G J.

■■= ( ade¡ ) = 0,
R2:
I d~j ■■= (ad f¡) =0 for all i # j.
324 Contragredíent Lie Algebras

A matrix A ^ satisfying CMl-3 is called a Caitan matrix.


In the literature the expression “generalized Cartan matrix” is often used
instead. We refer to a Cartan matrix satisfying CM l-5 of 3.3 as a Cartan
matrix of finite type. The relations R1 and R2 are familiar from Section 3.7.
However, the situation here is far more general. In Section 4.2 and subse­
quent sections we will show that a contragredient Lie algebra g can be
constructed whose structure matrix is any prescribed matrix A for which
CMl holds. If A satisfies CM l-3, then g can be constructed to be inte­
grable. In the exercises we outline a generalization of integrable contragredi­
ent Lie algebras that relaxes CMl.
The terminology “integrable” comes from the fact that each representa­
tion of j e J, in g (given by the adjoint action) can be integrated to a
group representation

: 5L2(K) ^ A ut(g),

as we have seen above.

A contragredient Lie algebra (g, is called a Kac-Moody algebra if it is


integrable. We say that g is of finite type if A is of finite type and of affine
type if A is of affine type. Otherwise, it is said to be of índefiníte type. These
Lie algebras were introduced in [Kal, Mol].

Let (g, be contragredient. We recall (Proposition 2.7.5) that the


radical r = rad(g) of g can be characterized as the maximal ideal of g
intersecting trivially. Since [e,, /¿] = e \ {0}, we see that

r n ( 0 »</,j e ^ e (0 Ke,)| = (0).


Let ~ : g g /r denote the natural mapping. By the Remark 1 of Section
2.7, g /r has a triangular decomposition ^ = (1^, g+, Q^, cr), where we have
identified and and written a for the anti-involution induced by cr on g.
Furthermore the display {e¿, , /J /e j rise to a display {é¿, a'^, of
g, showing that g /r is contragredient. If ad is locally nilpotent, so are ad
and ad /^. Thus we have a natural homomorphism

^:<^ad(s) ^ Gad(g)

given by

exp ad exp ad ,

exp ad í/¿ exp ad í/¿,


4.1 Contragredient Lie Algebras 325

whenever ad (and hence ad f¿) is locally nilpotent. Note, however, that it


can easily happen that ad é¿ is locally nilpotent, although ad is not; in
particular 6 need not be surjective. Likewise we evidently have a homomor­
phism of Weyl groups,

W{Q) ^ W(Q),

which need not be surjective. However, this map is injective^ since W( q) can
be viewed as a group of linear maps on Í) and i) and can be identified.
Thus we can simply identify W( q) as a subgroup of W^(§). Similar considera­
tions apply to the algebra § •= 9 /t if t is any ideal of 9 such that t e r .
More precisely

Proposition 10
Let (g, be a contragredient Lie algebra, and let § = g /t, where t is an
ideal of g contained in rad{%) such that o-(t) = t. Then the natural mapping
“ : g ^ g is injective on th^subspace ®^ ® Identify­
ing 5 and ^j_we have that ^ = (^, g+, cr) is a triangular decomposition of
g and (g, is contragredient. Furthermore (g, and (g, have the
same structure matrix, and g is integrable if g is integrable. We have

^^A(g) c^^A(g), PL(g) c iy(g)

with equalities if g is integrable.

Proof. After the remarks above it only remains to prove the statements of
the last sentence. If a e'^^A(g), then a = waj for some «y e '‘^n. Under the
identification of IF(g) inside W{%), the same equality holds relative to g,
so a e '’^A(g). If g is integrable, then = < r/ | ; e J> = W{%) and
^^A(g) = =^^A(g). □

Proposition 11
Let ( g , r ) be contragredient with structure matrix Л = y^j* Define
g = g/rad(g), and let ~ ^ q be the canonical map. Then

(i) I f A¿: e Z_ or if A¿: =Aj¿ = О for some i Ф j, then


(adey) and d¿j = ( a d / y ) belong to rad(g);
(ii) For ad éy {and hence ad /y) to be locally nilpotent, it is necessary and
sufficient that for all i ¥=j, A¿j e Z <q ^^d that A¿j = 0 => Aj¿ = 0;
(iii) g is integrable if and only if A is a Cartan matrix.

Proof, (i) Suppose that A^j e Z < q for some i ¥-j. Consider the §1^2^^-sub­
module M of g (under the adjoint representation) generated by /¿. We have
[ej,f] = 0 and [af,f¿] = - A ^ j f . By part (ii) of Proposition 2.4.2 (with f
326 Contragredíent Lie Algebras

replacing z;+, - a , replacing A, etc.), either d~j = (ad = 0 or M is


a Verma module and dj] generates a maximal proper submodule. In either
case

ad ej{d¡j) = 0.

Also

(a d e ,)(a d /,)"^ '^ " 7 , = ( a d = 0.

Indeed, If A¿j < 0, this is clear, and if A¿j = 0, then also Aj¿ = 0 and
[fj, h¿] = Aj¿f¿ = 0. Finally, it is obvious that ad ( a d = 0 if
k ^ i, j. Thus d'j is a highest weight vector for g in the adjoint represen­
tation. We know then that the module M' that it generates is
ad(U(g_))(ad 9_, and hence M' is an ideal of g whose intersec­
tion with is trivial. Thus M' c rad g, and we conclude that d~ e rad g.
Since rad g is cr-invariant, d~^j e rad g proving part (i).
(ii) If ad éj is locally nilpotent, then the A¿j are as desired by Proposition
8. (Recall that g and g have the same structure matrix, see Proposition 10).
Conversely if ^4^y e Z A^ - = 0 => Aj¿ = 0, then

- A a + 1
(ad ij) e¿ = 0 for all i j

by part (i). Thus ad Cj is locally nilpotent (part (i) of Proposition 8).


(iii) The proof follows from part (ii).

Proposition 12
Let (g, r ) be a contragredient algebra, and let = Then

Dg = g_e e g^..

Proof. For any a e A\{0}, [1^, g“] = g“, which together with [e„/;] =
shows that Dg = g_+(^ n Dg) + g+D g_+ + g^. It remains to show
that n Dg c . To prove this, we need only to show that [g_ , 9 + ] n ^ c
Q ^, which we can reduce to looking at a single product [ j:, y], where
X e g““ and y e g", a e A+. We use induction on ht a. If ht a = 1, then
a = a,, X = c/„ y = de^ for some c, d e i e J, and [x, y] e Oth­
erwise, ht a > 1, and since g_ is generated by the /y, j e J, we may assume
that X has the simple form [/ y, z], z e g“^, ¡3 e A+, ht = ht a — 1. Then
[jc, y] = [[/;, z], y] = [fjiz, y]] - [z,[fj, y]]. Now [z, y] € g“', and [fj, y] e
g^, so the proof is finished by the induction. □
4.1 Contragredíent Lie Algebras 327

Corollary
¡f ^ = Qk ^ ihen g is perfect. □
Remark 6 The decomposition

Dg = g_0 !2k ® 9+

(or more precisely (g+, Q^, o-log)) is not in general a triangular decom­
position. The problem is that it is possible (and always happens in affine
algebras) that there are roots 5 0 for which g^] = 0. Thus is not
in general large enough to serve as a diagonal algebra. This phenomenon,
which can occur only if the structure matrix is singular, is largely responsible
for the necessity in Section 4.2 of defining a realization of the matrix as a
preliminary to constructing contragredient algebras with a given structure
matrix.

Let (g, be a Kac-Moody algebra with root system A.

Let a and let 0 A. The a-root string through ^ is

5(j8,a) = A n (/3 + Za).

The next result is the fundamental fact about root strings.

Proposition 13
Let ( q, be a Kac-Moody algebra. Let a 0 '^^A. Then for any /3 e A there
are nonnegative integers d, u so that 5(/3, a) is an unbroken sequence

RSI p - d a , . . . , p , . . . , p + ua,

and furthermore

RS2 d - u = </3,a^>.

Proof Since a g“ + [g“, g““] + g”“ = ^1^2^ is isomorphic to ^l2(IK).


Furthermore g affords an integrable representation of (Proposition 6).
Let N ‘= E^ez9^^^“- Then N is an integrable submodule, and its weight
system P(N) is precisely S(p, a). From Proposition 2.4.7, N decomposes
into finite-dimensional irreducible submodules, each of which has the form
K= where dim = 1 and ad is scalar multiplica­
tion by 5 - 2/, / = 0 , .. ., 5.
Since ad a is multiplication by <j3, + 2k, we have

K S —2Í C + 0 , o : ' ' » /2 ] - i } a


328 Contragredíent Líe Algebras

This implies that s = <)8, (mod 2). Now we use the Freudenthal mid­
point argument [FdV]. The midpoint of the set of roots ¡3 + {^(5 — (¡3, a'^))
- ¿}a as i runs from 0 to 5 is

1/ \ is-(p,a^y
( 2 ^ l« + ^ + ( ^ )“

= /3 - -</3,a'^>a,

which is independent of s. Actually /n is a root of the string only if


<13, a^> e 2Z, but the point is that all the irreducible submodules K of N
share the same midpoint and they all involve fi or fi j a according to the
parity of </3, a^}.
Using the Weyl group W we can replace {jS, a} by wa} for some
w ^ W if necessary so that we can assume a is II. Now the fact that
elements of A do not have mixed signs shows that 5(j8, a) is bounded either
above or below, and now the midpoint argument shows that it is bounded
both above and below. An irreducible component K for which s is maximal
gives us a sequence of roots p — d a , . .. , jS,. . . , )8 + ua which necessarily
contains all the others (again by the midpoint argument). Finally, the mid­
point of this sequence is + ^(u - d)a, from which d — u = </3, a^>. □

We finish with a prescription to inductively compute the entire set of


positive roots of a Kac-Moody Lie algebra:

Proposition 14
Let ( q, be a Kac-Moody algebra with root system A. Let

Q+{n) = (a e ¡2+ lht(a) = «}, A+(n) = A+n Q ^{n), n ^ l .

{For n < 0, Q^{n) = A+(n) = 0.)


Then

(i) 0 ^ ( l ) = { a j / e J} = A^(l);
(ii) for a e Q^{n), n > 1, we have a e A+(n) if and only if one of the
following two conditions hold:
(iia) there is an / e J such that (a, a f ) > 0 and r¿a s
A+(n - ( a , a f ) ) ,
(iib) for all i e J, (a, < 0 and for some / G J, a —a¿ G
(n - 1).

The roots that occur in (iib) are imaginary, and a root a occurring in (iia) is
imaginary if and only if the corresponding root r^a is imaginary.
4.1 Contragredient Líe Algebras 329

Proof, (i) Clear.


(ii) Let a e Q^(n); n > 1. First, we show the conditions are necessary.
Assume that a e A+(/t). If condition (iia) fails to be the case, then from (4.6)
we see that <a, < 0 for all j e J. Moreover since g“ # (0) is spanned by
products , . . . , ] with + • • • + a . = a, we must have (0) for
some j G J.
As for the sufficiency, if condition (iia) holds, then by (6), r^a e A+, and
hence a e A+. If condition (iib) holds, choose / e J so that a - e A+
and {a, ayy < 0. By RS2 the a-string through a - a, is {a - a, + fca, 1- d
< k < u}, where d — u = {a — = {a, —2 < 0. Thus m > 0 and
hence by Proposition 13, a e A.
Finally if the root a occurring in condition (iia) is imaginary, then so is r^a
[Proposition 5(iv)]. If a occurs in condition (iib), then to show that a e'^'A,
we must anticipate Proposition 5.2.6. Indeed, if a then In
particular a ^ = ^ 0? and hence

( a , a ' ' ) = E y e j c / a ,a / > < 0,

contrary to <a, > = 2. □

Corollary
Let t be any ideal contained in rad(g) such that or(t) = t. Then A (g/t) = A(g)
{see also Proposition 10).

Proof The algorithm of the proposition determines the same subset of 2 ^


for g and g /t. □

We finish this section with a description of the Lie algebra of derivations


of a contragredient Lie algebra g.
Any linear mapping e : g/D g Z(g) lifts to a linear map e : g Z(g) g
annihilating Dg, and s is clearly a derivation. Identify the space of these
derivations with Hom|,^(g/Dg,Z(g))

Proposition 15
Let Q be a finitely displayed contragredient Lie algebra with dim < dim g+.
Then

D erg = H om K (g/D g,Z (g)) + a d g .

In particular, if Q is perfect or Z(g) = (0), then

Der g = ad g.
330 Contragredient Lie Algebras

Proof. Using Proposition 2.1.4 it will suffice to show that a derivation 5 e Dq


belongs to Honi(|^(g/Dg,Z(g)) + ad g. Let a : / , j be a finite display.
Then 8a^= d([ej, fj]) c K a / and 28ej = Cj]) = [5 a /, ej] + [a /, Scj]
shows that 5 a / = 0. It follows that if 8ej = XjCj, then 8fj = -Ay/y for some
Ay ^ IK.
Choose /i e so that aj(h) = Ay for each j. Then 8 — aid h kills Dg. Also
for any h' ^ ij and any x e Dg, we have 0 = (5 — ad h)[h', x] =
[(5 - ad h)h', x] showing that (5 - ad h)(h') centralizes Dg. Since it lies in
1^, it also centralizes ij, hence g. Thus (5 - ad /z)g c (6 - ad AXi) + Dg) c
Zg, and finally 5 e Hom(g/Dg, Zg) + ad g. □

4.2 REALIZATIONS OF CONTRAGREDIENT LIE


ALGEBRAS

In Section 4.1 we saw how every contragredient Lie algebra gives rise to a
structure matrix. In this section we reverse the process and show how to
construct a contragredient Lie algebra with a given structure matrix A. When
A is ai Cartan matrix, this leads to the construction of a Kac-Moody algebra
with structure matrix A.
As a preliminary step it is necessary to construct from A a space 1^, which
will eventually be the diagonal subalgebra of our Lie algebra. This involves
the idea of a realization of A.

Let J be a nonempty set, and let A = (A¿J), ¿,j e J, be a matrix with


coefficients in IK(more properly A is an element of I K By a realization of
A over IK, we understand a triple
/? = ( 5 , n , n ^ )
consisting of a vector space over IK together with two subsets II = {a, | i e
J} c ij* (the dual space of 1&) and II^= { a/ | / g J} c such that

Rl: each of the sets II and II ^ consists of linearly independent elements;


and if we denote the natural pairing of 1^* and l^by<*,->:l^*xf)-^IK,
then

R2: <a¿, a /> = A,j for all /, j g J.


The realization is finite dimensional (of dimension dim 1^) if 1^, and hence
is finite dimensional.

Example 1 Consider A = ^ a,b, ^ K . We distinguish two cases:

1. ab ¥= A. We may take to be a two-dimensional IK-space with basis


n ^ = {aj^, a /} and then define II = {a^, a2) c if* in the only possible
way, namely (a¿, a /> =-^4¿y; 1 < /, j < 2. That H is linearly indepen­
dent follows from the fact that det(^) = A — ab 0.
4.2 Realizations of Contragredient Lie Algebras 331

2. ab = 4. Introduce a third symbol d, and define

= Ka^,

Let n [a^, «2 }, and II = {a^, «2} c 1^* be defined by

a /> =Aij, 1 < i, j < 2,


( a i , d ) = i.

Then R '= (]^, n , n ^ ) is a realization of A.

Example 2 Let J = Z, and define A e by

An = 2,
A^j= and | / - ; | = 1,
Aij = 0, otherwise.

Let ]^ = ® and define e 1^* for all i e Z by

=A^ j , i , j e Z.

By setting n = {a, I / ^ Z} and 11^ = | i e Z}, we obtain a realization of


yl.

Proposition 1
Suppose that card(J) = / is finite. Let 1 = I -\- corank(y4).
Given nonzero elements , S/ o f IK, we claim that there exists an
I X l-matrix A with the following properties :

mini: Aij = Aij for all 1 < i, j < 1.


min2: det(^) 0.
min3: A^j = AjiS^ for a l l \ < i < l < j < l andA^j = Aji for all I < i, j < L
min4: If A is a Cartan matrix and all belong to then A is a Cartan
matrix.

Proof. To see that A exists, first observe that if corank(^) = 0, then A = A,


and we are done. Otherwise, we construct A in corank(^)-steps as follows:
Consider an (/ + 1) x (Z + l)-matrix A^ of the form

^1 =
SiOi
2
332 Contragredient Lie Algebras

Let S := { ( c j , e = 0 for all j}. This is a nontrivial


subspace o f Thus, if H — {(a^ . . . , aj) | Ej^iCjCjaj = 0 for all (c^,..., Cj)
e S}, then H is a proper subspace of Let R be the row space o f A. Then R
is a proper subspace o f IK^ Hence there exists (a^,. . . , a|) e ZL\{R U H}. By
construction the matrix defined by this choice o f a^ satisfies corank{Af) =
corank(A) — 1. Indeed its first I rows have row rank = rank(A) + 1, and the
last row is evidently independent of these first I rows. We now repeat this
argument for A ^ with respect to e\ , , . ., where e\ = and = 1- After
corankiAysteps we obtain a matrix A satisfying min(l)-(3). Suppose that the
hypotheses of min(4) holds. Then A is a Cartan matrix provided that each
Eia^ e Z. Thus must be chosen to cancel the denominator o f which is
easily arranged. □
Remark 1 Condition min(l) states that ^ is a submatrix of A. Condition
min(2) will be used to explicitly construct the realization of A. Condition
min (4) allows us to assume that ^ is a Cartan matrix whenever A is. (We
just make a suitable choice of Condition minO) is important later on
when we define symmetrizable matrices (Section 4.4). It says that if A is
symmetrizable by (e^,. . . , then A is symmetrizable by (e^,. . . , 1, . . . , 1).

Proposition 2
Let A G be a matrix. Then

(i) a realization of A exists',


(ii) if card(J) = I is finite, then ij can be taken to be o f dimension
I + corank(^), and this is the smallest dimension possible.

Proof (i) Define to be a C-space admitting the set {af, | / e J} as a


basis. For each i e J, let =f|* be defined by

<o:¿,a/> =Aij,

Then I I : = { a J / e J } is clearly linearly independent, and with II


{af \i ^ J}, R — (]^, n , n ^ ) is a realization of A.
(ii) Using Proposition 1, let ^ be a Cartan matrix which is 1 X 1, where
/ = / + corank A, which contains A as the submatrix (A¿J)l^¿ J^l, and which
satisfies d e t (^ ) ¥= 0. Let f) be an /-dimensional (K-space. Fix a basis
a^, . . . , ai of Define « i , . . . , ^ if* by a / > = A^j. Then the a¿ are
linearly independent because d et(^) ¥= 0. It is clear that if II — {a^,..., a/}
and { a^ , . . . , af}, then (]&, II, 11^) is a realization of A of dimension I
Finally, let R = 0),YI,U^) he any finite-dimensional realization of A.
Choose bases B and B* of f) and 1^* so that they begin with a f , . . . , a / and
ai , . . . , a i , respectively. Since the matrix M = has
4.2 Realizations of Contragredient Lie Algebras 333

linearly independent rows, in particular its first / rows are linearly indepen­
dent. Now, if S is any k X k matrix of rank s and S' is any (k) X (k 1)-
matrix of the form 5' = [5 I *], then S' has rank < 5 + 1. Applying this to
the submatrix A which occupies the top left corner of M, we see that for its
first / rows to be linearly independent, the number of columns must be at
least / + (/ —rank(^)) = / + corank(^). Thus / + corank(>l) is the mini­
mum dimension for if in any realization of ^4. □

If J is finite and R •= ( / z , n , n ^ ) is a realization of A ^ with


diml^ = / + corank(y4) then we say that R is a minimal realization of A.
Let A and = (f), 11,11^) be as above. We construct a realization R^ of
the transpose of A. Set

R^ = ( f ) *, n" ',n )
where II^c]^ ]^** under the canonical identification of f) inside 1^**. Let
< • , • > : f)* X f)** ^ K

be the canonical pairing. (Note that this pairing is written with the factors in
the reverse order that one might expect. We do this so that we may think of
the new < * , • > as being an extension of the original pairing under the
canonical identification of i) in 1^**), Then

(aj,ar)=A,=Ajj.

Since both n ^ c f)** and II c 1^* are linearly independent, it follows that
is indeed a realization of A^. We say that is the realization dual to
R.
We now begin to construct Lie algebras out of matrices. Let A ^ be
a matrix and R = (1^, I I ,11'^) be a realization of A. View as an abelian Lie
algebra, and let X = {e^, | / e J} be a set of symbols. We let g = be
the free Lie algebra on X and define
f ^*g

to be the free product of f) and g. Let R be the subset of f , consisting of the


following elements for all ¿,j e J and e 1^:

(1)
[h,fj] + (aj , h) f j .

The first Lie algebra that is of interest for us is then defined by


334 Contragredient Lie Algebras

where I(R) is the ideal of f generated by R. We call u(A, R) the universal


algebra of the realization R of A.
Remark 2 Let {h^ I i e 1} be any basis of Then u (^ , /?) can be described
as the Lie algebra generated by the set ej, fj | i e / , ; g J} subject to the
relations [ e j , f J - 8j^a]',[hi,ej] - ( a ^ , [ / i „ / ^ ] + (aj,hi)fj,
where i , k ^ I and j, m g J.

We now look more closely at the Lie algebra u = n(A, R), For conve­
nience the notation for the canonical map f ^ u will be suppressed when­
ever no confusion is possible. Thus we will use expressions of the form
“ Cy G u ” where, strictly speaking, this is to be understood as meaning
€j H- I(R) G u. For the time being, we will denote the image of fj under the
canonical map by that is, + I(R) c u. Later we will see that
is actually an injection, and we will be able to identify i) and
Via the adjoint representation acts on u. Combining this fact with the
canonical map ^ 1^^, we see that u has a natural 1^-module structure, and
under this action h • Cj = a j ( h ) e j , h • f j = —a j ( h ) f j , and h • if ^ = (ff) for all
/z G ]^. In other words, each ej and fj and all the elements of are weight
vectors of the l^-module u. Their weights are aj, -Uj and 0, respectively.
Since u is generated by these elements it follows from Proposition 2.1.2(ii),
that u admits a weight space decomposition relative to 1^. The weights all lie
in the group
Q := 0 l a j c i)=*

and
u = 0 u “,
a^Q
where
u" := (jc e u I[h, x] = ( a, h ) x for all h g

We define

e+==(V;e0j /
\{o},
Q- = -Q ^,
£tnd the corresponding subalgebras

u+= 0 u “.

We also define, in the usual way, the height function


h t : 0 ^ Z,
ht: £ c ,a ,. L c, .
4.2 Realizations of Contragredient Lie Algebras 335

Finally, set

Qk •= Ф the K-span of П in 1^*.

In the same way we define QX, QX, and giK inside 1&.
Our intention now is to show that u = u _ 0 ^ ® u + and that u admits a
regular triangular decomposition. We begin by establishing the following
proposition:

Proposition 3
There exists a unique anti-involution a o f n satisfying cr(cy) = fj, o-(fj) =
and c r I = 1.

Proof The uniqueness is clear since u is generated by together with the


e/s and f/s. The map e^ -e^ extends to an automorphism of
the free Lie algebra Since is abelian, the map x —jc is an automor­
phism of ]^. By Proposition 1.10.4(ii) there exists an automorphism (o of
f = ]^ * g satisfying (o(e^) = - f , caifi) = and <o\f^ = - 1 . The defining
ideal of u is stable under o). For example,

[(o{h),<a{ej)\ - (aj,h}(o(ej)
= e I(R).

Hence (o induces an automorphism (also denoted) co of u. It follows that


(T '= -0) is our desired anti-involution. □

Proposition 4
Let the notation be as above. Then

(i) is generated by {ej I y e J} and n_ is generated by {fj I ; ^ J} (as


Lie algebras);
(ii) u = u_® ® u+.

Proof The argument is precisely the same as that in Proposition 4.1.1. □

Proposition 5
Let the notation be as above. Then

(i) The sum YKf j + + ElKcy is direct and the restriction to it o f the
canonical map f u й injective;
(ii) and u_ are free Lie algebras, freely generated by {ej I ; ^ J} and
{fj I j ^ J}) respectively.
336 Contragredient Líe Algebras

Proof, Let A be the free associative algebra on the set X = {xj \ j e J}. Let
A e ]^* be arbitrary. We proceed to show that A can be made into a highest
weight u-module with highest weight A and highest weight vector 1 ^ A,
Define a representation of ]&on >1 by
h ' 1=
h-Xj^ Jk

for a l l ; i , . . . , ^ G J,/i el^.

Next we make A into an g-module as follows. Since S is freely generated by


{ej, fj I j e J}, it suffices to define the action of each Cj and fj in A. (Recall
the correspondence between ^-modules and left U(g)-modules and the fact
that U(3) is the free associative algebra on {ej,fj\j e J}. See Proposition
1.10.2). We define
f j ' X = XjX for all y e J, jc e y4.

The action of the e/s is defined recursively on each monomial Xj^... Xj^ as
follows:
ey • 1 = 0,
Xf = X:(e: • JC, . . . X: )
Jk Jl^ J Jl Jk'

+ djj^X - aj^ ---- ''Jk'


[The action of e, is defined in such a way as to be compatible with
= fjl ■ ^ifj = fi^ +
We have given A both an and an g module structure. Thus there is a
unique f = * ^-module structure for A compatible with the above (part (ii)
of Proposition 1.10.4). Straightforward computation shows that the ideal
I{R) of f lies in the annhilator of this module and hence that A inherits a
u := f//(jR)-module structure. For instance, consider /y] -

[e,,/;] • 1 = C,- •/; • 1 - fj • •1


= C,. • Xj
= Xj{e^ ■1) + 5,/A,a>'>l
= 5,/A,a,y>l = 5,,ar- 1.
Similarly
[e,, fj\ ■Xj ^... Xj^ = e, • ... x¡^ - x ¡ ■ e, • ... x¡^

= 5„<A - a , , ----------a . , a ^ ) X j . . . x ^ Jk
4.2 Realizations of Contragredient Lie Algebras 337

Now let e ]^ \ {0}. Then for the u-module A we have


(h + I { R ) ) • 1 = /i • 1 = <A,/i>l.
Hence h + I(R) = 0 (in u) only if (A, /z> = 0. Since we are free to choose A
arbitrarily, we may assume that <A, /i> ^ 0. It follows that +/ (/? ) ^ 0. In
other words, the canonical map i) ^ is injective.
The representation of u_ on is just the left regular representation of A
on itself; that is, each fj acts on A as left multiplication by Xj, This gives us a
homomorphism U(u_) A with fj Xj. Since A is free, the inverse map
with Xj ^ fj exists, and U(u_) —A. Thus the Lie subalgebras of Lie (U(u_))
and Lie {A) generated by the fj and Xj, respectively, are also isomorphic.
But the latter is free on the [xj] (Proposition 1.10.1). Thus u_ is freely
generated by the set [fj | j e J}. Finally, a interchanges u_ and u+, and
hence part (ii) is proved. From part (ii) of Proposition 4 and the freeness of
u the sum ElK/y + + LKcj is direct. □
Remark 3 The u-module A described above is evidently the Verma
module of type (1, A) for u.
Remark 4 Part (i) of Proposition 5 allows us to identify with its image
in u. We assume this identification henceforth.

Proposition 6
Let the notation be as above. Then

(i) for all a ^ Q \ {0}, dim u “ < oo;


(ii) for all i e J, dim u “' = dim u " “' = 1;
(iii) if J is countable and ^ is of countable dimension, then u is of
countable dimension.

Proof. We know that u “ = (0) unless a e ± 0 + or a = 0. Take a e Q^.


By part (i) of Proposition 4, u “ is spanned by the products
ej^ ...]], where + ••• = a, of which there are only
a finite number. For the special case a = u “' is spanned by e^. Using a,
parts (i) and (ii) now follow. Part (iii) is immediate from part (i). □

Putting everything together, we have

Proposition 7
Let u = Vi{A, R) be the Lie algebra defined by a matrix A and a realization R.
In the notation established above.

(i) ,5^ := (]^, u+, 0+, O') is a regular triangular decomposition o f the Lie
algebra u;
(ii) if An # 0 for all i e J, then (u, ^ ) is contragredient.
338 Contragredient Lie Algebras

Remark 5 The structure matrix S of the contragredient algebra (u, *9")


above need not equal A, In fact, assuming all ¥= 0, S = A if and and only
if all ¿, Aii = 2.

Example 3 Let A = {d) be a 1 X 1 matrix, a 0. Then with — Kh and


a e ]^* defined by (a , h) = a, we obtain a realization R = (^,{o:},{M) of A
and u = u(A, R) = K f 0 K/z 0 Ke with [e, f ] = h, [h, e] = ae, [h, /] =
-af.
It is easy to replace / , h, and e by suitable multiples (see Section 4.1) and
obtain the standard §l2(IK)-triplet

[e,f]=h, [h,e] = 2e, [A,/] = - 2 / ,


from which we conclude that u = ^I2(1K). The structure matrix of u is (2)
and of course any minimal realization R' = (]^',{a},{a^}) of this matrix (i.e.,
(a, a ^ } = 2) yields u((2), R') = n(A, R).

Let (g, be contragredient. Since later we will work extensively with


such algebras, we will set once and for all the notation to be employed in this
context.

•^ = (^, Q+, Q+, o-) Triangular decomposition


n = Fundamental roots

n '- Fundamental coroots

< 2 = 0 Zaj Root lattice


jsj
0 Za/
;s j
Coroot lattice

<2k = 0 IK«; c

= 0
yej-
e += 0 0 _= - Q ^
yej
The structure matrix of ( g , ^ )
< -.-> :rx 5 ^ K The natural paring
J /jr}yej A display of (g , (in particular
cr(ej) = fj for all j e j )

Consider the triple R == (1^, II, 11^) arising from (g, 3^). It is immediate
that R is a realization of the structure matrix ^ of (g, (note that we
4.2 Realizations of Contragredient Lie Algebras 339

need CT3 for this). We call R the natural realization of A in the context of
(g, ^ ) . Let us now construct the universal Lie algebra u = \i(A, R) as
above. Set

and

f=
Then
u = f//(i^),

where I(R) is the ideal of f generated by the elements of the set R defined
in (1). Recall (Proposition 5(i)) that

(2 ) I 0 K/,.j ® ^ e I © cf

can be identified with a subspace of u. Accordingly we henceforth identify


with a subalgebra of both u and g and think of ej and fj as elements of both
of these algebras. (It will always be clear to which algebra we are making
reference.)
As we have seen above the algebra u admits <5^ as a triangular decompo­
sition, where
= (f |,u^ ,G^ ,o-).

Moreover (u, ^ ) is contragredient (Proposition 7(ii)). Because of the way in


which f is defined (i.e., as a free product), there exists a (unique) Lie algebra
epimorphism

satisfying

Hfj) =fj’
(¡/(h) = h for all e ]^, 7 G J.

Since i/r(/(i?)) = (0), there exists an induced epimorphism

q.

This allows us to realize g as a homomorphic image of u. We refer to


340 Contragredient Lie Algebras

(u, Ty, if/) above as the universal covering of (g, «^). Evidently

(3) ^(u^)=g« for all a e g .

Proposition 8
Let ( g , be contragredient, and let (u , , i/r) be its universal covering.
Then

(i) ker((/r) c rad(u),


(ii) (/i(rad(u)) = rad(g).

Proof. In view of Proposition 2.7.5, we have t ^ ( u ) = m^^(u) = rad(u) and


= my^Cg) = rad(g). Since is an isomorphism, it follows that
ker(i/i) n ]^ = (0), whence we have part (i). Similarly (/r(rad(u)) is an ideal of
g which intersects trivially, while i/i”Krad(g)) is an ideal of u which
intersects trivially. This establishes part (ii). □

Let A G and let R he a realization of A. Define

Q= q ( A , R ) := u (^ ,i^ )/ra d (u ).

By (2.7(i)), rad(g) = (0). We call g (^ , /?) the radical free algebra of the pair
(A, R). Note that if A^^ ¥= 0 for all i e J, then g is contragredient, and in
this case we call q(A, R) the radical free contragredient algebra of the pair
(A,R\
We have seen that any contragredient algebra (e, with the natural
realization R has a universal covering (u, i/r). If furthermore e is radical
free, then by Proposition 8 ker if/ = rad(u), and it follows at once that e is
canonically isomorphic to g (^ , jR).

Proposition 9
If e is a radical free contragredient algebra with structure matrix A and
realization R, then there is a canonical isomorphism e g (^ , i?).

Our interest is primarily in the case that ^ is a Cartan matrix. In that case
q( A , R )
= u (^ ,/? )/ra d (u ) is integrable, and hence, a Kac-Moody algebra,
by Proposition 4.1.11. However, there may be intermediate Kac-Moody
algebras, between n(A, R) and g(v4, R).

Proposition 10
Let A be a Cartan matrix with realization R, and let n{A, R) be the corre­
sponding contragredient algebra. Let J{R) be the ideal o f vl{A, R) generated by
4.2 Realizations of Contragredient Lie Algebras 341

the elements

\dtj = {&áe¡) i*j


(R2)
1^.7 = i* i

Then R) •= u(yl, R)/ J{R) is a Kac-Moody Lie algebra. Furthermore for


any Kac-Moody algebra % with realization R, there exist surjective homomor-
phisms

q\ A , R ) ^ q ^ q( A , R )

{with the obvious effects on their displays) under which

ra d g ^ ( ^ , i? ) rad(g) ^ r a d g ( ^ , i ? ) = (0).

Proof. That g^(^, R) is integrable follows from Propositions 4.1.8 and 4.1.10.
The existence of the map q^{A, i?) g follows from factoring the covering
homomorphism u{A, R) ^ % through q^(A, R) by Proposition 4.1.9. The
rest follows from Proposition 8. □

This proposition is a natural complement to Proposition 4.1.9. In the


sequel (Section 4.6) we will find that for the cases of most interest Q^(A, R)
= %{A, R).

Example 1 A = bring the notation in line with our usual


notation for §Í2(IK) in the Appendix on A^^\ we will take II = {«(„«1},
n'^= a^} with

<«,-,«/> =A¡j

and set

h = IKif + + IKa^.

Using R = {§, n , n ''} , we define u(A, R). It consists of two free Lie algebras

^ +~ u _ —( / o , / i ) ,

and the abelian subalgebra h:

u = < /o> /i> ® < eo,C i> .


342 Contragredient Lie Algebras

The algebra u is contragredient and simple considerations with free Lie


algebras show that its root system is

A(u) = A(u)_u{0} U A(u) + ,

where

A(u) + = [ka^ 4- ma^ | f c , m > 0 , A: + m > 0 ,


and fc(resp. m) = 1 if m (resp. k) = 0}
A ( u ) _ = “ A(u) + .

The root space multiplicities (dim u “, a e A(u)) can be computed by Witt’s


formula (Proposition 2.5.5).
Since ^l2(IK) is contragredient with structure matrix A there is homomor­
phism

i/i : u êl2(IK)

with

e,
ar^h„

I = 0,1 (the fact that corresponding elements u and § 12(11^) have the same
name should not cause confusion).

This example continues in the Appendix on (see Section A.3.)

4.3 EMBEDDINGS, FIELD EXTENSIONS, AND


DECOMPOSABILITY

Let J c J', and let A g be a submatrix of A g Suppose that


R = (1^, n , n'^) and R' = (^', II', 11'^) are realizations of A and A , respec­
tively, where II = II'^= n' = and 11'''= ej-.
An embedding of i? in i?' consists of an ordered pair 17 = (•>7^, 77^), where 77^
(resp. is an injective linear map

(resp. -» 1^')
4.3 Embeddings, Field Extensions, and Decomposability 343

such that

(i) 77^:11^11' with 7]^a^) = a- for all i e J;


(ii) n'"" with for all i e J;
(iii) for all a e f)*, for all f| e i|, t7^(/z)>' = (a, h);

where ( • , • ) X ij and <•, •>' : f)'* X i)' are the canonical


pairings.

Note that since <«;, = A]j = A^j = (a^, a j ) for all /,; e J the con­
ditions (i) and (ii) are compatible with condition (iii). At the lattice level
there exists induced Z-linear maps Vl ‘ Q ^ Q' • Note
that for ¡L e

( 1) ^ V l ( Q ) if and only if supp(/i') c J.

Remark 1 The most obvious and usual example of an embedding occurs


when one takes f) to be any subspace of 1^' containing such that
{a-lygj, if restricted to functions on 1^, is a linearly independent set of 1^*
(e.g., 'tj = 1^'). Define re s: 1^'* -> 1^* to be the restriction map arising in a
natural way from the inclusion c 1^'. Then with II — [a\ I / ^ J} and
n^:= {a'^ I i e J}, we see that /? = (1^, II, n ^ ) is a realization of A. Fur­
thermore, with the maps

inclusion map,
^ any linear map that lifts the elements of 1^* to linear
functions on ij' (i.e., res • rjj^ = id) and such that T7^(res(a^)) = for
all i e J,

we have that 17 == (17^, is an embedding of R in R \ We call R a natural


realization of A in the context of R \
Remark 2 We have assumed in our definition that J c J' and that ^ is a
submatrix of A'. It is rather easy to extend the definition to an embedding
(p: J -> J', and matrices A and A' indexed by J X J and J' x J', respectively,
satisfying = ^ ij for all i , j e J. Results involving 17 have obvious
generalizations to this case.

Proposition 1
With the notation as above let rj = ( v l ^Vr ^ tm embedding o f the realization
R of A in the realization R! o f A . Then there is a canonical embedding {also
denoted 77)

77 : u := n ( A , R) u' := n { A , R')
344 Contragredient Lie Algebras

such that

(0 7j(]^) = ■njiiif) C 1^';


(ii) T7(e,) = e'i and rjifi) = for all i e J;
(iii) T7U" = for all a ^ Q \ {0};
(iv) a'(rix) = riicrx) for allx e u, where a and a' are the anti-involutions
on u and u', respectively.

Proof Let X ■■=[ej, fj}j ^ j and X' ~ {e'j, fJ}j ^ j. and define f and f' to be
the free products ii * and 1^' * that are used to define u and u',
respectively. We define a Lie algebra homomorphism

i/i: f ^ f' i

by

= Vr ,
il>(e¡)=e’¡, , ¡ , { f ) = f ¡ . for all i e J,

and observe that the relations R involved in the definition of u are annihi­
lated by ip. For example,

- «0«") = [<’ fj] - 0-

Thus ip factors through the ideal /(R) to give 77: u u'. Since 17(6^) = e)
and and are freely generated by the {e^} and {e'¿}, respectively, we see
that 7] embeds u+ into u+. The same applies to u_ and u'_. All the
statements of the Proposition are now obvious except perhaps for the
equality in part (iii). For that, observe first that for all a ^ Q \ {0} we have
u “ = (0) = unless a ^ Q+.
Assume then that a ^ (the Q_ case is similar). Then u" is spanned
by products [Cj^,. . . , CjJ with j) in J and aj^ + • • • +ay^ = a e Q+. Next
^rvLM jg spanned by products [e'y,. . . , e'j^] with in J' and a'j^ + • • • =
i7^(a). By (1) {j\ , . . . , ^} c J, and hence each of these products lies in t7(u").

Remark 3 The most common situation in which the last proposition arises
is the one in which A is some submatrix, say, ij-'l i,j < kl
of some matrix
A = that has some realization R! = (f)', II', 11'^). For the real­
ization R one may take (f)', II, 11^), where II — [a\ , . . . , a'j^} c II' and —
. . . , a '/ ) c (see Remark 1). Then u = u(yl, R) embeds into u' =
Vi{A, R ) by ^ fi ^ fi, a\^ so that u may be identified with a
subalgebra of u' in the most obvious way imaginable.
43 Embeddings, Field Extensions, and Decomposability 345

Another common and equally obvious embedding occurs when A, as


above, is nonsingular (even though A' may be singular). Then we may use the
realization R = ((2k >n , n'^), where 2 ^ == 0 <=i
* ''

Example 1 Let be a realization of A g and let 7 g J be fixed


with Ajj # 0. Let = (Ka/,{a^},{a/}). Then, as above, is a realiza­
tion of ( A j j ), and we have the canonical embedding u((Ajj), R^^'>) -» u(^4, R).
As we see in Example 4.2.3, u ( ( > l ^ . p , = gljClK). Now u((y4^.p, is
isomorphic to the subalgebra of \i(A, R) generated by Cj, fj, a / , which we
know also to be isomorphic to SljiK).

Proposition 2
Let Tj be an embedding o f a realization R in realization R’. Let 17: u(A, R) -»
\i(A, R') be the corresponding embedding o f Lie algebras. Then

(i) 77(rad(u)) = rad(u') n tj( u );


(ii) the induced homomorphism q(A, R) g(A', R') is injective-,
(iii) 77^ A = A' n t]j(Q), where A and A' denote the root systems o f q(A, R)
andgiA', R'Xrespectively.Moreoverforalla G A \ {0},77(g(y4, i?)“) =
g(A',
(iv) if g(A, R) and g(A', R') are Kac-Moody algebras, then

-nr.C^A) =^^A' n 7,^(2),

77^(""A) =""A' n 77^(2).

Proof Let J and J' denote the indexing sets for A and A', respectively. Let
a' ■■=Ttifa) G 2'+. where a g A+, and let x g u “. Assume x g rad(u). To
show that y ~ t](x ) g rad(u'), we must verify that given y^ g g ' " '
Vk ^ S -Pi where ..., ^ 2'+ and -I- • • • +iS'* = a', then
[yi. • • •, y*. y] = 0. Since supp(a') = U f=i supp(j8-) it follows from (1) that
supp(j8p c J for all 1 < i < A:. Thus = t7^(/3,) for some /3, G 2+ , and
hence y, = t7^(x,) for some x, g u “^' (part (iii) of Proposition 1). Thus
0 = 77(0) = T7([x i,. . . , X*, x]) = [yj,. . . , y*., y]. In the same way 77(rad u)“ c
radu' if a G A_.
On the other hand, if x i rad(u), then it is evident that y € rad(u'). Thus
77(rad(u)“) = rad(u')’’i-(“) n t7( u ). Now parts (i) and (ii) follow at once.
It is clear from (ii) and Proposition 1 that TjjfA) c A' n rjjfQ). For the
reverse inclusion it suffices to take a' g A'+n r]j(Q). Then a' g -q^fQ^), and
so a' = ■ +a-^ for some i j , . . . , G J. Since a' g A'+, for some
permutation of i j , . . . , say, the trivial one, [e^, ■■■, e^] * 0. Thus
[e,y...,e,^] # 0 and 6' ■= T7^(a,^ -!-••• +a,^) ^ t7^(A). Furthermore, since
346 Contragredient Lie Algebras

q(A', R')“' is spanned by the products of the form [e\^,. . . , eJJ,


riiqiA, i?)“) = q(A', R')“'. This proves part (iii).
According to Proposition 4.1.14, for n e Z+, n > 1,

(2) a e" A +( n ) => 3; e J, such that (a , a j ) > 0,

a - { a, a' j ) aj - < a ,a / > ) .

Similarly

(3) e"A '+(n) ^ 3; e J', such that <7,^«, < > > 0,

rtL*^ - e''^A'+(n - <17^«, a}'")).

Now supp(i7^(a)) e {a -1 i e J}, and hence a!^) > 0 is possible only


if j e J. Thus in (3), J' can be replaced by J. Now part (iv) follows by
induction on n, starting from the obvious fact that ry^C^Ad)) = t7^(A(1)) =
A'(l) n 71¿ Q ) =^"A'(1) n 71¿ Q l □
Remark 4 If we are in the cases described in Remark 3, then part (iii) of
Proposition 2 assumes the simple form

(4) A = A' n E

Remark 5 In this proposition q(A, R) and q(A', R') can be replaced by


q\ A , R ) and q \A ',R ').

We now look more closely at derived algebras. Let g = q(A, R) be the


radical free Lie algebra defined by a matrix A and realization R. The proof
of Proposition 4.1.12 (which is for contragredient algebras) carries over
directly to show that

(5) Dg = g_0 0 g+.

Here we will give an entirely different description of Dg that shows that it is


dependent only on A (not on R).
Let Л be arbitrary, and let b be the Lie algebra defined by the
generators e^, (/ e J), together with the relations

(R')
43 Embeddings, Field Extensions, and Decomposability 347

where h¡ ■= [e,, /,]. We begin by noting that

[ht,hj] = [ hi , [ e j J j ] \ = [Aj^ejJj] + [cj, - A j J j ] = 0 ,

which shows that a == is an abelian subalgebra of b. We can grade b


by (the set of mappings of J into Z). Let be defined by
«/(;) = ^¿j, and set
deg = a^,
deg/, = -a,.,
deg hi = 0.

Then the relations (R') are homogeneous, which suffices to determine the
grading on b.
We have need only of the subgroup Q' of consisting of functions a
with finite support. Let Q'+ = {En,o:, e Q' \ > 0} \ {0}, and let Q'_ =
Then with b+:= we have by a trivial adaptation of Proposition
4.1.1,
b = b_0 a © b ,
and
b += e J>,
b_= </- 1/ e J>.

Evidently there is a surjective homomorphism

</): b ^ Du c u ( ^ , jR)

defined by = e^, (/>(/) = / , and hence with (f>(hi) = a / , since the


corresponding elements in u satisfy the relations (R').
Since u+ is freely generated by 1/ e J}, <f>:b+ —^ u+. Likewise
(j>: b_ —^ u_. Since { a/ | / g J} c c u are linearly independent, we have
(f>: a , and we conclude that </> is a Lie algebra isomorphism.
Furthermore there is an isomorphism </>: Q' —^ Q with a, and

Define

(R'') r := maximal homogeneous ideal of b intersecting a trivially.

Since r is graded, <f>(x) = n u “) from which it is clear that


(j>(x) is invariant by ad 1^. Thus <^(r) is an ideal of u (not just Du), and hence
(/)(r) c rad(u). On the other hand, rad(u) c D(u), and <^"^rad(u)) is a
graded ideal of u intersecting a trivially. Thus r = <^“ Krad(u)), and we have
348 Contragredient Lie Algebras

finally

b /r = Dg.

This gives the alternative description of Dg by (R') and (R").

Proposition 3
Let A e be a matrix, and let R b e a realization o f A,

(i) Dg(yl, R) —b /r , where b is the Lie algebra defined by the presenta­


tion (RO and r is the ideal defined by (R"). In particular Dq(A,R)
depends only A.
(ii) Dg w perfect if and only if A has no row consisting entirely of zeros.
(iii) I f A is a Cartan matrix then Dg^(^, R) (see Proposition 4.2.10) has
the presentation with generators e^, /¿, h^, i e J, and the relations

(R l)
[e¿,fj] = 8¿j hj for all i j ;

i(adc,)'^'^"^e, = 0,
(R2)
\ fj) V/ = 0 for ii^j.

Proof (i) The proof has already been established.


(ii) and (iii) The proofs are left as exercises. □

Corollary
Let (t, and (c', ^ ' ) be radical free contragredient Lie algebras admitting
the same structure matrix. Then De = Dc'.

Proposition 4
Let A e and let R be a realization o f A. Let g = g(yl, R). Then the
centers of g and Dg are given by

Zg = (/i e I h) = 0 for all i e j},


ZDg = {A e G k K«/, A> = 0 for alii e j).
43 Embeddings, Field Extensions, and Decomposability 349

Proof, Let z be central in g or Dg. Then writing z = ^ 9“?


see immediately that each z" is central. Proposition 2.7.3 shows that z “ = 0
if a e A+U A_. Thus z e ]^, and the result is immediate. □

Let (g, be contragredient, and let {e^,ay,f^} be a display of g and


R = {]^,n, n^} the natural realization of the structure matrix A of g. The
situation when J is finite and i? is a minimal realization is the most common
and most important. Assuming that J is finite, we put together a number of
conditions which are equivalent to R being a minimal realization. Write
^ = i2iK ® ^ for some subspace t of i). Then each f e e t induces a linear
functional on <2iK by a <a, fc>, and we obtain a linear mapping

i \ i ^ Q\IK-

Proposition 5
Let ( g , r ) be contragredient with natural realization R and structure matrix
A= Assume that J is finite. Then the following are equivalent.

(i) R is a minimal realization o f A.


(ii) Zg c Dg.
(iii) Zgce„^.
(iv) For every supplement t o f in 1^, the corresponding map i : t Qt
is injective.

Proof. Since Zg c and Dg n = 0ii< >it is clear that parts (ii) and (iii) are
equivalent.
(i) => (iii) We have dim finite by Proposition 4.2.2. Suppose that z e
( ^ \ G k ) Zg. Then {ai , z } = 0 for all i e J, and II induces a srt II
of line^ly independent functionals on 'ii/Kz. The realization R •=
{i)/Kz, n , n'^(mod Kz)) contradicts the assumption of (i).
(iii) (iv) Let t be a supplement of in f). Then

ker t = {z e 11 (a,, x) = 0 for all / G j}

= t n Z g c t n ( 2 K = (0)-

(iv) ^ (i) Let t be a supplement of in f). Since i : t -» <2k is injective


and dim 0 k is finite, dim (| = dim + dim t is finite. Thus through the
pairing <• , • >, induces all of ^**; in particular ii induces all the
functionals of 0K- Thus we have the exact sequence

Zg ^ ^ 0K 0.
350 Contragredient Lie Algebras

Now, by assumption, t n Zg = (0). Since t was arbitrary, Zg c Q'^. Thus


Zg = {/z e | <a^, h) = 0 for all i e J}. Writing out the equations {a^^h)
= 0 in terms of the basis {a/} of , we see that Zg corresponds to the null
space of A and hence dimZg = corank(^). From the exact sequence we
have dim = dim + corank(^), and part (i) follows.

A contragredient Lie algebra (g, is minimally realized if

MRl: it has a finite display (i.e., J is finite)


MR2: any one of the conditions i-iv of Proposition 5 hold.

Proposition 6
If (ind (g', ^ ' ) are minimally realized radical free contragredient Lie
algebras with displays S ' = a / , j and S) = , fj)j^j tmd with
the same structure matrix A, then g and g' are isomorphic by a (noncanonical)
isomorphism <p such that cp(ej) = e'j, (pifj) = / / , (p(af) = a'j^ for all j e j
and <p(i)) = f)'.

Proof Let R and R' be the realizations of A defined by (g, and


(9', Since they are radical free, g = q(A, R) and g' = q(A, R '\ and
furthermore by Proposition 3 Dg = Dg' by an isomorphism, call it
mapping S into in the required way.
Write 'i} = 0 t, where t is some subspace of 1^, and let t : t be
the natural injective map of part (iv) of Proposition 5. Let be a
basis of t, and let k'^ be chosen (the choice is not necessarily unique)
so that kj induces the function on Q'^ corresponding to kj; that is,
(a^, kj) = (a'i, kj)for all / e J. These k'j are linearly independent and
EIKA:y n = (0) (since t n = (0)). Set t' == EIKA:), and extend (p to 2l
linear isomorphism of g to g' by defining (p(kj) = A:', j e J. Then is our
desired Lie algebra isomorphism. □

Corollary
Let (g, r ) be a minimally realized radical free contragredient Lie algebra, and
let r be the Coxeter-Dynkin diagram o f its Cartan matrix A. Then any
automorphism y o f T lifts to an automorphism y o f ^ such that ye^ = e^^j^y
yfi = fy^iy and y { af ) = i e J {where we have expressed y as a permuta-
tion of the index set o f the nodes of Y). □

Since a Cartan matrix A can always be embedded as a submatrix of a


larger Cartan matrix A , we can likewise embed g ( ^ , R ) into %{A, R) for
suitable R and R . This opens up a number of techniques for proving results
4.3 Embeddings, Field Extensions, and Decomposability 351

about representations that are not available in the finite-dimensional theory.


For our purposes the following result will be sufficient.

Proposition 7
Let be a minimally realized radical free contragredient Lie algebra, and
let R be the natural realization o f its structure matrix A = [so that
g = of A, i?)]. Let J' be a finite set with J c J', and let A' be any Cartan
matrix indexed by J' X J' with A as a submatrix o f A' in the obvious way. Let
R' be a minimal realization o f A'. Then R embeds into R' compatibly with the
set inclusion J c J' and q(A, R) embeds into (¿{A', R') so that e^ e',
a/ ^ / / for all i e J. Moreover, if A and A' denote the root systems
of <^{A, R) and of A', R ), respectively, and we view L as a subset o f Q then
A = A' n

Proof Let R = be a minimal realization of A'. Set IIi —


{a-ji e J} and ^ {a\^ \i e J}. It is easy to find a subspace of 1^'
containing so that R^ = (l^j, IIi^) is a minimal realization of A.
Then R^ embeds in R trivially, and our result follows from Propositions 2
and 6. □

Let jR = ( ] ^ , n , n ^ ) b e a realization of a matrix A over a field K, and let


u(A, R) and ^(A, R) be the universal and radical free Lie algebras of (A, R),
respectively. Suppose that K' is a field extension of K. Replacing by
:= K' and identifying 1^* inside in the obvious way, we obtain a
realization

R^, := ( ^ ^ , , n , n ^ )

of A over K'. We then form the Lie algebra u(A, R^>) over K'. It is routine
to verify that

(6)

the isomorphism being canonical:

i)K' ^ K' ® ^ (identity m ap),


e, 1 ® e¡,

fi ^ 1 ®/i-

In addition by Proposition 2.7.4,

rad(K' ®|)^u(/l, i?)) = IK' ®||^ rad u ( A , R),


352 Contragredient Lie Algebras

and hence through the isomorphism (6),

rad(u(^,i?K')) = K' radu(/l,i?),

which leads to

(7) q{A, = IK' i?),

with

IK- ^ IK' ®K ( identity m ap),


1 ® e¡,

^ 1 ®/i-

If a contragredient Lie algebra q(A, R') over the field IK' has a subalgebra
g (^ , R) over IKsuch that q(A, R') = IK' ® R), then we say that q(A, R)
is a IK-rational form of g(yl, R').
Given a square matrix A e IK^^'*, we define the skeletal graph of A to be
the graph F whose nodes are indexed by J and in which the nodes i and j
are joined by an edge from i to j iff A¡j 0 oi Aj¡ ^ 0. (In Chapter 5 we
define the Coxeter-Dynkin diagram, which is a more elaborate version of this;
see also Section 3.4.) The connected components of F correspond to the
minimal nonempty subsets of J such that A¡j = A¡¡ = 0 whenever i e
j e p ^ q. The submatrices A^^^ defined by the subsets of J
are called the indecomposable blocks of A. A is indecomposable if F is
connected (so that there is only one block).
In all cases of any importance to us we will find that A¡j =/= 0 <=»Aj¡ ^ 0,
and these simple notions of connectivity are sufficient.

Proposition 8
Let R b e a realization o f a matrix A e IK-*^-*, let F be the skeletal graph of A,
and let g = g(yl,f?). Let J = be a partition of J so that the
subgraphs F^^^ corresponding to the subsets are mutually disconnected
(i.e., A¡j = 0 = Aj¡ whenever i e j e p q). For each p ^ P, let
R*'”^ be a realization o f the submatrix A^‘’^ ■■=(A¡j)¡ yej«’» so that R^'’^ is a
natural realization o f A^'’'>in the context o f R {see Remark 1). Identify the
radical free Lie algebra g^^> == q(A^’’\ R^>’'>) as a subalgebra o f g by Proposi­
tion 2. Then

(0 9^= X ^ ^ ^ g V ’), g_=


(ii) D g = X^^^Dg<^>,
(iii) g = Eg^^^ + where is the diagonal subalgebra o f g.
43 Embeddings, Field Extensions, and Decomposability 353

Proof. Let p, q e P, p ^ q, and let i e j e Then [/„[e,., effi =


-[«y, Cj] = -AjiCj = 0. Similarly [fj, [e„ Cj] = 0, and it follows at once that
[e,,Cy] e rad(g) = (0). Since the a„ i e J, are linearly independent in we
see that [g(f> n = (0). We have just seen that [g^/\ = (0) if
p ^ q, so we have proved the first part of (i). The other follows in the same
way (or one may use cr).
Since Dg^^^ n = 0 and since Dg n i) = <2k =
we may use Dg g"©(Dg n 1^) © g"^ to see that part (ii)
follows from part (i). Now part (iii) is immediate.

We call a square matrix A combinatorially symmetric if for all i and j in J


we have

(8) A:: ¥= 0

and

(9) Aji = 0 whenever = 0.

Proposition 9
Let A be combinatorially symmetric, and let R = be any realization
of A.

(i)
I f A is indecomposable, then any ideal o f q(A, R) is either inside ij or
contains Dg.
(i) If J is finite and f) = Q ^, then the decomposition o f ^ = ^
defined by the indecomposable components o f A is a decomposition of
g into simple ideals. In particular g is simple whenever A is indecom­
posable.

Proof (i) Let a be an ideal of g. Then a = ea", where a“ = a n g" c


a and a" ¥= 0 for some nonzero root. Without loss of generality we can
assume that there is an jc e a" with a e A+, x ¥= 0. Since rad(g) = 0, there
is a sequence of elements ..., where fj. e g~“'j, i^,. . . , i^ ^ J, such
that 0 [fi^,. . . , fi^, x] e i). Then [f^^,, , , , f- x] e (IK \ {0})ei^ e a. Let
Jq := {j e J |ej G a). Since [ej,[fj,ej]] = [aj'^jeJ = A^ej it is evident that J q
indexes a union of connected components of r ( ^ ) . Thus J q = J. From ej e a
we obtain [fj, ej] = - ci f , and [fj, = Ajjfj are both in a. Thus fj e a for
all j e J, and hence a d Dg.
(ii) Since ^ = Q k , we have g = Dg. Assume first that A is indecompos­
able. Since J is finite, dim(0K*) = card(J), and evidently II is a basis for
354 Contragredient Lie Algebras

0IK* = 5*- Then, given e f), h 0, there exists an i e J, for which


[h,e¿] = ce^ 0. Thus any ideal of g containing h contains some root space
g“‘ and with what we have proved in (i), this shows that g is simple.
In general let J = t>e the decomposition of J into subsets
labeling the connected components of the skeletal graph of A, For each
p e P we have

Kaj

and

:= (a e r I a = 0 for all q ^ p) = .

Set = {i>:y};sj(p) c There is a natural em­


bedding of == ( 2 k n<^>, into R. With Q^P'> ■■=
g((^,y), ye we have = Dg^^\ which is simple by part (i). Now by
part GO of Proposition 8, g=xg<^>. □

We finish this section by relating Kac-Moody algebras to the finite-dimen­


sional semisimple Lie theory.

Proposition 10
Let A be a Cartan matrix o f finite type, and let Rbe a minimal realization of A,
Then

(i) the Lie algebra g = %{A, R) is, up to isomorphism, independent ofR;


(ii) g is finite dimensional',
(iii) g is semisimple with simple ideals corresponding to the indecomposable
blocks o f A; in particular g is simple if A is indecomposable.

Proof See Exercise 4.1. □

Let A be an indecomposable (resp. decomposable) Cartan matrix of finite


type. Let jR be a minimal realization of A. Then g = g(-^4, R) is called the
split simple (resp. semisimple) finite-dimensional Lie algebra over K with
Cartan matrix A.
Remark 6 It is a fact that if IK is algebraically closed then every finite­
dimensional simple Lie algebra over IK is of the form g(-^4, R) A indecom­
posable, as above. Over nonalgebraically closed fields there are (in general,
many) other possibilities.
4.4 Invariant Bilinear Forms 355

4 .4 INVABIANT BILIN E AR FORM S

We wish to find necessary and sufficient conditions for a contragredient Lie


algebra to carry a nontrivial symmetric invariant bilinear form. We formulate
an existence theorem in the general context of Z-graded Lie algebras.
Let 9 be a Lie algebra over IK graded by Z, g = ® * ^ k eN
define

q{k) = 0 9'.
Uls/t
A bilinear form

(• 1•) : 8 (* ) X g(*) ^ IK

will be called locally invariant if

( [ y , ^ ] U ) + ( x | [ y , z ] ) = 0,
or equivalently
([^ ,y ]U ) = (xl[y ,z])
for all x , y , z ^ g(fc)
for which [x, y] , [ y, z] e g(fc).

Proposition 1
Let g = ^ Z~graded Lie algebra over IK, and suppose that there
exists h ^ such that for both e = 1 and e = —1, ^ generated by
Suppose that

(1) Xg(/i)

is a symmetric locally invariant bilinear form satisfying

(2) ( qM qO "" (^) ^ j ^ Q for alii J with 1/|,|;| < h.

Then (* I *) extends uniquely to a symmetric invariant bilinear form on g


satisfying (2) for all i, j e Z.

Proof We extend (• | • ) to g(n) X g(n) for all n e Kl, az > /z by induction on


n. Assume that (• | • ) is defined on g(n - 1) and is symmetric and locally
invariant. We have to define ( I*) on g(n). In view of (2) we define
(g"*"" I gO = (0) and (g* I g^") = (0) unless i = +«. It remains to define
356 Contragredient Lie Algebras

(• I *) on g - " X At this point we interrupt the proof to establish the


following
Claim Let / , k, I be integers satisfying

(3) 0< |i|,l;U A :U /l < n,

(4) i + / = ±n, k + I = +n.

Let u‘, v \ w*, z ‘ be elements of g', g^, g*, g^ respectively. Then

(5) I z ') = ( m‘ | [ u^ [ h'* , z ']]).

Proof (of Claim). Notice that i + j + k + I = 0, and hence |i + j + A:| = 1/|


< n and \J + k + l\ = III < n. Thus both sides of (5) make sense. Also notice
that |i + ¿1 = I; + /| < n and that \j + k| < n. We have

([[«', |z') = ([[u‘,H'*],y'] \z‘ ) + ([«', [u^,vv*]] |z')

= ([[u',H'*^]|i;^z']) + (M'|[[i;^w*],z'])

= {u‘ \ [ w \ [ v f z ‘]]) + {u‘ \ [ [ vf w' ^ ] , z ‘])

= (^г‘ |[ t^ ^ [ н '^ z '] ] ) .

This establishes the claim. □

Now let X e g^", y e g“®", where e = ± 1. We want to define (jc | y). By


assumption for some k , m ^ Nv/e can write

x= T , [ x ' j , x ' ; ] , y = H[ y ' j , y " ] ,


;=i ;=i

where the x), x] e ©"^j g^‘ and the y], y'J


Define

( 6) ( x \ y ) ■■= Z { [ x , y ' j ] \ y j ) .
y=l

To see that (6) does not depend on the way in which y is decomposed, we
4.4 Invariant Bilinear Forms 357

observe that

(7)
k \ m k,m
E { x' i , x'l]\y ( j: I y ) = E ([x, y'j] I yj) = E ( [ [ ^ :, x'[], y'j] I yj)
/=1 ; =1 i,j=l

k,m
= E (^il[<.[y;,y;]]) (by the Claim)
i.j=l

= E (jc;i[< .y])
/=1

and hence that (6) depends only on y as desired. To see that (• h ) is


symmetric, we reason as follows:
m
( ^l y) = E [by (6)]
y=l
m
= - L (y" I[y). ^]) [by symmetry on %{n - 1)]
; =i

= E ([y;,y;]i^) [by (7) interchanging x and y]


; =1
= (yU ).

It remains to be shown that (• | • ) is locally invariant on g(n). For this it


will suffice to show that ([x, y]| z) = (x l[y, z]) for homogeneous elements
X, y, z, where 0 < |deg(x)|, |deg(y)|, |deg(z)| < n, degix) + deg(y)
+ deg(z) = 0, and at least one of these degrees is ±n. If deg(x) = ±n and
0 # deg(y), 0 ^ deg(z) the invariance follows from (6). The same applies if
0 ¥=deg(x), 0 deg(y), and deg(z) = ±n. There remain three cases:

Case 1. 0 < |deg(x)|, |deg(z)l < n, deg y = ±n, deg([jc, z]) = Tn,

( [ A t , y ] U ) -----( [ y , x ] | z )
= -{y\[x,z]) [definition (6) on [;c, z]]
= (y\[x,x])
= [by (6)]
= (x,[y,z]) (by symmetry).
358 Contragredient Lie Algebras

Case 2. deg(jc) = ±n, deg(y) = Tn, deg(z) = 0. We may assume that x


has the simple form [ jc' , x "], 0 < |deg(x')l, |deg(A:")l < n. Thus

( j : | [ y , 2 ] ) = ([j c',JT "]l[y ,z])


= {x'\[x",[y,z]]) [by (7)]
= ( x ' \ [ [ x " , y ] , z ] + [y,[^:",2 ]])

(by induction assumption and case 1)


= ([x', [x", y]] I z) + ( [ [ a;', y], x"] 1z)
(by induction assumption)
= ([^>y]lz).

Case 3. deg(jc) = ±n, deg(y) = 0, deg(z) = H-n. Then we have

( x | [ y , 2 ]) = - ( [ a: , z ]| y) = ( z |[ jc ,y ] ) = ( [ j c , y ] | z ) ,

using case 2 twice and the symmetry of (• 1• ).

This completes the induction step.

We turn our attention to contragredient algebras. Let (g, be contra­


gredient as in Section 4.1 and maintain all the notation that was established
there and in Section 4.2.

Proposition 2
Let g be contragredient as above, and let (* I *): 9 X g ^ IK be an invariant
bilinear form. Then

(0 ( 9 “ I 9^) = (0) for all a, p Ei Q such that a + j8 ^ 0;


(ii) for ( ' \ ' ) to be nondegenerate, it is necessary and sufficient that the
restriction o f to 9“ X g~" be nondegenerate for all a ^ Q;
(iii) (• I • ) w symmetric on Dg X Dg.

Proof, (i) Let a, /3 e and let jc e g“, /i e 1^, and z e g^. Then

( [ x , h ] \ z ) = ( x \ [ h , z ] ) <=> (/3 + a)(/z)(jc I z) = 0.

Since h is arbitrary, this shows that (jc | z) = 0 unless ¡3 = - a , thus estab­


lishing part (i). Now (ii) follows from (i).
4.4 Invariant Bilinear Forms 359

(iii) Let jc e g", y e g“"* for some a e II. Then [x, y] e Ka ^ and

([ji :,y ]| a'' ) = (y l[a'^,Jt]) = 2(y U )

and

( [ ^ . y ] ! « ' " ) = ( ^ l [ y , « ' ' ] ) = 2( j ; | y ) .

Let (• 1• )' be defined by (x \ y)' = (x | y) —(y 1x). Then (• I * )' is invari­


ant and vanishes on g“ x g““ for each a e II. By part (i) we see that both
g“ and g““ lie in the radical of (• | • )'. Since the radical is an ideal of g, it
follows that ( • ! • ) ' vanishes on the ideal generated by the g“ with a e ±11.
This ideal is precisely Dg (Proposition 4.1.12). □

An invariant bilinear form (• | • ) on a contragredient Lie algebra (g, is


called proper (relative to if (• 1• ) restricted to g“ x g““ is nondegener­
ate for each a e II.
A contragredient Lie algebra (g, is called invariant if there exists a
proper invariant bilinear form (• 1• ) on g relative to Notice by part (iii)
of Proposition 2 that such a form is necessarily symmetric on Dg. Normally it
is quite irrelevant whether such a form is symmetric on all of g. Invariant
Kac-Moody Lie algebras are often referred to in the literature as symmetriz-
able Lie algebras.

Proposition 3
Let (g, c^) be invariant contragredient with proper invariant form (• | • ) and
structure matrix A. Then relative to any display {e^, ay, j>

(«<• I = («; Ifj)Aji for all i, j e J.

Furthermore the quantities (e, | /¿) are nonzero.

Proof. We have

= (« M [« ;./;])

= ([«/>//] I[«;>/;]) = {ei \ f i ) Ai j .


By the definition of proper and CTl we see that (e, | /^) =5^ 0. □
360 Contragredient Lie Algebras

A matrix e is said to be symmetrizable if there exists a family


such that for all i and j in J,

Notice that if J is finite, then this can be interpreted as stating that


A • diag(£y) is a symmetric matrix. This concept generalizes the one defined
in Section 4.3.
Restated, the last proposition says that a contragredient Lie algebra is
invariant only if its structure matrix is symmetrizable. The structure matrix of
such an algebra may be symmetrized using e¿ = | for all i.
Remark 7 If ^ is symmetrizable then evidently = 0 if and only if
A ji = 0. Thus A is combinatorially symmetric (see Section 4.3). If A is
indecomposable, then the set {ej} is uniquely determined up to a scalar
multiple. Furthermore the {gy} may be taken in the smallest subfield of IK
containing all of the A¿j/Aj¿, Aj¿ ¥= 0. If the A¿j all lie inside a subfield of R
and for all i and j both A¿j and Aj¿ have the same sign, then all the e¿ may
be taken of the same sign, for instance, positive. As a matter of convention
we will always assume that the > 0 in these circumstances. In particular
this holds for Cartan matrices. If A is decomposable, then these remarks
apply to each of the individual blocks of A.

Let (g, be a contragredient algebra with structure matrix A. Our aim


now is to show that if A is symmetrizable, then g is invariant. Suppose then
that A is symmetrized by {fylyej-
Define a surjective linear mapping

( 8) (•) :Ö ,
by linear extension of
a¡ af ■■=e¡a'^.
By means of (• )^ we define a bilinear mapping

as follows:

(9) (a® I/i) := <a,/i>, /z e f).


Notice that
( a ? I tty) = <a¿,o:y> = < a,-,e y a />

= A¡jSj = A j ¡ £¡ = [aj I a ? )

and hence that ( • | • ) is symmetric on X Q^.


4.4 Inyariant Bilinear Forms 361

We may define (• | • ) on X by (9) and symmetry. It is irrelevant how


we define (• | • ) on the rest of X 1^. We take 1^' to be any supplement of
¡2k in and define (• | * ) on 1^' x to be symmetric but otherwise arbitrary.
Remark 2 We note that the nondegeneracy of < • , • > implies that for each
a« # 0 in Q^, (a° | i)) # (0).

Example 1

' 2 -1 0
A = -2 2 -2
0 -1 2

is symmetrized by

/1
diag{ 1,2,1} =

Since the row rank of A is 2, we extend A using Proposition 4.2.1 to

' 2 -1 0 o'
-2 2 -2 0
0 -1 2 1
0 0 1 2J

Write

ft = IKai'^e IKa^® Kd^

and

ft'^= K«! ® lKa2 ® IKtts ® Kd,

with the last matrix defining the pairing. In the notation above.

«? = = 2«2", «3 == «3"

• ) in the basis {aj, a ^,d

' 2 -2 0 O'
-2 4 -2 0
0 -2 2 1
0 0 1 *,

where * is the arbitrary choice for id '' |


362 Contragredient Lie Algebras

We now return to our contragredient algebra g. Define a Z-grading on g


by setting g" = ®ht(a)=/z9“* definition of contragredient, g is gener­
ated by g“ ^ 0 0 g^ = g(l). We introduce a symmetric bilinear form (• | •)
on g(l) by

(10a) ( • 1•) on defined as above.


(10b) (g“ l 9^) = 0 ifa,p^Q ,a + p^0.
(10c)
[x, y] = (x I y)a^ ( ^ i)) whenever a e g , jc e g“ , and y 0 g"“.

Of course at this point (10b) and (10c) only say anything if |ht(a)| = |ht(/3)| <
1. It is easy to see that (• | • ) is symmetric and locally invariant on g(l). For
local invariance there are only two distinct nontrivial cases that need to be
checked: Let a: e g“, y e g““ (a # 0), and let h

Iy) = ~{a,h){x Iy) = (x |[/î, y]),


([jc ,y ]l^ ) = (jcl y ) ( a ° |/i) = ( j:| y ) { a , h )
= (J il[y ,/i]).

Proposition 4
Let ( q, be contragredient, and suppose that its structure matrix A is
symmetrizable by Let (O ^ i G ik ^ be defined by (8) above. Then
there exists a proper symmetric invariant bilinear form (• I *) on g with the
following properties:

(i) (a? I a?) = A^jSj for all i, j ^ 3 and (a® \ h) = ( a , h ) for all a s
and h
GO [* I * ] : 9“ X g““ -> f o r a l l a ^ Q a n d [ jc, y] = ( x \ y ) a ^ f o r a ll
e g“, y e g"“.
(iii) (9“ I 9^) = (0) if a + p 0,
(iv) (• I • ) a-invariant.

Furthermore, if rad(* I • ) == {a: e g l(x 1 g) = (0)}, then rad(* | • ) w a graded


ideal of g, and

ra d (- 1•) n 0 g“ = rad(g),
rad(* I •) n f) c Z (g).

Thus rad(* | • ) c n(g) {see Section 2.7).


4.4 Invariant Bilinear Forms 363

In particular, if g is radical free, then the restriction o / (• 1 • ) io g“ X g ^ is


nondegenerate for all a ^ A \ {0}.

Proof The existence of (* I • ) on g satisfying part (i) follows by defining


( I ) on g(l) using (10) and then applying Proposition 1. Part (iii) is
immediate from Proposition 2. To establish part (ii), note that with a: e g“,
y e g"“, [x, y] e (Proposition 4.1.12). Thus for all h œ Î).

{[^^y] - { x \ y ) a ^ \ h ) = { x \ [ y , h ] ) - ( x l y ) ( a ^ |/ i )

= <a,/z>((A: I y) - ( x \ y ) ) = 0

Remark 2 then shows that [ a:, y] = ( jc | y)a°. Now part (iv) follows from parts
(ii), and (iii).
Set ê = {;c G g |(jc I g) = (0)}. The invariance of (• | • ) implies that ê is an
ideal of g. In particular ê is graded. Let 5“ e ê be homogeneous of degree
a, a e Q. If a ¥= 0, then [5“, g““] = (5“ | g“"")u:° = (0). Now, if the ideal
ad(U(g))i“ of g generated by s"^ meets \ {0}, then there is an element
u e U(g)"“ such that ad(w)5“ ^ ^ \ {0}. Then ad(M)5“ e Ô kXÎO}. Choosing
h si} with (ad(w)i“ | /z) ¥= 0 and using the invariance, we have (5“ | g““) ^
(0), which contradicts the above. Thus 5“ s rad(g). On the other hand, if
a = sO, then for all )3 e A, <)8, 5“) = (/3° | 5“) = 0, and hence [5“, g^] = (0).
Thus 5“ e Z(g). This shows that ê c rad(g) + Z(g). Finally, if x s rad(g) is
of degree a ¥= 0, then [;c, g”“] = (0), and therefore x s ë by part (ii). □
Remark 3 It is remarkable that g“ X g““ is one dimensional for nonzero
a when A is symmetrizable. This need not happen if A is combinatorially
symmetric but not symmetrizable [Sg].

Proposition 5
Suppose that (g, «^) is a minimally realized contragredient Lie algebra and
that its structure matrix is symmetrizable. Then any invariant bilinear form
(• I • ) on g defined by Proposition 4 is nondegenerate on f). In particular, if g
is radical free, then (• 1• ) w nondegenerate on g.

Proof Let h s i ) . Then, using Proposition 4, (i)l5) = (0) => /1 e rad((-l • )) =>
h s Z(g). Since the realization is minimal, h s (Proposition 4.3.5). Thus
we can write h in the form for some a in Finally, (0) = (/z 11^) =
(a° I ]^) = (a,i}) by (9), and hence a = 0, Thus h = 0,

If A is symmetrizable with
364 Contragredient Lie Algebras

then is symmetrizable with

= {A^)ns7^^

We may apply the construction of ( • I * ) on carried out above to the dual


realization

/? ^ = (r,n \n ),

which is a realization of Æ . Thus we begin with a linear map

defined by

(a,y)

Evidently this is just the inverse of the operator ( •)^ we defined above. We
denote it again by ( •)^, since no confusion usually results:
(.)0
( 11) Qk - ^ Q k -

We now define (• | • )* on There are several ways in which we can do this.


Following the same process as above, define (• | • )* on 1^* x by

( 12)

Extend (• I • )* to <2k ^ t)y symmetry and to all of 1^* X 1^* in some


arbitrary way as before.
However, if is finite dimensional and (• | •) is nondegenerate on 1^ (for
example, if (g, ,5^) is minimally realized) then we can use (• | • ) to construct
an explicit isomorphism a> between ^ and 1^*, namely

(ifcl-)

We can then use this isomorphism to transfer (• | •) to 5*, thereby obtaining


a nondegenerate bilinear form (• | • )* on This is precisely the way in
which the Killing form is moved over to in the finite-dimensional case.
Let us see that this process produces (•!•)* satisfying (12). Let a e Q^.
Then a° ^ and under the isomorphism a>, (a ° | •). From (9),
(a° I /i) = { a \ h } for all h showing that o)(a°) = a. Then co\q^ is the
inverse of (-)°, which we had agreed to also call (-)°. We now extend ( )®to
all of Í) by declaring (-)° = w. Thus

(1 3 ) ( k° , h} = ( k \ h ) for all A:,/i e ]^.


4.4 Invariant Bilinear Forms 365

If we also agree to denote o).-1^ by


Kxr (*)^, then for all a, ¡3 ^ if*.

(p\af =

= (^° 1a**) by definition of (• | ' )*


(14)
= <j800,«o> by (13)

= <i3,a°>.

From this (12) is obvious.


Whenever dim is finite and (• | •) is nondegenerate we will assume that
(• I • )* is defined in this second way. In particular (13) and (14) hold. In any
case we always have

(13') (k^,h) = ( k\ h) forall e fi,

(14') (jS la)* = </3,a0> for a ll/3 e fi*, a e 0 k -

Henceforth we write ( • I •) instead of ( • 1• )* [thus both ( )° and ( • 1•) now


have double meanings]. We summarize our conventions and notation in the
following:

General case dim 1^ finite, (* | -) nondegenerate


(•)“ w n (•)“
IBFl 0 K - 0 ^ , : « ? = s , V
IBF2 for all A: e 0;^, A e for all k , h
( k\h) = (k°,h) (k\h)=(k^,h)
IBF3 for all a e 0,^, /3 e for all a, p ^i )*
(/3| a) = (/3|a) = <^,a°>
IBF4 (a,-1a,) = 2e,- a,^= 2af/(a,-1 a,);
IBF5 ( a r |« r ) = 2 e r ' a, = 2 (« r)V (« r 1« r);
IBF6 Si > 0 for all i
if the matrix A permits this
IBF7 Si e Q for all i if the matrix
A permits this

A selection of the diagonalizing constants e, for a symmetrizable matrix A


together with a symmetric bilinear form (• | •) on g and another one [also
denoted by (• | •)] on for which the conditions IB Fl-6 hold is called (by
abuse of language) a standard invariant symmetric bilinear form on g.
366 Contragredient Lie Algebras

Proposition 6
Let (9, be an invariant contragredient Lie algebra, and let (• | • ) be a
standard invariant symmetric bilinear form on g. Then

(0 Gad acts as a group of isometries o f g relative to


(ii) Whenever aj rj acts as the reflection in a j (resp, af) on
{resp, 1^*) relative io (* 1• ), that is,

rj{h) = h -

2{a\ a¡)
forallh^íj,a^íl'^;

(hi) W acts as a group o f isometries on {resp. 1^*) relative io (• | • ).

Proof Gad is generated by elementary automorphisms, and these act as


isometries on g relative to (• | • ) by Proposition 1.7.1. For we have

Vjih) =h - { a j \ h ) a J = h - («9 I V IBF2)

2 ( a / 1
I / )
=h - (bylB FlandlB FS).

and, mutatis mutandis, ry(a) = a —2((a | aj)/(aj \ aj))aj. This proves part
(ii). Now part (iii) follows from parts (i) and (ii), together with the definition
of W (Section 4.1). □

From IBF7 we see that (ala) e Q for all a ^ Q.

Proposition 7
Let (g, «^) be an invariant Kac-Moody Lie algebra, and let (• | • ) be a
standard invariant symmetric bilinear form on g. Then

(i) (a I a) e {(«. | a,) | / e J} c Q_^ for all a


(ii) ( a |a ) < 0 / o r a / / a e ^ '”A,
(iii) for all a e A, a e'"*A <=> (a | a) < 0.

Proof Part (i) is clear from the PT-invariance of (* | • ) while part (iii) then
follows from part (ii). Let a e'^^A. We may assume, without loss of general­
ity, that a e^'^A^. Since Wa c^'^A^. (Proposition 4.1.5), there is no loss of
4.5 Casimir-Kac Operators 367

generality in further assuming that a is of minimal height. Then for all i e J,


ht(r^a) > h t(a) => ( a ^ a ^ ) < 0 => (a | a¿) < 0,
so clearly (a | a) < 0 since a e A+. □

Let i? = (]^,n, n ^ ) be a realization of the Cartan matrix A, An element


p e ]^* is called a minimal regular weight if for all i ^ J, ( p , a ^ ) = 1.
If is a basis of 1^, then of course p is unique. Minimal regular weights
appear many times in semisimple Lie theory and in many places later in this
book (e.g., in Section 4.5). The terminology comes about as follows: In the
most important representations of a Kac-Moody algebra the weights are
integral: elements A g 1^* satisfying ( A , a ^ ) g Z, for all i g J. Of these,
the dominant weights are those satisfying (A, a ^ ) g N for all i g J, and the
regular weights are those satisfying, for all / g J, (A, 0. Thus the
“minimal” (dominant integral) weights are those satisfying ( A , a ^ ) = 1 for
all i G J.
Suppose that A is symmetrized by ^ j and (• | • ) is a bilinear form on
1^* satisfying the conditions IBF. Suppose that p is a minimal regular weight.
Then
2a,- \ 2(p 1a,.)
(15) l = < p , a r > = P l
(a, I a,) / (a,. | a,) '
Thus
2(p I a,) = (a, I a,) for all i e J.

Remark 4 Let A be an indecomposable Cartan matrix of finite type, £nd


let g = g(y4, R) be the corresponding split simple Lie algebra over K. If IK is
the algebraic closure of K, then by Section 4.3
g(.i4,/?jj^) = (K®j^g(^, i?),
and hence g is central simple (Exercise 1.7). Hence any standard invariant
bilinear form on g is a multiple of the Killing form.

4.5 CASIMIR-KAC OPERATORS

I have heard statements that the role of academic research in innovation is


slight. It is about the most blatant piece of nonsense that it has been my fortune
to stumble upon.
Certainly, one may speculate idly whether transistors might have been discovered
by people who had not been trained in and had not contributed to wave
mechanics or the theory of electrons in solids. It so happened that the inventors
of transistors were versed in and contributed to the theory of solids.
368 Contragredient Lie Algebras

One might ask whether basic circuits in computers might have been found by
people who wanted to build computers. As it happens, they were discovered in
the thirties by physicists dealing with the counting of nuclear particles because
they were interested in nuclear physics. One might ask whether there would be
nuclear power because people wanted new power sources or whether the urge
to have new power would have led to the discovery of the atomic nucleus.
Perhaps—only it didn’t happen that way, and there were the Curies and
Rutherford and Fermi and a few others.
One might ask whether an electronics industry could exist without the previous
discovery of electronics by people like Thomason and H. A. Lorentz. Again it
didn’t happen that way. ...
Or whether, in an urge to provide better communication, one might have found
electromagnetic waves. They weren’t found that way. They were found by Hertz
who emphasized the beauty of physics and who based his work on the theoretical
considerations of Maxwell. I think that there is hardly any example of twentieth
century innovation which is not indebted in this way to basic scientific thought.
—Henrick Casimir

Throughout this section (g, will be assumed to be invariant contragredi­


ent and radical free. We endow g with a standard invariant symmetric
bilinear form

(• I •): 9 X g

We also recall that

(■ I ■)l8“xg-<'

is nondegenerate for all a e Q \ {0}, since g is radical free. As in Section 4.4


we also denote the corresponding bilinear form on 1^* X f)* by

The category <^(g, 5^) of Section 2.5 will be denoted simply by 0 ,


We plan to construct what, in the language of categories, would be called
a natural transformation from the identity functor to itself. In plain language
this means that

NTl: for each module M < 0 , we have a g-module homomorphism (or


operator)
4.5 Casimir-Kac Operators 369

NT2: for all M and N in ^ and for every g-module homomorphism


f : M - ^ N , the following diagram commutes

M -^M
fi i f
N -^N

This leads to a family {F^}m of operators that plays a crucial role in most
of the subsequent theory. The construction of these operators depends on
the existence of the symmetric invariant form (* I • ), and this accounts for the
considerably less developed theory of Lie algebras q(A, R) when A is not
symmetrizable.
Recall that dim(g“) < oo for all a ^ Q \ {0} (Proposition 4.1.1). The
nondegeneracy of the pairing g“ X g““ implies that if = [x^,. . . , is a
basis of g“, then there exists a unique basis = [y^,. . . , y„} of g~“ such
that I yj) = 8^j, The bases B^ and B_^ are then said to be dual to each
other.

Lemma 1
Let a Ei Q \ {0}, and let {x^,. . . , and {y^, ...,y^} be bases o f g“ and g"“,
respectively, so that these bases are dual to each other. Define y “ — ®
G g 0 g and T“ = g U(g). Then

(i) y“ and T“ do not depend on the choice o f dual bases o f g“ and g""^.
(ii) I f u E and V E g““, we have
n
M= L
1=1
n
v = E IX i ) y t ,

(«k) = E
1=1

(iii) Let a, a' e <2 \ {0}, and let r “ and y “' be written in the forms
T,%\yj ® Xj and ® x'k, respectively, with respect to some pairs
of dual bases. Then for all z e 9“ “ Itave

E [yit, z] ® 4 = E yy ® [ ^ . 4 ] 9®9
^=1 7=1
370 Contragredient Lie Algebras

and

'L{y'k^AA= in U (g ).
k=\ ;= 1

Proof. Part (ii) is an immediate consequence of the fact that the bases
{jCj,. . . , jc„} and {yj,...,y„} are dual to each other. If and
are any other two such bases then, by means of part (ii), we
obtain

HY¡®X¡= E E y¡{Y¡ Ix¡) ® E (yj


i= \ ¿= 1 \ ; = 1 k= \

= E E(5^-
j , k = \ \i = l

= E hkyj ®x„ = ' Lyj ^Xj ,


j,k = l j

thereby showing that y “ is independent of the choice of dual bases. Next we


use the linear map g <8> g U(g) given by x <8) y to finish the proof of
part (i).
For part (iii) we define a symmetric bilinear form (*, • ) on g ® g by

(x (8>y, x' <S>y') = (x Ix ') ( y I y ').

Because of analogous properties of (• I • ) it is easy to see that for all


a, a', j8, j8 'e (2 \ {0},(*, • ) determines a nondegenerate pairing between
g"" <8) g^ and g“' <S> g^' if a + a' = 0 = + j8' and (g“ (S> g^, g“' 0 g^') = (0)
otherwise.
It follows that in order to establish the first assertion of part (iii), it will
suffice to show that

E [y*> z] E y¡ ® [ z, Xj ] , a ® 6| = 0
k=l j~l
for all a e g" and e g
4.5 Casimir-Kac Operators 371

By definition, the left-hand side of the above equals

k=l J=l

m
= E
k = l

n
+ (yy I I[z, b]) [because of the invariance of ( • | •)]
y= i
= { [ z \ a ] \ b ) 4- ( a \ [ z , b ] ) = 0.

For the second assertion of part (iii) we again apply the natural map
g ® g U(g) with x y xy to the assertion we have just established.

Let p e be a minimal integral regular weight fixed once and for all (see
Section 4.4).
Let M e We define the operator discussed in the introduction by

r — + V'

where and are two operators defined as follows:

Definition of Let M = be the weight space decomposition


of M. Then is defined so that F^ acts on each weight space as
multiplication by (p, + p I p + p) —(p I p) = (p + 2p | p).

Definition of F^ View M as a left U(g)-module. Then for each a e 0 \ {0}


the element F“ e U(g) acts on M. We denote this operator by F^ and
define

Although the sum in the definition may be infinite, it makes perfect sense
as an operator. Indeed, if A e P(M ), then the set

{a G (2+ I A + a e P ( M ) }

is finite (Lemma 2.6.1). Hence there exists only a finite number of a e A+


such that
r “(Af^) # (0).
372 Contragredíent Líe Algebras

Thus is well defined on each weight space of M. Notice, however, that


need not be determined as an operator by some element of U(g) if A+ is
infinite.
Finally, we observe that both and are graded operators; that is,
they apply into for every weight space of M. Thus + F^
also has this property. The operator is called the Casimir-Kac operator of
the g-module M.
Of course, depends on the initial choice of p but for our purposes a
single choice of p suffices, so we do not record it in the notation.

Proposition 2 [Ka2]
Let M e Then F^ commutes with the action of g on M.

Proof Since acts on each weight space of M as scalars and


graded, it is clear that commutes with the action of on M, Thus it
suffices to show that the action of F^ + Fj[^ commutes with that of each
and /y for all j e J. For simplicity we omit the subscript M in the rest of the
proof.
Let V e M^. We have

(1) ( r % - e^r°) ■V = { { f i + aj + 2p\n + aj) - (p. + 2p\ p)}ej ■ v

= {{p + 2p I aj) + [aj I p + aj)]ej ■v

= {2{p I a,.) + {aj I aj) + 2{p \ aj)]ej ■v

= 2[p + aj I OLj)ej • V,

where we have used (4.4.15): 2(p 1 ay) = (ay | ay).


Similarly

(2) (r° // - /yr°) ■V ^ [{p - aj + 2p \ p - aj) - ( p + 2p \ p ) ] f j ■v

= - 2 { p \ a j ) f j ■V .

Next we compute the commutator of F' and Cy, in two parts. We have

F“^ = efej

where ef e g is such that {ef \ Cj) = 1. Since af = Cy«/ = [-Sjfj, Cj] =


i - £ j f j I CyX-tty) (see IBF and Proposition 4.4.4), we conclude that ef = ejfj.
4.5 Casimir-Kac Operators 373

Then

(3) 2 ( r “>e^. - c / “/) • V = 2ej{fj • ej ■ej ■v - Cj • fj ■ej ■y)

= - I s j a y - €j ■V

= -2ej(iJ. + a j , a j ) c j ■v

= - 2 [ i i + Uj I aj)e^ • V.

Following along similar lines, we show that

(4) 2(F“^X - fjT-j) ■v = 2{^\ aj)f, ■v.

Comparing (1) and (3) and (2) and (4), we see that to establish the proposi­
tion we need only show that

E r-
ae A+\{«;}

commutes with the action of ej and fj for all j e J. It is useful to note that if
a E A+\{ay}, then either

a — aj ^ A or a — e A+\{ay}.

since 2«y ^ A.
Let {x^ f^} and be bases of g“ and g““ that are dual to each other.
Here k e K ia), where K(a) is some finite set. It is convenient to define
K(a) = 0 if a E A^_. We have

(5) ( E T“e , - e , E
' aeA q: e A

= E E (y-a,k'Xa,k- ej-V V)
a sA AK > k s K (c c )

= E E +
aeA+\{ay} k^K{a)

where we recall that only a finite number of nonzero terms occur in this sum.
By part (iii) of Lemma 1 with a' = a + aj and z = we see that

E = E y-c.kUj^X^j,].
k '^ K ia ') k ^ K ( a )
374 Contragredient Lie Algebras

Thus for any a ^ the (finite number of) terms involving roots in a + laj
cancel out in pairs. After cancellation, any remaining terms involve roots a
for which a + ay ^ or a —ay ^ A I n either case the offending terms
are clearly zero.
This shows that the left hand side of (5) equals 0 as desired. A similar trick
can be used to show that

L T“
A+\{ay}

commutes with the action of /y. This finishes the proof of the proposition.

Coming back to the two properties that we wished T to have, we see that
NTl is now proved. As for NT2, suppose that f : M ^ N is g-module
homomorphism. Then / : for all weights ii, and for x e it is
elementary to check that /(T ^(x)) = T^(/(x)).

Proposition 3
Let M e T) be a highest weight module with highest weight pair
and let be its Casimir-Kac operator.

(0 acts on M as the scalar (A + 2p 1A).


(ii) I f N is a submodule of M and p is a primitive weight relative to N,
then
(A + 2p I A) = (p. + 2p 1p ).

(iii) I f p. and [ M : L(p)] > 0, then (A + 2p | A) = (p + 2p | p).

Proof Let u e U(g). By our previous proposition we have

T^u • V = uTm • V = u ■{ r ^ - v ) = ( A + 2p\ A)u • v.

Since U(g) ■v = M, part (i) follows.


For part (ii), choose w e with w ^ N but w (z N. Then w == w +
N is 3. highest weight vector of weight p for M/ N. It generates a highest
weight submodule U(g) • w of M/ N, and hence by part (i),

ru(8)i5>(^) = = ( p + 2 p lp )w .

On the other hand, if ir: M -» M / N is the canonical map then


= ( A + 2p I A)ir(w)
by part (i) and NT2.
4.6 The Radical Theorem 375

Finally, if fji and [ M : L ( ) ] > 0, then given any local composition


ijl

series {MJ for M at L(/i) occurs as some factor Any weight


vector in Mj that is the preimage of a highest weight vector of L(/a) is a
primitive vector of weight /x (Lemma 2.6.3(iii)). Then (A + 2p | A) = (/x +
2p I /x) proving part (iii). □
Remark 1 For finite-dimensional semisimple Lie algebras, Casimir opera­
tors refer generally to elements in the centre of U(g). On representations of
g they thus produce operators commuting with the action of g. For a
description of the center of U(g) one can consult [Bo4]. The most common of
the Casimir operators is the quadratic Casimir operator F == LfT[^xfXi,
where { x j , { a : * } are dual bases of g relative to the Killing form. For
the infinite-dimensional case Kac reorganized the sum using basis elements
from root spaces and “normal ordering” the products so that the vectors
from positive root spaces precede (in order of application on modules)
those from negative root spaces. This produces an expression which, as we
have seen, makes sense as an operator on modules from category

4.6 THE RADICAL THEOREM

In this section, unless mentioned to the contrary, we let A = (^¿y)/ye j ^


symmetrizable matrix and R a realization of A. We denote by u the
universal algebra of (A , R) and set g = u/rad(u). All the notation of
Sections 4.2 and 4.3 is maintained throughout. We identify h with a subalge­
bra of both u and g and endow both of these algebras with invariant bilinear
forms ( *I • ) u and ( • | * )g in such a way that (• ), = (• • )u when restricted
to ^ X t
We denote this common restriction by (• | • ) and use the same notation for
the dual bilinear form

Our intention is to establish an important result due to Gaber and Kac [GK]
that states that rad(u) is generated as an ideal by some of its homogeneous
components rad(w)“, where the a ’s appearing as generators are character­
ized in terms of the bilinear form (• I • ). In the case where ^ is a Cartan
matrix, this leads to a characterization of rad(u) as the ideal generated by the
that were defined in Section 4.1. This is called Serre’s theorem when A is
of finite type.

Lemma 1
Let abe a Lie algebra, and let Vi{a) be its universal enveloping algebra. Define
U'^(a) to be the two-sided ideal o f U(a) generated by a viewed as a subspace
376 Contragredíent Lie Algebras

of U(a) [jo U(a) = K1 © U'*‘(o)]. Then


(i) U+(a) = U(a)a = aU(a);
(ii) a n (U^(a)U+(a)) = [a, a];
(iii) § n SU'^ía) = [§, §] if § is a subalgebra o f a.

Proof
(i) Both U(a)a and aU (a) are two-sided ideals of U(a) containing a
(because 1 and a generate U(a) as an algebra). This clearly implies part i.
(ii) Let IT; U(a) -» U (a/[a, a]) be the homomorphism induced from the
canonical map a a/[a, a]. Since a/[a, a] is abelian the enveloping alge­
bra U (a/[a, a]) is isomorphic to a symmetric algebra S = (Section
1.8). It is clear that ir(a) c and that -n-(U‘^(a)U‘''(a)) c
rr(a n U‘^(a)U^(a)) = 0, and therefore a n U'^(a)U'^(a) belongs to the
kernel of the restriction of ir to a; this is [a, a]. The reverse inclusion,
namely [a, a] c a n U(a)'^U(a)'^, is clear.
(iii) By Corollary 2 of PBW (Section 2.8), we know that U(a) is a free left
U(§)-module admitting a basis of the form {1} U { x ¡ ^ , Xj ^ ¡ ^ < ... A: > 0,
where (xj + §};ej is some totally ordered basis of the quotient space a / l
with Xj e a. Thus (U(§)A:y|... Xj^) n § = (0), and hence
§ n § U (a )'" c § n § U (§ )‘^.
We can now use part (ii) with a replaced by i. □

Via the canonical mapping - : u ^ g any g-module may be viewed also


as a u-module. In addition, the corresponding associative algebra homomor­
phism - : l l ( u ) ^ U ( g ) allows us to view U(g) as a right U(u)-module,
namely q • u ■= gU for all g e U(g), u e U(u).
Proposition 2
Identify if with a subalgebra o f both u and g, let A e ^*, and let M(A) and
M ix.) be the Verma modules of highest weight X o f n and g respectively. Then

(i) U(g) ®u(u)M(A) is a highest weight q-module of highest weight A.


(ii) The natural surjection f : M(X) -* U(g) ®u(u)M(A) is an isomorphism
i f is the surjection given by the universal nature of MiX); see Section
2.2).

Proof
(0 Clear.
(ii) We look for an inverse for /. Think of M i X ) as a u-module via . That
is,
X • V = X ■V for all X © u , y ^ Mi X).
We then obtain a u-module surjection
- , onto ^ .
M i x ) ^ M(A)
4.6 377
The Radical Theorem

given by the universal nature of M(A). This extends to a linear mapping

which is easily seen to be a U(g)-module homomorphism. Consider the


multiplication map
U(g)<8> M (A )^M (A )
U (u)

given by
X ®V ^ X •V for all X e U( g ), í; e K.

By composing this map with the above, we obtain a U(g) homomorphism


(and hence g-module homomorphism)

Ar:U(g)® M(A) ->M(A).


U(u)

It is immediate that / and x inverses of each other. □

Let us fix a minimal regular weight p.

Theorem 3 [GK]
Let A he a symmetrizable matrix, R a realization of A, and u = \x{A, R) the
universal algebra of iA , R), Let rad(u) = r_ 0 r+ be the radical o f u. Then
r+ {resp. r_) is generated as an ideal by the elements of {resp. r ““), where

(0 a e (2+,
(ii) (2p \a) = (a \ a).

Proof Let Miff) be the Verma module for u with highest weight 0, and let
¿5+ be a highest weight vector of M(0). Consider the unique maximal proper
submodule iV(0) of M(0). Since

ejfj • u += a f ' v^= 0 for all j G J,

we see that each /^ • ¿5^ g MO) and that these elements generate MO) as a
module.
Since highest weight submodules of Verma modules are themselves Verma
modules we conclude that the submodule of M(0) generated by f j ' v+ is the
Verma module M i —af), Thus

M 0 )= L m - a j ) .
;s j
378 Contragredient Líe Algebras

Recall that u_ is freely generated by the f / s (part (i) of Proposition 4.2.4),


and hence that U(u_) is the free associative algebra on the f / s (Proposition
1. 10.2).
It then follows that for all choices of U j g U(u _), j g J, the family
{mJ JJ'
f }-ci
J J
c U(u_) is linearly independent
^
and hence [from the construction
of M(0)] that the sum of the M i —af) is direct,

( 1) iV(0) = 0 M{-aj).
yej

By Proposition 2 we have a g-module isomorphism

U(g)® M(-aj) = M{-aj)


U(u)

which together with (1) gives

( 2) U(g)0 iV(0) s 0 M ( - a ) ,
U(u) ;e j

where each M i.-a/) is the Verma module of g with highest weight -Uj.
Consider the map (p: r_-> U(g) ® given by

for all X G r_.

We claim that ^ is a u-module map. (Recall that r_ is an ideal of u and


hence an u-module under the adjoint action. Similarly g is a u-module by
means of the canonical map “ : u g.) Indeed, if m g u and x g r_, we
have
<p{[u,x]) = 1 ® [ m, x ] • 5+
= 1 ^ U ' X ' Vj^— 1 ^ X ' U ' V^

= U ® X ' V^—X ^ U '


= u ® X ' V^ (since 3c = 0)
= u(p(x).

This calculation also shows that cp annihilates [r_, r_]. Taking into consider­
ation (2) we obtain an induced linear map

0 Af(-a^).
/sj

Moreover r_/[r_, r_] has a natural g-module structure, and <p above is then
4.6 The Radical Theorem 379

a g-module homomorphism (this is easy to verify). We claim that <p is


injective. Let x e r_. We use the embedding r _ ^ U(u_) to write

' L uj f j
;e j

with Uj e U(u_)'^ (for recall that r_ intersects 0^-^j(K/y trivially, see


Proposition 4.1.10). Then

<p{x) = 1 ® T,UjfjV+
= ®fjV +

= L “;( l +
Under the isomorphism (2) we may identify v{.-= 1 ® fjV+ with a highest
weight generator of M( - a j ) . Thus from

(p{x) = J^Uj- 0 M{-aj)

we see that <p( a:) = 0 <=> Hy = 0 for all j ^ J. This last implies that x ^ v_n
r_U(u_)'^= [r_,r_] by Lemma 1. This shows that <p is injective as con­
tended.
The g-module r_/[r_, r_] belongs to the category ^ since we can identify
it, via with a submodule of ^ ^ show that
^ & directly).
Let V be any primitive vector of r_], say, of weight - a , a e Q^.
Then there is a submodule N of U(g) *^ such that U(g) *v / N = L{ —a).
Using the Corollary of Proposition 2.6.13 we find that

0 < [ r _ / [ r _ , r _ ] : L ( - a ) ] < L [ M ( - a j ) : L( - a ) ] ,

and hence [M (-o:,): L ( - a ) ] > 0 for some i e J. Then L ( - a ) is a proper


factor in M i —ai) and by Proposition 4.5.3, ( —a + 2p I —a) = ( —a , +
2p I —a,). However, (2p | a,) = (a, | a,) by (4.4.15) and thus 2(p | a) = (a | a).
Let
i? = (a e e + I(2p I «) = ( a I a)}.

Since r_/[r_, r_] is generated as a Q_-module by its primitive vectors


(Proposition 2.6.5), we obtain
t _ / [ x _ , x _ ] = £ U ( g _ ) r l “ mod[r_,r_]
a^R
= U ( u _ ) n “ mod[r_, r_],
a^R
380 Contragredient Lie Algebras

and hence r_/[r_, r_] is generated as a u-module by the elements of


rZ“ + [r_, r_], a ^ R.
Finally, let §_ be the ideal of u generated by {rZ“ | a e i?}. We want to
show that §_= r_. By the above §_+[r_, r_] = r_. Thus

[§_ + [ r _ , r _ ] , r _ ] = [ r _ , r _ ] ,

and hence

§ _ + [ [ r _ , r _ ] , r _ ] = r_.

By induction it follows that

(3) §_+ r " = r_,

where r ” = [ r " “ \ r _ ] . Now the weight spaces of r" lie in


-(¡2 ++ *** +j2+X«-times), and hence

n 1:-= (0)
n= 0

(also see Proposition 2.7.5). Then (3) implies that §_= r_, as prescribed by
the theorem.
To establish the result for we use the anti-involution a. □
Corollary
Suppose that the entries o f the matrix A are real numbers that are either all
positive or all negative. Then \x{A, R) is contragredient and radical free. In
particular %(^A,R) = vl{ A, R) and and g_ are free.

Proof. Since A¿^ # 0 for all i, the algebra u is contragredient (Proposition


4.2.7). Let S = (5,y) be the structure matrix of u. Then (ibid. Remark 5)

2A,j
c = ___jL
^ij A *

In particular 5,,. > 0 for all i, ; e J and 5 is synunetrizable with positive


£,’s. (If {epyej symmetrizes A, then symmetrizes 5.) We conclude
that the universal covering algebra u of g is synunetrizable. We can there­
fore apply Theorem 3 to u. If a e A^. and we write a = Ec,a„ we have (see
Section 4.4)

2 ( p \ a ) = 2XXp,c,e,a,^> = 2X)c,e,-
4.6 The Radical Theorem 381

and
( a | a ) = 2 £ c f 6 ,. + Y,CiCjSjSn
/ i

(all of these sums have finite support). Thus

2(p 1a) - (a I a) = 2£(c,. - cf)si - Y^CiCjBjSij < 0.


i i*i
The theorem then shows that rad(u)+= (0) and hence that u is radical
free. Finally, since u is the universal covering of 9, we conclude by Proposi­
tion 4.2.8 that 9 itself is radical free; in fact u = 9. □

Theorem 4
Let (9, be contragredient and suppose that its structure matrix is a
symmetrizable Cartan matrix^ A = j. Let R be the natural realization
of A in the context o f (9, and let u(A, R) be the corresponding universal
algebra. Let J(R) be the ideal o f Proposition 4.2.10. Then

(i) J(R) is the radical o f m{A, R)\


(ii) rad(9) is generated as an ideal by the elements

dtj =

dj: = (ad fj) Vi for all

In particular 9^(^, R) = %{A, R).


(iii) If 9 is integrable (i.e., if 9 is a Kac-Moody Lie algebra), then 9 is
radical free.

Proof. If (i) is true then (ii) follows at once by Proposition 4.2.8 and (iii)
follows from the fact that 9/rad 9 is radical free (see Exercise 2.18).
We have to show that j {r ) = rad(u). Using Proposition 4.1.11(0 we see
that J{R) c rad(u). To show the reverse inclusion, we let ¡jl'.vl 9^ •=
\x/J{R) be the natural mapping. Then it will suffice to show that 9^ is radical
free, for then, using Proposition 4.2.8 again, rad(u) c ker()Lt) = J{R).
According to Theorem 3, rad(u) is generated by certain elements x^
where
(4) “ s Q+ and2(p|a) = ( a |a ) .

Let be such a generator and suppose that ¥= 0. Then ^


rad(g^)“. By Proposition 4.1.10,

^^A(g^) =-A(gVrad(g^))
382 Contragredient Lie Algebras

and since real root spaces are always one dimensional (Proposition 4.1.5) we
see that rad(g^)“ = 0 if a e
Thus we may confine our attention to those a for which e (g^)“
and a e ^'"ACg’O. Assuming that our bilinear form on g^ is chosen so as to
conform to (IBF) of Section 4.4, we conclude from Proposition 4.4.8 that
(a |a ) < 0, and thus by (4) that (p|a) < 0. However the definition of p
(see(4.4.15)) together with a ^ gives (p|a) > 0. This contradiction shows
that ¡1 kills all the generators of the radical of u and hence the entire
radical. □
Remark 1 Is the assumption that A is symmetrizable necessary to obtain
the statements of Theorem 4? At the time of writing this remains an
outstanding open problem in this subject.

4J HERM ITIAN CONTRAGREDIENT FORMS

Let (g, be an invariant contragredient Lie algebra carrying a standard


nondegenerate symmetric invariant bilinear form (• 1*). Using the anti-in­
volution (T we can construct a new form
( * I*) : 9 X g IK
by
<x I y> = ( a: lo-y).
We observe the following properties of < • | • >, which are trivial conse­
quences of the definition and the results in Proposition 4.4.4.

Proposition 1

(i) ( • I • > is symmetric and contragredient (see Section 2.8).


(ii) g“ _L g^ with respect to ( • | • > unless a = jS.
(iii) The pairing < • I • > lg“xg« w nondegenerate for all a.

The object of this section is first to prove that if K = IR, g is an invariant


Kac-Moody algebra, and (• | • ) is if scaled correctly, then < • | • > is positive
definite on g“ X g“ for all a e A \ {0}. We can use this to define a hermitian
contragredient form that is positive definite on root spaces when IK = C.
Later we will use this form to establish a positive definite hermitian contra­
gredient form on any irreducible highest weight module of g over C.
For the time being we assume only the conditions on g announced above.
We define an operator similar to the Casimir-Kac operator of Section 4.5:
r g_,
r(jc) == E E jc]] for all a: g g"“,
0<fi<a i

where for each a e A+, {e^-^ and are dual bases relative to (• | •).
Let p e be a minimal regular weight.
4.7 Hermitían Contragredíent Forms 383

Proposition 2
For a e A^. and x e g““, r(x ) = 2((p | a) —(a | a))x.

Proof. Let M ■■=M(0) be the Verma module of highest weight 0 for g, and
let denote a highest weight vector of M(0). It suffices to show that for
r e g-“, a e A+,

T ( a:) • v+= (2(p I a ) - ( a | a ) ) x •

From the Casimir-Kac operator K we have

(1) 0 = X• • y+) = r ^ ( x ■v ^ ) + T¡^(x ■y+)


= { - a + 2 p \ - a ) x ■V++ 2 Y, ^ ' ^ +-
peA+ i

Using the fact that for /3 e A+,

e f - x - v + = [ e ^ ¿ \ x ] - v + + x - e f - v ^ = i ^ ^ ^ ’ ^^ if^< a,
otherwise.

From (1),

( 2 ( p \ a ) - ( a \ a ) ) x • v += 2 Y ‘ ^+
0<P<a i

- E EFi>».FS>.*ll
0</3<a i

(2 )
+ E '^ 0 < / 3 < a
Z W - A ‘%
i

- E 0<fi<a i '

Observe that the set {)8 e A+ | j8 < a} has the symmetry p a - p. Thus
we may rewrite the terms inside the braces as

E
0<p<a
Z [ 4 ‘- A ‘ % -
i
E E«'-(.-»,[«.e,l,
0<p<a i

which then cancel completely by part (iii) of Lemma 4.5.1. Finally using (2),

(2( p | a ) - ( a | a ) ) j r • i; + = Y L [4> ^]] ‘ ^ +-


0<P<a i

which proves the proposition. □


384 Contragredient Líe Algebras

Lemma 3
Let g be an invariant Kac-Moody algebra. Assume that its Cartan matrix is
symmetrized by ^ Q+- Assume (• I *) is defined
according to the conditions IBF o f Section 4.4. Then
2( p l a ) > ( a l a ) for all a

Proof We use induction ht(a). For a = a, of height 1, <p, = 1 =>


2(p I = (a^ I «¿) > 0. Suppose that ht a = A: > 1. If a then (a | a)
< 0 by Proposition 4.4.8, whereas 2(p | a,) > 0 for all i => 2(p | a) > 0. If
a then there is a ; e J so that rja = a — <a, a/>ay has height
k' < k. Hence

2 (p | a ) > 2 (p | a - <a,a/>ay)
= 2(p 1rjo) > [r^a I rjo) = ( a I a ) . □

Proposition 4 [KP2]
Let g be an invariant Kac-Moody algebra over U and assume the hypotheses of
Lemma 3. Then < • I * ) w positive definite on all nonzero root spaces of g.

Proof Because of a it suffices to show this for root spaces g““, a e A+. For
-a ^ we have

(f i I//> = (e,-1fi) = ef' > 0


(see Proposition 4.4.4 and IBF4 in Section 4.4).
Suppose that ht a = A: > 1 and the result is proved for g“^ with 0 < ht )3
< ht a. For each jS e A+ with ht )3 < ht a, let {e^!}^} be a self-dual basis
(relative to < • | • >) for g"^. Then for

[e% \ e^p) = {e% \ e^l\) = 5,^,

which shows that and {ep) are dual with respect to (• 1• ).


For any X e g~“ \ {0),

(2(p|o;) - (o:|a))<A:U> = D L ( [ « - W [^i8^ •']] I


0</3<a i

= E E([«p^^]
0<p<a i

0</3<a i
by the induction hypothesis. Thus <jc | x> > 0. □
Exercises 385

Remark 1 Proposition 4 remains true if the given field 1 : is replaced by any


ordered field. □

Let be a Kac-Moody algebra over IR defined by A , and let < • 1• >


be the corresponding symmetric contragredient form on qJ . A \ which is
positive definite on nonzero root spaces.
Let g(y4) := C 0 Then q( A ) is a complex Kac-Moody Lie algebra
and we can extend < • 1• > to be a hermitian form on g (^ ) uniquely
(depending on which variable is to be antilinear), and it is positive definite on
nonzero root spaces. Summarizing we have

Proposition 5
Any invariant Kac-Moody Lie algebra g over C carries a hermitian contragre­
dient form < • I • ) that is positive definite on nonzero root spaces. The form
{• \ - ) is unique up to positive scalar factors on each o f the indecomposable
factors o/ Dg (see Proposition 4.3.8).
Remark 2 Propositions 4 and 5 can be extended to include other radical-
free contragredient Lie algebras carrying nondegenerate symmetric invariant
bilinear forms. The only requirement is the condition

(PC) 2(p I a ) > ( a 1a:) for all a G A+ and > 0 for all i e J.

EXERCISES

4.1 Let ^ be an / X / Cartan matrix of finite type.


(a) Show that A has an /-dimensional realization R = (i\,U ,U ^).
Let g = q(A, R). This Lie algebra is up to isomorphism indepen­
dent of the choice of R in (a).
(b) Show that g is a semisimple Lie algebra and that g is simple if
and only if A is indecomposable. (In this case g is called the
simple Lie algebra of type A over K.)
(c) Let (• I • ) be standard symmetric invariant bilinear forms on g
and i)*. Show that (• I * ): X is positive definite when
restricted to Q q X (rational span of the roots).
(d) Show that A U {0} (Proposition 4.4.8) and that '^^A c 1^* is a
finite root system with Cartan matrix A and base II.
(e) Show that dim^^ g = / + T^A| < oo.
(f) Compute dim g for all indecomposable Cartan matrix of finite
type.
Remark For algebraically closed fields IK the only finite-dimensional sim­
ple Lie algebra are the Lie algebras g(-^4) when A is of type Ai, Bi, Ci, D^,
^4’ ^rid G2.
386 Contragredient Líe Algebras

4.2 For each of the finite type Cartan matrices A of rank 2 construct a
two-dimensional realization R of A and explicitly compute a basis for
g(A,R).
4.3 Let A be a Cartan matrix of type A¿, and let R be an /-dimensional
realization of A. Prove that g = (A, R) - §!(/ H- 1, K).
4.4 Let K be a 2/ + 1-dimensional IK-space, / > L Fix a basis
{ljq, Ü2i} of V. Let be the (nondegenerate sym­
metric) bilinear form on V whose matrix with respect to the above
basis is
' -2 10 !0 )
1 1
0 11 111 I
0 1I 1 1
21
where / is the / X / -identity matrix. Let V' •= © and let be
¿=1
the restriction of j8 to V'.
(a) Show that g()8) (see Exercise 1.1) consists of all matrix of the
form
01^ 1 b \
1 1
2b^ 1 11 B
1
1c 1 -A^j
where A is an arbitrary / X / matrix, B and C are arbitrary skew
symmetric / x / matrices, and a and b are arbitrary 1 X / matri­
ces.
Conclude that g(j8) has dimension 21(1 + 1) and that g()3') is a
subalgebra of g()8) of dimension 2/(/ - 1).
(b) Let
a E , - E,/+/,/+/> 1 < / < /.

Show that ]^ := 0 is an abelian subalgebra of g()8').


¿=1
(c) Let be the basis of b* dual to , c b- For 1 :< i <
j < I define
“ E qj +¿
^ ~'^E i +¿ q — E qj
^ ^ij ~ E^i+jj+i
“ ~^j,i ~ ^ l +i,l+j
= ^¿J+j ~ ^ j j +i
Exercises 387

Let
A == [±2s¡, ± £, ± Sj\\. < i < / < /} and
A' != {±£,- + £y|l < i < / < / } .
Show that
9(jß) = e 0 K X^ and
a e A

9(j8') = ^ ® 0
aeA'
(d) Show that [h, X^] = a(h)X^ for all h and a e A. Conclude
that g()8) is a Lie algebra of type Bi and that g()3') is of type D^.
4.5 Let F be a 2/-dimensional IK-space; / > 1. Consider the matrix
0 I,
/ =
-I, 0
where /, is the identity / x / matrix. Let ¡3 be the bilinear form
on V defined by /3:
P{v, w) = v^Jw.
(a) Show that /3 is alternating and nondegenerate.
(b) Let g(/3) be the Lie algebra defined by /3 (see Exercise 1.1). Show
that g(j8) consists of all x = - a^ ' ^ square
/ X / matrices such that B = and C = C^.
Conclude that dim|^ g(/3) = /(2/ + 1).
(c) Let 1 :< I, j < 21, be the matrix with 1 in the i, j position and
0 elsewhere. Let
ay = Ea - E,+ij^i, 1 < / ^ /.
Set = E|.=iKa,^= ©/^jlKa/. Show that is an abelian subal­
gebra of g(/3).
(d) Let {£,}j s / sz c: ]^* be the basis dual to {an<=i,z ^ For 1 < / <
j < I define
^2ci ~ ^i,l+i’
^ - 2 ., = 7+/,/’
AT. = F. ,• - E,
^ — e¡ + ej ^ l + i, l + j ^

+ ~ ^ij+j ^j,l +i’

Show that each of these 2/^-matrices belongs to g(jß).


388 Contragredient Lie Algebras

(e) Let A = {2s,, ± e, ± 11 < / < ; < /} c 1^*. Show that g(j8) = ij
0 [h, x j \ = a(h)X^ for all a ^ A and
h
(f) Let A be the Cartan matrix of type Q , and let q(A, R) be the Lie
algebra defined by some /-dimensional realization of A. Prove
that q(A, R) = g()8).
4.6 Let ^ be a Cartan matrix and R a minimal realization of A. Show that
q( A , R ) is finite dimensional if and only if A is of finite type. In
particular all finite-dimensional Kac-Moody algebras are invariant.
4.7 Let (g, be a finite dimensional Kac-Moody Lie algebra, and let
(• I • ) be a minimally realized positive definite invariant bilinear form
on g (see Exercises 4.6 and 4.1). Let M e ^ (g , <^) be a highest
weight module
(a) If jjL is such that [ M : L(/i)] > 0 show that
(filfi) - 2 ( A + p Im ) = 0 .

(Use Proposition 4.5.3).


(b) Conclude that M has a finite composition series.
4.8 Let g be a minimally realized finite-dimensional Kac-Moody Lie
algebra.
(a) Show that p = is the unique minimal regular weight.
(b) Show that g is a semisimple Lie algebra and that g is simple if
and only if A is indecomposable.
4.9 Let ^ be a symmetrizable Cartan Matrix, i? = ( l ^ , n , n ^ ) a realiza­
tion, and '^^A the corresponding set of real roots of g (^ , R), Prove that
for all a e'^^A, a^ = 2a^/{a^ \ a^) = 2a^/{a \ a) (see Proposition 4.1.6
and Section 4.4).
4.10 Let A ht 2i Cartan matrix, R a realization. Prove that Dg(y4,/?) is
centrally closed. [Take a universal central covering algebra of D g(^, R\
and lift back a set of generators. Show that these can be adjusted to
satisfy the relations (Rl) and (R2) of the original set of generators of
D g(^, R\]
4.11 In this exercise we obtain an explicit realization of the Lie algebras
g(^,i^), where A is an affine Cartan matrix of type A^P, Cf-\
D p , E p , EP, EP, FP, or o p and R is a minimal realization of A.
Let Ah&di Cartan matrix of type A^, C^, D^, E^, E^, Eg, F^, or
G2, and let A be the corresponding finite root system with base
n = {«1, . . . , a/}. Let (f> = L\=ln¿a¿ be the highest root of A, and let
A be the affine Cartan matrix constructed from (f> in Lemma 3.5.3. Let
g = g(v4, R) be the simple Lie algebra of type A based on some
minimal realization R = (1^, II, 11^) of ^ over the field IK.
Excercises 389

Let (• 1• ) denote standard invariant bilinear forms on g and also on


1^*, and assume that it is so scaled that ( a \ a ) = 2 for long roots.
Finally, let lK[i-^] denote the ring of Laurent polynomials
1 e K, = 0 for all but finitely many i e Z}.
(a) Let § — IK[i-^] 0|)^g 0 Cc be a one-dimensional vector space
extension of the Lie algebra IK[t -^] (over K). Define a new
bracket [ •, • ] ~ on § by bilinear extension of
(i) [ g , c ] ~ = o
(ii) w ® X , <8> y]~ = ® [x, y] + q( x I y)c for all i, j
e Z, X, y e g.
Prove that (§,[•, • ]~) together with the mapping

TT: g h± 1] ® 9,

(S>X a0 <S>X,

is a central covering of lK[r ® g (see Section 1.9).


(b) Let res: IK[i-^] -> K be defined by res(E«y t^) = a Prove that

[ f ® x , g ®y ] ~ = fg ® [ x , y ] + r e s |g ^ j ( j : I y)c

for all / , g e IK[i*^],A;,y e g.

Henceforth we will simply write [•, • ] instead of .


(c) Let {e^, <!>'', f^} be an §l2-triplet with e g’*’, e g“ '*’ (Pro­
position 4.1.6). Prove that = 2<f>°/(<l>° \ <j>°) = 'L'l^iftyay,
where n / = n,(a,-1 a,)/2 .
(d) Let e„ /i,, /„ i = 1,.. ., / be the standard set of generators of Dg,
and identify them with the elements 1 ® 1 ® /i„ 1 ® /, of g.
Define

Cq == 1 ® /^ , /o == 1 ® /iq ~ C ~ 4>'^■

Prove that {cq, / io,/o) is an §l2-triplet of g, and prove that


{e„ hi, /, 11 = 0 , . . . , /} satisfy the relations (Rl) and (R2) of
Proposition 4.3.3 for the affine Cartan matrix A.
(e) Let §* := ^ © Kc © Kd, and set fl'^== {a^, a ^ , . . . , a/ } c ^ -l-
Kft with {«i'',. . . , a /} c i) from the realization R and uq — ¡t -
Define n = {ao, « i , . . . , a,} c (§*)* by

<a,,d> := 5,0-
390 Contragredient Lie Algebras

Prove that R •= n , n ^ ) is a minimal realization of A and


that R is naturally embedded in R.
(f) Let § := q(A, R), Show that there is a natural surjective homo­
morphism
:Dg ^ D§.

(g) Using Proposition 4.3.9, show that <p is an isomorphism.


(h) Show that the imaginary roots of § are Z8, where 8 •=
«0 — 1, and
if A: ^ 0,
dim § fed
+ 2 if A: = 0.
Show that = A + Z5.
(i) Show that the reflections act on A as follows:
r^8 = 8 for all / = 0, . . . , /,
r^(a + k8) = r^a k8 / = 1, . . . , /,
fQ(^oi + Ac5) = + (Ac + for all oc ^ A, k ^ Z,

where (f) is the highest root of A.


(j) Prove that § is centrally closed and that hence § is the universal
covering algebra of
4.12 Let R = be a minimal realization of the Cartan matrix
A = Let n = . . . , CD;} c ]^* satisfy (a)¿, a /> = S^j,
1 < i, j < I, where II^= {a]^,. . . , a/}. Let P = Cl Q, Prove that P,
and also P+-= {fi ^ P\ ( , cr^^> e N for all / = 1,..., /}, span 1^*.
jjl

4.13 In this exercise we consider a generalization of contragradient Lie


algebras that allows one to treat the following very natural question:
Let g be a contragredient Lie algebra, and let f c g be a subalgebra
such that I ^ the diagonal subalgebra of g) and cr(i) = f, where
a is the anti-involution of g. What sort of Lie algebra is I? It is, up to
a central factor, one of these generalized contragredient algebras. This
type of generalization was first introduced by R. Borcherds (Gener­
alized Kac-Moody algebras, /. ^4/g. 115 (1988): 501-512) in order to
answer such questions. He assumed invariance and positive definite­
ness (as in Section 4.7). The presentation in this exercise follows our
presentation in Section 4.4 and does not assume invariance. The
exercise involves reworking most of Section 4.4, taking note of how to
cope with the extra generality. At the end we return to the motivating
subalgebra problem. It is noteworthy that even if the original contra­
gredient Lie algebra g is finitely generated, f need not be, and hence
the potential that the index set J is infinite is very real.
Exercises 391

Let g be a Lie algebra over a field K. We say that g is generalized


contragredient triangular or simply generalized contragredient if there exists
a triangular decomposition (g+, 1^, Q+, a) of g satisfying the following
conditions:

GCTl If ej c is the given basis of 0 + according to the axioms of


triangular decomposition, then the Lie algebra g^-*^ generated by g“>
and g““j is three-dimensional for all j e J.
GCT2 The elements of g“^ g”“', j e J, and generate g (as a Lie
algebra).
GCT3 The sum Eye 9“^] is ^ direct sum.
(a) Show that dim,^ g“^ = dim^^ g~“^ = dim J g " “^ g““^] = 1. If a /
is a basis of [g“j, g““^], show that two cases are possible:
Case H: [a^^, g“'] = (0) in which case we say that j e Jyy,
Case S: [ a /, g“'] = g“' in which case we say that j e J^.
Write J = Jyy U J5. Show that ; e Jyy <=> g<^> is a three-dimensional
Heinsenberg algebra and / e <=> g^-'^ s §I2(IK). Conclude that
(g, 5^) is contragredient iff Jyy = 0 .
(b) A display of (g, is a set ~ (ej, a j , /y)ye j such that
Dl: g(^> = IK/y ® K a / e Key,
D2: [ey,/y] = a / , [a/,ey] = 2cy, [ a / , / y ] = - 2 / y whenever
j e J5.
D3: <re, = /¿.
Show that displays exist. Show that if {cy, a / , /y} and {e'j, a '/, //}ye j are
two displays of (g, ^ ) , then there exists families (tj)j^ j, (sj)j^ j, (kj)j^ j
m such that

e;. = ryey, a^=kjar, fj = t^fj,

with kj = 1 fy = +1, whenever j e J^.


(c) Define the structure matrix A = A ( 3 ) of relative to a
display 3 = {ey, a / , /yly^ j of J by

Aij ^(x^yOij ),

Show that Ajj = 2 if ; e J5 and Ajj = 0 if ; e Jyy. Describe how


A changes if 3 is changed.
(d) Generalize all the theory developed in Proposition 4.1.1 to
Proposition 4.1.7 to generalized contragredient Lie algebras (Let
" J : = { ; ■ e J5 I ad Cj is locally nilpotent}. We say a e A is real if
a = w u j for some w ^ W and j e " J .
392 Contragredient Lie Algebras

(e) Let A be the structure matrix of (g, relative to a display. Fix


j ^ Js- Show that the following hold.
(i) j iff e Z <0 for all i e J, j # ; .
(ii) If j g " J and i ¥=j, then == (adcy) = 0.
(iii) If j then for all i e J,

Aij = 0 « [ e , , c j = 0,
A,j = 0 ■Aj, = 0.

(f) Let (g, be a generalized contragredient Lie algebra, and let


g = g/t, where t is a o--invariant ideal of g contained in rad(g).
Then the natural mapping ~ : g ^ g is injective on the subspace
^® Identifying 5 and 1^, we have that
*^=_(^5S+,G+,cr) is a triangular decomposition of g and that
(g, is a generalized contragredient. l^ rth erm ^e every display
3 of (g, induces a natural d isp l^ 3 of (g, and relative
to 3 and 3 both (g, and (g, 3~) have the same structure
matrix. Finally,

"^A(g) c^"A(g) and IF(g)cIF(g).

(g) Let (g, be generalized contragredient with structure matrix


A = Define g — g/rad(g), and let " : g g be the
canonical map. Then for j e
(i) If A¿J e Z_ or if A¿J = Aj^ = 0 for some i ¥=;, then d^j —
(ad and d~j -= ( a d b e l o n g to rad(g).
(ii) For ad 6j (and hence ad fj) to be locally nilpotent, it is
necessary and sufficient that A^j ^ Z < q that A^j = 0 =>
Aj^ = 0 for all i ^ j\
(h) Let (g, <5^) be generalized contragredient. If 3 and 3 ' are
displays of ( q, 3~) and A and A' are their respective structure
matrices, show that A is symmetrizable iff A' is. Follow Proposi­
tions 4.4.1 and 4.4.4. Define the concept of invariant generalized
contragredient Lie algebra. Show that the proper form (• | •) on
g can be chosen to satisfy IBFl, IBF2, IBF3, and
IBF4 (a¿ I ay) = A^jSj for all i, j e J, and a^ = 2aJ/(aj \ ay) if
j ^ J5.
IBF5 (a,^ I a / ) = A^jS~^ for all i, ; e J and a,. =
2 ( a / ) V ( < I < ) if j ^ h
A form satisfying IBF1-IBF5 is called standard.
(k) An ordered generalized contragredient Lie algebra is a triple
(g, 3 ) consisting of a generalized contragredient pair (g, 3^)
and a display 3 relative to which the structure matrix A of
Exercises 393

(9, satisfies

ORD all entries of A lie in an ordered subfield F of IK


and for all ; e J the set
Rj -.=
has the following properties:
If j e Jff, either Rj c Fg.o or Rj c F<o-
If j e J j, either R j (z í ^ q Ot R j c Z
Define J * by y e Rj c F ^ q> J \ J"^ so that

/ = J -U J+.

If (g, 3 ) is radical free and A has the property that A¡j = 0 <=>Aj¡
= 0, then show that
^7 = J + n J5,

'■'”/= = J \ " J = J -U ( j + n j „ ) .
(Use part (g) above.)
(1) Let (g, y , 3 ) be an invariant ordered generalized contragredient
Lie algebra, and let A be its structure matrix relative to the
display 3 . Show that the family isj)j^j syimnetrizing A can be
chosen so that

IBF6 > 0 if j e J"^,


1Cy < 0 if j e J “ .

(m) Let (g, y .^ ) be as in (1). Show that g carries a standard


invariant form satisfying the following:
(i) The restriction of (• I•) to (2 ^ 2 is F-valued.
(ii) (a, I tty) 0 if i, j e J, I 9«=j.
(iii) ( t t y I a y ) > 0 if y e''*J.
[Choose a form as in (1). First consider the case rad(g) = (0), and use
(k). For the general case use (f).]
Let p e 1^* be such that

<P, a/> = l if y e Js,


<P,<> = 0 if y ^ J/Í-

(iv) If J'^ n Jo ='^*J show that


394 Contragredient Lie Algebras

(0 (ay I ay) < 0 for all j Conclude that (a | a) < 0 for all a e

(ii) 2(p Ia) —(a Ia) e F^o for all a e A+. Moreover 2(p | a) =
(a I a) if and only if either a = ay for some ; ^ J or a =
+ • • • +Cy^ay^, where /: > 2, {j\ \l < i < k ) and ^y = j =0
for all 1 < p < ^
(n) Define and recover the results about Casimir-Kac operators for
radical-free invariant generalized Kac-Moody Lie algebras.
(o) Let (g, D) be an invariant ordered generalized contragredient
Lie algebra, and let A be its structure matrix. Show that the
radical of g is generated (as an ideal of g) by the elements

+1 ( J^
dr:= (a d /y )

A , J = A . = 0.

Show that u modulo the above relations satisfies J^, and


use the Gaber-Kac Theorem).
(p) Let (g, «5^) be a contragredient Lie algebra. Let i be a subalgebra
of g satisfying 5 c f and o-(f) = i.
(a) Prove that there exists a generalized contragredient algebra
together with a surjective homomorphism tt : ^ i,
such that ker(7r) c centre ( ^ ) c and ^ is invariant if g
is. (Order the positive roots of g+ in a way respecting height.
Choose generators of f in a-symmetric pairs by ascending
height. Note: The number of generators will in general be
infinite even if g is finitely generated.)
(b) Prove that if (K = [Rand g is invariant, then ^ can be chosen
with ^ = f .
Chapter Five

The Weyl Group and Its


Geometry
A m a n t r a v e l i n g a c r o s s a f i e l d e n c o u n t e r e d a t ig e r . H e f l e d , t h e t ig e r a f t e r h im .
C o m in g to a p r e c ip ic e , h e c a u g h t h o ld o f a v in e a n d s w u n g h im s e l f d o w n o v e r th e
e d g e . T h e t ig e r s n i f f e d a t h i m f r o m above. T r e m b lin g , th e m a n lo o k e d d o w n to
w h e r e , f a r b e lo w , a n o t h e r t ig e r w a s w a i t i n g t o e a t h im . O n l y t h e v i n e s u s t a i n e d
h im .

T w o m i c e , o n e w h it e a n d o n e b l a c k , lit t le b y lit t le s t a r t e d t o g n a w a w a y a t t h e v in e .
T h e m a n s a w a l u s c i o u s s t r a w b e r r y n e a r h im . G r a s p i n g t h e v i n e w it h o n e h a n d , h e
p l u c k e d t h e s t r a w b e r r y w it h t h e o t h e r . H o w s w e e t i t t a s t e d !
—A Buddhist parable—from T h e W o r ld o f Z e n , Nancy Wilson Ross (New
York: Vintage Books, 1960)

The Weyl group W of a contragredient Lie algebra (g, was introduced in


Section 4.1. As we saw, it is possible to gain valuable information about g by
means of W. For example, the automorphisms of g of the form n^(t),
a permute the root spaces g^ according to the rule n^(t ): g^ ^
where r^ ^ W is the reflection in a defined in Section 4.1. We used this to
conclude that if a is real, then g“ is one-dimensional.
Evidently we can think of W in relation to automorphisms of g, and we
can also think of PL as a purely geometrical object that permutes the set
Ac Information obtained by looking at PL as a group of permutations of
A can be used to understand g, and vice versa. These considerations are
quite vital for our later work on g and its representations. However, abstract
root systems are quite important in areas of mathematics outside Lie theory,
and we have consequently tried to make this chapter somewhat self-con­
tained. Thus our approach to PL is axiomatic and, on the whole, free from
the Lie algebra context. In fact we do need a little Lie algebra theory to
establish one of the basic facts about root systems right at the beginning
(Proposition 5.1.4) and again in the section on imaginary roots.

395
396 The Weyl Group and Its Geometry

This chapter then is devoted to a detailed study of Weyl groups and


abstract root systems. The framework in which we work is a structure called a
set of root data. One of the main problems in dealing with infinite root
systems is that although they may have finite bases in the sense of finite root
systems, their subroot systems may only have infinite bases. Thus from the
very outset one needs to consider infinite bases. On the other hand, to prove
that subroot systems even have bases and then that bases of root systems are
conjugate, it seems necessary to create a chamber geometry similar to the
well known finite theory (see [Bo3]). Our definition of root data is designed to
encompass all root systems and subroot systems of Kac-Moody Lie algebras,
to allow construction of a chamber geometry, and to work over arbitrary
fields of characteristic zero. It is based on earlier work of J. Tits, E.
Looijenga, and Moody-Yokonuma. The presentation follows [MoPi2] quite
closely.
After establishing the notion of root data in Section 5.1, we have three
sections dealing with basic facts on the Weyl group and its action on root
systems, including the length function, the relationship with Coxeter groups,
and the Bruhat ordering. The central section of the chapter is Section 5.6 on
the chamber geometry of a set of root data. This is followed by a section on
subroot data in which we prove the important fact that any subroot system of
a set of roots may indeed be viewed as a set of root data in its own right.
Section 5.8 is on imaginary roots. This contains the Kac characterization of
imaginary roots and shows how the three basic types of root systems, finite,
affine, and indefinite, are distinguished by their imaginary roots. The final
section shows that any two bases of an indecomposable set of root data are
conjugate up to an overall sign by the Weyl group. This shows immediately
that the Cartan matrix is an invariant of the root system and is an important
first step in proving that it is also an invariant of any invariant Kac-Moody
Lie algebra (Chapter 7).

5.1 ROOT DATA

Let K be a field of characteristic 0. By a set of root data over IK we


understand a 6-tuple

consisting of

RDl: a Cartan matrix A = where J is an index set (see


Proposition 4.1.9);
RD2: a pair of vector spaces V and V ^ over IKtogether with a nondegen­
erate pairing ( • , • } : V X V IK;
RD3: subsets II = {a,},ei ^ ^md 11'^= c such that
(a,, a / > =Aij for all i , j e J.
5.1 Root Data 397

Furthermore we require that

RD4: Q := and are free abelian groups admit­


ting bases of the form {%}, ^ Q and ^ i2 where the
and the are linearly independent over IK and satisfy

H e ® Kly. and n^c ©


¿el ¿el

The last of these axioms is familiar from the axioms of triangular decomposi­
tions.
Note that the sets II and II ^ of RD3 need be neither finite nor linearly
independent. The standard examples of root data are formed from Cartan
matrices A and realizations R = (1^, II, II of ^ over K by setting V •= ij*,
F^-= and < - , - > : ] ^ * X] ^ ^ [ K the natural pairing. We call ^ =
(A, n , n • » the standard root data constructed from R,
We define groups W c GL(K) and IT GL(K in the usual fashion:
For each ; e J define Vj e GL(F) and r / e GL(K by

r- : a ^ a — ( a , for all a ^ V,

rj^ : h ^ h - (aj, h)a^^ for all /i e F

It is clear that rj and are reflections in and , respectively (see


Section 3.2). Let

IT:= <r^-|y e J> and IT < r / | ; e J>.

As usual we also define

2 := i r n and 2 ^:=IT^n^,
which are called the set of roots (resp. coroots) of 2). We also call 2) a set of
root data for 2. Actually the set 2 corresponds to the real roots in the case
of Kac-Moody Lie algebras, but since there are no other roots to be
considered for quite some time, it is more convenient to simply call them the
roots. One observes that the definition of root data invites one to consider
the dual set of root data in which the roles of 2 and 2 ^ are reversed:

where < • , • : F ''X F ^ IK is defined by


398 The Weyl Group and Its Geometry

Note that a ^ ) = 2 for all j e J. Thus aj 0 and 0, and hence


0 2 and 0 ^ 2 ^.
Before establishing our first result we lay down a list of familiar looking
objects

2 + == 2 n X)

/e j
X_:= - X + and X:f== - X ^ ,

Q = L Z«;, Gk == Z “
;e j ;e j
e^== E z a / , G ,r= = E K a/.
;ej ;ej

Let D be a set of root data, as above, with Cartan matrix A and root
system 2. To begin understanding the nature of 2, let us consider a minimal
realization R = (ij, B, B ^ ) of A, To avoid confusion with the objects of our
root data, we change our standard notation for realizations and write

B= j} c r and B ''= { p / \ j (Zfi.

Then, if < • , • > denotes the natural pairing of and 1^, we see that
(A, B, • » is a set of root data. Axiom RD4 is satisfied with
% := Pi and py^ for all i e J. Note that 2 (resp. 2 is the set of real
roots of the Lie algebra q(A, R) (resp. %{A^, R ^)). Let

# = = <r. I; G J> c G L ( r ) and G - ©

be the Weyl group and root lattice of R. As usual let

:= B®.

Proposition 1
Let Wq and Wq be the groups of automorphisms of Q and Q obtained from W
and W by restriction to Q and Q respectively. Then

(i) there exists a unique group epimorphism r : Wq Wq satisfying

r : r, for all j ^ J;
5.1 Root Data 399

(ii) if ф: Q Q is the unique T-linear map satisfying = ay for all


j G J, then for all w g Wq we have the commutative diagram

Q Q
W i iT(iv)
Q Q

(iii) ./-rA) = X;
(iv) ker i/f c {)8 G 0 I<j8, = 0 for all j g J}. In particular, if i is
finite arulA is nonsingular, then iff is an isomorphism.
Remark I We will see below that the above restrictions to Q and Q are
actually faithful. We will also see later that W and W are isomorphic.

Proof, (i)-(ii) If T exists, it is clearly unique. Let i, j g J. Then

•■pj ■/3, ^ - (Pi, = Pi - A^jpj,

while

0■ ^ “ i “ AijOj.

This shows that for all e J,

(1) ■■■r p j i = E c,.y^y ^ ry, • • • ry a,. = E c,7« r


ysj

(The reverse implication is not immediate because the c,y on the right-hand
side may not be unique.) From (1) it is easy to conclude that t exists and that
the diagram of part (ii) commutes.
(iii) We have

2 = = Wil,{B) = i!)WB = ipCA).

(iv) Let j8 G g be such that i/f(/3) = 0. If ; g J, we have

0 = rji[,(l3)

= H^p^P)

= ^{p - (B ,p p p j)
= -(P,B/}aj.

Thus (p, p y ) = 0. Finally, if J is finite and A is nonsingular, then since R is


400 The Weyl Group and Its Geometry

minimal, we have 1^* = and shows that

In view of this result we call ® == (A, B, B 1^, 1^*, < • , • » a covering of


5), in fact the universal covering of S). The term universal is justified in
Section 5.5 where we discuss morphisms.

Lemma 2
There exists a unique group isomorphism * :W satisfying r f = ry.
Moreover, if we identify W and W ^ via this isomorphism, then

( a , h ) = {wa, wh) for all a ^ V , h ^ V ^ , w ^ W,

Proof If a e F and h then for all j g J,

(rja,h) = { a ~ { a ,a y ) a j,f^

= { a , h -{aj,h)aj^)

= (a,r/h).

We claim that for all j \ , . . . , j^, i^, ...Jrn ^ J, we have

r, • • * r,Jn = r, *• • r <=» r,-}\V • • • r,JnV = r, V • • • r:

Indeed, by the nondegeneracy of < • , • > , we have


r,Jl • • • r,Jn = r,M • • • r, <=> r,Jn • • • r,J\ = r, • • • rM

« </•;„ • • • r^a, h) - <r,^ • • • r,. a , h) = 0

for all a e K, A e F'"


« < « , / • / ••• ••• r , l h ) = Q

for all a e F, /i e F
«=> rJ\^ • • • rJm
y = M
• • • r^^.
^n

It follows that there exists a well-defined map * :JV ^ W'^ given by

(r.
V ;i
■■■ r ) * = r ^
Jn^ Jl
••• Jn
.

It is dear that * is unique and that it is a group homomorphism. Similarly we


5.1 Root Data 401

can find a homomorphism * : PF ^ JV satisfying

(o r ■ ■ ■ o r ) * - o , ' " o . -

Thus W = via *. The rest of the lemma is clear. □

Henceforth we identify IV and PF ^ by the isomorphism of Lemma 2.


Thus, for example, we write X ^= {W a/ | j e J}.

Proposition 3 (Weak intersection property)


Let the notation be as above. Then

Zai n Yh ^ >0^; ^ (0) for all i e J,


j^i
WIP:
Z a ^ n i ; Z ^ o < = ( 0) for a l l / e j .
j^i

Remark 2 These conditions are a weak form of linear independence on


the elements of 11 and II ^ and are necessary to proceed further. Their proof
requires RD4.

Proof. Suppose that -nf^a^ = 'Lj^^njCtp where nj > 0 for all j ^ k. If


nj^ < 0, then we obtain a contradiction by computing < • , > on both sides.
If > 0, then 'Lj-rijaj = 0, where each nj > 0 and not all of these are zero.
But by RD4 we can write aj = where the c¿j > 0. Combining these
gives a contradiction to the independence of the {yj}. The case of H ^ is
similar. □

Proposition 4 (Wonderful union property)


Let the notation be as above. Then

WUP: 2 = and

where the unions are disjoint.

Proof. Using the notation of Proposition 1 and recalling that


(see Section 4.2), we have

X = U i ^C®A_) c X + U X _ .
402 The Weyl Group and Its Geometry

Now WIP shows that

(2) if ^ = 0 where all e Z t h e n every = 0.


/e j

Hence, if where < 0 and > 0 for all i, then by


moving everything to the right-hand side and using (2), we find that =
0 for all i. This proves disjointness. □

Example 1 (Maxwell’s demon [Mx])t Let

/ 2 - 2 0^
A = \-2 2 -2
I 0 -2 2

Let V = R^o ® R«! 0 R o:2 be a three-dimensional real space, V* its dual,


and < • , • > : K X V* R the natural pairing. Let II = {ao> “ 2) ^
and define II {uq , a^, a^] c V* by <a,, a^y) = Ajj, 0 < i, j < 2. With
the Weyl group IF = <r, | i = 0 ,1 ,2> defined in the usual fashion and identi­
fying IF ^ with IF, let

/3a = ('•2''i)* “ o and Pk = (.r2r y a ^ .

Claim 1 = «0 + 2k^a^ + 2(/c^ + k )a 2 and

< + 2*V+ + ic)a^ forall k e Z.

To establish this, reason by induction on k.


Claim 2 For all p, q ^ Z, p ¥=q, we have

< 0,

< /3 „ ^ ;> = 2.

Indeed

( Pp, Pg) = ( ( ' - 2''l)'’« 0.('-2''l) ‘'«o')

= ( « 0»(''2''l ) * < )
= ^ oiQ, (Xq + 2k^0LY + 2(^k^ + k)oL2^
= 2 - Ak^,

^This is another Maxwell and another demon.


5.1 Root Data 403

We return to our main construction. Let

2 -2 -1 4 -3 4
-2 2 -2 -1 4
-1 4 -2 2 -2
-3 4 -1 4 -2 2

Thus defined, B is a generalized Cartan matrix (infinite, of rank 3). Let

and define

fl := Wr^T and n ^ := W^T

where

W^:= CZW,

The set is an example of a subroot system of X (see Section 5.7). We claim


that

is a set of root data with root system ft and Weyl group Wr^. RD1-RD3 are
obvious. For RD4 we may simply take

To = « 0. T i = 2tti, 72 = 4«2,

T o" =< > T l" = 2 a l^ 72'^=4a2'^.

The reader may verify that ® ® that

The interesting lesson in this example is that it demonstrates that even if


we begin with root data based on a 3 X 3 Cartan matrix, we may find inside
the corresponding root system sets of roots that are based on infinite Cartan
matrices. It might occur to the reader that ft has some finite subset H,
unknown to us, that could be used to replace the infinite set T in the sense
404 The Weyl Group and Its Geometry

that

where C is defined analogously to B using H. But this is not the case. We


will prove in Section 5.8 that B is an invariant of its system of roots. This
example shows why it is necessary to allow the index set J to be infinite in
order to be able to discuss subroot systems.

Proposition 5
Let if/ :Q Q be as in Proposition 1. The restriction o f i/f to {resp. ^^A_)
is a bijection onto (resp. 2 _).

Proof Suppose that = il/(w2pj) for some ^ ^ i ,j s J.


Then

lll{w2 = l/rOy) = aj.

Now

Lc*/3*,
k^J

where the have all the same sign. Then

OLj ^
A:ej

violates WIP unless = 1 and = 0 for all k ¥=j. Thus = Pp


this shows that i/f is an injection. The proposition now follows from
c S + and WUP. □

Let q: e S. Then by the last proposition there exists a unique p =


Lj^jCjpj e'^^A such that ij/ip) = a. We have

« = E
yej

which we call the natural expression for a in terms of the {ay}. The quantity
EyejCy is called the height of a and is denoted by ht(a):

ht(a) ==
yej
5.1 Root Data 405

The next five results have already been established for root systems of Lie
algebras. Here they are proved in a Lie algebra free context.

Proposition 6
Let q: e S, and let j e J.

(i) Qa n 2 = {+ «}.
( ii) r / 2 A { « ;} ) = 2 A (« y }.

Similar results hold if we substitute 2 ^ for 2.

Proof (i) It suffices to show part (i) when a = for some i e J. Suppose
that e 2 for some p, q in Z. There exist integers Cj, j e J, all of the
same sign such that

1/7
\ ‘1 /

Cross-multiplying by q and applying WIP, we see that c¿ = p /q Ei Z, Now, if


na^ G 2 for some n e Z, then na^ = waj for some w and j e J. Then
w = (1 /n)aj e 2 , and we can reason, as above, to conclude that n = ± 1.
(ii) Let a e X+\{aj}. Because of part (i) we have a = where all
Ci > 0 and Cl 0 for some / # j. Now r^a = a + kaj for some A: e Z. Thus

r ja = (cj + k ) a j + Y ^ C iO C i.
i*j

Because of WIP we see that TjU # Uj, so it will suffice to show that rjOL ^ X+.
Suppose not. Then by WUP,

{Cj + k)aj + L c,a,. = £ d,a,,


i* j / €J

where d¡ < 0 for all i. Thus

[d j - {Cj + k ) ) a j = £ (c, - d,)a,.


i^j

By WIP we have

(3) E (c, - d ,)a, = 0.


i* J

But Cl — d i > 0, and (3) then contradicts (2). This shows that ry(2+\{ay}) c
(X^\{aj)). Since r? = 1, equality follows. □
406 The Weyl Group and Its Geometry

Lemma 7
Let i, j e J, and suppose that w ^ W is such that = aj. Then

(i) wr¿w~^ is a reflection o f both V and in af,


(ii) wriW~^ = rj on {in Proposition 10 we prove that wr^w“ ^ = ry).

Proof Let p Ei V. Then

wr^w~^l3 = — (w~^P, ay) ai )


= /3 - (w~^p,a^}aj.

This shows that wr¡w~^ maps «y into —«y and pointwise fixes both the
hyperplane {wy | y e F and <y, a ^ } = 0} of V) and the hyperplane [wy \ y
e <2k and <y, a /> = 0} of This establishes part (i).
Next we write © = rywr^vv *. Then

(4) 0/3 = /3 + (<w - </3,ay'^>)ay.

We conclude that ©ay = ay and that if k ¥=j then ©a*, e 2+ (for this we
use WIP). In turn we have © 2 + c 2+, © 2 _ c 2_, and hence ©2+= 2+.
We conclude that ©II = II. For suppose that ©a* = where
c, > 0 and that at least two c, are strictly positive. Since ©“ *2+= X+, we
obtain
a* = E d , a , ,
/SJ
where d^ > 0 and at least two d^ are strictly positive. This contradicts WIP,
and hence ©II = II, as desired.
If k E J, then by (4), ©a^ = for some n e Z. But ©II = II.
Thus naj = ai for some / e J. By WIP we conclude that I = k and
n = 0. This shows that ©a^ = Thus ©Ig^^ = id and (ii) follows. □

Lemma 8
Let i , j E J, and su p p o se th a t w e W is s u c h th a t w r -w ~ ^ = r^ o n Q ^ . Then

wa^ = ±ay.

P ro o f We have
—a j = rytty = >vr^iv” ^ay

= w{w~^Uj — <w” ^a:y,


= ay — <H^“ ^ay, ay)wa^.
Thus iva^ E Qay, and we can now apply part (i) of Proposition 6. □
5.1 Root Data 407

Proposition 9
Let i, j e J, and let w ^ W . The following are equivalent:

(i) wai = Ci j .
(ii) w a y = a ^ .

Proof. If wa^ = t t j , then wr-w~^ — r, on (2k (Lemma 7). Thus on


wr/iv"* = rf" (identification of W and W'^), and hence ± a / (the
last lemma applied to IT instead of IT). Consider now the identity (4) of
Lemma 7. For arbitrary ^ e it reads

/3= 13+ (<w'-‘i 3 , a , " > - < ^ , < > )« ,.

= /3 + (<j8,wa,>'> - <j8,a/>)a:y.

Suppose that w a^= Then by setting j8 = we get - 4 a j = 0. This


contradiction shows that w a ^ = a ^ , as desired. The converse follows along
similar lines by dualizing. □

Proposition 10
Let i, j ^ }. If w ^ W is such that wa^ = aj, then wr{w~^ =

Proof Using Proposition 9, we can replace a / > of equation (4) by


{p,wa^^} = (p, a / > , so now it reads 0)8 = p for all )3 e F.

Proposition 11
Let a, )3 e n , a ¥= j8. Then the order o f r^r^ e W depends only on the
product := <a, )8 ^ ><)8, a ^ > and is given by

0 1 2 3 > 4

2 3 4 6 00

Proof Let X ^ V and consider the vector space U(x) •= IKx + Ka + K/3.
This is clearly invariant under r^r^. Assuming that x ^ Ka -\r Kp, the matrix
of r^r^ \u(x) relative to the ordered basis (x, a, /3) of U(x) is of the form

1 0 0
* (P,a^y
* - ( a , 13"^} -1
408 The Weyl Group and Its Geometry

and its characteristic polynomial is

( A - l ) ( A ^ + ( 2 - A / ' „ ^ ) A + l).

If X \a we drop the first row and first column and obtain the
characteristic polynomial + (2 - + 1.
For the cases = 1,2,3, + (2 - + 1 has roots that are conju­
gate pairs of primitive third, fourth, or sixth roots of unity. Thus r^r^\u(^x) has
distinct eigenvalues, so is semisimple and clearly has finite order 3, 4, or
6, respectively. Since x was arbitrary, has order 3, 4, or 6 on V.
For - 0, = r^r^, and since For > 4,
2- < —2 and A^ + (2 —A/^^)A + 1 has no roots of unity as roots. Thus
has infinite order. Finally, for = 4, r^r^lc/(0) has eigenvalues 1,1,
and since r^r^\u^o^ # 1, it has infinite order. □

The above proof shows the following:

Corollary
Let denote the group generated by r^ and r^. Then (r^,r^) acts
faithfully on Ka Kp. □

5.2 THE LENG TH FUNCTION

We continue with the setup and notation of Section 5.1. In particular we


assume that we have root data ® = {A, FI, II K, K < • , • » with root
system X and Weyl group W,
Given an element w e IF, iv 1, we can write

( 1) W = r:

where ..., ^ J (these need not be distinct). Expressions like those of


(1) are called words (of length n) of w. [Strictly speaking, words should be
taken as sequences (r^^,. . . , r^) satisfying (1). Fortunately the less cumber­
some notation that we use is not likely to lead to any confusion.] There is
nothing unique about the way in which an element of IF can be expressed as
a word. As a trivial example.

rj = rprj^ for all ; e J.

There is, however, a smallest positive integer n that permits w to be written


in (1) as a product of n of these reflections. This number is called the length
of w and is denoted by /(w). A word of w of length /(w) is called reduced.
We define /(1) = 0. If w = r, r:Jn‘. then w~ a r , , from which it is
5.2 The Length Function 409

evident that /(w) = /(vv“ 0. In general there is more than one way to write w
as a reduced word.
The length of an element w of the Weyl group is closely related to the
number of positive roots that are mapped into negative roots by w '^. More
precisely for each w e PT we define

5^ = (a G 2+ I G 2_}

We will show later (Proposition 3) that

l{w) = card(5^).

Lemma 1
Let j e J. Let w ^ W, and let be as above. Then

(0 =

(ii) S^. = {«y};


(iii) ry(5^ \ {tty}) = \ {tty};
( iv ) I f aj i S^, then Uj G and
Srjy. = rjS„ U {tty} (disjoint union);

if aj G 5^, then aj ^ and

(v) card(5^) < Kw).

Proof
(i) The proof is clear.
(ii) This follows from Proposition 5.1.6.
(iii) Let a G \ {ay} c Then

(2) r y tt G

by Proposition 5.1.6. On the other hand,

(rjW)~^(rja) = w~^a e S_.

Combined with (2) this shows that

The reverse inclusion now follows by replacing w by rjW.


410 The Weyl Group and Its Geometiy

(iv) By definition,

Uj G 5„, <=> w~^aj e X_,

^ ^rjw ^ - w - ^ a j G 2 _ .

Since by WUP exactly one of these holds, part (iv) now follows from part (iii).
(v) By part (iv), card 5^.^ < card 5^ + 1. Thus part (v) follows by induction
on /(w). □

Lemma 2
Let w = r, • • • r, be a reduced word o f w ^ W. Then

r,-^1 • • • r,Jn-l^(a ,Jn^) ^ 2 +


+.

Proof. Suppose the result is false, and let m be the largest positive integer
such that

Om *** ^
Set a ’= r, • • • r, (a, ). Then a e 2+ and r, (a) ^ X_. Thus a = a.- by
part (ii) of Lemma 1. Set w' — r^^^^ • • • As we have just seen, =
a,Jm , and hence w' = r,Jm iv'r,Jn by Proposition 5.1.10. We can now find a word
for w of length n — 2, namely
w = r: • • • r,JmwV,Jn = r,J\ • • • Jm
r,- —l w' = r-Jl • • • rJm • • • r,-Jn —1

(the hat indicating that this term is omitted). This contradicts /(w) = n. □

Proposition 3
Let w = rj^ • • • rj^ be a reduced expression of an element w ^ W . Then

(i) 5^ = {aj^, rjaj^,. . . , ... rj^_aj}, and the elements of displayed


above are all distinct’,
(ii) card(5^) = /(w).

Proof Each of the expressions rj^,r^rj^,... ,r^^ ••• ^ is reduced, and


hence each of the roots displayed above is positive, thanks to our last lemma.
Also, if 1 < m < n, we have

w~^r:J \ *• * r,Jm —\ a,Jm = r,Jn • • • r,Jm a,Jm


= —r-J n • • • rJm + l a,Jm

(with the obvious conventions when m = 1 and m = n \ and this last root
5.2 The Length Function 411

belongs to 2_, again by the last lemma. We conclude that

••• ^■5«-
Suppose now that two of the elements listed above are equal. This leads to
an identity of the form
a,Jm = r,Jm • * r,J p —l a,J p

for some 1 < m < p < n. Applying rj^ to both sides gives
—a, = r ‘ •r a, ,

which is impossible since the right-hand side belongs to X+ by Lemma 2.


This shows that all the elements in question are distinct. Since there are n of
them, both parts (i) and (ii) now follow from card(5^) < l(w) = n [Lemma
l(v)]. □

Corollary 1
Let w ^ W . Then
5^ = 0 ^ IV = 1,
5«, = {a,} = Vj.
Corollary 2
W acts faithfully on Q.

Proof If w \q = 1, then 5^ = 0 and iv = 1 by Corollary 1. □

Corollary 3
W = W Q a n d W = Wq. (see Remark 5.1.1)

Proposition 4

(i) Let rj^ • • • rj^ be a reduced word for an element w ^ W .


If j e J, then

l{rjw) = l{w) + 1 «=> ay « 5^ H'-i(ay) e


l{rjw) = l{w) - 1 <=> tty e <=> ^ “ ^(ay) e 2_.

In particular l(rjw) = l(w) ± 1.


(ii) Let rj^ • • • rj^ be a (not necessarily reduced) word for w. Then
n = l(w) mod 2.
(iii) I f w and w' are conjugate elements o f W, then l(w) = /(iv')mod2.
412 The Weyl Group and Its Geometry

Proof, (i) This follows by combining Lemma l(iv) with Proposition 3(ii).
(ii) Use induction on n . l i w = rj^. .. then, for any j\ Krjw) = l(w) ±
l = n ± l = n - \ - l mod 2.
(iii) This follows from (ii). □

If S is a finite subset of the root lattice 0 , we define

<5) = L r ^ 0 -
y^S
The sums turn up as an important ingredients of the character formulas
later on. If the space V contains an element p satisfying

(3) <p, a /> = 1 for all ; e J

(for instance, if II ^ is finite and linearly independent in K ^ (see minimal


regular weights defined in Section 4.4), we have the following simple formula
for <5^>.

Proposition 5
Assume that p e Vsatisfying (3) exists. Let w,w^,W2 ^ W. Then

(i) (S ^ ) = p - w p ,
(ii)

Proof, (i) We reason by induction on l(w). The result being clear for
/(w) = 0, we assume that w = rjw' for some w' ^ W satisfying /(w) = 1 +
l(w') and that part (i) holds for w'. Then
p — wp = p — rjw'p

= P - rjp + r^(p - w'p)

= P - {p - (p^oij^)oij) + rjiS^r)
= oLj + r^{S^>)
= <5.>
(this last equality by Lemma l(iv) and Proposition 4).
(ii) We have
= p - W^W2p
= p - WiP + H^i(p - W2p)
=
as desired.
5.2 The Length Function 413

Corollary

(i) wp = p «=> w = 1.
(ii) <5„,> = Wj = W2.

Proof, (i) wp = p => p - wp = 0 => (5^> = O= > 5 ^ = 0 = > > v = l


(Corollary 1 to Proposition 3).
GO (S „) = <5^^> =» p - w ip = p - W2 P = » W i“ V jp = p =» w r ‘ w 2 = 1

(by part (0) =►Wj = W2*

Proposition 6
There exists a unique bijection

/:2 ^ 2 ^

satisfying f(waj) = /or all w ^ W. Moreover

a e S +« / ( a ) e

Proof It is clear that if / exists, it is unique. Suppose that = ^ 2^]-


Then W2 ^^1^/ = hence v= a / (Proposition 5.1.9). Thus
Wiuy= ^ 2« / . This shows that there exists a well-defined surjection / from
2 into satisfying way >-* way. Using Proposition 5.1.9 again, we see that
this map is injective.
Let a e X, and write it as a = w^^ay for some ; e J, w e IP. Then using
Proposition 4 applied to the root system X and the coroot system X'^
together with the usual isomorphism of W and W

a = w '(tty)eX+«»/(ryw) > /(w)

« w-iay'^e X^

« /(a)

Nidation The bijection / : X -» X satisfies /(ay) = a / . There is no ambi­


guity then in denoting f simply by to the extent that (ay)'^= ay. With this
notation

(w a,)'"= w ( a / ) ,
(4)
a e X+<^ a'^e X
414 The Weyl Group and Its Geometry

Note also that if a e 2 and a = then

(5) = {wa^. way) = {a-^,ay) = 2.

For each a e 2 define a reflection e G LiV) by

: jc jc — {x, a ^>a.

Proposition 7
Let a, e 2 e PF. TTzen

(i) on/y if a = ±]8,


GO

(iii) e PF, moreover, = r, /or all j e J,


(iv) r^h = h — {a, h )a ^ for all h ^ V

Proof (i) The proof is obvious from the definition and from Proposition
5.1.60). We have

wr^w~^x = w{w~^x — {w~^x, a"^)a) = x — {x, wa' ^}wa.

By Proposition 6, wa ^ = ()vo:)^, and part G O follows at once. Choosing w so


that w~^a = a^, we obtain wr^w~^ = r,, which shows that Oii) holds. Finally,
part (iv) follows from part (ii) and Lemma 5.1.2. □

Proposition 8
Let a, p ^ 1,, a ±p.

(i) The following are equivalent:


(a) = 0.
(b) r^a = a.
(c) = r«r^.
(d) ( P , a ^ ) = 0 ,
(ii) > 0 <=> > 0.

Proof (i)
(a) => (b) Obvious.
(b) => (c) Proposition 7(ii).
(c) => (b): r^r^ = r^r^ => r^^^ = r^a = ± a => r^a = a, since
a — kp = —a is impossible.
(c) <=> (d) Use symmetry.
(ii) Suppose this is false. Then by (i) we have

(6)
5.2 The Length Function 415

for some a and )3 in S. By Lemma 5.1.2 there is no loss of generality in


assuming that in (6) a e n . Replacing by —/3, if necessary, we may assume
that Now by the last proposition

r^a = a - <q;,)8^>/3 while r^a^= a ^ -

By (6) these two have different signs by WIP and this contradicts Proposition
6 given that (r^a) ^ □

Let I c J be arbitrary. Let

III •= \i ^ 1 } and := ( a / U' ^ l},

and let AI denote the submatrix of A corresponding the index set I. Define

Wi := clLcG L(K ).

For convenience we set = {1}. Subgroups of W of the form are called


parabolic subgroups. (The terminology comes from the theory of finite­
dimensional simple Lie groups.) What follows are some basic facts about
parabolic subgroups.
Define

S, := 2 ,^:=

We also define the sets 5^ and the relative length function /j on Wi in the
obvious way. Consider the six-tuple

®,:= { A , , U „ U , \ V , V \ ( - ,• }).

It is obvious that the axioms RD1-RD3 of root data hold. RD4 is also obvious
in the most standard case in which the sets Hi and are linearly
independent. In Corollary 1 to Theorem 5.7.1, we will also show that RD4
always holds when dimF is finite. In any case, if 2)| is a set of root data,
when Wi is its Weyl group, 2 i its set of roots, and its set of coroots.

Proposition 9
Assume that S)| is a set o f root data and let Wi be its parabolic subgroup of W.
Then

(i) S i = 5 , for all w G W,;


(ii) 11 and I coincide on \
(iii) if *rj^ • • • rj^ is a reduced word for w g PLj in W, then g I.
416 The Weyl Group and Its Geometry

Proof. Let IV e W^. It is obvious that /i(w) > /(w). Let

S i :=

Clearly 5^ c 5^. We now apply Proposition 3 to conclude that

/i(w) > /(w) = card 5^ > card 5^ = /i(iv).

This proves parts (i) and (ii). Let w be as in part (iii). By Proposition 3,
aj^ ^ 5^ = 5^, and hence aj^ = > 0. By WIP, j\ e I. Induction
on n now completes the proof. □

Given I c J as above, we define a subset IT ^ of IT as

W^:= [ w ^ W \ wa,. e 2 + for all i e l).

By convention IT^ = IT.

Proposition 10
Let w G IT. There exist unique elements iv^ e IT* and W2 ^ ITj such that
w = WiW2- Moreover /(w) = + Kwf)-

Proof We reason by induction on liw) to show that iv = IV1IV2 for some


Wi e IT* and IV2 ^ with /(w) = /(wi) + /(1V2). If /(w) = 0, then w = 1
and 1 = 1.1 is as desired. In general if w e IT* we are done, since we can
write w = Wj. Suppose that w ^ IT*. Then there exists i e I such that
wa, ^ X_; that is, a, e S^-i. By Proposition 4,

l(wr^) = /(r^iv“ *) = /(w“ *) — 1 = /(w) — 1.

By the induction hypothesis


wr,- = IV1W2.

where iv^ e IT*, W2 ^ + ^(^2) + 1 = Kiv), and therefore iv =


W1IV2''/’ where e IT* and 1V2''/ ^ as desired.
We now show that the decomposition is unique. Suppose that
w = W1W2 = W1W2,

where iVi,Wi e IT* and iV2,W2 ^ Then

Wj = W1IV2IV2

If 1V2W2 * ^ 1, then w'2^2 ^ S i _ c 2_ for some i e I (Corollary 1 to


5.2 The Length Function 417

Proposition 3). Then

e w'iSi_c X_

contrary to e Thus W2W2 ^ = 1 and hence = w[. The proof of the


proposition is now complete. □

Proposition 11
Let I c J be arbitrary. Let a e X, and suppose that in natural form {see
Section 5.1), a = Xy^j^y^y Then

a e Xi <=!►Cy = 0 for all j ^ I.

Proof (<=): We maintain the notation of Section 5.1. By assumption ==


Replacing a by —a if necessary, we can assume that
We reason by induction on the height of p to show that ¡3 •= WiB^. If
htijS) = 1, then = jS, for some i e I. Suppose that ht()8) > 1. Write

yej

where dj > 0 [see (4)]. Then by (5)

2 = (p,p''}=

Since < 0 whenever 7 ^ I we conclude that <)8, > 0 for some


i e I. Then ht(r^j8) < ht()8) so that r^,^ by the induction hypothesis.
Thus p e rp^^Ai as desired. Finally,

a = ilf(p) G i/r(^"A,) = = PT,n, = X,.

(=>): Conversely, if a e Xi, then a = wa^ for some i e I and w g IF|. Let
^ * * * '*a,y «zV • • • ’ ^ ''^/1 ' * * T h e n P := =
^ k ^ i^ k ^ k is the natural expression of a. □

Theorem 12
Let A = {A¿f) be an I X I Cartan matrix of finite type. Let 3 ) =
(A, • , • }) be a set o f root data over R for A with dim V =
n < 00. Let X and be the roots and coroots for 3 . Then card(X) and
card(X^) are finite. In particular X and are finite root systems with
associated Cartan matrices A and A ^ respectively.
418 The Weyl Group and Its Geometry

Proof. Since A is of finite type, the Coxeter-Dynkin diagram of A lies in the


list of Theorem 3.5.4. In particular A is symmetrizable, say, by e =
diagicj,. . . , ej), where > 0 and As is positive definite (Proposition 3.6.9).
Since A is nonsingular, II = { a ^ , i s a linearly independent set.
Define (* I • ) on A" := c K by (a, | aj ) = A^jSj, so (• | • ) is positive
definite on It is immediate that each reflection r, is an isometry of
(• 1• ) (see, for example. Proposition 4.4.6), and hence S = WYi
U • • • U 5^ when Sj is the sphere of square radius (aj \ aj) in X. Since Sj is
compact and 2 c is closed and discrete, Sj n 2 is finite, and hence
2 is finite.
In precisely the same way 2 ^ is finite. Of course we can identify V ^ and
the dual space K* of K by < • , • >, and then we easily see from 2 c 2Za^,
2^c2Za^^ that e Z for all a , e 2. Also ( a , a ^ ) = 2 for all
a e 2. Then 2 and 2 ^ are seen to be finite root systems (Section 3.2), and
hence II is a base for 2 in the sense of finite root systems (Section 3.3). Thus
A is the Cartan matrix of 2. Similar remarks apply to 2 11 and A^.
This completes the existence part of Theorem 3.5.4. □

5.3 COXETER GROUPS AND THE EXCHANGE CONDITION


Let 5 be a set, and let oo be a symbol. A Coxeter matrix (on the index set S)
is a matrix R = (M^^O indexed by 5 X 5 satisfying

2>2 Li if s s'.
m ss' for all s,s' ^ S
= 1 for all s ^ S.

Our intention is to attach to the pair (5, R) a Coxeter group. This is done as
follows.
Let G = G(5) be the free group on the set S (see [Ja2]). We recall that a
typical element of G can be written uniquely in the form

cj?2 . . .

where for all i we have 5, e 5, e Z \ {0}, and


Let H be the normal subgroup of G generated by all elements of the
form
53 Coxeter Groups and the Exchange Condition 419

such that ^ (note that oo ^ Z+). The Coxeter group C(5, R) of the
pair (5, R) is now defined by

C ( S , R ) := G/ H.

A group C is said to be a Coxeter group if C = C(5, R) for some pair (5, R)


as above.
Remark 1 If = oo, then ss'H is of infinite order in G/ H. This justifies
the choice of symbol oo. Since = 1, ( sH Y = 1/f for all 5 e 5 (i.e., C is
generated by involutions).

The following important result provides us with an endless list of interest­


ing examples of Coxeter groups:

Theorem 1
Let W be the Weyl group o f a set o f root data 2) as above. Let

S = be a set indexed by J.

Define for all i, j e J,

' order o f r jj whenever r^r^ is an element o f finite


order o fW ( see Proposition 5.1.11)
) otherwise.
Let
R := (m ). . .

Then W is isomorphic to the Coxeter group C = C(5, R) o f type (5, R). In


particular W is a Coxeter group.

Proof We keep the above notation. For convenience we write Sj instead of


SjH (i.e., Sj is the image of Sj in C = G/ H) . It is clear that there exists a
(unique) homomorphism
W
satisfying
for all j e J.

Moreover if/ is surjective. Our intention is of course to show that ip is


injective. We do this by following an adaptation of an argument of Steinberg
[St].
420 The Weyl Group and Its Geometiy

Suppose that if/ is not injective. Then there exists an element

( 1) SjJn # 1

such that

( 2) n = 1-

Furthermore we assume that n above is chosen so as to be minimal. It


follows from (2) and Proposition 5.2.4(iii) that n = 2m is even. Moreover
m > 1 (and hence n > 4) because of the way in which C is defined. We can
therefore write
y. *. * jr, = jr. • ••
Jim Jm+2

This shows that r, • • • r,Jm + l is not a reduced word and hence that there
exists a smallest positive integer k < m such that
W := r^*+1 Jm + \

is a reduced word but r, r, • • • r, is not. By Proposition 5.2.4(i) we have


aj^ e 5^. We can now apply Proposition 5.2.3(i) to conclude that

a,Jk = r,
Jk + 1
• • • r.ia,
Jt^ Ji + \J
) for some k < t < m,

Then by Proposition 5.2.7(ii)


/-,Jk = r,Jk + \ • • • r,Jt r,Jt + l r,Jt • • • r,Jk + l

so that

0* r.Jt = r-Jk+l Jt+l'

We claim that k = 1 and that t = m. Otherwise,

Sj. = 5;,.. Jt + l

by the minimality of = 2m. By replacing the above left word by the above
right word in the original word, we obtain

SiJn =S:Jl SjSj


Jt Jt+2,

By applying if/, we conclude that the right-hand side of the last equation
equals 1 (this because of the minimality of n) and hence that Sj^ • • * Sj^ = 1.
This contradiction shows that k = 1 and that i = m, as desired.
5.3 Coxeter Groups and the Exchange Condition 421

It follows that
r,J1 • • • r,Jm ГJ2
: • • * r,Jrrt + l

and that

(3) a,
-^1 = r}2 • • • Г: ( a ,
Jm^ Jm+i^) .

Now Sj^ *• • SjSj^ (otherwise Sj^ • • • Sj^ = 1) and • • • SjSj^) =


\l/(SjSjSj^ • • • SjJSj^) = 1. Hence we can apply all of the above reasoning to
this element to conclude that
r,Jm + 1 = Г:3 *‘ * ГJ:m +2
j

We rewrite this expression in the form

r,J3 r,J2 ГJ3


: • • • ГJm
: + l r,Jm+2 ГJm
: +\ r,J = 1
a

to see that

S:Jm + l S:J m +2 S,Jm + \, Su) = 1-


But
S, -'/71 + 1
S:
-'/7 1 + 2
S: ,
•'/71 + 1
S,J Ф 1.
a

Otherwise, if we multiply on the left by we obtain

5,J m +2 = 5,-J\ S:J S:J 3 a


s,Jm+\ ■
"
m + 2 symbols m symbols
and we can then substitute the right-hand side of this expression into.our
original word 5 S,J2m to shorten it, thus contradicting the minimality
of n.
Here then is a nontrivial word 5,J3 5,J2 5,J3 S:Jm. . .+. .1S,-Jm + 2 S,Jm + 1 S:J O f C o f a

length n whose image under i/r is 1. If we apply the above reasoning to it, we
conclude just as in (3) that

a,
J3
= ГJ:2 ••• ГJrr^
: ( a ,Jm + l )' = a, .
Jl

Thus = У3. Cyclically permuting the product, we conclude that /3 =


= **■ ^ h m - v the same fashion we show that j’2 = j 4 = •• = ; 2m-
Thus our original relation was

(O /y.)” = 1-
422 The Weyl Group and Its Geometry

But then m is divisible by the order rrij^^ of and hence Sj^... Sj^ =
= 1. This contradiction establishes our result. ”□

Next we investigate the so-called exchange condition. Let ® =


{A, n , n F, F ^, < * , • ) ) be a set of root data with root system 2 and
Weyl group W = <Ayly e J>. We maintain the notation of Section 5.1.
Suppose that w = is a reduced word of W. If + i e J, then by
Proposition 5.2.4 either

Krj„^w) = l ( w ) + 1

or

=l(w) - 1 .

In the first case r, r, • • • r, is a reduced word for r, w, while in the


second it is not. The exchange condition allows us to find a reduced word for
0„+i^ when Krjw) = /(w) — 1. More precisely it says that there exists some
1 < k < n such that r, w = r,- • • • r, ' ’ r , , where, as usual, " means
deletion.
We introduce a change of notation that will avoid the use of so many
double indices. (The new notation also parallels the standard notation
employed in other texts when dealing with these topics.) Define

S := (o 1y e J} = [ r j a e n}

and

T '= [ wsw M *5 e 5 and w ^ W).

With this notation every element of W is of the form s„ for some


5i,. . . , e 5. Note that by Proposition 5.2.7,

r= { rjaex } .

We have already seen that the map

given by
a ^

is a bijection (ibid.).
5.3 Coxeter Groups and the Exchange Condition 423

Proposition 2 (Strong exchange condition [Ve2], [Del])


Let w = ^ W, and let t = ^ T, where a e

(i) The following conditions are equivalent:

SECl: w~^a e 2_,


SEC2 l{tw) <l { w) .

(ii) If either of the conditions (i) hold, then

SEC3: tw = Si ' ' ' ' ' ' s^ for some 1 < i < n.

(iii) If Si • • • is a reduced word, then SEC3 is equivalent to SECl and


SEC2. Furthermore the value o f i appearing in SEC3 is unique,
(iv) I f t Ei S and Si • • * is reduced, then the word 5^ • • • 5^ • • • o/
tw prescribed by SEC3 is reduced.

Proof We first show that SEC2 SECl => SEC3. Assume SECl. Let 1 < i
<n h t such that s^_i • • • SiU e 2+ and SiS^_i • • • Sia e 2_. By Lemma
5.2.1(ii) we have s^_i • *• Sia = aj, where ; e J is such that s^ = rj. It
follows that *• * s^_iSiSi_i • • • 5^ (Proposition 5.1.10) and hence that
r^w = Si • • • s¿ • • • s„. Thus SECl => SEC3. This argument applied to a
reduced word of w shows that SECl => SEC2. If we now replace w by r^w,
we obtain from SECl => SEC2 that

w V„a e 2 _=> l(r„r„w) < l(r^w),

or equivalently that

l{r^w) < l(w) => w E 2 _,

which is precisely SEC2 => SECl. This establishes parts (i) and (ii).
If is reduced, it is clear that SEC3 => SEC2 and hence that all
three conditions are equivalent. In addition, if

^1 *• • s,. • • • S„ = 5l ■■■ Sj ■■■ S„,

where i < j, then • • ■ Sj = Sj ■■■ Sj_i. Thus

*^1 *** ‘^/‘^/+1 • • ■■■ S„ = s^ ■■■ Si ■■■ Sj ■

which is impossible since Si • • • is reduced. This finishes the proof of part


(iii). Finally part (iv) follows from /(^vi^) < l(w) => l(sw) = l(w) - 1 (see
Proposition 5.2.4). □
424 The Weyl Group and Its Geometry

Remark 2 The implication SEC2 => SEC3 is referred to in the literature as


the strong exchange condition and as the exchange condition in the case
when i e 5. The exchange condition was first established by H. Matsumoto
and the strong exchange condition by D.-N. Verma. The term “exchange” is
due to the following: If ts^ • • • then

tSi • • • Si_i = Si • • • Si,

which says that we can “exchange” t on the left with s, on the right of
5i • • • Si_i. Note that then

tSi Si S ^ = Si

and hence we have the corollary:

Corollary 1
I f s ^ S and Ksw) < liw) then w can be written as a reduced word of the form
SS2 • • •

Corollary 2
Let w = Si • • • e PT. Then there exists 1 < ¿i < ¿2 < ' " < n such that

and Si^ • • • Si^ is a reduced word for w.

Proof If Si • • • is reduced, the result is obvious. Otherwise, there exists a


largest j such that Sj+i • • is reduced but SjSj+i ' • s^ is not. By the
exchange condition, we have

•^1 *’ * ~

where j < i < n. By repeating this argument, the corollary follows. □


Remark 3 Let 5 be a subset of a group G. Assume that S generates G and
that the elements of 5 are of order 2. Since 5 generates G, there is an
obvious concept of a length function. Suppose that the “exchange condition”
holds. Thus, if Si, , . . , s^ e 5, then

/(^1 ••• <1(S2 ••• =>^i 5/ s„ for some i.

It can then be shown that G is a Coxeter group. (This was first shown by H.
Matsumoto [Ma]; see also Bourbaki [Bo]). Therefore, if a group G is
5.4 The Bruhat Ordering 425

generated by involutions, G is a Coxeter group if and only if G satisfies the


exchange condition.
Remark 4 It is known [Mt] that in a Coxeter group G = (C, S) any word
S1S2 ' " that equals 1 can be reduced to 1 by successive applications of the
following type of substitutions:

Replace by •••
m,, terms
Replace by

Remark 5 The Coxeter groups that appear as Weyl groups are precisely
those for which the entries s s' of the Coxeter matrix are 2 ,3 ,4 ,6 ,00.
The restriction that e {2,3,4 ,6 ,00} for all s s' is called the crystallo­
graphic restriction. Thus Weyl groups are, as Coxeter groups, precisely those
that satisfy this crystallographic condition. The strong exchange condition is a
theorem that applies to all Coxeter groups. However, to prove this one needs
to introduce a more general form of root system introduced by [Deodhar].
For a recent treatment of this see [Hu2].

Proposition 3 ([St])
Let 2) = • \ • )) be a set o f root data with root system S
and Weyl group W. Let G be the group with the presentation

generators: a: e 2
relations: R^R^R~^ = R r ^i^-

Then the mapping G W defined by R^ ^ r^ for all a is an isomor­


phism of G into W.

Proof See Exercise 5.13. □

5 .4 THE B R U H A T ORDERING

Let W be the Weyl group of a Kac-Moody Lie algebra. In this section we


define and study a partial ordering on W called the Bruhat ordering (or
perhaps better the Bruhat-Chevalley ordering). The Bruhat ordering has
numerous applications to the study of Weyl groups, weight systems, flag
varieties, Verma modules, and combinatorics. Although we treat only Weyl
groups here, the results in this section apply to all Coxeter groups [Hu2].
As in Section 5.3 define

5 — {r„ I a G n }
426 The Weyl Group and Its Geometry

and

r== и = {r„ 1«

Define a partial order < in Ж as follows: Given w and w' in Ж set >v < w'
if and only if either w = w' or there exists a sequence Wq, w ^, ... of
elements of W such that

BOl: Wq = w and = w',


B02: Kw^) > /(vv^_i) for all 1 < i < p,
B03: w¿ = with e T for all f = 1 ,..., p.

A sequence Wq, w ^, . •., as above satisfying B 01-B 03 is called a chain


of length p joining w to w'. In this situation it is convenient to write
w' = tp ... t^w and call this a chain of length p joining w to w\
It is rather easy to see that < is indeed a partial ordering, the Bnihat
ordering in. The following diagram depicts the Bruhat ordering for the Weyl
group of type C2.

(r a^'r a2'

T«1T«2r«1 r
0^2r «1r «2

r
«1r «2 r
0^2r «1

“1 «2

For example, to see that

r
O'!r «2 —
< r r
<^2r
notice that
5.4 The Bruhat Ordering 427

Remark 1 Since => e it is clear


that B03 could be replaced by
B03': with e T for all ¿ = 1 , , n.
Thus the “left” and “right” Bruhat orderings are the same.
We record the following simple facts. Our approach to the Bruhat order­
ing follows Deodhar [Del].
(1) Ifw G Wand t ^ T, then tw < w l(tw ) < l(w ).
Indeed l(tw) and /(w) have different parity (Proposition 5.2.4).
(2) w<l<=>)v=l;l<w for all w ^ W,

Lemma 1
L e t w ,w ' and s ^ S, If w < w ', t h e n e it h e r s w < w ' o r sw < s w '.

P ro o fWe may suppose that w ¥= w'. First suppose that tw = w' for some
t Ei T. If s = t, then sw = tw = w', so that we can also assume that s ¥= t.
We want to show that sw < sw'. Since (sts)sw = sw', it will suffice to show
that l(sw) < l(sw'). Suppose not. Then l(sw) > l(sw') = l((sts)sw). Let w =
be a reduced word of w. Then ss^... s^ is a word for sw and by the
strong exchange condition there exists some 1 < / < n such that stsss^ • *• s^
= • • • Si_i. (Note that because s t the right-hand side is not the empty
word.) It follows that
S tS S S ^ ••• ‘

Thus
w' = tw = Si Si s^.
and this contradicts the fact that l(w') > l(w).
To establish the lemma in general, we let w' = • • • t^w be a chain of
length k from w to w' and use induction on k. Let w" = tj^_^ • • • t^w. Then
w < iv" < w' and by induction either sw < w" or sw < sw". In the first case
we are done: sw < w'. On the other hand, w' = tf^w" so that either sw" < w'
or sw" < sw', and the second case follows. □

Proposition 2 (Subword condition)


Let w, w' E W, and let w' = s ^ ... s^ be a reduced word for w'. Then w <w ' if
and only if w = Si^ *• * for some 1 < • < i^ < n. Furthermore, if
w < w', then the subword S: **• S: can be chosen to be reduced.

Proof. Suppose that w < w'. Let w' = t ^ .. . t^w be a chain of length k from
tv to tv'. We show that tv can be written in the form s, ... s, by induction on
428 The Weyl Group and Its Geometry

k . l i k = 1, then by the strong exchange condition SEC3,


^ • • • 5,. • • •
To establish the general case, we repeat this argument. Conversely, suppose
that ... Si^, We show that h' < w' by induction on the length of w'. If
Z(w') = 0, the result follows from (2). We now distinguish two cases

# 1. Then by induction
W= <52 • *• 5„ = S^w' < w' (by faCt 1) .

¿1= 1. Then by induction


5,^ • • • 5,^ < 52 • • • 5„,
and hence by Lemma 1 either w = S:*1 • • • 5,‘■m< 59^ • • • 5„^ < or >v =
5; • • • 5: < 5; 5o *** S„ = W'.

For the final statement use Corollary 2 to Proposition 5.3.2. □

Proposition 3 (Z-lemma)
Let < be the Bruhat ordering o f (IT, 5). Let w, w' e IT, and let s E: S.
Suppose that l{sw) < l{w) and that l{sw') < /(u^'). Then the following condi­
tions are equivalent:

Zl: w < w',


Z2: sw < sw',
Z3: sw < w'
.w'
w

sw

Proof Write w' as a reduced word s^ • • • 5„ with s^= s (Corollary 1 to


Proposition 5.3.2).
Z l => Z2: By the previous proposition w = s^^ • • • 5,^ for some 1 < ¿^ <
• • • < ¿^ < n. If ¿1 = 1, then sw = 5^^ • • • Si^ so that 5w < 5w' = 52 • ** s^
by Proposition 2. Ijf ¿1 ^ 1, then we use the exchange condition to conclude
that sw = 5;, • • • 5; • • • 5, for some j. Now this is a subword of sw' =
52 ... 5„, and hence sw < sw',
Z2 => Z3: The proof is clear, since sw' < w'.
Z3 =» Zl: By Proposition 2, sw < w' => sw = s¿^ • • • s¿^ for some 1 < i^
< • • • < i^ < n. Then w = 55^j • • • 5^^, and this is always a subword of
5. • • • 5„. Thus w < w'. □
5.4 The Bnihat Ordering 429

Proposition 4
Let w,w' ^ W be such that w < w ',w ^ w'.

(i) There exists a chain of maximal length joining w to w'.


(ii) In any chain w = Wq < < ••• < = w' o f maximal length we
have /(w^) = iw¿_f) + 1 for all 1 < i < p. In particular, all maximal
chains are o f length Kw') — l{w).

Proof (i) The proof is obvious from the definition of chain.


(ii) The result will follow if we can show that given as above then

(3) there exists a chain w^, Wj, • • •, joining w to w' such that l{w^ —
= 1.

If /(w') - /(w) = 1, then w' = tw for some t e T because of B02 so that the
result is clear. Thus if the proposition is false, then (3) fails to hold for some
pair w < w' with Z(w') - l(w) = N > 2 ,B y choosing N and l(w') as small as
possible, we may assume that for sll a,b ^ W with a < b wo have

INDl: (3) holds for a and b whenever 1(b) — 1(a) < N,


IND2 (3) holds for a and b whenever 1(b) — 1(a) = N and 1(b) < l(w').

Note that if w = 1, then the result is obvious. We henceforth assume that


w 1.
Let be a reduced word for w'. By Proposition 2 we may assume
that w admits a reduced word of the form w = 5^^... 5,^ for some nonempty
sequence 1 < • < i^ < n.

Case 1. 1. By Proposition 2

IV = • • • 5,-^ < ^2 • • • = W'.

Then INDl implies the existence of a maximal chain joining w to s^w',


and this with one more step gives a chain from w to w' satisfying (3).
Case 2. = 1. Then /(ijw) < l(w) and /(5iw') < l(w'). By the Z-lemma
we have s^w < s^w', and by IND2 we can join SiW to s^w' with a chain.

(4) w -s.w
p-l
430 The Weyl Group and Its Geometry

with /(w^) = /(iv,_i) + 1. Notice that p > 2, since Ks^w') — Ks^w) =


l(w') — l(w) > 2.

There are two subcases

Case 2a. By the Z-lemma

5iW'

we conclude that w < < w'. If w = s^w^, then = s^w, and this
contradicts p > 2. Similarly ¥= w'. Thus l(w) < < l(w') so
that by INDl we can join w with and with w' (hence w with
w') in the desired fashion.
Case 2b. < w^. By the Z-lemma

s^w

so that w < < s^w' < w\ But l(w) = /(^ivv) + 1 = Hence w


< => w = Wj. Thus (4) shows that we have a chain from w = to
w' as required. □
Remark 2 Proposition 4 shows that B02 can be replaced by the stronger
looking

B02': l{w^) = /(iv^.j) + 1 for all I < i <p.

Many authors define the Bruhat ordering in this way.

5.5 MORPHISMS OF ROOT DATA: SUBROOT SYSTEMS

Throughout this section 3 = (A , II, H K, K ^, < * >* )) will denote a set of


root data over K. Let 3 ' = (A \ U \ H ' V', ,• >') be another set of
root data over a possibly different field IK'. We use ' to denote the objects
5.5 Morphisms of Root Data: Subroot Systems 431

of 3 '. For example,

n- - { a » ,.,,.

A morphism of into is a pair (<p, ^ of Z-linear maps

such that

MORI: (p(U) c n ' and c


MOR2: <x, y) = ((p(x),(p^(y)y for all x e Q and y ^ Q^.

The concept of isomorphism and automorphism between sets of root data


is defined in the obvious way. Note that no assumptions are made on the
relationship of V to V' and to F ' In fact 3 and 3 ' can be isomorphic
without V and V' being defined over the same field. We denote by Aut(.^)
the group of automorphisms of 3 ,

Example 1 Let 3 = {A, B, B 1^, 1^*, < • , • » be the universal covering of


3 (see Section 5.1). Recall (see Proposition 5.1.1) the unique Z-linear maps

ilf:Q-^Q and

given by

ijj \ ßj ^ aj and (/f ^ : ßj^ ^ .

Then 3 3 is a. morphism (said to be canonical).

Proposition 1
Let ((p,(p^): 3 ^ 3 ' be a morphism.

(i) For fl// a G n and a' e II',

(p{a) =«'<=> ( p ^ ( a ^ ) = a '^ .

In particular <p(a)^=
(ii) = <<p(a), for all a, e II.
(iii) (p :U ^ TL' and <p^ : II II' ^ are injections.
432 The Weyl Group and Its Geometry

(iv) There exists a unique homomorphism r : W ^ W' satisfying r(r^) =


^<p(a) all a ^ II, Furthermore r is injective, and the diagrams

V J L . g ,v
(1) Q' Q

t(w) and ^ r(w)


1 1
Q Q'

commute for all w ^ W .


(v) (p maps S {resp. S_) injectively into S' {resp. S'+,S'_). Similar
statements hold for cp^ and S

Proof
(i) Assume <p(a) = a'. Then by MOR2, 2 = ( a , a ^ ) = (a', cp^ia'^))'.
Thus [otherwise, <a', <p^(a ^)>' < 0]. The reverse implication
follows along similar lines.
(ii) The proof is obvious from part (i).
(iii) Assume that a, ß ^ U, a ¥=ß, and that <p(a) = <p(ß). By part (i) we
have <p^ia^) = cp^(ß^l
Thus by MOR2,

0> ( a , ß ' ' ) = { ( p ( a ) , ( p ' ' { ß ^ ) ) ' = { < p ( a ) , ( p ( a y y = 2.

(iv) and (v): By Theorem 5.3.1 and Proposition 5.1.11, W is 2l Coxeter


group, and the orders m^ß of the products depend only on
(a, ß ^ ) { ß , a ' ^ ) . Similarly W' is a Coxeter group, and now part (ii) shows
that the map r^ ^ a e n extends uniquely to a homomorphism

T iW ^W ',

Let X ^ Q and a e II. Then

= <p(x - ( x , a ^ } a )
= <p{x) - {x,a^}<p(a)

= (p{x) -{<p(x),(p(ayy<p(a)

Introduction on length shows that the diagrams (1) commute for all w e IF.
5.5 Morphisms of Root Data: Subroot Systems 433

From this it is clear that <p(2) c S', and hence

<p(S+) = <p( E z ^ o « n s ] c ( E n S ') = 2'^,


a'ell'

and similarly <p(S_)cS'_. With this we can show that t \ W ^ W is


injective. Suppose that w w^ and r(vv) = 1. Choose a e n with
]va G S_ (Corollary 1 to Proposition 5.2.3). Then <p(wa) g ^(S _ ) c S'_,
whereas <p(wa) = Tiw)<p(a) = <p(a) e II' and II' Pi S'_= 0 . This concludes
the proof of part (iv).
Finally suppose that cp(p) = (p(p') for some e S. Write = wa and
P' = w'a' for some w,w' e W, Let v •= w~^w'. Then T(u)(p(a') = <p(ua') =
(p(a) by part (iv), and hence by Proposition 5.2.7(ii)

Since T is injective, it follows that

vr^'V ^ = r^.

and hence by Proposition 5.2.7(i), (ii) va' = ± a , so w'a' = ± wa. But w'a' =
-w a is impossible, for then

c p ( w a ) = c p ( P ) = <p{ P' ) = < p { w ' a ' ) = c p ( — w a )

and hence <p(wa) = 0 ^ S'. Thus /3 = wa = w'a' = j8' as desired.


The argument for (p ^ is similar. □

Corollary
// I c J, I =?i= 0 , ¿znd is the subgroup o f W generated by the reflections
r,, i G I, then Wi is the Coxeter group with generators {r^ I i ^ 1} and the
relations (r^ry)'”'-' = 1, where m^j = is given by Proposition 5.1.11.

Proof Let '= (Aj, n ,, IIi'^, F, K ^, < • , • » be the set of root data ob­
tained by restricting A, n , and 11^ to the index set I. Then we have the
obvious morphism of into 3 , and hence by part (iv) of the proposition,
the Weyl group of 3 ^ (which is a Coxeter group) is mapped injectively onto
Wj, □
Remark 1 If the pair (cp,(p^) is an isomorphism of root data, then
(p: S S', ^ ^ : S S ' ^ are bijections and r : IP ^ IT' is an isomorphism
of IP and IP'.
434 The Weyl Group and Its Geometry

Remark 2 Proposition 1 makes it clear that the essential information in a


set of root data is contained in the mai^ix A. _ _
Let K be a field containing K. Let F = and_F ""= tK® The
pairing < • , • > extends uniquely to a (nondegenerate) IK-pairing

< • , • >r : F X K.

For all a e n J e t a == 1 ® a e K ^milarly define a ^ 1 ® a e V for all


a e n'^. Let n := { a l a e n} and n^== ( a | a n T h e n

is a set of root data oyer IK said to be obtained from ^ by extension of the


base field from IK to IK.

Obviously ^ and are isomorphic as root data via a a, a ^ a It


is also quite easy to restrict the field. We begin again with Let F be a
subfield of IK. Let ^ Q ^nd ^ be the K-linearly indepen­
dent sets in V and V ^ whose Z-spans are Q and Q ^ as given by RD4. We
claim that there exists a set M d I and F-linear spaces Fp =
Fp^ = 0^^j^Fj8y^, and a nondegenerate pairing

<-,->F:nxFp"

such that

</3,,i3/>F = for all i , j e I.

For instance, set M == lU I', where I' = {i' 11 e 1} is a set equipotent to I.


Form vector spaces Vf and Ff'^ with bases {j8,}, eM ^^d {/3j}, eM> ^nd define
< • , • >F by

(^¡yPj''>r = <%■>')'/>>
<i3,,/3/>F = <j8,,/3/>F = 5,y,
</3,-, j8/>F = 0 for all i , j e I.

Now set

Q = 0 Z i 3 ,c F F , Q'':= 0 Z ) 8 J c F f^
/e l /el

Then Q - Q, Q ^ - as Z-modules via y-, respectively,


and we can thereby identify II and II ^ as subsets of Q and Q After doing
5.5 M orphism s o f Root Data: Subroot System s 435

this, we evidently have

a/>iF = Aij for all /, j e I,

and we see that

• , -»F

is a set of root data over F isomorphic to


In particular we can construct rational root data in this way and then
by extension obtain root data isomorphic to ^ over any field F of characteris­
tic 0. This allows us in Section 5.6 to use F = R and thus take advantage of
the topology of R" in proving results about root data.

Let be a set of root data with root system 2. A nonempty subset ft of 2


is a subroot system if for all a, e 2,

CK,)S e ft => e ft.

In view of the remarks we have just made, the fact that ft is a subroot system
of 2 is independent of the particular set of root data in which 2 is
manifested.
We study subroot systems in Section 5.7.

Proposition 2
Let .^ = (.4, n , n ^ , K, < • , • >) and = { A , ft', F', F ' \
be isomorphic sets of root data over fields IK and K' via a pair of
maps (cp,(p^) for which the corresponding isomorphism o f their Weyl groups is
T\W ^ W \ Then w EiW is a reflection in the space V if and only if rw is a
reflection in V'.

Proof Let Q q (resp. ^ q) be the rational span of Q in F (resp. Q in F')-


Clearly <p extends to a Q-linear isomorphism ^ : (2(q ^ i2oi-
From RD4 we have the natural IK and IK' linear isomorphisms

( 2) <S) <2q = C k C K; ® Q q - Q k' ^ V'-


Q Q

Let w G IF be a reflection. By restriction to Q q we obtain the nontrivial


(by the Corollary of Proposition 5.2.3) involution w on Qq. Now w has
eigenvalue - 1 with multiplicity exactly one on F and w has - 1 with
multiplicity at least one on Qq. Because of (2) w has eigenvalue - 1 with
multiplicity one also, and so vv is a reflection (see Section 3.1), say, with a
hyperplane H of fixed points.
436 The Weyl Group and Its Geometry

By means of the diagram (1) we see that = tw \q'^ and hence that
tw \q'^is a reflection in the hyperplane H' —
Let us see that this implies that w' '= tw is itself a reflection of V'. Let

h'^H'

By (2), is a hyperplane of and hence

= K'y e H i .

where y e Q'^r \ {0} is chosen so that w'y = - y .


Let F denote the space of fixed points of w' in V'. We observe that by the
definition of w' as a product of a'j e II' (since w' e W'), one clearly has
w'x G X + for all X G V', Thus writing
X — w'x X + w'x
X = ------------------ + ------------------
2 2

we see that V' c + F, Thus

V' = K'y H i . - \ - F = K'y 0 F,

and therefore w' is the reflection in y with F as its hyperplane of fixed


points. □

5.6 THE GEOMETRY OF A SE T OF ROOT DATA

Let ^ = (A, n , n V, V • » be a set of root data over U. We main­


tain all the notation of the previous sections except that now IK is replaced by
R. Throughout the section we also assume that dimgj(F) = n is finite.
Because of RD2 and RD4 we have

dim,R(F^) = n,
J is at most countable (though not necessarily finite),
Q= and Q ^=

for some {%} c Q and c g ^ linearly independent over R.


Henceforth we will think of both V and V ^ asjopological vector spaces
by identification with R". If 5 c R", then 5^ and 5 will denote the interior
and closure of S, respectively.
In this paragraph we study the action of the Weyl group W on V from a
geometrical point of view. The principal object of interest is a certain convex
cone 3£, the Tits cone, in V. The Tits cone is decomposed into a collection of
5.6 The Geometry of a Set of Root Data 437

smaller cones by a set of hyperplanes H, one for each positive root of the
root system 2 associated to 3 , The Weyl group operates so as to leave 3£
stable and to permute these smaller cones amongst themselves.
Since V and V ^ can be interchanged by replacing ® by ® there is
another Tits cone X ^ in F Both are used in the theory of contragredient
Lie algebras. (Warning: Some authors use X and X^ with exactly the
opposite meanings to ours. We prefer to keep “checked” objects in V ^.)
Much of what follows concerns hyperplanes of V. Any such hyperplane is
of the form

/ / = {x e V \{ x, h ) = 0},

where /i e F It decomposes V \ H into two open half-spaces

i / ± = [x e V \ ( x ,h ) e ±R^}.

Two points X, y e V \ H are on the same (resp. opposite) side of H if x , y


lie (resp. do not lie) in the same one of these two open half-spaces. H
separates x and y if x and y lie on opposite sides of H.
If h,h' ^ V ^\{0} define the same hyperplane H, then h! = ah for some
nonzero e K. The reason is that if V = H ® Ru and a is defined by
( m, W) = a{u, h), then <jc, h! — ah) = 0 for all x e F.
Given any set H of hyperplanes of F we can define a relation h on
F by X y if andonly if for each hyperplane / / e H either jc, ye / / or x, y
lie onthe same side of H, It is obvious that ^ is an equivalencerelation. If
X, y e F, X y, we denote by [x, y] (resp. ]x, y[) the closed (resp. open) line
segment joining x and y. Thus

= ( tt + (1 - i )yl 0 < i < 1},


= {a: + (1 - i ) y | 0 < i < 1}.

Let X and X be the set of roots and coroots of respectively, and let
JV be its Weyl group. Thus X = WTl, X ^= WU'', and for all x ^ V ,

{wx,h} = {x,w~^h}.

For each a ^ e X define the hyperplane

/f„v:= { x e = 0}.
438 The Weyl Group and Its Geometry

Since {a, a ^ ) = 2, w e have

V = Ra 0

Clearly is a reflection that interchanges the two half-spaces and has


V as its hyperplane of fixed points (Section 3.2).

Note that H^v= a^= (Proposition 5.1.6) <=> r ^=


±jS ^ (this last by Proposition 5.2.7). Set

H ==

The equality

{w~^x, = <jc,)va^> for all jc e K, w e W, 2^,

shows that

wa

and hence that W acts as a group of permutations of H.


The fundamental chamber (relative to H ^) is defined as

f := |x € V\(x, a / > > 0 for all ; e j | .

For each subset of I of J define the set

Fj = (jc e F |< jc ,a /> = Oif / e l a nd ( x , a ^ ) > Oif i e J \ l } .

Note that F = Each of these sets Fj is convex, and each is a cone in the
sense that
X G Fj => (R+x c F ,.

Since the elements of n ^ are not necessarily linearly independent, there is


no guarantee that Fj ^ 0 . Notice, however, that

Oeif^==
¿ej
5.6 The Geometry of a Set of Root Data 439

Given two subsets A and B of V, we say that A supports B (and that B


supports A) if A and B have the same affine span.

Proposition 1

(i) # 0 . Moreover F °= F.
(ii) For all k ^ J we have that F^|^^ spans, hence supports, the hyperplane

(iii) F = U , c j ^ i -

Proof, (0 Let y / be as in RD4. The set A = {x V \ ( x , y n > 0 ,


! < / < / } is open and nonempty. Since 2 + c ©/ = ; it is clear that
A c F. Thus F° 0 . We claim that for all 0 < i < 1 we have iF + (1 -
t)F^ c F^. Since the^eft-hand side is open, it will suffice to show that it lies
in F. Now, if X ^ F, then <jc, a /> > 0 for all j G J. Let y e F^, and
0 < i < 1. Then (tx + (1 — t)y, aj^) > 0 for all j G J. This establishes our
claim. Finally, if a : g F and y g F®, then ]x,y] c F° so that x g F®.
(ii) Let B be an open ball in F. For each x E: B, rj^x satisfies the
inequalities
<r^x, a / > > 0 j ^ k.

The same goes for every point of the interior U of the convex hull of
B\jvj^B, Now U riH^v is a nonempty subset of ■f(*} from which part (ii)
follows. _____
(iii) Let X G Fp If_y g F, then ]x, y [ c F . Thus j c G ] x , y [ c F . This
shows that U i c The reverse inclusion is clear. □

Proposition 2
For each subset I c J for which Fj ^ 0 , Fj is an equivalence class of jj.

Proof Let a : G Fj, and let x denote the equivalence class of x in V,


Evidently X c F|, since Fj is defined by a subset of of the relations defining
X, Now suppose that y g F j, and let a ^ G 2"^. Since //^ v = we can
suppose that a ^ g and write a Ey^ with Cy > 0. Then <jc, a
= Eye ^ / ) ^ 0 with equality if and only if ; g I whenever Cy # 0. It
follows that <y, > 0 and <y, = 0 <=> <x, a^> = 0. Thus either x
and y both lie in or both lie in This proves y x. □

Recall the parabolic subgroups PT|, I c J defined in Section 5.2.

Proposition 3
Let I, I' c J, and let w e W, I f wFj n FJ =?^= 0 , then I = I', wF| = F^> = Fj,
and w e W^. Moreover w fixes F| pointwise.
440 The Weyl Group and Its Geometry

Proof. We use induction on the length Kw) of w (relative to {rfi e J}). If


l(w) = 0, the result is clear. Suppose that /(w) > 0. For any i e J, we have
the following chain of equivalences:

l{r^w) < l{w) <=» w e 2^ (Proposition 5.2.4 applied to 2


« e 2:
> 0 for all X e F
<=> {r^wx, a f ) > 0 for all j: e F
<=» c

( 1) <=> wF c H~y.

Thus l(r¿w) < l(w) => wF Cl H~y gives

(2) ^ ^ C\F

Now suppose that x e jv F jJl Fy # 0 . Choose / e J so that Kr^w) < Z(]r).


Then X e wF^ fl Fy c wF C\F a by (2) and Proposition l(iii). As a
result X = r¿x e and x e FJ. Hence x = r^x e r^wFi H Fy, The induc­
tion assumption implies that I = I', r^wF^ = Fy = Fj, and e PPj. How­
ever, X G H^y n Fy <=> Z G I' = I. Hence iv g PFj and wFj = F|. That w fixes
Fj pointwise follows from w ^ □

Since W permutes the set of hyperplanes H, W preserves the equivalence


relation and hence permutes the equivalence classes of ^ h - particu­
lar each of the sets wFj, I c J, is an equivalence class if it is not empty. We
call these particular equivalence classes facettes. The union of all the facettes
is the Tits cone

3E := U { i v F i | w G P T ,I C J } = \j{wF\w G W}.

Since each Fj is a cone, it is evident that 3£ is indeed a cone. Furthermore 3£


is PT-invariant. Since F^ ^ 0 , the interior of 3£ is nonempty. If two
facettes vwFj and w'Fy intersect nontrivially, then by Proposition 3, I = I'.
Thus each facette has a well-defined subset I of J associated with it, called
the type of the facette. The facettes of type 0 , which are of the form
wF0 = wF, are called the chambers of 3£. These are precisely the facettes
that have a nonempty interior. The facettes of type {/}, with i g J, are called
faces of 3E.
From Proposition l(ii) the face

^0) = ^ V \ ( x , a y y = 0, > 0 fo r; G
5.6 The Geometry of a Set of Root Data 441

dearly supports (some authors say “spans”) the hyperplane Thus every
face supports a hyperplane, and it is elementary that faces are the only
facettes supporting hyperplanes. _
The faces / e J, all lie in F. Indeed these are the only faces lying in
f , for if X ^ F, then a: e Fj for some I c J. If also x e wFy^ for some
w ^ W and j e J, then by Proposition 3, {;} = I and wFy^ = Fyy The faces
lying in wF for any w are then {wF^^^\i e J}. These are called the faces of the
chamber wF. The hyperplanes wH^y supported by these faces are all called
the walls of the chamber.

Proposition 4
Let the notation be as above, then

(i) 3£ = {a: e V\{x, a ^> < 0 for only a finite number o f a ^ 2+},
(ii) 36 and 36^ are convex cones.

Similar results apply for 36

Proof (i) Let 36'Jbe the set in question. If y e 3£, then y = wx for some
w and X ^ F. Let a; e 2+. Then by Proposition 5.2.6,

<y, a^> <0«=> (x,w~^a"^y


<=> w~^a e 2_<=> a e 5^.

Since is finite, this shows that 36 c 36'. Conversely, suppose that x e 36'.
Let Sj^ = {a e 2+Kx, a ^ ) < 0}._We show that x ^ 3i by induction on
card (Sjf). If Sj^ = 0 , then x ^ F (z X. If 0 , then there exists j ^ J
such that ( x, aj ^ ) < 0. ByProposition 5.1.6 it is easy to see that =
rj(S^ \ {aj}). By induction rjX e 36, and hence x e /^36 = 36.
(ii) Let X, y e 36. For 0 < A < 1 and a g 2+, the inequality

<Aac + ( 1 - A ) y , a ^ > < 0

holds only if ( x , a ^ ) < 0 or <y a <0. By part (i) there are only finitely
many such a, and then by part (i) again. Ax + (1 - A)y e 36. This proves
that 36 is convex. It follows at once that 36^ is convex. □

Proposition 5
Let x ,y G 36. The following are equivalent :

(i) There exists w such that both x and y belong to wF.


(ii) There exists no hyperplane o f H separating x and y.
442 The Weyl Group and Its Geometry

Proof, (i) => (ii) The proof is clear.


(ii) =►(i) Consider the open interval Jjc, y[. We claim that if p, q ^]x, y[,
then p q. If not, we may assume that there exists / i e H such that either
(a) p e and q e H~
or
(b) p ^ H and q e H~,

and it follows in either case that H separates x and y :

Since ]x, y [c X, it follows that ]jc, y[ is included in one equivalence class of


3e. Thus
]x, y[ c wFi for some w ^ W ,1 (z J.
Hence [x, y] c wF^ = wFj c wF. □

Proposition 6
Let ;c, y e 36.

(i) There exists only a finite number jt(x, y) of hyperplanes o f H separating


X and y.
(ii) I fy = wx for some w e IT, then tt(x, y) < /(w) with equality ifx lies in
a chamber.

Proof. Recall that the elements of H are in one-to-one correspondence with


the dements of X+. Since W stabilizes both 36 and H, we may assume that
X e F and y = wz for some z e F and e PT. Let a X + . Then
separates x and y
^(x,a'^)> 0 and (y,a^)<0
<=><x,o:^>>0 and <z,iv“ V ^ > < 0
<=>(x,a^)>0 and
<=><A:,a'^>0 and a e 5^,
Both parts (i) and (ii) now follow from the fact that card S^ = l{w). □
5.6 The Geometry of a Set of Root Data 443

Proposition 7
Let x , y ^ di. Then the closed interval [x, y] lies inside the union of finitely
many facettes.

Proof For two points p and of 3E to lie in different facettes, they must lie
in different equivalence classes. Thus there must be a hyperplane if of H
that either separates p and q or passes through one of these two but not the
other. With this observation in mind let us analyze the closed segment [x,y].
Let Zi,. . . , be the consecutive points at which the hyperplanes (finite in
number) separating x and y cut [x, y]:

By the above observation each of the open intervals ]x, z^[, ]z^, Z2L .. •, ]z^, y[
lies in a single equivalence class. The proposition now follows. □

Proposition 8
Let jc e 36^. The set

:= {if e HU e if}

is finite.

Proof Let y e 3E®\ U (This is possible since U / / e n ^ is of the


first cat^ory.) Then y x. Let B be an open ball centered "^at x whose
closure B lies inside X®. The line through x and y meets the frontier of B at
points p and q. Then p and q belong to but not to any if e Also if
if e Hj,, then i f separates p and q, since x = |( p + ^) e if. By Proposi­
tion 6, Hj, is finite. □

Proposition 9
Let di be the Tits cone, and let F be its fundamental chamber

(i) If w and l ( Z j are such that wF^ C\ F 0 , then w Πand w


fixes FI pointwise;
(ii) I f w EiW fixes jc e 3E, then w pointwise fixes the facette containing x\
444 The Weyl Group and Its Geometry

(iii) W acts simply transitively on the set of chambers, and the following
condition holds

FR: I fx G 3£, there exists a unique z e F such that x e Wz,


In particular F is a fundamental region for the action o f W on 3£.

Proof
(i) This is Proposition 3, since F = U Fj.
(ii) Let jc G 3£ be such that wx = x. Let w' El W and I c J be such that
X G w'Fj (this is a typical facette containing x). Then w'~^ww' fixes a point
of F|, and hence all of F| because of part (i). Thus w fixes w'Fi pointwise.
(iii) W is transitive on the chambers by definition. If wF = F = then
vi^ G = {1} by part (i), showing that W is simply transitive on the cham­
bers. FR follows from part (i). □

Proposition 10
Let / / G H, then FT fl spans, hence supports, H.

Proof The reasoning is the same as in Proposition 1. □

The picture we now have of 3£ is that of a convex cone partitioned into a


set of disjoint cones {wFi), the facettes, by the set of hyperplanes //^v with
a^G The facettes with nonempty interiors are the chambers wF, w
and the Weyl group acts so as to permute them in a simply transitive way.
In the most common situation II and II ^ are linearly independent sets in
V and V respectively. Suppose in addition that II spans V, Then II ^ spans
V ^ , since d i mK^ = dimF. Define co,, i g J, by ( cd,, « /> = 6^ y. Then it is
easy to see that F is the cone generated by the simplex
5 := (x G V\x = Lc^iOi, Ci > 0, Ec^ = 1};
that is,
F = {R+x|x G 5} = (x G Fix = Ec,o)^, c, > 0}.

Thus F is a “simplicial cone,” and hence so too are all the chambers.
5.6 The Geometry of a Set of Root Data 445

The face ties I J, are also simplicial cones (of smaller dimension),
namely Fj is generated by the simplex

= (x e Vjx = = 1,
>0 if / ^ I, = 0 if / e l } ’

If we continue to assume that II and II ^ are linearly independent but no


longer assume that II spans K, then we have to take into account the
subspace

n n
This space is pointwise fixed by W, and every facette is stable under
translation by :

vuFj = wFj + .

Consequently 3£ = + H . If we let

V /H ^

be the natural mapping, then W acts uniquely on V /H so as to make a


IT-equivariant map:

w(iJLx) = jjl( wx) for all w e IF , jc e V.

One may then consider the convex cone /¿(3E). It is the disjoint union of
the “facettes” {w/i(Fi)}, The “chambers” are then the simplicial cones
{wii(F)}. We make no further use of this in what follows
If F is a minimal realization of A, then II and H ^ do consist of linearly
independent elements and dim H ^ = corank A. Thus for A of finite type we
can arrange for H ^ = (0) and emerge with the first picture above. However,
for A affine we can never do better than dim H ^ = 1, and the second picture
is appropriate. The examples at the end of this section illustrate some of the
possibilities.
We return to our general development. If T is a nonempty subset of V
define

H y := { / / e H | y c / i } ,

:= Iw e W\wy = y for all y e y ) ,


[ y ] := the subspace of V spanned by Y.
446 The Weyi Group and Its Geometry

Proposition 11
Let Ybe a nonempty subset o f 3£, and let K •= C\ h ^ ^
There exists w ^ W and I c J such that

(i) = ^ w W ^w-\
(ii) =
In particular the three following groups are equal:

(iii) The elements of W that fix Ypointwise.


(iv) The subgroup o f W generated by all reflections r^ such that r^ fixes Y
pointwise.
(v) The subgroup of W generated by all reflections r^ such that Y c

Proof Let yj , . . . , e y be a basis of [y]. Because 3£ is a convex cone the


set {E/ = 7,^A^yjA^ > 0} is included 3£. This shows that 3£ contains a nonempty
open subset of the topological subspace [Y]. Each hyperplane i f e H \ H y
intersects [y] in a proper subspace. Since the number of such hyperplanes is
countable (see beginning of this section), there exists x g [y]ri3£ such that
X ^ H whenever / / e H \ Hy. Thus

(3) y

Let wFj be the unique facette to which x belongs. It is convenient now to


translate the whole problem by \ replacing Y by w ^y, jc by iv ^jc, and
K by x~^K. Then x e Fj, and hence by Proposition 9

(4) = Wi = W'"'.

Suppose that Hy # 0 , let H e Hy, and let a e 2+ be such that H =


Since X G [y], we have < x , a ' ^ > = 0. Writing a ^ = Cja/, c,. in
natural form (Section 5.1) and using the fact that jc g Fj, we conclude that
Cj = 0 if ; ^ I. Thus a^G (Proposition 5.2.11). Conversely, if )8^ g X f,
then <x, = 0, and hence x g FT^v. By (3), FT^vG Hy. It follows that

H y=

Of course, if Ny = 0 , then 1 = 0 , and this last equality still holds. In


particular r, pointwise fixes K for all i e I and hence by (4),

IV'"' c JV^.
5.6 The Geometry of a Set of Root Data 447

Finally, noting that we have

czW ^c:W ^= c JFW =

and whence ^ = W^. Translating back by w gives parts (i) and (ii).

Corollary

I f r ^ W i s a reflection, then r = for some a e 2.

Proof Let H be the hyperplane of fixed points of r. By Proposition 10,


if n 3£ spans H, By Proposition 11, is generated by all the reflections r^
for which H c //^v. Thus r is a product of some reflections r^ for which
H = Hi^v, By Proposition 5.2.7, is unique up to sign, and it follows at
once that r = r^. □

Proposition 12
Let X e 3£°.

(i) There exists an open convex set B in X® containing x and satisfying the
local flniteness condition:

forallH ^H .

(ii) If B <zdi^ is open and convex containing x and if B satisfies (LF) then
for all w ^ W,

wBCiB^0<=^w^

Proof (0 Let / j , . . . , /„ be closed intervals in linearly independent direc­


tions of V lying entirely in 3£® and containing x in their interiors. Each Ij
lies inside the union of finitely many facettes (Proposition 7) and each of
these facettes intersects Ij either in an open subinterval or in a point (since
facettes are convex). Let Zy^,. . . , Zjku) points obtained in this fashion.
Note that if / / e H intersects the interior of Ij at a single point p, then
p= for some 1 < / < fc(;). Let /y be an open subinterval of 7y chosen so
that

Jj
448 The Weyl Group and Its Geometry

and
Zji ^ Jj for all 1 < / < A:(y) whenever Zy, ^ x.

The above discussion shows that for all / / e H,


/f n /y ^ 0 <=>jc e //.

Then the convex hull B of the set /i U • ♦• U is by construction a set


satisfying LF.
(ii) Let B be an open convex neighborhood of x satisfying LF.
Assume that fl 5 ¥= 0 . If ¥= x, then there exists if e H separating x
and wx. Indeed, if not, by Proposition 5, x and wx lie in the closure of a
common_facette F'. There is no harm in assuming that F'czF. Then
x,wx ^ F = Uicj^i* X and wx lie in facettes of the same type and
hence x, wx e F^ for some I. Now Proposition 3 implies that x = wx, which
is false. By assumption then, B and wB lie on opposite sides of H, Thus
wB f] B = 0 . It follows that wx = x and hence that w e □

We now consider the case when A is of finite type in some detail.

Proposition 13
Let 2) = (A , • , • )) be a set o f root data over R. The follow­
ing are equivalent:

(i) The Cartan matrix A is o f finite type.


(ii) There exists w ^ W with wF = —F.
(iii) 0 g 3E^.
(iv) 3i = V.
(v) The number o f facettes is finite.
(vi) The number o f chambers is finite.
(vii) W is finite.
(viii) S is finite.
(ix) X is a finite root system.

Proof, (i) => (ii) if A is of finite type, then 2


base n (Proposition 3.3.6). Since -1 1 ^ is another base for (check the
definition in Section 3.3), there exists an element w ^ ^ W with WqU. -II ^
(Proposition 3.3.5). Now

o F = {x e > 0 for all


= {x e F|<x,vvoa;'^> > 0 for all a II
= {x e F|< - x , a ^ > > 0 for all n'^} = - F .
5.6 The Geometry o f a Set of Root Data 449

(ii) =* (iii) 0 lies in the interior of the convex hull of F and —F (Proposi­
tion 1), hence 0 e 36®.
(iii) <=» (iv) The proof is obvious.
(iii) =» (v) Since 0 e 36® and 0 ^ H for all /f e H, Proposition 7 shows
that H is finite. Let C be a facette, and let x e C. For each //„vS H,
a'^e X+, assign a label A^(x) g {-1-, - ,0} by

i -I- if ( x , 0! ^) > 0,
a::( x ) = - if<x,a^xo,
[o if = 0.

Since A^(x) is independent of the choice of x e C, we can define A^(C) =


A^(x). Now facettes C and C are equal if and only if A^(C) = A^(C') for all
a'^e X+. Since the number of possible labels is finite, so is the number of
facettes.
(v) =» (vi) Since chambers are facettes, the number of chambers is finite.
(vi) => (vii) W is simply transitive on the chambers.
(vii) => (viii) Since if and only if a = +/3, X is finite.
(viii) => (ix) X satisfies the axioms of a finite root system (with ambient
space the linear span of X in F).
(ix) ^ (i) Let us show that II is a base of X in the sense of finite root
systems. It will suffice to show that II consists of linearly independent
elements. Let V be the linear span of X, and let (-| •) be a positive definite
IF-invariant bilinear form on V (see Proposition 3.2.5). Suppose that we have
a nontrivial relation E, s jC,a, = 0, c, g R. If all the c, > 0, then restricting to
the set I for which c, > 0, we see that the matrix (^/y),,;^/ has a null root
and hence decomposes into affine Cartan submatrices. Suppose that (^,y),ysK
is one of these affine components. Then E, eK^<“ / = 0, c, g Z+. This contra­
dicts WIP. Thus we have E, ejC,o;, with mixed signs. We can now finish the
proof by following the argument of Proposition 3.3.1 (part (ii) of the proof),
but for this we must know that (a ,|a p < 0 for all i # ;. To see this, note that
r, is an involution of V that pointwise fixes the hyperplane n V and
maps a, into -a ¡. Since it is also an isometry of (-| • ), it is the reflection

X ^ X ~ 2(x\a¡)/(a¡\a¡)al

Thus for all j we have 2(ay|a,)/(« ,. I«.) = A j i , and since >1 is a Cartan matrk,
it follows that (ayla,) < 0, as desired. We have shown that II is a base of x!
Accordingly A = ((a,^'^))a,;3en is a Cartan matrix of finite type. □

Proposition 14
Let X ^ X. Then x e implies is finite. In particular the stabilizers of
any facette in 36® are finite. I f U. is finite, then x e if and only if IPF}
finite.
450 The Weyl Group and Its Geometry

Proof. Suppose that x e Let B be an open ball with x e B c 3£° and


satisfying LF of Proposition 12. Any chamber with x in its closure meets B,
and it follows from Proposition 8 that the set of such chambers is finite. Since
permutes these, and the action is faithful, is finite.
Next we assume that II is finite, and prove that is finite implies
X G 3£°. We can assume that x e F, for some I c J. We let II, == {a,|i e I},
n , ^ := G I}. Let A i ■= (v 4 ,y ),y G I be the submatrix of A obtained
from I. Then

© , : = ( ^ „ П „ П , ^ F , F ^ < • , • >)

is a set of root data (see Section 5.2 for a discussion of parabolic subgroups).
The Weyl group of S), is the parabolic subgroup IP,. For S), the
fundamental chamber is

F<*> := <y G F|<y, a ^ } > 0 for all i e I }

and its Tits cone is

(5) XO>= U WF®


M 'S » ',

Suppose that IPF) ¡g finite. Since IPF) = = IP, (Proposition 11), it


follows that IP, is finite. Then by Proposition 13,

( 6) = V.

Furthermore, since x G F „ we have <x, a' ' > > 0 for all « ' ' g II''\II,'^ and
since IP,x = X,

<x,wa'">>0 for all w G IP,, n ^ 'X n ,''.

Hence

X G C == { y e F | < y , w a ' ^ > > 0 for all w G I P „ a ''G n '^ \ n ,'^ }.

Since C is defined by finitely many inequalities, there exists an open ball B


about X lying entirely in C.
Let y ^ B. By (5) and (6) there exists a w g IP, with wy g F^*l Thus
<wy,a''> > 0 for all a'^G H,'". Since <vyy,a'^> > 0 for all a g
(because y g B), we have w y g F c X. Thus y e X, and in conclusion
F c X and X G X°. □
5.6 The Geometiy of a Set of Root Data 451

Proposition 15
W acts properly discontinuously on 3£^. In other words, given any pair of
compact subsets U and V of 36^, {w e W\wU ClV ^ 0} is finite.

Proof We may cover each of U and V with a finite number of closed balls
lying in that satisfy the hyperplane condition LF of Proposition 12; that is,
H 0 X ^ H, In particular, by Proposition 8 only finitely many
hyperplanes of H meet each of these balls. It suffices then to prove that if U
and V are closed balls and only finitely many hyperplanes of H meet U and
V, then {w e W\wU f \ V 0>) is finite. Now in this case the labelling
argument of Proposition 13(iii) => (v) shows that U is covered by finitely
many facettes. The same goes for V, The relation wU fl V indicates the
existence of facettes C, C with

cnc/^0 C 'riK ^0, wC = C ,

The same pair can occur for only finitely many w ^ W , since w'C = C w
and IT^ is a finite group (by Proposition 14).

Example 2 Let

^12
A =
‘21

be a 2 X 2 Cartan matrix. Let S) be the root data of A obtained by


considering a minimal realization of A (Example 1, Section 4.2). We illus­
trate the contents of this section for such S).

A of finite type: A minimal realization of A is given by

R =

is a finite root system (Proposition 13), and ij* = Ua^ ^ Ua2 can be
identified with by means of a IT-invariant positive definite bilinear form
( I • ) (Proposition 3.2.5). Under this identification a / = 2 a ',/(a ja ,), i = 1,2.

Figures 3.2 and 3.3 show drawn in the euclidean plane defined by
( I • )• The fundamental chamber F is given by

F = (jc G >0 / = 1,2}

= {x e f|*|(x|a^) > 0 i = 1,2);


452 The Weyl Group and Its Geometry

Figure 5.1.

since { x , a ^ } = 2(jc|a^)/(aJa^). Thus the walls of F are the hyperplanes


[ a j - ^ , and F lies on the same side of as a^. The resulting decomposi­
tion into facettes is now easy to see. The schematic view of the root system of
type A 2 in Figure 5.1 shows how the PT-orbits of the facettes appear.
Facettes with the same letter are in the same orbit. Figure 5.2 depicts the
fundamental region C and its PF-translates for the finite root system of type
^2-

«1

Figure 5.2.

2. A affine: There are two cases


2 -2
A =
-2 2
2 -4
5Cf>: A =
-1 2

In either case in a minimal realization

we have
h* = Ud* ® Rap © R«!,
5.6 The Geometry of a Set of Root Data 453

where d* can be chosen so that

< d * , 0 = l, < d * , a / > = 0.

The null and conull roots are given by

5 = «0 5 = “o 2«,,
and

5 < + «i'" 5 2a^ + < ,

according to the two cases. Since <5, a / > = 0, / = 0,1, we have WS = 8.


Since ( d * , a ^ y = 0 = ( 8 , a ^ ) ,

//„V = Ud* +

Similarly

- n ( ‘i* - y ) + RS.

Picture ]^* with the subspace Ud* + Rag as the plane of the page and R8 in
the direction orthogonal to this plane. Every hyperplane contains R5:
indeed R5 is the subspace

H*= n

defined in the discussion following Proposition 10.


The intersections of the with the open half-space

D^:= >0}

appear like the pages of an infinite book (see Figure 5.3). The fundamental
region F is an open wedge, and

3E = U = D^\JU8.

We illustrate this for the case Identifying Rd* -l- Rao with the quotient
space ^*/R 5 by the mapping

ad* + ba^ + c8 ^ ad* -t- buQ.

The line {x,S'^} = 1 in Rd* + Roio is stabilized by W and is met by the


454 The Weyl Group and Its Geometry

hyperplanes //^v at the points + k / l a ^ , k ^ Z, k odd (see Figure 5.4).


The requisite calculations are

+ - a , \ = d* - I - + 1

''ll + 2^0 j = d* - - « 0 +

d* - «0 - 1 /2 «0 d* d* + 1 /2 «0 + «o

3. A hyperbolic:

2 A^2
A =
All 2

with ^ 12^21 > 4


5.6 The Geometry of a Set of Root Data 455

We simply state some results and sketch the resulting Tits cone 3£ (Figure
5.5). Set a - = ^ 21» ^

/?== {Ua¡' +

be a minimal realization of A, We have

ij* = Rofi + Ua2,

which we coordinatize by (x, y) xa^ + y a 2*


The Tits cone is that part of the third quadrant bounded by the pair of
lines

y —ab ± y/ab(ab — 4)
X 2a

The fundamental region is defined by the inequalities

lx ay > 0,
¿X + 2y > 0.

There are infinitely many chambers that crowd in around the two boundary
lines of 3£.

It is more convenient in this hyperbolic situation to look at —3£ which lies


in the first quadrant. Figure 5.5 illustrates the case for the matrix

2 -3
-2 2

The open region bounded by the two lines

y 3 + \/3
X

is “ 3£, and the region between the lines y = x and 2y = 3x is —F, The
figure also shows some of the real imaginary roots. The integers associated
with these are the root multiplicities.
Up to this point in this section we have looked at sets of roots data over R.
This has allowed us to develop a “chamber geometry” by means of which
important information has been gained. For a set of root data over K 9^= [R
many of the results in this section would be meaningless (e.g., the Tits cone
456 The Weyl Group and Its Geometry

y ---3 +^/3
a

cannot be defined over fields that are not ordered). Yet Section 5.5 suggests
that some statements proved only for U should hold for arbitrary fields. We
summarize some of the most important of these in

Theorem 16
Let 2) = • » be a set o f root data over IK such that
dim^(VX= dim ^iV ^)) is finite,

(i) Let Q q == and let S ■= {x ^ < 0 for only a


finite number o f a For a subset Y o S we have

IV^ ■= [w e W\wfixes Ypointwise} = ( r J ( Y , a ' ^ ) = (0)>.

I f x e 5, there is a unique z e F fl 5 with x e Wz.


5.7 Subroot Systems 457

(ii) The only reflections in V which belong to W are those o f the form r^
with a e 2.
(iii) The following are equivalent:
(a) A is a Cartan matrix o f finite type,
(b) W is finite,
(c) X is finite,
(d) X is a finite root system.

Proof, (See Section 5.5.) We first construct the rational set of root data
and then its realization 3 q , Then 3 and are isomorphic. Now to
we can apply the theory developed in this section. Then parts (i) and (iii)
follow. For part (ii) we use Proposition 5.5.2. □

5.7 SUBROOT SYSTEMS

h&t 3 = (A , n , n F, F ^, < * , * ) ) be a set of root data over IK. In this


section we assume that dimj^F = dimj^F n is finite. We maintain all
previous notation and terminology for 3 , Recall that a nonempty subset il
of X is called a subroot system if for all

The main result of this section is:

Theorem 1
Let 3 = {A, n , n F, F < • , • » be a set o f root data over IK. Let X be
the root system of 3 , and let Ci czX be a subroot system. Then there exist a
Cartan matrix B and a subset a o f X+ such that if we define H ^ = {a e H},
then (B,S, H F, F j* )) Is a set o f root data with root system Cl,

The proof of this theorem will be given later as a consequence of a series


of preliminary results. Because of the results in Section 5.5 it will suffice to
establish Theorem 1 in the case when IK = R. We will henceforth assume that
this is the case.
Define an equivalence relation ^ on F as in Section 5.6, but now using
the set of hyperplanes e il}. Since c H, we have
^ ^ H^y for all x , y ^ V ,
The set of chambers relative to ii is defined as follows:
:= {C|C is an equivalence class of n and there exists a
nonempty open subset U of V such that J7 c C fl *}.
Note: If F' is a chamber of 3 , then F' cz C for some equivalence class C
of ^ n on F. Clearly C is a chamber relative to fi (Proposition 5.5.1). In
particular, 0,
458 The Weyl Group and Its Geometry

A face of a relative chamber C is an equivalence class of ^ on K that


lies in C and supports a hyperplane In this case we also say that
is a wall of C. Note that if O = S, then all these concepts coincide with
those defined before. The first aim is to prove that walls exist.
Let := <r„|a G n>. Since JV^ stabilizes and we see that ¡V^
stabilizes Given jc, y e K, we let

3^) = card{// e IH separates x and y}.

Lemma 2
Let C and C be elements of The cardinal #^{x, y) with x C, y G C\ is
finite and independent of the choice of x e C and y g C'.

Proof Fix a: G C, and let and y2 belong to C . Let H g If H


separates x and y^, then a: and y^ lie on opposite sides of H. Therefore, if
H does not separate x and y2, we conclude that either y^ and y2 lie in
opposite open half spaces defined by / / or y2 e H. In either case, this is a
contradiction, since y^ n3^2- This shows that the set of hyperplanes in
separating x and y is independent of the choice of y ^ C and, mutatis
mutandis^ for a: g C. Finiteness follows from Proposition 5.6.5 by taking
a : G C n X and y g C' fl

In the sequel we will denote this finite number by C'),

Lemma 3
Let C e and let x ^ C f\?i. Let ^ 3£ \ // e y U
Then there exists an
open ball B about y in x such that # q( a:, y) = #^{x, z) for all z ^ B.

Proof Take B to be an open ball about y in S satisfying LF of Proposition


5.6.12. □

Recall that i f jc , y g K, a: =7^ y , the (open) ray from x through y is defined


by

y) — [x + t{y - x)\t e R>o}.

Lemma 4
Let C El Let_x g C fl 3£, and assume that y e 3£ \ C. I f B is any open ball
about y in 3£ \ C, then the cone of rays ^ ( x ) = \J ¡^^^R(x,b) cuts at least
one face of C in an open subset of that face.
5.7 Subroot Systems 459

Proof. Let Z c 5 be a set of representatives of the rays of ^ ( x ) ; that is,

G 5 => 3z G Z such that R ( x , b ) = R ( x , z ) ,


Zi,Z2 ^ Z; ¥=Z2 R{x, Zj) =5^ R{x, Z2).

If z G Z, the closed interval [x, z] is covered by finitely many classe^of


^ (Proposition 5.6.7), one of which is C (since x g C). Clearly C fl
[jc, z] = [jc, c(z)] for some c ( z ) g [ ; c , z]. Moreover c(z) g X, and there
exists a ^ ( z ) g such that c(z) g H [To establish this last assertion,
note that if c(z) is off every hyperplane of H^, we can consider a ball as in
Proposition 5.6.12 to find points in C further away from x than c(z)].
For each o:^ g let Z ( a ^ ) •= [c(z)|z g Z and c(z) g H^ v). Now
^(x), which is open, is the countable union

= U «Ve^v (.^ (jc) n [affine span of a: and Z ( a ^ ) ]

of subsets of affine subspaces. Thus by Baire category Z(a generates


as an affine space for some a ^ g Let c(zf),.. . , c(z„) g Z(a be an
affine basis of If 5 denotes the open sirnplex with the c(z,) as vertices,
then S is open in H^v, S c H^v fl 3£ and 5 c C. □

Corollary
Walls in exist. Moreover, if C and C are distinct chambers of then
there exists a wall separating C and C .

Proof Let C G (recall that we know that ^ 0). To show that walls
exist, it will suffice by Lemma 4 to show that S \ C ^ 0 . Let x g C fl 3£, and
let a G il. Then g 3£, but r^x ^ C. Indeed, if x g if^v, we can choose a
ball B about rj^x) such that B c H~w.
Finally, if C ¥= C and x g C, y ^ C , then x Because neither x
nor y lie in a hyperplane of at least one hyperplane of separates x
and y. By Lemma 2 this hyperplane separates C and C . □
460 The Weyl Group and Its Geometry

Choose C G to be the unique relative chamber containing F (see


Section 5.6). According to the last corollary C has at least one wall. Let H be
defined as follows:

a '= [a is a wall of C and > 0 for all x e C}.

Note that by the definition of F, a Let

:= < r ja e H>.

Proposition 5
Let the notation be as above. Then

(i) W-s acts transitively^ on Cq ,


(ii) JV^B = n and ft
(iii) « a, )8 ^ ^ e 2 is a Cartan matrix.

Proof, (i) Let C' be a chamber relative to ft. We reason by induction on


#(C, C ) (see Lemma 2) to show that C = wC for some w e W>s-
If #(C, C') = 0 then C = C' by the last corollary. Assume #(C, C') =
> 0. Let U be an open subset of V such that U c C' fl 4. Fix jc e C fl
By Lemma 4 we can find y ^ U such that the open ray R(x, y) meets a face
of C. Let H^v, Of e H, be the wall of such a face. Let [x^,. . . , jc„} e]jc, y[ be
the distinct points at which a hyperplane of cuts [x, y]. We order these
points starting from x as

Let z e]jci, X2L where X2 •= y if n = 1. Then z e r^C (this is because


is the unique hyperplane of going through x^, since x^ lies in a face).
Since [z,y], hence [r^z,r^y] is cut by N — 1 hyperplanes, we have
#(C, r^C) = N - 1. By induction wr^C = C for some w e
(ii) Let )8 e ft. We claim that we can choose x e fl S such that
X ^ H whenever H e From Proposition 5.6.10 we can find a
relatively open subset 5 of with 5 c S . Then 5 ¥= U H
since ft is countable. Our claim then follows.

^Once we have established Theorem 1, it will follow from Proposition 5.6.9(iii) that the action of
on is simply transitive.
5.7 Subroot Systems 461

Let X be chosen as above, and consider an open ball B about x satisfying


LF (see Proposition 5.6.12).

Let 5"^ be one of the open semiballs defined by Now, by the choice of x
and B, B^< zC for some C' e From this we conclude that is a wall
of a relative chamber C'. By part (i) we can write C' = w” ^C for some
w e PFc. Then wH^v= is a wall of C. Since it follows by
definition that ±w)3 e H. We conclude that W-sB = ft.
(iii) Let )8 e H. We claim that is the unique hyperplane of
separating C and r^C. Suppose that separates these two relative
chambers. Then

(1) <.^,7''> > 0, <x,y'^> - <0 for all x e C.


(We may have to replace y by - y if necessary.)
Choose X inside an open subset of V lying in C fl X. Then we can find an
open ball B about r^x with B <^r^C, The interior of the cone of rays from x
to this ball B cuts in an open subset S of Choose z e 5 such that
z^ If we now let x approach z in (1), we reach a contradiction. This
establishes our claim. In particular, if a e H, a then does not
separate C and r^C. Thus for all jc e C.

(2) > 0, and > 0.


If we now let x approach a point of not in we conclude that
<j8, a ^ > < 0, as desired. □

This last proposition contains most of the information needed to establish


Theorem 1. However, we still need a result about lattices to eventually show
that RD4 holds.

Lemma 6
o
Let L be a lattice in R" and C a convex cone. Suppose that the interior C 0.
Then C contains a basis o f L.

Proof Since C contains balls of arbitrary large diameter (being a cone), we


have L n C 0 . Because C is a cone, it follows that there exists e L fl C
such that is primitive. Extend to a basis v ^ of L. For N
sufficiently large, Nv^ + v^ ^ C for all 2 < i < n, and it is clear that
{^1, Nvi + i;2> • • • ? is a basis of L lying inside C. □
462 The Weyl Group and Its Geometry

Lemma 7
Let L be a lattice in IR", and let Cbe a closed cone in IR". Suppose that the dual
cone has nonempty interior C^, Then there exists a basis .,., of L
such that the real cone generated by this basis contains C.
Remark 1 The hypothesis that has a nonempty interior is true [Br] if
Cn - C = (0), i.e., C is pointed.

Proof. Let L* be the Z-dual of L. Then the dual cone to C is

C^-= {jc e IR"|jc.i; > 0 for all u ^ C }.

By assumption 0 . By the previous lemma there exists a basis


, . . . , ¿;*} of L* contained in C^. Let T be the real cone generated by this
basis: r — L"=;lR>oi^f5 and let be its dual. Then

r^ = E
¿=1
where u^} is the basis dual to {i;*,..., u*}. Now T c C‘^=>
C'^'^= C, while {i;f, . . . , y*} is a basis of L* => . . . , i;„} is a basis of L.

Proof of Theorem 1. To the subroot system ft of S we have attached the


six-tuple (B, H, H K, F * ))? which has been shown in Proposition 5
to satisfy RD1-RD3 and is known to have ft as its root system. All that
remains to be shown is that RD4 holds. Lets 7 i , . . . , y/ be as in RD4 (for 3J)
so that
/ /
< 2 = 0 Zy^ and n c 0 I^y..
i= l i= l

Let C be the real closed cone of V generated by {y^,. . . , y/}. Set

<2' = Za (the root lattice of ft),

and let C' — V' fl C, where V' •= U Note that C' is a pointed cone
[i.e., C' n ( - C ' ) = (0)], of F ' and that <2' is a lattice of F'. By Lemma 7, Q'
contains a basis [y\,. . . , y^} that generates a real cone containing C'. Finally
Hc C, and hence S c C'. Thus

Hc ^ L R^or;|ne zy;I = e Ny^.


Similar considerations apply to H
5.7 Subroot Systems 463

Corollary 1
Let 2) = » be a set o f root data over IK. Let II = {aj\j e
J} and let I c. J be a nonempty subset. Let IIj, 2 i, etc., be defined as
in Section 5.2. Then
(i) S i is a subroot system of 2
(ii) 2 ^ is a set of root data with root system Sp

Proof (0 Let a = ua^, f3 = vaj, be elements of 2 i where a^,aj g Ui,u,


V G W^. Then r^p = ur^u~^vaj g W P^i II i = Sj, proving (i).
(ii) We can assume that K = R. According to the theorem there exists a
subset H of 2i+ which serves to construct root data for Sj. Furthermore,
following the proof of the theorem, H is constructed from the walls of the
unique relative chamber C (relative to -- that contains the fundamental
chamber F of 2 . We will be done if we show that H = Hi.
If a G I I I ^hen / / ^ v is a wall of F. This means that some open subset U of
//^v that supports lies in some equivalence class of of F. Then U
serves to show that / / ^ v is a wall of C. Also if jc g F c C, then (x,a"^) > 0
and so, by definition, a g H.
Conversely let a g H c Si+. Let x ^ F (z C. Then x and r^x are sepa­
rated (relative to Si) only by the wall On the other hand the line
segment [x,r^x] = [x,x — ( x , a ^ ) a ] passes through some wall of F, say
I f ' ^ J \ I then ia, oi^y < 0 (since a g 2i+) and then both {x,
> 0 and ( x - ( x , a " ^ } a , a y y > 0, a contradiction. Thus i g I and is a
wall of C by what we have already shown. It follows that H^v= FT^y and
hence a ^ = a / (note, a^G 2i^+). This proves that H c III and hence that
3 = Hi as required. □

Corollary 2
{See also [De 2], [Ku 2]) Let 5 c 2, and let '= {r^\a g 5). Then is the
Weyl group o f a set o f root data. In particular is a Coxeter group.

Proof. Let fl := W^S. Then ft is a subroot system of 2 with Weyl group W^.

Corollary 3
{See also [FTT]) Let A g F and let

2^ : = { a G 2 | < a , A > G Z } .

Then 2^ is the set o f roots o f a set o f root data.


Remark 2 We do not know under which conditions 2 ^ admits a finite
base.
464 The Weyl Group and Its Geometry

5.8 IMAGINARY ROOTS

Let 2 be a root system determined by root data (^,11,11^,


< • , • » . If e ¡2, a e X , and r^cp = <p + ua for some u E: Z, then the root
string between cp and r^cp is defined as

{(p, (p + a , . , . , (p + ua) if w > 0,


[cp, (p — a , . . . , (p ua) if w < 0.

A subset <I> of (2 is said to have the root string property relative to X if

RSPl: X c and
RSP2: whenever <p e and a e X, then [<p, r^cp] c 4>.

There is a unique minimal subset A = A (^ ) of V with RSP relative to X.


In fact, if we define inductively

A o-2,
A„ — {P\P ^ [<p, r^(p] for some a e X, where cp e A„_ i },

then Aq c Ai c ... and A == U A„ has the root string property and is


minimal. Henceforth A denotes this minimal set with RSP relative to X. We
call A the root string closure of X. We denote A \ X by ''"X.

Proposition 1
A is W-invariant.

Proof. We show that A„ is IT-invariant by induction on n. This is clear for


A q . For A„, let p E [(p,T^(p], where <p e A „ _ j , a e X. Let w e W. We have
jS = 9 + ka, where k e [0, u] and (cp, = —u. Then

W'[iP,r„<p] = ^w<p]

= [w<p,r^^w(p] c A„

(by the induction assumption w<p e A„_ i ). It follows that =

W(p + kwa E [w(p,wr^(p] c A„. □

Let ^ be a set of root data as above, and let 3 be its universal covering.
We maintain the notation of Proposition 5.1.1. In particular il/:Q ^ Q is iht
covering map.
5.8 Imaginary Roots 465

Proposition 2
Let A and A be the root string closures o f % and S, and let {A„} and {A„} be
the sets as above that define A and A, respectively. Then

fK = for all n

i/rA = A.

Proof = S = Aq. Suppose that = ^ n - v For A


Q, A ^ Q with i/rA = A and for a g 2, a e S with <pa = a, we have

A A

X + sd A + sa

and i/f([A, r¿A]) = [A, r^A].


Thus, if /3 ^ A„, p G [A, r^A] for some A e A„_i, a e 2, and ^ =
{¡/(p) G [A, r^A] c A„. Conversely, if j8 e A^, then p g [A, r^A] for some
A G A^_i, « ^ 2, choosing preimages A e A„_i, a g 2, we find that p g
[A, r^A] with il/P = p. This shows that (^A„ = A„ for all « g N and hence
that i/rA = A. □
Remark 1 The map ijj'.A A need not be injective. For example, consider
the Cartan matrix

2 -2 -1 4 -3 4
-2 2 -2 -1 4
B' :=
-1 4 -2 2 -2
-3 4 -1 4 -2 2

which, as we saw in Example 5.1.1 is of row rank 3. With the notation of


Example 5.2 the set T = {p^, p^, P2, P^) provides us with root data =
{B', r, r , V , V ^ , ( • I • » . Call the resulting root system 2', and observe that
^0 - 3/3i + 3 p 2 - p 3 = 0. ^
Let with root system 2', be the universal covering of W , and let A
- -
and A be the root closures of 2' and 2'. We know that 2' -> 2' is injective.
..iff - ^ «
However, A A is not injective. Indeed y •= Pi P2 Pa con­
nected support and satisfies <7, p ^ ) < 0, / = 1, .. ., 4. Now this implies that
y G A, as we will see in Theorem 6 below. For the same reason 33f ± n(pQ
- 3pi + 302 — P 3 ) ^ A for \n\ < 10 and all 21 of these roots have the same
image in A.
466 The Weyl Group and Its Geometry

The next proposition shows that imaginary roots for abstract root data are
as expected in the Kac-Moody case:

Proposition 3
If X A for some root system A o f a Kac-Moody Lie algebra, then the root
string closure of 1, is A,

Proof Let <t> denote the root string closure of 2. We already know that
A d '^^A and that A has RSP. Thus c A. We show that A ^ \ ^ = 0.
Similarly A _\ O = 0 .
If A+\ <I> ^ 0 , choose e A+\ <I> of minimal height. Note that ht > 1.
Then by Proposition 4.14 either

1. there exists a e n such that 0 < ht < ht in which case r^p e


so also e <I) by Proposition 1, or
2. for all a e n , <)8, a^> < 0 and for at least one a e II we have
P - a ^ A^. In this case - a e <&, and since (p — a , a ^ ) < -2 ,
P ^ [ p - a,r^(p - a)] c <|). □

Corollary
In the notation o f Proposition 2, ”2 and ( / r - i r 2 ) n A = ' ' ”2 .

Proof We can interpret 2 as the real roots of a Kac-Moody algebra and


hence view ^"*2 as its imaginary roots. Suppose that y e " ”2 and y — ij/iy)
e 2. Using W, we can suppose that y = e II. Writing y = all Cj of
the same sign, we have y = 2cy«y = which contradicts WIP.
Conversely, let y e'"*2. If = y and y e 2, then y e 2, which
contradicts ^"*2 = A \ 2. □

In the remainder of this section we study some properties of imaginary


roots. We will do this only in the context of Kac-Moody Lie algebras.
Let A = be a Cartan matrix, and let R = (1^,11,11^) be a
realization of A, We let Q = A, R) be the corresponding Kac-Moody
algebra over IK. This gives rise to the set of root data

whose root system 2 is the set of real roots '^^A of g and whose root string
closure A is the set of all roots of g.
We will maintain all the standard notation for 3 ,

Given a = 2cyo:y e Q, the support of a, supp(a), is defined by


supp(a) :={; eJlcy #o}.
Clearly supp(a) is a finite set. The set of nodes of the Coxeter-Dynkin
diagram T that is indexed by supp(a), together with all the edges of T that
join the nodes of supp(a), form a subdiagram T(a) of T. The support of a is
5.8 Imaginary Roots 467

connected if r ( a ) is connected (as a graph); (see Section 3.4). Alternatively


this means that for all i, j e supp(a) there exists a set i =
= j of elements in J such that for a\\ k = 1 , n, ^ 0.

The first result is completely elementary but very useful:

Proposition 4
If a ^ then suppia) is connected.

Proof We can assume that a e A+. Then by Proposition 4.1.1, g" is spanned
by elements of the form [e^^,. . . , e^^], where + ••• = a. If supp(o:)
were not connected, then for such an expression for a there would be a
maximum p < k, for which supp(a^ + • • • +«/^) is not connected. Then
[e,y. .. , e,.J = ad , e,J) = 0 (or ad = 0 if p = k - 1),
since e, ] = 0 for all r > p (Proposition 4.1.8). □

Since it is awkward to always distinguish between the nodes of T and the


indices by which we label them, we usually use them interchangeably. Our
definition of the connectivity of supp(a) is an example. Similarly we can
speak of supp(o;) being of finite type, the connected components of supp(a),
and so on.
Let
[a e Q qKo:, > 0 for all i e j},
[a e Q q K«, a ^ } > 0 for all i e j | .
If our base field is R and J is finite, then is the fundamental chamber of
the Tits cone 3£ (See Section 5.6).

Proposition 5
Let a Then

(0 Wa
(ii) Wa n ( — is a single point {p} and ¡3 is the {unique) element of
minimum height in Wa.

Proof, (i) was proved in Proposition 4.1.5(iv). Let p e Wa be chosen of


minimal height (in view of part (i) such elements exist). Then
ht(r^)3) = ht(P - <)S,a/>a,) = ht()8) - <)8,a:,^> > ht()3) for all i e J,
and hence < - /3, a ^ ) > 0 for all i. Thus e
As for the ui^ueness assume j3 and j8', both in A, are such that
j8, G WaC{{ - ^ ) - Then this situation arises inside 3 i for some finite
subset I of J. Looking at the real version of we obtain )3 = from
Theorem 5.6.16. □
468 The Weyl Group and Its Geometry

Define

^conn _ ^ R j s u p p ( a ) is connected}.

By Propositions 4 and 5 the unique point of minimum height in each orbit


Wa, a lies in Thus +c Our main result reverses
this inclusion.

Theorem 6 [Ka4]

Let A be the root system o f a Kac-Moody algebra. Then ''”A^_=


We prepare for the proof of this by a number of results of independent
interest. Suppose that A is indecomposable (hence J is connected). Let
a e (2^, and write a = Then for j e J, (a, a / > = 'Lx^A¿J, Sup­
pose next that a e 0 ^ fl so that

(1) Ex,y4, > 0 for all ; e J.

Then supp(a) = J, for otherwise we can find a e J \ supp(a) with


( a , a k ) < 0. In particular card J < oo. Using Proposition 3.6.5 (applied to
A^, which is of the same type as A) we conclude that either A is of finite type
and at least one inequality in (1) is strict or A is affine, an ^ all the
inequalities in (1) are equalities. In the same way if a e C)(— = R+,
then

( 2) E XiAij < 0 for all j e J.

If J is finite then either A is indefinite and at least one inequality in (2) is


strict or A is affine, and all the inequalities in (2) are equalities. If J is infinite
then letting K — supp(x) we see that (^z7)/,yeKis indefinite or affine and
then A itself is neither finite nor affine.

Proposition 7
Suppose that A is indecomposable.

(i) I f A is of finite type, then R ^= 0 . In particular ^"*A = {0}.


(ii) I f a ^ Q+ C i^cin d (a , a / > ^ 0 for at least one i e J, then A is of
finite type. □

We will need the following result about root strings. We prove here the
minimum that we can get away with. There is a more elaborate discussion of
these facts in Section 6.2.
5.8 Imaginary Roots 469

Lemma 8
Let jS e A, and let a If < 0, then ^ + a is a root.

Proof This proof follows from Proposition 3. □

Proposition 9
Suppose that j8 ^ A+ and for all i g J, /3 + ^ A+. Then

(0 iSe^^A,
(ii) supp(p) is of finite type and is a connected component o f the Coxeter-
Dynkin diagram T of A.

Proof Let K denote the connected component of F containing supp()8), and


let B := (^/y)/,yeK- Note that B is indecomposable, A(^) = AflEyeK^«;?
and j8 e A^_(^) (see Remark 4.3.4). Let be a Kac-Moody algebra based
on some minimal realization of B. Then, using Proposition 4.3.7, the assump­
tion )8 + ^ A for all i e K shows that

« == H Q b ) ■9 | = U ( 9 b _ ) • U ( ^ ) ■ U ( 9 g ^ ) • g g c E 9 |-
=(2+U{0)
(• denotes the adjoint action). Since a is an ideal of uDDg^D
E«eA(B)+9s Proposition 4.3.9. Thus ^+(B), hence A(B), is finite showing
that B is of finite type (Proposition 5.6.13).
If supp(/3) K, then we can find j e K\ su pp( j8) with <)8, a /> < 0.
Then )3 H- ay e A by Lemma 8, contrary to hypothesis. Thus part (ii) is
proved. Since B is of finite type. Proposition 3.6.5 shows that for at least one
0. Repeating this argument, if necessary, on rj
that there is 2lw ^ W with wp e A_. Then by Proposition 5(i), e'^^A. □

Proof o f Theorem 6. It suffices to show that c"'”A+. If A is of finite


type, then R+= 0 (Proposition 7), and there is nothing to prove. Assume
that A is not of finite type. Let a e Choose )8 e A+ of maximum
height subject to a - e 0 ^ U{0}. We assume by way of contradiction that
a - jS e (2+. Write
a = "Lm^ai i e J,
/3 = Ilk ¿aI i ^ J.
By assumption m^ > for all i. We proceed by steps.

1. supp(a) = supp(j8): Otherwise we can find an i e supp(a) \ supp()8)


with i joined by an edge of F to supp(j8). Then <j8, a,'^) < 0, so, by
Lemma 8, /3 + a, ^ A+, contradicting the choice of j3.
Define
E := {/ e J|m^ = A :J .
470 The Weyl Group and Its Geometry

2. 0 ¥=E J: By assumption E ¥= J, Since => p ^ A+,


Proposition 9 shows that if E = 0 , then supp(j8) is a connected
component of finite type in F. But then a e R ^(B ) = 0 , where B is
the submatrix of A defined by supp()8) (= suppCo:)). Thus E ^ 0 .
Let f/ be a connected component of supp(a) \ E

supp(a) has the schematic form E U' E U E

Define

i^U

For i ^ U, (p , a ^ y > 0 by Lemma 8 and (p', a /> < 0 (since p' has
zero coefficients on U), and hence

(3) ) > 0.

Also there is at least one i e U for which the inequality in (3) is strict.
Indeed supp(a) is connected, and some i e U must be joined
by an edge to some vertex of su p p (a)\i7 . Then (p',a^^y < 0 and
( P u , a n > 0 . We have then, by Proposition 7(ii) applied to the
submatrix of A corresponding to i/,
3. U is o f finite type:
Set

i^U
4. <a', a / > < 0 for all i e U: For i ^ U we have <a', a /> = (a -
P , a ^y , since i7 is a connected component of s u p p (a )\^ . Since
a ^ R ^ , ( a , a ^ y < 0. We already have (p, a / ) > 0. Thus (a', a ^ y <
0.

Since U is of finite type and a' is a positive integral combination of the a,,
i G U, step 4 is impossible by Proposition 7(i) [applied to (A^j)^ j^ц], This
finally proves that a e A+. However, also a e => 2a e => 2a e
A+. Thus a and so a e"'”A+. □

Corollary, (i) a e^""A => Qa n i2


(ii) For a G A \ {0}, a e'^^A <=> Za fl A = {a, 0, —a).

Proof (0 We may assume that a Then a = wp for some p e R^f'


If q ^ Q+ with qa e Q_^, then qP e R conn + c A+, and hence qa =
w(qp) Also - q a G^'”A_.
(ii) Use Proposition 4.1.5(i) and part (i). □
5.8 Imaginary Roots 471

Proposition 10
Let A be indecomposable, and suppose that J is finite.

(i) A is o f finite type = {0}.


(ii) A is o f affine type = Z8 where 8 is the null root o f A.
(iii) A is indefinite <=> there exists a e^"*A+ for which supp(a) = J and
{a, a ^ ) < 0 for all i e J.

Proof Since "”A^_= it suffices to examine We use Proposition


3.6.5 (actually applied to A ^) to conclude that

A not of finite type <=> # 0 <=>"'”A ^ {0}.


A is indefinite <=> there exists a e R^^^^ with supp(a) = J and {a, a y ) <
0 for all i e J.

In the affine case R^^^^ can only contain null roots (otherwise, we are in the
indefinite case). By Proposition 3.5.1 there is a unique ray of null roots. Thus
j^conn ^ z^8 , where 8 is defined as in Proposition 3.5.1. Since W8 = 5,
'■'”A+= Z+5, whence ^'"A = Z^. □

Corollary
Suppose that A is a Cartan matrix and that J is countable. Then the following
are equivalent:

(i) ^'”A = {0}.


(ii) Each of the connected components o f the Coxeter-Dynkin diagram
r(y4) is o f finite type or lies in the list
^A ^ . . . O — O — O — O — O - --
A „ 0 — 0 — 0- -
O <= O — O • • •
C„ O => O — O • • •
0
1
O — O — O •••
O

Proof Any root has connected finite support. It follows that ^'"A = {0} if and
only if every subgraph of r(yf) defined by a finite subset of the nodes of
r ( ^ ) is of finite type. In particular, (ii) => (i). Conversely, if ^"*A = {0} and if
there are connected components of T (^ ) with infinitely many nodes, then
there are infinite sequences of nested connected subgraphs of finite type,
each one containing one more node than its successor. But only the finite
Coxeter-Dynkin diagrams of types A, B, C, D admit such nesting, and then
only in the ways shown. □
472 The Weyl Group and Its Geometry

5.9 CONJUGACY OF BASES

In this section we prove that for indecomposable root systems, bases are
conjugate by W augmented by {± 1}. We assume throughout that dim F < oo,
but we do not assume that the basis are finite. The proof uses imaginary roots
as a key ingredient.

Let
{ A , n , n \ V , V \ { • ,• »
be root data with root system S. A subset B of 2 is a base of 2 if with
B=
we have that

is a set of root data with corresponding root system equal to S.


Thus with the notation of the definition we have

5 is a Cartan matrix;
2 = IFsH, where IT5 — <r^|a e H>;
when the base field is U, V admits a fundamental chamber C relative to H
and C® ¥= 0 ; and
n is a base of 2.

It is also clear that

n ^ and H ^ are bases of 2 where 2 ^ is viewed as the set of roots of


the dual set of root data 3 ^ (see Section 5.1).

Example 1 For w g IT,

( ^ , w n , w n ^ , F , F ^ , < • , • >)
is a set of root data (we have to replace the {%}, of RD4 by {w%}.

Similarly
-^ := (^ ,-n ,-n ^ F ,K \< -,-> )
is a set of root data. Thus —wH and wU are bases for all w ^ W.

Let 2 be the root system associated with root data 3 . A subset 5 c 2 is


decomposable if
5.9 Conjugacy of Bases 473

and for all e 5^, 72 ^ ^ 2^ (7 ^7 2 ^ = 0 (<=> {72,71 ) = 0 by Proposition


5.2.8G)). Otherwise, 5 is indecomposable.

Let ^ = U , n , n \ F , F \ < » be a set of root data for 2, and let H


be a base of 2.

Proposition 1
n indecomposable 2 indecomposable <=> Eindecomposable.
Proof. Suppose first that H decomposes into H i U S 2, where { a , ^ ^ ) = 0
for all a e Hi and e H2. Then with the obvious notation we have
W= the elements of and commute, fixes H2 point-
wise and 1^22 pointwise. Thus
X = »^h(H) = »^'h.(Hi )UW^3,(H2)
is a nontrivial decomposition of 2.
Conversely, if 2 decomposes as 2 i U 22, then set Hi == H H 2„ i = 1,2.
If H2 = 0 , then < 2 , 2 ^ ) = (W-s(BiX 2 ^ ) = (0), which is impossible (a e
22 { a , a ^ y # 0).
Since n is a base, the result applies equally well for H. □

Theorem 2
Let ^ be root data as above, and let 2 be its root system. Assume 2 is
indecomposable. I f
isomorphism r o f
H H'and are two bases o f 2, then there exists an
onto 3's' (see Section 5.5) such that t H = H'. Moreover
T can be taken so that it is induced by w or —w for some w belonging to the
Weyl group of 2^.
Before proving this we indicate two important consequences.

Corollary 1
The Cartan matrices o f any two bases o f an indecomposable root system 2 are
equal (up to re-indexing the rows /columns). In particular the Cartan matrix is
an invariant of X. □

Let 2 be the root system of the root data 3 . By an automorphism of the


root system 2 we mean a pair (a, cr e GLiQu^) x GLiQ^^) satisfying

AUTl: 0-2 = 2, cr ^ 2 ^ = 2 \
AUT2: (aa = for all a e 2,
AUT3: { x , y ) = <o-x,o->> for all x y

The group of all automorphisms of 2 is denoted by Aut(2). If cr e Aut(2),


then it is immediate that
474 The Weyl Group and Its Geometry

is a set of root data and a determines a morphism of 3 into Clearly


—\ y and each w determine automorphisms of S, so we can identify
{ —\y} and W with subgroups of Aut(2). Also A u t(^ ) determines automor­
phisms of S, but these are very special since by definition they must stabilize
n and n Such automorphisms are usually called diagram automorphisms,
since they clearly correspond to the automorphisms of the Coxeter-Dynkin
diagram of A.

Corollary 2
Let S be the root system of the set of root data 3 . Then

(i) W < A u tO .\ W n A u t { 3 ) = {!},


(ii) Aut{X) = W y \A u t{ 3 ) if 2 is finite,
(iii) Autfl,) = {±1} X W X A u t ( 3 ) if 2 is infinite.

Proof of Theorem 2. Using the results of Section 5.5, we can assume that our
base field is R. Let F and C be the fundamental chambers for II and H. It
suffices to show that WC n(±i ^) ^ 0 , for in that case, seeing that both wC
and F (or —F) are equivalence classes for wC = ± F for some w ^ W .
The walls then being identical, ± w a = II.
If A is of finite (resp. affine) type, then the interior 3£(II)® of the Tits cone
is V (resp. an open half space of V). In either case 3£(n)^nC # 0 or
3 e( n) ^ n( -C ) ¥= 0 , and hence WF n ( ± C ) ^ 0 .
We can suppose then that A, and similarly the Cartan matrix B of 3>^, is
of indefinite type. By the last proposition, A and B are indecomposable.
Introduce a set in 1-1 correspondence with H, and write

H = {oLi\i ^ J h}-

Let 3 ' and 3 be sets of universal root data for 3-^ and 3 with correspond­
ing root lattices Q and Q. Then using the obvious notation there are linear
maps il/', {¡z for which we have the following picture:

«A'
Q'u Qu Qu

n n
— —
t' 2 t

‘■'"A' ‘■'"A ‘"■a


5.9 Coiyugacy of Bases 475

Note that the surjectivity of and ^'"A ^"'"A is by the Corollary to


Proposition 5.8.3. Choose I c Jg with the following properties:

I is finite and is indecomposable.


B y is indefinite.
El == {«¿1/ e 1} spans (although and may be infinite dimen­
sional, is finite dimensional).

To see that an I exists in the case that is infinite, choose I such that

{«¿}, ei spans (2r,


I is connected (i.e., Bj is indecomposable),
card I > dim Q r + 2.

Then B y can be neither of affine or finite type, for in that case B y would
have dimension either (/ + 1) X (/ + 1) or / X /, where / = dim

Since B y is indecomposable and indefinite, there is by Proposition 5.8.10


an imaginary root y' e'"*A'+ (positive relative to H) such that

supp(y') = I,
(y',a'/y < 0 for all ; e I.

Let =• y e^'^A, and choose y e"'”A with il/(y) = T-


Now by the construction of f',

<0 forall; e J s ,

and hence

<r,a/> < 0 forall; G J s,

with strict inequality on I. However, we claim that

<7, a /> < 0 for all j e J^.

For suppose that <y, a /> = 0 for some ; ( e Jg \ I). Write

y ' = 'Lciá'i, C,. > 0 .


/el
Then

r =E
/el
476 The Weyl Group and Its Geometry

and

0 = <r,a/> = => (ai, a / ) = 0 for all i e I.


¿el

But aj lies in the span of Hi, so 2 = (aj, a / > = 0. This contradiction shows
that —y e C.
Since no hyperplane meets C,

( 1) <y, ^ 0 for all a e 2.

Now y (relative to fl). Suppose that y Choose w ^ W so


that htn(wy) is minimal, so

<wy, <5^> < 0 for all ¿ e f t .

Thus

<)vy, a^> < 0 for all a e n ,

and hence

(wy, a ^ ) < 0 for all a e n

because of (1), and hence —wy e F.


Thus -w y e Ff]y^C. If y we conclude similarly that for some
w ^ W, wy e F n —wC. This is what we wanted to prove. □

For Corollary 2 we note that there is no w e IF with wF = —F (in the


nonfinite case), since then w2 += 2_, contradicting the finiteness of the
length function l(w).
Remark 1 Maxwell’s Demon (Example 5.1) shows that Aut(H) can be
infinite. More precisely one can show that for the Demon, Aut(H) -
A u t( S ) /± W is infinite dihedral.

EXERCISES

5.1 Let be a set of root data and 2 its set of real roots. For each finite
subset 5 of 2 recall <5> = and <0> = 0. A finite subset 5 of
2+ is called perfect if there exists unique distinct ...,
such that <5> = jSj + • • • Show that the following are equivalent:
(a) S is perfect.
(b) S = for some w ^ W.
Exercises 477

5.2 Let 2) = ( ^ , • » be a set of root data. Assume that


A is an I X I matrix. Let S be its root system and W its Weyl group.
An element c e WKis called a Coxeter transformation if there exists an
ordered base B = ..., of 2 such that

Cr "/3/*

Assume that the Coxeter-Dynkin diagram of ^4 is a forest.


(a) Show that all Coxeter transformations are conjugate (see Exercise
3.10).
(b) Assume W is finite. The common order of all Coxeter transfor­
mations is called the Coxeter number of W and is denoted by h.
Verify that h is as follows: (W of type *, ft = - X A i J + 1),
{Bi, 2 l \ (Q , 2/), (D,, 2(/ - D), (Eg, 30), (E^, 18), (E^, 12), (E4,12),
(G2,6).
(c) Show by direct calculation that for all finite indecomposable
reduced root systems |A| = /z/ (an instructive noncomputational
proof can be found in [Bo3]).
5.3 Let A := {+ (e^ —ej)\l < / < ; < / + 1} c be a root system of
type AI (see Sections 3.1 and 3.2). Let II — [a^ ■= e, —£/+ill < z < /}
(a base of A), and identify W with 5/+^ via = (z, z + 1).
(a) Show that ( 1 , 2 , . . . , / + 1) = and that w E:W is a
Coxeter transformation if and only if vv is a cycle of length / H- 1.
(b) Suppose that p — / + 2 is an odd prime number. Assume that 2
is a generator of Z/pZ ^. Let tt e be the (p — 1) cycle
(1,2,4,..., 2^"^). Show that / ( 7 7 ) with respect to II is (p^ -
l)/8 . Conclude that (p^ - l) /8 is odd.
5.4 Let C = C(5, R) be a Coxeter group. We maintain the notation of
Section 5.3. Attach a labelled graph F = (5, E, M) to (5, R) as follows:

vertices: S
edges: E — ((^,*s') ^ 5 X 5|m^y > 1},
labels: for all ( 5 , 5 ') ^ E.

Let E' = {(5,5') e E|M^y = 0mod2}. And consider the skeletal


graph r = ( 5 , E \ E ' , M ' X where = 1 if ( 5 , 5 ' ) e E \ E ' and
ML = 0 otherwise. Define an equivalence relation on 5 by

s s' s and s' belong to a connected component of F'.


478 The Weyl Group and Its Geometry

Let [s] denote the equivalence class of s ^ S, and let 5 / ^ =


Let G be the group that is the direct sum of card(I) copies of
Z /2Z and let {1, be the copy of Z/2Z inside G corresponding to
i e I.
(a) Show that there exists a unique surjective group homomorphism
if/:C G such that ilr.s whenever 5 e 5 is such that s e
[5,].
(b) Let C' be the derived group of C. Show that C /C ' = G. (Ob­
serve that C' D ker i/r, and use the obvious inverse map G ^ C /C
arising from the defining relations of G.)

In the next two exercises we see how to explicitly compute the Weyl
group, the minimum coset representatives for the coset and double
coset decompositions W/W^ and \ W/Wi, I, K c J, and the Bruhat
ordering.
5.5 Let 3 = (A, U , U ^ , V , V • » be a set of root data over IR, W the
corresponding Weyl group, and 5 = {rj\j e J} the set of generating
reflections. Let F be the fundamental chamber, and let I c J. Let
A e F|. As far as this exercise goes it does not matter how A e Fj is
chosen, but the canonical choice is

<A, a / } = 1 if / ^ I
^0 if i G I.
We define a (coset) graph F(A) == (V, E, M) as follows:

vertices: V = {wX\w ^ W},


edges: E = {(wA, 5wA)|w e IF, 5 e 5, l(sw) = l(w) + 1, swX ¥= wX],
labels: for ijl, v ^ W\, = {j} if (¡i, v) ^ E and v = = (j>
otherwise.

(a) Show that V is in 1-1 correspondence with W/Wj.


(b) Show that (V , E) can be constructed as follows: Define c V,
E^ c E, d = 0,1,2,... inductively by
(i) Vo = {A},E^ = 0 ,
(ii) = {rjfJLlfi e V^_1, (fjL, a / } > 0}, E^ = {(/X, v)\fjL e
X/_i, V = rj^i e V^}, d = 1, 2 , . . . ,
(iii) V = U2=oY/,E= U2=oE^.
(c) Prove that the sets are disjoint (hence also the sets E^). For
/X e V, /X is of depth d if /x G V^.
(d) Let IV e IF and let = (A, r A), ^2 = . . . = (rJ k - \
r,JiA,’ rJk rjXX be a minimal chain of edges connecting A and
f l = w \ = rj^ r,J \ A. Show ^that w == r,Jk r, is the minimal
coset representative of ivIFj (see Proposition 5.2.10).
Exercises 479

(e) Prove that up to isomorphism the graph r(A) is independent of


A e F|. We denote it by r(iT,I).
(f) Show that the graph T(W, 0 ) essentially constructs W by induc­
tion on l(w) with 93 ^ determining all elements of W of length d.
(g) Let A e F. Show that for w ^ W, j ^ J,

l{rjw) > /(w ) <=> {wX, ay) > 0,

l(rjw) < l(w ) <=> <wA, a j y ) < 0.

(h) Let K c J be an arbitrary subset, and consider the double coset


decomposition Prove that each double coset has a
unique representative of minimal length in W and that these
representatives may be determined from the vertices v of the
graph TOV, I) that have no upward A:-edges, A: e K (i.e., edges
(¡JL, v) with ^ e K), from them.
(i) Let /1 = wA be a vertex of r(W^, I) with no upward /c-edges,
A: e K. Prove that

where K q == e K\(ix, = 0}.


5.6 Show that the following procedure constructs the Bruhat ordering on
W by induction on length on the graph r(W, 0 ) of Exercise 5.5. Given
w ^ W, v/c determine all predecessors w' < w of w. We can assume
l(w) > 1.
(a) Let 5 e 5 be chosen so that /(w ) < l(w). Show that for w' ^ W
with l(w') < l(wX one has w' < w iff either l(sw') < l(w') and
sw' < sw or l(sw') > l(w') and w' < sw (use the Z-lemma).
(b) Show that it suffices to determine all w' ^ W such that w' <w
and l(w') = l(w) — 1. Show that with s as in (a) and l(w') = l(w)
— 1 we have w' < w iff either l(sw') < l(w') and sw' < sw or
w' = sw.
5.7 Let = (A, n , n F, F < • , • » be a set of root data over U with
dim^V < 00, and let F be its fundamental chamber. Suppose that A is
indexed by J X J. Let K c J, and let R = where F =
U icj^l-
(a) Prove that for i e J, iv e PF, l(r^w) > l(w) <=> wF c H^y.
(b) Prove that for all i ^ K, R (Z
(c) Prove that R is convex.
(d) Show that any union of sets of the form wFj^, iv e PF, K c J is
closed in the Tits cone 3£.
480 The Weyl Group and Its Geometry

5.8 Let ^ be a set of root data over U corresponding to an indecompos­


able Cartan matrix A of finite type with base II, and finite root system
A. Let (f) be the highest root of A with respect to II.
(a) Show that <f>^ F,
(b) Let A G h* \ {0} be such that A e F. Write A = e
U. Show that > 0 for all i (join </> to A).
(c) Show that all the entries of A~^ are positive.
5.9 In this exercise we show how to realize the Weyl group W of an affine
Cartan matrix A of type ( X = A, B , . . . , G) as a group of affine
linear transformations on a euclidean space E of dimension /. We will
also obtain a description of the Tits cone for a set of root data
based on a minimal realization of A. This exercise uses parts of
Section 4.4 (that are not explicitly about Lie algebras). We will assume
that A is an indecomposable Cartan matrix of finite type II =
{aj,. . . , Ui) a base for the corresponding finite root system A with
highest root </> = E-=i«,a,, and that R = (1^,11,11^) is a minimal
realization of A over the real field R.
We form the extension 5^ = h ® 0 Rd, n ^ = {«0^..., where
= (t - (/)'', n = {«o,. . . , c with
ioLi^d) =
Then R := (r* , n , n ^ ) is a minimal realization of A. We assume
that (*1 • ) is a standard invariant form on f)^*, scaled so that (a^|a,) = 2
for all the long roots of 11 c E. In particular (</>!</>) = 2. The element
5 := «0 + (/> is null; ( 5 |a /) = 0 for all / = 0 , .. ., /.
We set V = r * , V Then 3 := {A, F, F E, E < • , • » is
a set of root data for A. Let its Weyl group be W == (r^\i = 0, .. ., /),
and let W '•= (r^\i = 1 ,..., /> be the finite subgroup which is (isomor­
phic to) the Weyl group associated with the finite root system A.
(a) Prove that 8 and i are PF-invariant vectors and hence that
F^:={xeF|<x,(t>>0},
¿ : = { x e F | < x , ( t > = l},
are IF-invariant.
(b) Let 7t:F F/R S be the natural inapping. Show that V/U8
carries the natural structure of a IF-module that makes tt a
PF-map. Show that (1 *) induces a PF-invariant bilinear form on
F/R 5.
(c) Show that E == 7t(E) is a PF-invariant affine subspace of V/U8
and that PF acts as a group of affine linear mappings on E (see
Section 7.3 for a definition of affine linear mappings).
(d) Prove that by defining the distance between x , y in E by (x - y\
X —yY'^^, E becomes a euclidean space and the elements of PF
act as isometries of F.
Exercises 481

(e) Show that = E-^iIRa, can naturally be viewed as the group of


translations of E.
(f) Let t := e W. Show that t is the translation x ^ x (f> and
that the normal subgroup T oi W generated by t is the group of
translations
jc JC + /X,

where ¡ runs through the lattice L generated by the long roots of


jl

A.
(g) Prove that L = under the mapping / jjP (see the IBF
jl

table in Section 4.4). (Show that any indecomposable root lattice


is generated by its short roots.)
(h) Let d* e 5* = be chosen so that <d*, a ^ ) = S^ q. Show that
Tr(d*) e E and that 7r(d*) is a fixed point for the subgroup W of
W.
(i) Prove that W r [ T = l , T < W , 2i n d W = W T = W \ j < T .
(j) Show that the fundamental region F oi 3 lies in and hence
that the Tits cone 3£ lies in U{0}.
(k) Using the convexity of 36 and (f) prove that 36 = U{0}.
(l ) Show that each facet wFj, / c J, intersects E. Show that the sets
iriwFj) determine a simplicial decomposition of E and that W is
the group of isometries generated by the reflection in the sides of
any simplex of maximal dimension in this decomposition.
(m) Explicitly draw these simplical decompositions for the cases
^2*
5.10 [Cp] Let (g, r ) be a Kac-Moody algebra. An imaginary root a is
called strictly imaginary if

P e'^^A => a p and a —j8 e A.

Let

^^"*A := (a e^'^Ala is strictly imaginary}

and

(a) Show that


(b) Recall := [a e 0 + n ( —t^)lsupp(o:) is connected}. Show
that if a fl and g A+, then a + /3 e A+ [reason
by induction on ht()8)].
(c) Show that is a semigroup.
482 The Weyl Group and Its Geometry

5.11 Let A = <i, j<n be an indecomposable Cartan matrix with


Coxeter-Dynkin diagram F. Let R = ( l& ,n, n^) be a minimal realiza­
tion of A over R, and let 3 be the standard root data constructed
from R. We assume that A is symmetrizable and (1 • ) is a standard
PF-invariant symmetric bilinear inform on 1^*. We know from
Proposition 2.6.8 that ( | ‘ ) is positive definite if and only if A is of
finite type and positive semidefinite if and only if A is of affine type. In
this exercise we are interested in the case where ( I • ) has signature
{n - 1,1). We say then that A is of hyperbolic type. This term has also
been used in a more restrictive sense of level 1 matrices (see below.)
We define the level of the Cartan matrix ^ to be / if / is the least
nonnegative integer for which removal of any / nodes from F leaves a
graph whose connected components are of finite or affine types. Thus
finite and affine Cartan matrices are of level 0.
(a) Prove that if A is of level 1, then A is hyperbolic.
(b) Prove that if A is of level 1, then the closure 3£ of the Tits cone
satisfies

3E= {x e ^ *|(jc |x ) <0}

and 3£ c — >o0^r
(c) [Mx] Prove that if A is of level 1, then

"A^= {/3 0 j(/3 |j8 ) ^ 0 } .

(d) Show that if A is of level 2, then A is hyperbolic.


5.12 Let ^ = {A, • » be a set of root data, and let S be
its root system. Let ft be a subroot system of 2.
(a) Let ft+:= 2 + n ft == Show that there exists a well
ordering < in K so that ht()3,) < ht(pj) => i < j for all i, j e K.
(b) Let H be the subset of ft+ obtained by the following prescrip­
tion.
(i) B, where is the smallest element of K.
(ii) If k k^, then e H if and only if

^ © Npj.
}< k

Show that H is a base of ft.


5.13 The purpose of this exercise is to prove Proposition 5.3.3. (Show that
R^ = R_^, Use Proposition 5.2.7(iii) and the Coxeter group presenta­
tion of W to conclude that there exists a well-defined map R^ ,
Show that this is surjective.)
Chapter Six

Category & for Kac-Moody


Algebras
T h u s d o th sh e , w h en fro m in d iv id u a l sta te s
S h e d o th a b s t r a c t th e u n iv e r s a l k in d s
W h ic h th e n r e c lo t h e d in d iv e r s n a m e s a n d f a t e s
S t e a l a c c e s s th ro " o u r s e n se s to o u r m in d s .
—Coleridge

In this Chapter we investigate in detail the nature of the representations


afforded by category ^ for a Kac-Moody algebra g. An important new
ingredient that does not exist in the generality of Chapter 2 is the existence
of a “Lie group” G associated with g. We have already seen in Sections 2.4
and 4.1 the advantages of associating the group 5L2(IK) with ^I2(IK), in
analyzing the structure of g. Along the same lines we wish to associate a
group G with g. This is somewhat complicated by the fact that there are
really a number of different, but closely related groups, that one might wish
to use. A nice survey on this is to be found in J. Tits’ paper [Ti6]. We have
chosen the approach due to Peterson and Kac [PK], which is fairly straight­
forward and by its construction allows G to be represented on all the
“integrable” modules of g. It is a historical irony that Lie algebras arose as
an aid to understanding Lie groups, whereas here we find ourselves reversing
the process and constructing Lie groups in order to understand Lie algebras.
Just as Dg turns out to be slightly too small for a healthy representation
theory (because it cannot separate the weight spaces that we want) so too we
need to enlarge G to a group G. This is particularly important for the
conjugacy theorems of Chapter 7 where the groups play a key role. In the
classical case (Cartan matrix of finite type) G G.

483
484 Category á for Kac-Moody Algebras

In Section 6.1 we develop the theory of integrable modules for g and the
groups G and G hand in hand. Section 6.2 investigates the nature of the
weights and weight spaces of integrable modules, although the best results on
this direction occur in Section 6.4 where we obtain the important Weyl-Kac
and Weyl-Macdonald-Kac formulas. Among other things, these enable one to
explicitly compute the weight multiplicities for irreducible modules in the
category ^ (when g is invariant). As the names suggest, these formulas were
originally discovered by H. Weyl in the classical finite-dimensional theory.
In Section 6.3 we establish that, somewhat like g, G also has triangular
decomposition (or Birkhoff decomposition). This and the related Bruhat
decomposition are important in understanding the structure of G and in the
classical situation have important uses, ranging from the study of differential
equations to the algebraic geometry of Schubert varities to the construction
of abstract geometries on Lie groups. In Section 6.5 we prove the complete
reducibility theorem for integrable modules in category <^, and in Section 6.6
we give the explicit formula for the Shapovalov determinant for invariant
Kac-Moody algebras.

6.1 INTEGRABLE MODULES

In the classical theory of Lie groups and Lie algebras the connection between
a Lie algebra and its Lie group is made by the exponential map exp: g -> G.
Ordinarily one expects that representations of g lift (or integrate) to repre­
sentations of G (although there are some subtle points here). We wish to
create a group and a class of representations of g for which there is a similar
relationship. The primary problem is that for infinite-dimensional representa­
tions 7T one can hardly expect that exp ir(x) will be defined for arbitrary
X 0 g.
There are two conceivable ways out. One could topologize the spaces
involved and require that exp(7r(jc)) converge, or one could restrict the set of
a : in g and the class of modules under consideration so that 7t ( jc ) is locally

nilpotent, and hence exp(7r(A:)) makes sense. For arbitrary Kac-Moody alge­
bras it is only the latter for which there is a well-developed theory. The x for
which we will define exp a: are those in U«ere^g", and the representations
are those that admit weight spaces and for which each : g a U ^^ts
locally nilpotently (integrable modules).
The group G is essentially defined as the smallest group admitting the
action of exp7r(A:), x e Later in this section we enlarge G to a
group G that stands in relation to G much as g stands in relation to Dg.
Let g be a Kac-Moody Lie algebra over a field IK of characteristic 0 with
structure matrix A and triangular decomposition (5,g+, We
use the standard notation of Section 4.2.
6.1 Integrable Modules 485

A g-module ( tt, M) is said to be integrable if

IMl M has a weight space decomposition M= 0

relative to

IM2
For all j e J, the elements of 7r(g“>) and 7r(g““>) are locally nilpotent.

We speak of tt as an integrable representation of g. The integrable


modules form a full subcategory ^ of the category of all g-modules. It is
immediate that ^ is closed under the formation of direct sums, tensor
products, submodules, and quotient modules.
Let £ be a subgroup of 1^* such that 0 c £. We define ^ ( E ) to be the
full subcategory of cX whose objects satisfy

IM(E) P{ M) (ZE,
Of course ^ 0 ) * ) = We also use the full subcategory ^fJ<E) of ^ { E )
consisting of modules M for which

IM(fin) dim < 00 for all Л e 1^*.

Lemma 1
Let Ml and М2 be K-spaces, and let e End(,^(M,), i = 1,2, be nilpotent.
Define a linear map

X :H o n iK (M i, М2) М2)
by
X--f^X2f-fXi

for all f e Homj^CMi, М2). Then x is nilpotent and

(ехрл :)/= (ехрх2)/(ехрдс,)"*.

Proof. Let L (resp. R) denote left (resp. right) multiplication in


Hom^iMj, M2) by X2 (resp. Xj). Then
X=L - R and LR = RL.
Since L and R are nilpotent it follows that x is nilpotent and that expx =
exp L exp(—i?) = exp L(exp (Proposition 1.5.2). □
486 Category ^ for Kac-Moody Algebras

Proposition 2
Let % be a Lie algebra, and let tt : g ^ gI(M) be a representation of g on
some K’Space M. Let jc e g te such that both ad x and 7t( jc) are locally
nilpotent. Then for ally e g,
-1
Tr((expad Ac)(y)) = (exp-n-( j:))Tr(y)(exp i7( a:))

Proof Consider K jc as an abelian Lie algebra. Then

U(IKa:) = IK © K x © IKjc^ © • •

Fix V ^ M and y e g. Define

M l •= IKi; H- IK7r(x)i; + IK7r(x) + • • • = U(IKjc) • v


and
M2'^= U( 1Ka:) •7T(y)Mi.

Both Ml and M2 are finite dimensional, and

7r(y) : Ml M2.

Let Xi := 7t( jc)Imi and X2 — 7t( jc)Im2- These are nilpotent. By the previous
lemma we have

(expad 7r(x))(7r(y)) = exp 7r(jc)7r(y)(exp 7t( x )) in Hom(M i,M 2).

Now 77 °ad jc = ad(7r(x))° 77, and hence

77 ((expad x ) y ) = (expad 7 7 (x))(7 7 (y ))

= exp 77(x)77(y)(exp77(x))"^
as maps on M^ Since v ^ M was arbitrary the result follows. □

Proposition 3
Let (7 7 , M) be a %-module, and suppose that M admits a weight space
decomposition relative to Then M is integrable if and only if the elements of
77(g"*') are locally nilpotent for all a

Proof Let a Then a = wa^ for some 7 e J and w e PF. By Proposi­


tion 4.1.4 there exists an elementary automorphism n = n{w) of g such that
wg"> = g“ and n is a product exp(ad ... exp(ad where each ©
U a en u (-n )9 “‘ Let y e g“^ and use Proposition 2 to conclude that Tr{ny) is
locally nilpotent. Thus all the elements of 77(g“) are locally nilpotent. □
6.1 Integrable Modules 487

Fix a subgroup E of 1^* with Q c.E. We now construct a group


attached to g and E that acts on every g-module of ^ ( E ) in a way
compatible with the action of g.
We begin by recalling the concept of the free product of a family of
groups. Let A be a set, and let be a family of groups. By a free
product of the family {G}^^j^v/t understand a pair consisting of
a group G and a family of mappings :G^ ^ G such that

FPl: Each : G;^ G is a group homomorphism


FP2: Given any group H and any family of homomorphisms H,
A e A, there exists a unique group homomorphism f : G ^ H such
that f ' ix = fx A e A.

As one would expect free products exist and are unique up to isomor­
phism. Furthermore the homomorphisms ¿x are injective and the subgroups
ix(Gx) of G generate G. The construction and basic facts about free products
of groups are covered in [Ja2].
The uniqueness of free product up to isomorphism allows us to speak of
the free product of the family {Gx)x e a- denote it by

G = *AeA<JA-
Notice that since the ¿x are injective we can identify each Gx with a subgroup
of G.
We now return to our Lie algebra g. For each a consider the
one-dimensional space g"". We think of each g“ as an abelian (additive)
group and define

G = *

We denote the injective homomorphism g" ^ G of FPl by exp^. We also


write the group G multiplicatively so that for all a and x, y e g“

exp„^(x + y ) = exp„^(x)exp„^(y)
and

e x p ^ ( - x ) = exp„^(x) \

Since the subgroups exp^(g^) generate G, a typical element of G is a finite


product of the form

(1) exp^"^(^i)---exp^^^(x,)

for some r e where /3, e"A and x, e g^' for all 1 < i < r. If /• = 0, then
(1) is understood to be the identity element of G.
488 Categoiy ^ for Kac-Moody Algebras

Let ( tt, M) be a g-module in Then for each a and for all


X e g“, exp(7r(jc)) is a well-defined element of GL{M) (Proposition 3). The
mapping exp^: g" GL{M) given by x exp 7t( x) is a group homomor­
phism, and hence by FP2 there is a unique homomorphism

tt:G ^ GL{ M)

such that
exp^
r GL{M)

ex p ¿

commutes.
Let iC := n ker tt where the intersection is taken over all modules ( tt, M)
of g which are in ^ ( E ) . Define

:= G /K ^ .

We call the group G^; the derived (Lie) group associated with the Lie algebra
g and £. This terminology will be justified in Proposition 15 below.
Convention In general, whenever E is understood from the context, we
simply write G instead of G^ and refer to the elements of ^ { E ) as
“integrable.”

Let v :G G be the natural quotient mapping. For each x e g“ and


a we define exp x e G by

exp X = r^(exp^ x).

If 77: g qKM) is integrable, then there exists a unique group representa­


tion

TT: G —> GL(A/)

such that the diagram

GL(M )
6.1 Integrable Modules 489

commutes. In this notation we have for all a and for all x e g“,

(2) exp 7t( jc) = TT exp x.


A particularly important case is that of the adjoint representation
a d :g ^

By the above this formula leads to a group representation a d : ^ GL(g).


We write Ad instead of ad in keeping with the standard notation in Lie
theory:
A d :G ^ ^ G L ( g ) .
Equation (2) reads

(3) exp(ad x) = Ad(exp x) for all x e g", a

Since exp(ad x) e Aut(g) and G^ is generated by exp(g“), a e'^^A, we see


that
A d: —>Aut(g)

and Ad(G£) = G^^ as defined in Section 4.1. In particular AdCG^) is inde­


pendent of E. We will simply write Ad(G). Using Proposition 2 we have: if
(ir,M) e jr(E ),

(4) X ^ Q, and g ^ Ge then tt (A d (g )x ) = Tr(g)7r(x)Tr(g)

Remark 1 Let jc e G be such that ttCx) acts like the identity for every
integrable representation tt of g in <y(E). Then x = Indeed the
definition of is designed precisely to ensure this.

The category <^(g, of g contains numerous integrable modules. The


next two results show what the integrable highest weight modules look like
and how to construct them. Before starting this we point out that elements of
g+ always act as locally nilpotent transformations on modules in ^ (g ,
(see Proposition 2.6.1).

Proposition 4
If M is an integrable %-module and A e P { M \ then (A, > e Z for all
y e J.

Proof Fix any j e J. By Lemma 2.4.4 M is completely reducible as an


^I ^-module and the irreducible components are finite dimensional. In
particular is diagonalizable on M and its eigenvalues are integers. Thus,
if X G for some A e P ( M \ • jc = <A, a / >jc => <A, a / > G Z. □
490 Category ^ for Kac-Moody Algebras

Proposition 5
Let L be a highest weight module with highest weight pair (A, v^) and weight
system P(L). The following are equivalent:

(i) L is integrable.
(ii) For each / e J, L is the direct sum o f integrable {hence finite-dimen­
sional) irreducible -modules
(iii) For each j e J, <A, }^ N • u^= 0.
(iv) For each j e J, (A, and there exists a positive integer Uj
such that f f j ' v^= 0.

Proof (i) => (ii) See 2.4.4.


(ii) => (iii) The ^1^2^^-module M generated by is finite dimensional and
hence by Lemma 2.4.1 (ii), we have dim M = <A, ) -\- 1 > 1. This shows
that <A, a / > e N, and also that 0.
(iii) => (iv) Obvious.
(iv) => (i) Since L e the elements Cj, j e J, act locally nilpotently on
L. We show that

L •= [x e Llfor each ; e J there exists nj e Z+ with f p • jc = 0}

is all of L. But i; +e L ' by assumption, and so it suffices to prove that


f L c L' for all i e J. For this we use the following simple identity in U(9):

(5) m = J r

in U(g). This is easy to show by induction on n. Assuming (5), let x e L', and
let us show that /¿ • x e L'. Let j e J, and choose mj e I\1 satisfying f^ j •
jc = 0. Let n = mj + \A ^ Then for each summand in (5) either k > \A¿^\ + 1
or n - k > mj, and hence fff¿ • x = 0 by Proposition 4.1.8(iii) (applied to fj
and f¿). □

Using Proposition 5 we can construct infinitely many non-isomorphic


integrable highest weight modules for g.

Proposition 6
Let A e g* satisfy (A , аУ } e N for all j e J. Then there exists a {unique up
to isomorphism) integrable highest weight module L^^JiA) o f g with highest
weight pair (A, i;+) with the property that, for every other integrable highest
weight module L of q with highest weight pair (A,¿;+), there exists a unique
homomorphism L ^ ^ J ^ A ) L satisfying + z;+. Moreover the irreducible
highest weight module L{A) is integrable. Both L{A) and L^^^(A) lie in
^fiJ<E), where E is the subgroup o f generated by A and Q.
6.1 Integrable Modules 491

Proof. Let M(A) be the Verma module with highest weight pair (A, z;^). Let
Xj • v+, y ^ J. Then Xj is a highest weight vector in M(A).
Indeed for i ¥=j, e¿ • Xj = • ¿5^= 0, while for i = j this result
follows from Proposition 2A1 (ii). Set N ■='Lj^jVH^)xj. Then N =
E^.^jU(g_)xy c M(A). Set L ^ ^ ( A ) ’•= M (A )/N and v + v^-\- N. Then
Lj^^(A) is integrable by Proposition 4. Furthermore, if L satisfies the
hypothesis above, then by the universal properties of M(A) there is a
surjective homomorphism cp: M(A) L. Its kernel contains N by Proposi­
tion 4. Thus there exist a unique induced surjective homomorphism
9• ^ ^ the desired properties. Since L(A) is a homomorphic
image of L^^(A), it is integrable. Clearly L(A) and L ^ ^ ( A) lie in

Remark 2 The foregoing can easily be applied to obtain analogous results
about the lowest weight modules in the category ^fin-

We now return to the study of the group G = G¿;. We wish to impose


some restrictions on the subgroups jE of 1^* that we will consider. These
restrictions are partly suggested by Propositions 4 and 5.
For each i e J let (o¿ e 1^* be any element such that a / > = 8¿j. Then
the set
ft := [(o¿\¿ e j}
is called a set of fundamental weights of (g, ^ ) .
Remark 3 Any set of fundamental weights is K-linearly independent.
A subgroup P of is called a weight lattice of (g, if it satisfies the
following properties:

WLl: e c P.
WL2: <P,!2^> ^
WL3: There exists a minimal regular weight p in P (see Section 4.4).
WL4: P contains a set ft = ft(P ) of fundamental weights of (g, <9^).

If in addition P satisfies

WL4R: P is a free Z-moduIe with a basis 0 consisting of K-linearly


independent elements and such that O contains a set of fundamental
weights of (g, .9^), then we say that P is a restricted weight lattice of
(9, ^ ) .

Notice that in any case WL2 and WL4 give = Z.


Elements of P are called integral weights (because of WL2). An integral
weight fjL is dominant if
fjL e P+:= ( a e P| <A, a / > e N for all i e j}.
492 Category & for Kac-Moody Algebras

The standard weight lattice for (g, is

P== (A cZ }.

Remark 4 The standard weight lattice for (g, is clearly a weight lattice.
If J is finite, det(>0 4^ 0, and is minimally realized then P is the only
weight lattice for (g, In fact and 1^* = where
n = e J} is the (unique) set of fundamental weights of (g, Now
WL4 and WL2 combine to show that any weight lattice must be exactly
= P. This is clearly a restricted lattice and we have P^= Hj^jNcoj,
The situation for the cases when d et(^) = 0 is more complicated. The
Appendix describes the case of in some detail.
Remark 5 If J is finite and P is any weight lattice then the Z-span of is
all of P. Let \ ^ P, and define aj •= <A, ^ Z for all j e J. Let J_ be
the subset of J consisting of those j for which Uj < 0. Then ji ==
Eyej_~ ^ ^+7 ^Iso A + )Li, G P^, whence A = A + / i , —/a g P_^_-
'

As far as the representation theory of g is concerned, there is no need to


consider restricted weight lattices, and the standard weight lattice is the
natural choice. The representation theory of G is more subtle, and there are
instances when the standard weight lattice will be too large (see the introduc­
tion of G later in this section). For this reason we introduce restricted
lattices (which really are lattices).
We now choose an arbitrary weight lattice P and work in the category
^ { P ) with the group G = G(P). For each a let c G be the
subgroup generated by exp g“ and exp g"“ (cf. Proposition 4.1.6). We fix
once and for all an § 12-triplet [e^, a ' ^ , f j with e g“, / „ e g Clearly
= 5L^2““1

Proposition 7

(i) Let ( tt, M) be an integrable q-module, and let a e'^^A. Then there
exists a surjective homomorphism

k(;>:5L2(K) ^ ' it(5L<2“>)

such that for all t g IK,

0 1) ^ 1Í 1) = '»f(exp if«).

(ii) There exists a unique isomorphism

: 5L2(K) ^ 5L<2“^
6.1 Integrable Modules 493

such that for all t ^ K,

and

■ok<“) = k(«).

Proof (i) Identify with ^I2(1K) by <-> o ) ’ ^ (1 o)*


ir(e^) and 7t(/^) are locally nilpotent, M is an integrable §l2(lK)-module and,
using Lemma 2.4.4, tt integrates to a representation k^“^ of 5L2(IK) in which

®^(^(o 0 ) ) ^ (0 j) expT7(re„) = ir(ex p ie„)

and r(exp tf^). Evidently k^“X5L2(IK)) = tt(SL^2^\


(; ;)■
(ii) Let SL^2^ be the subgroup of G generated by exp^ and exp^^ g
By universal properties = (exp^ 9“)*(exp:i^^ g~“).
The homomorphisms exp^ |J ^ j, exp:^^ie_^ ^ ( t l)
mine a surjective homomorphism o) : SL^2^ SL2(K),
Let ( tt, M) e ^ ( P ) , and consider the mappings k^"^ of part (i) and
TT: G ^ GUM) . From

^(exp^^ te^) = exp(irie„) = k<“>| J JJ = k<“><w(exp„^ ie„)

and a similar calculation for exp tf^, we obtain the commutative diagram

SLi“^

5L2(K)

Thus ker -ir 3 ker ca. This being true for all (ir, M ) e ^ ( P ) , we have
1 ^ ^ 0 5 4 “^ =>kero) and hence an induced homomorphism k^“^: SL2(K)
SL^2 ^ with k^“^<«) = V on SL^2 ^ (see above for the definition of v : G -* Gp).
494 Category ^ for Kac-Moody Algebras

This leads to the commutative diagram

Now is surjective since v is. Hence it remains to show that is


injective. Suppose a e Choose A e so that <A, > e 2Z + 1 (this
is possible since \ a ^ ^ Q^X Then ^ ^ ( P ) , and if we let v+
denote a highest weight vector for L ^J iA ), M — E^l’o is the
irreducible §I2 (IK)-module of dimension <A ,a^> + 1. But the correspond­
ing action of 5L2OK) is faithful (Lemma 2.4, Corollary 6). If we denote this
representation of g on L ^ ^ (A ) by tt, then our conclusion is that k^“^ is
injective and hence also k^“\ If a e ''^A_ we use the same argument revers­
ing the roles of and /^. □

In the sequel, when dealing with representations ( tt, M ) in ^ ( P ) , we will


often suppress explicit mention of the mapping tt and write both the action
of g and G by an infixed * or even simply by juxtaposition. We will also refer
to the representation in J^(P) simple as integrable (omitting the explicit
mention of P which is assumed to be fixed in the discussion).
For each i e IK^ we define elements n(t) and h(t) of 5L2(IK) as follows:

[see (2.4.9) and (2.4.10)].


For each a we have a fixed il2-triplet {e„, f j and §1^“^= Ke^
® © IK/„ = §I2(IK). Define elements n„(0 and /i„(i) of by

n , ( 0 = k<“> (n (0 ),
( 6)
/»„(0 = k<“) ( /,( 0 ) .

From (2.4.9) and Proposition 6,

(7) n „(i) = ex p ie„ex p (-fY „)ex p (te„).

From (2.4.10) we have

(8) MO ="a(0««(l)“ ‘-
6.1 Integrable Modules 495

From the definitions

(9) n „(í) ^ = n ^ ( - í ) and ‘ = /i„ ( í ^).

We also have (Proposition 4.1.3)

( 10 ) A d (n ^ (í))a ^ = - a ^ .

We introduce two subgroups N and H oí G defined as follows:

N := {n^{t)\a t e 1K^>,

H '= {h^{t)\a t e K^).

Proposition 8
Let ( tt, M ) be an integrable representation o f g with weight system P(M ). Let
a e'^^A, and let A e P(M). Let be the reflection in a.

(i) Given V e there exists v ' e such that Tr(nJ,t))v =


for all t e K^, In particular ir{nJ^t))M^ =
(ii) For all t e IK^, Tr(/t^(i)): ^ and irihj^t)) acts as a scalar
multiplication by on M^.

Proof Fix V e M^. To keep the notation down we suppress the tt. Then
from (7), nj, t)v e the only power of t occurring in the
homogeneous term of degree Á + k a in njit)v is t^. Thus

k^Z

where e Computing *««(0 *v = nj^t) • (A d(n^(-i))a^)


• V = -nj<^t) ' • V (see (4)) in two ways gives

E(<A,a^> +

from which = 0 unless k = —(A ,a ^ ). Part (i) follows with v' *=


^A-<A,aV>a n = nJ^~t)M'^<^^ C M \]
From (9), /t^(r) = « « ( - 0 ” ^ n « (-l), and from part (i),
nj^ —t)v = { — Now part (ii) follows directly. □
496 Category û for Kac-Moody Algebras

Corollary
(i) H is abelian.
(ii) Every element o fH may be expressed uniquely in the form n¿ ^
where each t¿ ^ and t¿ = 1 for all but a finite number o f i.
(iii) For each a the mapping t hj^t) is an injective homomor^
phism of into G.

Proof (0 Let a, )8 5, i e K^, and let (v , M ) be integrable. Then for


each weight A of M, hj,s)h^(t) acts as on M^. Thus
Tr(hJ,s)hp(t)) = Tr(h^U)hJ,t)X and this being true for all ( tt, Af) e <X(P)
we obtain hj^s)h^{t) = h^(t)hj.s). This proves (i). We observe in the same
manner that hj^s)hj^t) = hj^st).
(ii) Let a e'^^A, and suppose that = Hi^jC¿a¡^ (where all but a finite
number of = 0). Then for any t and any weight A of (rr, M ) e ^ ( P) ,
hj, t) acts on as multiplication by

7<A,a/>

This is precisely how Ylh^it^O acts on M^. This being true for all ( tt, M), we
obtain

hÁ0 = n h 4 n ^

Thus every element of hj^t) may be expressed in the desired form. Since H
is abelian and is generated by the elements hj^t), and from our observation
in part (i), we see that every element of H can be expressed as required.
Applying h = Ylh^it¿) to the weight (Oj, we find that h acts as tj on
The uniqueness for the expression of h follows.
(iii) Let a e'^^A. We have seen that the mapping

5 h^{s)

is a homomorphism. We write = L c ia ^ , and observe that since is a


PT-translate of some is indivisible [i.e., gcd(c¿) = 1]. Now, if hj^s) =
1, we see from L(o)j) that s^j = 1 for all j, whence 5 = 1. □

According to Proposition 8, nj^t) induces a permutation on the set of


weights P(M) of any integrable module ( tt, M), namely A r^A. Since each
element of A/^ is a product of terms of the form n^(t) [see (6.12)], there is a
map

[\ : N
6.1 Integrable Modules 497

such that n(M^) = for all n N and for all A e P(M). Indeed, if
n = «ajiti) • • • then InJ = Thus IJ is independent of the
representation (ir, M).
For jv e IF we will use the notation fwl to denote any preimage of w.

Proposition 9
The mapping I J is a surjective group homomorphism with kernel H. In
particular H < N and N /H = W.

Proof. If InJ = w, tn'J = w', then for any (ir, M) e J^(P), nn'M^ = nM ”''^
= which shows that I J is a homomorphism. Evidently H lies in the
kernel of L J [Proposition 8(ii)].
We show that H is normal in N. Let a, ¡3 e''*A. Then we claim that

(1 1 ) ^ = hr^0Ís)

which obviously will imply the normality of H. To prove (11), apply both sides
to a vector u e for an arbitrary ( tt, M) e ^ ( P ) and A g P(M):

since <r^A,j8^> =A, and


At this point we have a surjective group homomorphism of N / H onto W,
Now we construct its inverse [St]. We use the presentation of W given in
Proposition 5.3.3:

IF = <w„|a e " A , = 1,

Let w^ ■= nJX)H, a e"A . Then by (8), nj. t) e HnJ^i) = which shows


that w^ = n J.t)H for all t e IK^. Thus w^ = nJ ^ t ) n J , - \ ) H c H, and this
shows that = 1 in N / H . Also

‘ = «a(l)exp Cp cxp( -/p)exp C^n„(l)

= exp(Ad n „(l)e ^ )ex p (-A d n„(l)/^)exp(A d n „ (l)e ^ )/f by (4).

Now Ad nj. t) e Aut(g) and so by Proposition 8(i), Ad n„(r) maps {e^, ,fp}
into an ilz-triplet for that is, {se,^, r„p'^, s~ ^frji for some s e
498 Category ^ for Kac-Moody Algebras

(see Proposition 4.1.6 and also Section 4.1, Remark 1). Thus

w,

It follows that there is a homomorphism W N /H with for all a,


and this is clearly the inverse of the induced homomorphism N /H W. □

Corollary
G is generated by the set S ■= {exp te^, exp tf^\i e J, i g IK}.

Proof. Let a and write a = for some j e J. Let


n¿^í) • • • 1). Then exp(Ad nitej)) = n exp(tej)n~^ shows that exp g“ is
generated by 5. □

As a direct consequence of the foregoing we have

Proposition 10
Let (t7, M) e ^ ( P ) .

(i) W stabilizes P(M).


(ii) I f n ^ N and [n\ = w, then nM^ = for all A e P{M). In
particular dim = dimM"^^. □

Proposition 11
Let ( tt, M ) be module in category 0 and category ^ ( P ) . Then for each
A ^ P(M) there exists a unique ¡jl e W \ such that

W \ n P + = ill).

In the case that IK = [R and J is finite we have ¡i = W \ n F, where F is the


closure of the fundamental chamber for i)* relative to II.

Proof Let A ^ P(M), and let D Mq = (0) be a


composition series of M relative to A (Proposition 2.6.8). Then there is a
subset I of (1 ,..,, fc} for which = L(A¿) for some A, > A and such
that (p > \ => (p ^ P(L(X¿)) for some i e I. Evidently each L(A¿), i e I, is
integrable, and hence A e P(L((p)) for some cp e P_^_. In particular A e
<pi Q ^ ( z P . Furthermore WÁ <z <p and hence there exists a w e IP for
which wX = (p — a — jjL with a e j2 +Li{0} and ht a minimal. Then rjfi < p,
and hence { p . a y ) e N, for all j e J. This shows that )l¿ e WX. If
6.1 Integrable Modules 499

v ^ P + n WX then L(v) is integrable and so WX In particular


V G WX ^ fi i Q+ and we see that necessarily v = ji. Thus P+Pi WX = {/x}.
In the case that IK = R and J finite, lies inside F. □

Let us recall that the derived algebra Dg of g is generated by the


elements e¿,ay,f¿, i e J, and that g = Dg + Í), Dg Pi Al­
though Dg is easier to define than g, g is far more convenient for the
purposes of their representation theory. This is primarily because i) can
separate the weight spaces of any module with a weight space decomposition,
whereas ' Li ^j Kay in general cannot.
The same problem arises with our group G: It cannot separate the weight
spaces of the integrable modules in general. Furthermore the subgroup
Ad(G) of Aut(g) is not large enough to produce the ingredients needed in
the conjugacy theorems that we wish to prove in Chapter 7. The solution is to
embed G in a larger group G. We will do this in such a way that G = (G, G)
and that G = GX, where is a group that acts diagonally on integrable
modules. In Section 6.3 we will see that if J is finite and the Cartan matrix of
g is nonsingular, then G = G whenever g is minimally realized, just as
Dg = g.
To introduce the subgroup A" in a satisfactory way, we will assume that P
is a restricted weight lattice (see WL4R) in 1^*. Let H G = G(P) be the
corresponding diagonal group. A typical element of H is h =
We have seen in Proposition 8 that for any module M of ^ ( P ) and for any
weight X of M, h acts on as multiplication by H /e ^ This leads us
to consider the mapping

/ / A H o m (P ,K ^ )

defined by

L(h)

where t(/z)(A) = ^ Since P contains a fundamental system of


weights, we see that t is injective. We will often identify / / as a subgroup of
the group Hom(P, of characters of P.
A subgroup X of Hom(P, IK^) is called a separating group of characters if

SGC: For all X, fjL P with X there exists x ^ ^ such that ;^(A)

Since P is a free abelian group, it is clear that Hom(P, is itself a


separating group of characters (a natural choice). In particular such groups
exist. However, H need not be separating. Indeed, if P contains weights
5 # 0 satisfying (8, = 0 for all i e J, then H cannot separate 0 and 8.
500 Category ^ for Kac-Moody Algebras

Let X c Hom(P, be a separating group of characters of the weight


lattice P. If ( it, M) is any g-module in ^ i P ) , then we obtain a homomor­
phism

A ut^iA /),
AT

where is defined by (see Section 2.1)

x J m>
^ ■= multiplication by a'( m)-

A simple calculation shows that for all a e ''^A, x e g“,

( 12 ) exp(rr(Arad(-*))) = Ai^exp( 17(01:)) at; ^

Remark 6 Let X c Hom(P, IK^) be a separating group of characters. Let


( tt, M) be any integrable module in ^ ( P ) , and let u E: M, Then the smallest
subspace of M containing Xj^u) is the 1^-submodule N generated by u. If
u= ^ In particular any jC-invariant subspace of M
is an ^-submodule. The argument to prove this is a variation of that of
Proposition 2.1.1.

Let G •= G(P). We are now going to define a map :X Aut(G)


through which we will enlarge G to a semidirect product G >^X. Let a' ^
For each a e '’^A let

be the group homomorphism

Ar„ : o: exp(Arad(-«)) = exp(Ar(a)o:).

By the definition of G this extends uniquely to group epimorphism

X:G ^G

satisfying X exp^ = a:,,.


Let us show that Kp = ker x. Consider a typical element

X = expp^(xi) • • • exp^^jxj
6.1 Integrable Modules SOI

of G. Then

;f(^) = X expp^( Jíi) ■■■ X


= e x p ( A T a d ( ^ i ) ) • • • exp(A T ad(

Let ( tt-, M) be an integrable representation of g in ^ { P ) . We have

='^exp(ATad(^i)) ••• 'irexp(;(^3d(^«))


= exp(TTArad(^i)) •••exp(TT;^rad(-^n)) [by (2)]
= Ar^exp(Tr(ACi))Ar"' • • • exp(ir(jc„));^:-i [by(12)]

= Ar^exp(ir(xi)) • • • exp(ir(x„))Ar;:*
(13)
= Ar^irexp(xi) ••• Trexp(x„))Ar“ ‘
= AT^irv expp^^(xi) ■■■ Ttv exp^„^(x„)Ar“ ‘

= AT>exp^^^(xi) ••• <ffexp^^^(x„)A^-‘

= x jr { x ) x : ;; \

From (13) and Remark 1 it follows that Kp = ker x, as desired. We therefore


have an induced automorphism of G making commutative the diagram

We define the group homomorphism

- : X - y A ut(G ) by

~ - X ^ X-

and form the semidirect product

G* = G XIZ.

By definition this means that in G * the composition law is given by

{ S i , Xi ) { 82, X2) = i 8 iXii82)>XiX2)-


502 Category ^ for Kac-Moody Algebras

Note that for j: g g“, a e "A we have

x{exp^ x) = iv(exp„^ x) = x(exp^


(14)
= Xa(^) = X a(^ ) = eXp(ATadAi).

Note also that G* is generated by the elements of the form (exp x ,x ), where
X e g“, a G "A, and x ^
Let ( tt, M) be an integrable g-module. There exits a unique representa­
tion (-ff*, M) of G* satisfying for all g G G, g X, and i; g M,

(15) = Tr(g)ox^(u).

Indeed for all ^ e g“, fe g g^, and o-, t ^ X,

IT* ((exp fl, o-)(exp fc, t ) ) ( z;)


= TT^(exp a{&exp b ) , a r ) { u )
= TTexp(fl) o Tr(i7 exp(fe)) o(o-r)^(i;)
= irexp(fl)o(exp(7ra-3d^?))°a-^°T^(i^) [by (14) and (2)]
= TTexp(iz) 0 0-^0exp 77-(fc)oT^(z;) [by (12)]
= Tr*(exp(fl),o-)(Tr*(exp(6),r)(i;)).

We now define the Lie group G = G(P, X ) (with respect to a restricted


weight lattice P and a separating group of characters J^) of a Kac-Moody Lie
algebra g as follows:

G — G^/Kp,

where Kp •= fl tt*, the intersection being taken over all integrable represen­
tations ( tt, M) of g in J^(P).
The group G = G(P) depends on the choice of weight lattice P. (Later in
Proposition 6.3.15 we will see that G is essentially independent of P,) In turn
G = G { P , X ) depends also on the choice of the separating group of charac­
ters X. For simplicity of notation we will write G and G in the remainder of
this section. As above, “integrable representations” will be understood to
belong to J^(P). Since no confusion will arise, we will still denote the
elements of G by (g, \ ) (instead of (g,
There are obviously homomorphisms

ic :G G^ G,

G* G.
6.1 Integrable Modules 503

We denote the image of in G by X, (Since G actually embeds in G, we


will continue to denote its image by G.)
^ Given an integrable g-representation ( tt, M) we have the mapping itx •
it Aut[|^(M) with X ^ Xtv;: Evidently factors through X to give itx '•
X Auto^(M). Given e ^ we will usually denote its image in A" by
then TTxix) = Xv’ which acts on by multiplication by ;^(^) as before.

Proposition 12

(i) The homomorphism Iq i G G is injective.


(ii) The kernel o f the homomorphism ix • X G is

{x ^ = I for all fi e P ( M ) , and for all e

In particular if {ix\ix e P{M), (ir, M) e J^(P)} spans the Z-module


P, then ix is injective. This always happens if J is finite.
(iii) Let ( it, M) be any integrable q-module. Then there exists a unique
representation (ir, M ) of G for which

TTIg = TT,
(16)
'ñ’lx = Xtt-

Proof Let ( tt, M ) be any integrable g-module. Let g e G. Then from (15),
iXy) = 'ir(g)(i;), so 'ir*(g, 1) = 'ir(g) and (g, l ) e / ^ if and only if
g = 1. This proves part (i). Similarly for x ^ X , ir*(A')lw'‘ = multiplication
by xil^X so -a^ix) ^ K if and only if x(p-) = 1 for all p e P(M) and for all
integrable modules M.
In the case that J is finite P^ spans P as a Z-module (Remark 5). Since
for all A e L(A) e ^ ( P ) we see that [p\p e P(.M), (ir, M ) e ^ ( P ) )
spans P. This proves part (ii). Now part (iii) follows from (6.18). □

We refer to (ir, M) as the representation of G obtained from (v, M). In


particular (ad, g) induces a representation (ad, g) of G. In the sequel we will
usually simply write ir f o r ^ . In particular, since ad = Ad, Ad will be
denoted by “Ad.” Note that Ad|c = Ad^.
Remark 7 X is a subgroup of Hom(P, K^), while H can be viewed as such.
Let h e H C\ X. We can view h as an element of G in two different ways,
namely
504 Category 0 for Kac-Moody Algebras

and

H C \ X ^ X - ^ G.

These in fact give the same element of G, In particular, ii x ^ H C\ X, then


in G,

( aT>1 ) = ( 1 >AT)-

Thus G n AT is not trivial in general. We look at this in more detail in


Section 6.3.

Proposition 13
Let q be a Kac-Moody algebra, and let G be its Lie group. Let ( tt, M) be an
integrable representation o f g, and let g ^ G and y ^ Q. Then

(0 ir C g M y M g ) - ' = ir((Ad iX y)),


(ii) 'ir(g)exp ir(y)'ir(g)~‘ = exp Tr((Ad |X y )) if Tr(y) is locally nUpotent.

Proof. It suffices to establish the result for § = (exp x, for x g g“, a e "A,
and X ^ Now

* ( exp X, AT) y ) -n-(exp X, AT) ” ‘

= ir(expx)Ar^Tr(y)Ar“ H'Tr(expx))"'

= 'ir(expA:)Tr(Arad>')'>r(expx)"‘ (12)

= ir(Ad(exp Jc)ATad(y)) [by (4)]

= ir(A d(expx,A ')(y)).

This establishes part (i). As for part (ii) one just has to expand expir(y). □

Remark 8 For each x e g“ and a e "A we have defined an element


exp X G G with the property of equation (2), namely ir(exp x) = exp tt(x) for
all ( t t , M ) g ^ ( P ) . At this point we cannot extend this definition to all
X G g for which wix) is locally nilpotent in every integrable representation
( it, M) since we do not know whether or not G contains an element that
induces exp tt( x ). In Chapter 7 we will show that, this is actually possible if g
is an invariant Kac-Moody algebra. For now, however, we can make the
following observation.
6.1 Integrable Modules 505

Proposition 14
Suppose that some x g g has the property that tt{ x ) is locally nilpotent for all
(tt, M) G cX(P), and suppose that we have an element exp jc e G satisfying
(2) (it is evidently unique). Let g ^ G. Then

(0 Tr((Ad gXx)) is locally nilpotent for all ( tt, M )7rexpjc = exp7r(x) G


J^iPX
GO there is a unique element exp((Ad gXx)) ^ G for which (2) holds {for
Ad g{x)\
Gii) exp((Ad gXx)) = g exp{x)g~\

Proof Part (i) follows from Proposition 13G). By Proposition 13(ii),

exp-n-((Ai/g)(A:)) = Tr(g)expir(;»:)'ir(g)"‘ for all {v,M )^J^(P).


Thus, if we define exp((Ad gXjr)) by part (iii), it follows that (2) holds [for
(Ad gXj:)]. Evidently (2) guarantees that exp((Ad gX;c)) depends only on
(Ad gXjr) and not on the choice of g and x. □

Here is an example. Using Proposition 14(iii), for all g e G and for all
a e IK,

exp((Ad g)ie„) = g exp te„g~^ e G for all a e "A+, t e IK.


In particular, if g ''^A, 5 g

hp(s)exp te„h^{s)~^ = exp(Ad

= exp(s<“’'^''>re^)
by Proposition 8(ii). Thus

hp{s)exp t e^hp{s)~\ exp tej~'^


(17)
= exp((i<“'^''> - l)ie„).

We have similar results for with a g '^^A_.

Proposition 15

(i) G < G, G = GX, G /G = X /G n X.^


(ii) (G ,G ) = G.
(iii) G is perfect [(G, G) = G)].

^In Corollary 6.3.12 we will prove that G n X c. H .


506 Categoiy ^ for Kac-Moody Algebras

Proof. Part (i) follows from G < G ^ and G^ = G X. Since G /G is


abelian, (G, G) c G. To prove part (ii), we need only to prove part (iii).
However, (17) with a = ^ and ^ = 5 gives exp(24fe^) e (G, G) for all a e "
and f e K. Since similar considerations apply to /^, and since G is
generated by the exp(g“), a g '^^A, part (iii) holds. □

We finish this section by studying in detail some representations of G.


These results will not be used until the conjugacy theorems of Chapter 7.

Proposition 16
Let be integrable representations in ^ ( P ) o f g, i = 1,2. Let M =
Hom|,^(Mi, M2), and let ( tt, M) be the representation o f g given by

7 r ( x ) f •= 7T2( jc )/ —f'TTi(x) for all X ^ g , / £ M.

Define a representation ( tt, M) o f G by

■w(g)/=

Then for fl// X G g“, a e ''^A we have

(18) exp 7r{x)f = Tr(exp x ) f .

In particular, if ( tt, M) is an integrable ^-module, then tt coincides with the


representation of G obtained from ( tt, M) by Proposition \2{iii).

Proof The argument is a variation of the proof of Proposition 2. Let


a e '■^A, X e g“. Let v ^ M^, and define M[ to be the finite-dimensional
space generated by the action of 7Ti( a:) on í; and M^ the corresponding space
for 7T2( x ) acting on fM[. Then /|^fj e Hom¡,^(MÍ, M ^, and applying
Lemma 1 with x^ -= irfx), we obtain

exp 7r(x)f = exp 7T2(x)/(exp 7Ti( jc)) ^ on M[.

Since V was arbitrary,

exp 7r{x)f = exp 7T2(x)/exp 7Ti( x ) “ ^ == Tr(exp(x))/.

Thus (18) holds.


Suppose next that M is integrable. We show that

X>n‘2f^'Tri Xvf‘
6.1 Integrable Modules 507

For this we may assume that / e M**. Then for mj e Mj we have

XtT2J ^TTi XtT2

j^(A + fi)
/( m i) [sin ce/(m i) e
at( a)

= a ( m) / ( " I i )-

Having established this fact, we see that

1-1
■rr(exp X, x ) f •= 'iT2(exp x, ;if)/Tri(exp x, at) '

= •rr2(exp x)Ar^^/A-;:/'iri(exp x)

= <iT2(expx)Ar^/iri(expA:)“ ‘

= ir(expA:)Ar^/,

which is precisely the representation of G obtained from (ir, M).

Proposition 17
Let (tti, M j) and (772, be integrable representations o f g. Then the tensor
product representation ( tt, M) == (77^ <S>772, ® M2) o f g is integrable, and
the corresponding representation ( tt , M ) o f G satisfies T rC g X m j <S>m 2 ) =
TTi(gXmj) 0 'TT2(gXm2) for all g G: G, E: Mj, ¿znd m2 ^ M2.

?roo/. Let jc e g“, a e Then

IT (exp x)(m i 0 m2) = exp 7r{x){m^ 0 m 2)

“ 77(x)”(m i 0 m 2)
n= 0 n\

= £ £ i ” ]^(^i(^)" ' w i ®'n-2(x)'m2)


„ = 0/ =0 V‘
= e x p ir i(A :)m i ® e x p 7 T 2 (x )m 2

= 'tTi(exp x ) ( m i ) ® 'iT2(exp x ) ( m 2 ) .
508 Category ^ for Kac-Moody Algebras

For e A", e M^ \ and ^ we have


'Tr{x){mi ® m2) = ® ^2)
= AT(Ai + A2)mi <S>m 2
= Ar(Ai)mi <S>x(^i)m2
= ^ i ( x) ( f n^ ) ® 'n-2(Ar)('W2).
We can now apply Proposition 12(iii).

6.2 WEIGHT SYSTEMS

In this section we study the weight systems P(M) of integrable highest


weight modules ( tt, M) of a Kac-Moody algebra g. We give two characteriza­
tions of P(M) as well as algorithms for constructing P(M) and the dominant
weights in P(MX
Throughout this section g is a Kac-Moody algebra over a field IK of
characteristic 0 with structure matrix A and corresponding triangular decom­
position T = (ij, ¡2+7 E denotes the standard weight lattice. We use
the notation of Sections 4.2 and of 6.1.
Let M ^ J^, the category of integrable g-modules, and let A lie in the
weight system P(M) of M.

For each a e the a-weight string through A is defined by


5(A, a ) = A, O') == (A + ka\ k e Z} Pi P ( M )
If M = g under the adjoint representation, then P(M) = A(g). If j8 e A
and a e ''^A, then S(p, a) is the a-root string through p. The terminology is
justified by the following (cf. Proposition 4.1.13):

Proposition 1
Let A e P(M ), and let a ■"A.

(i) Either
Sooi 5(A,o') = {A + koc\k ^ Z}
or
Sfin* a.) = [k ka — d < k < u] for some d and v in N.
In either case the reflection r^ reflects the string 5(A, a) about its
midpoint p := A - <A, >a. Moreover, if holds, then
SI: d — u = (A, a ^ ),
S2: the sequence {dim is increasing up to its midpoint and
then is decreasing, and
S3: the sequence (dim is symmetric about its midpoint,
(ii) If M ^ then only the case 5^^ can occur.
A sequence with the properties S2 and S3 is called unimodal.

Proof, (i) Since a e ^1^2^ = g“ © [g“, g"“] 0 g““ is isomorphic to


^l2(k) (Proposition 4.1.6X By Proposition 6.2.3, M is an integrable ^1^2^-
module. Let N ■= Then N is an §I^2“^"S^t>^odule of M and
hence also integrable. Furthermore P( N) = 5(A, a). By Proposition 2.4.4 we
can decompose N into irreducible ^1^2^^'Submodules. Each of these irre­
ducible submodules has the form K = where acts diago­
nally on by the scalar s - 2i and dim = 1 for each i (Section
2.4). Since acts diagonally on by <A, > + 2k.
K ^-2i ^ ))/2-i] a^

Thus already s = <A, a ^ >(mod 2).


We now repeat the Freudenthal midpoint argument of Section 4.1. The
midpoint of this set of weights, as i runs from 0 to s, is

5 - < A , q:'^> \ 5- ( X , a ^ y

jIA ------- -------- —5 a + A + ------- -------- a

= A - -<A,«^>a,

which is independent of 5. Note that /x is a weight of the string if and only if


(A,q:^> = 0 (mod 2). In any case all the irreducible submodules K involve
sets of weights with the same midpoint and they all involve / or all involve jl

¡JL + depending on the parity of <A, a^ > . If the dimensions of the


submodules K are not bounded, then 5(A ,a) = {A + ka\ k e Z}. If, on the
other hand, the dimensions are bounded, then taking K of maximal dimen­
sion, we have P(K) = P(N). Then 5(A, a) is finite and of the desired form,
and {dim is unimodal. Since

^(A + ua + A —da) = jjl = X — |(A ,a^)a,

we have d — u = (A, ).
(ii) Assume that M If the dimensions of the irreducible factors
used in part (i) were unbounded, then there would be an infinite number of
factors. The fact that each of these factors contributes one dimension to
either or (depending on the parity of <A,a'^>) contradicts
Me □
Remark 1 If M is a highest weight module for g with highest weight pair
(A, Ü+), then for any A e P { M \ A A there exists at least one a e II for
which A + a e P{M). This follows from the fact that Á¥^A=>Á = A — p,
P ^ Q+ and is spanned by elements of the form f ¿^... f¿^ • u+, where
* +a¿^ = This trivial remark is extremely useful for inductive
proofs on weight systems of highest weight modules.
510 Categoiy ^ for Kac-Moody Algebras

Let A and consider the Q+-fan A iQ ^ . We define the depth function


dj, : A i Q ^ ^ N
by

E c , a , .) = E cy .
;e j / yej

Let A e F, and let L be an integrable highest weight module with highest


weight A. Then according to Propositions 6.1.5 and 6.1.6, A ^ F+ and there
are homomorphisms L^J^A) ^ L -> L(A) [the last because of the definition
of L(A)]. The next result is an algorithm for constructing F(L).
We denote this common set of weights by F(A).

Proposition 2
Let L be an integrable highest weight module with highest weight pair (A, i;+),
A e P^. For all n let PJiL) denote the set o f weights o f depth n in P (L \
Then

(0 Po(A) = {A},
(ii) A e A i 2+ with d ^ k ) = n > 0, then A e F„(L) if and only if there
exists an / e J and a k ^ such that A + ka¿ e P„_^(L) and
(A + ka¿, ) > k.

Proof Evidently Pq(L) = {A}, since L = U(g_) *í;+. Let A e P^iL), n > 0.
By Remark 1 there exists an / e J such that A + a¿ e F(L). The «¿-string
S(k,a¿) through A has the form {A + Then (A + wa¿,«¿^> =
(k, 2u = d - u 2u = u d > u. Conversely, if A e A J. 2+ and
A + ka¿ e P^_¡fL) for some i e J and A: e Z+, where <A + ka¿, ) > k,
then 5(A + ka¿, a ^ ) = {A + ka¿ + pcíi)-d^p<u’ where d - u = (k ka^,
) > k . Thus d > k, and hence A e F(L). □
Corollary
Let L be an integrable highest weight module with highest weight A e Then
P ( L ^ J A ) ) = P(L) = P(L(A)).

We denote this common set of weights by F(A).


Proposition 3
Let A e ]^*, then rad(g)L(A) = (0).

Proof This is an application of Proposition 2.7.5(ii). □

Integrable modules arise directly in the study of the adjoint representation


of a Kac-Moody algebra. We will assume that g = R) is a Kac-Moody
6.2 Weight Systems 511

algebra with structure matrix A = (Ajj), i, j e J. Let 0 # K c J, and set


9k = e K> + ^ c g.
Then 9 k is a Kac-Moody algebra with structure matrix Ay^ ■= (.<4,^); y^K-
We view g as a gK-module by restricting the adjoint representation. Set
G k = ® 2 a, c Q,
i S K

and note that


9k = ® 9“,
ae Ak
where = A (^) n [s®® Section 4.3, Remark 2, and 4.13)]. Thus g^
admits == (^, ®„eAKnA+9“>G k ^ G+^o'lgK^ ^ triangular decomposi­
tion.
For each coset a e G /G k define
Ms = E 9 ^

We have the following facts.

1* 9 ®a€ Q/Qk^ ^ '


2. is an integrable If ^ suppose that
á = a + Qj^, where a e A. If in addition a e then a e and
M- c is an ideal of If a ^ A \ A^, then a + n A c A+,
and hence ^ 9 ± according to whether a e A +.
3. For all á for which ^ (0), either

is an ideal of g^, or
c g_ and e or
c g_^ and Af^ g ^ k)'

It is surprising how little we know about the modules M- when a e A \


A^; not even whether or not they are completely reducible (unless g^ is
invariant). However, the modules when / e J \ K are quite accessible.
For simplicity we treat the cases M_^..

Proposition 4
Let q be a Kac-Moody algebra with structure matrix A = (.^4¿y), /, j e J, and
let K be a nonempty proper subset o f J. Let / g J \ K. Then with the notation
above we have

(i) M_s. = ad u(9 k )/„


(ii) M_^. is an integrable highest weight module with highest weight pair

(iii) P(.M_¡¡) - ( —a, -1- Gk ) ^ M“ s = g“ for all a e


512 Category ^ for Kac-Moody Algebras

Proof. Set M := ad uCg^)// ^ Froni ad = 0 for all A: e K and


ad /t(/¿)= —{a¿,h)f¿, we see that M is a highest weight module with
highest weight pair ( —a¿,f¿). To show that c M , we consider any
-a¿ - a e A, a e and prove that g““'" “ c M. Certainly a e (Q k ^
Q- ) U {0}. Thus every element of g“"'“" is expressible as a linear combina­
tion of elementary words F = [ f , f j , where . . . , ^ {0 U K and i
occurs exactly once in each of these elementary words. However, if i = ip
then

^ ^ [//,’ •••’ [fij+i’ [fi’ ***]

- -ad/,,...a<l/,,.,ad([/,,.,sM .

Thus M = M_^p and all the results follow. □

We use Proposition 4 to obtain a characterization of P(L) for an inte­


grable highest weight module L. Let g be a Kac-Moody algebra with
structure matrix A, and let L be an integrable highest weight module of
highest weight A e In determining P(L), we can suppose that L = L(A)
because of the Corollary to Proposition 2. By Proposition 3, rad(g)L(A) = (0),
so we can suppose that g = of A, R) is radical free. We now carry out an
extension trick that allows us to view P(L) as a set of roots of a Lie algebra.
Increase J by adjoining a single element /; let j == J U {/}. Define a matrix
A = (A¿jX i, ; e J by

Ai¿ = < - A, G Z^o for all i e J,


A¿i arbitrary in Z < q subject to ^4/^ = 0 ' A ¿I = 0 for all i e J,
An = 2.

Then ^ is a Cartan matrix, and ^ is a submatrix. If A is symmetrized by {e¿},


then Á will be symmetrizable if we set Án = A^Si for each i e J.
Now extend to a realization R = (§, H, of Á in which R embeds
compatibly with the inclusion J ^ j. Then we can view g(^4, R) as a
subalgebra of qf Á , R) (Proposition 4.3.5). Denote g (^ , i?) by g, and let
J + § = 9 + § = 9 + § c g. According to Proposition 4, M_-
is an integrable gj-module of highest weight —a^ Restricted to g, it is then
an integrable highest weight module of highest weight A [since —<«/, a f ) =
- Á n = (A, a f ) ] and is generated by f¡. The weights of are the roots
of g of the form -Ui - 'Lj^jCjUj, Cj e [\1.

Let A e 1^*. Let A e A J, ¡2+? and write

A = A —a
6.2 Weight Systems 513

for some a e Q+u{0}. Let be the connected components of


supp(a) (Section 5.8), and write

“ = L “ 5,>
/ = 1, m

where supp(a^) = 5^. We say that A is connected through A if either A = A,


or for each \ < i < m there exists some j = ;(0 ^ such that <A, a / > =5^=0.

If A = A - a, as above, is connected through A and one draws a


diagram of supp(a), then all the connected components of supp(a) are
connected to each other “through” A

Proposition 5 [Ka5]

Let L be an integrable highest weight module with highest weight pair (A, u+).

(0 Every weight o f L is connected through A.


(ii) I f k ^ P +n(A i Q^X then A e P(L) if and only if A is connected
through A.

Proof (i) Suppose (i) does not hold. Then there exists a weight A of minimal
depth for which part (i) fails. By Remark 1 there exists a e j2+u{0} and
y e J such that A = A - a - and A — a ^ P(L). Because of the minimal­
ity of A, the weight A —a is connected through A. Thus aj ^ supp(a),
{a, a ^ ) = 0, and <A, a / > = 0. Therefore <A - a, a / > = 0 and a - aj ^
¡2+. Hence A - a + Qfy ^ P(L), and by Proposition 1(0 it follows that
A = A - a - ay ^ P(L), a contradiction.
(ii) By the extension trick we may take L = in some Lie algebra
g = 9j, where j = J U {/} and L certainly have the same weights even
if they are not isomorphic g-modules. All the weights of lie in
- ai - (j2 +L^{0}X We show that if a e -a ^ - (2+, { a ^ a ^ ) > 0 for all
y e J (a is dominant relative to J), a is connected through -a^ , and if we
choose the extended Cartan matrix A appropriately then a is a root of §. In
that case a e by Proposition 4(iii). “Connected through - a ^ ” is
514 Category ^ for Kac-Moody Algebras

equivalent to saying that supp(o:) is connected. By Theorem 5.8.6 we will be


done if we show that a is dominant relative to J (then —a e F).
Let a = —a i ~ ^ Now (a , a / > = —2 —'LcjAji, and by
assumption CjAji 0 for at^ least one j e J. Since the Aji can be chosen
freely, we can assume that Aji = - 2 for each j ¥= /, for which Aji ^ 0. Then
(a, a / > > 0, which is what we want. □

A subset 5 of the weight lattice P is said to be saturated if it satisfies the


following condition:

SAT: If A - S, a ®A, and k is an integer between 0 and <A, a ^ >, then


\ - ka : 5.
If A e 5 as above and a e then r^(A) = A - <A, a e 5. It follows
that saturated sets are IT-invariant. The concept of saturation occurred
previously in Section 5.8 where we discussed the root string property.

Proposition 6
Let M e Given a positive integer d let

P( M) a == {a e P( M) \ d i m^ P ( M) ^ > d).

Then P{M)^ is a saturated set. In particular P{M) = P{M)^ is saturated.

Proof. The proof follows from Proposition 1. □

Proposition 7
Let A e The following sets are equal:

(i) (ITA> n (A + <2X y^here (ITA> denotes the convex hull o f WK.
(ii) The weight system P(A) of L(A).
(iii) The smallest saturated subset 5(A) o fP containing A.

Proof (i) c (ii) Let A e <ITA> n (A + Q). Then A = "L^^^^c^wA with


> 0 and = 1. Thus A —A = 'Ly^^цrcJ^A — wA) e (2 +Li{0} since A -
wA e j2+u{0} and A — X ^ Q. Since <^A> n (A + 0 ) and P(A) are both
IT-invariant, there is no loss of generality in assuming that A is chosen in its
IT-orbit so that djfX) is minimal, or equivalently, so that A e P_^. Finally,
A = H^^iyC^wA is connected through A since wA is connected through A
for every w e IT. By Proposition 5(ii), A e P(A).
(ii) c (iii) Let 5(A) be the intersection of all saturated subsets of P
containing A. It is clear that 5(A) is saturated, it contains A, and is the
smallest subset of P with these properties. By Proposition 6, 5(A) c P(A).
6.2 Weight Systems 515

Suppose that 5(A) P(A), and choose A e P(A) \ 5(A) so that d ^ k ) is


minimal.
Since A e 5(A) by definition, A =5^ A. By Remark 1 there exists ; J such
that A + ofy e P(A). Let d — u = (A, a^) be as in Proposition 1. Then u > 1
and A + uaj 5(A) because of the minimality of d^(A). Then A + « a j -
ta: 5(A) for all 0 < t < <A + a / >. Since <A + uaj, a / > = u+
2M = d + w>w, A e 5(A). This contradiction shows that 5(A) = P(A), as
desired.
(iii) c (i) It is clear that <ITA> n (A + ¡2) is saturated because of convex­
ity, and that it contains A. Thus 5(A) c (WA) n (A + Q). □
Proposition 8
Let A G and let A e P(A). Then for all ¡jl e (WX) n (A + Q),

dim L( A)^ > dim L(A)"^

Proof Without loss of generality, we may assume that A e P_^_, Let


dimL(A)'^ = d. The set P(A)¿ of weights of multiplicity > d is saturated.
(Proposition 6), and it contains A. By Proposition 7, applied to A, {WX) n
(A + 0 = (WX) n (A + ¡2) is the smallest saturated subset of P containing
A. Thus <ITA> n (A + G) c P(A)^. □

Proposition 9
Suppose that ( q, ¿s an invariant Kac-Moody algebra and let (*| * ) be a
standard invariant bilinear form on 1^*. Let A e Then for all X, p ^ P(A)
we have

(A|/x) < (A|A)

with equality if and only if X = p ^ WA.

Proof Using The IT-invariance of ( | • ) we can assume that A e Write


A = A - a and p = A — p where a, p ^ Q + u{0}. Then

(X\p) = (AIA) - (A la) - (A|i8) < (A|A)

since (A|a) > 0 and (X\p) > 0.


Suppose that we have equality. Then (Ala) = 0 = (A|j8). Now A e P(A)
and so A is connected through A. But this means that (Ala) = 0 if and only if
a = 0. Thus A = A and now a similar argument gives p = 0. The result
follows. □
516 Category ^ for Kac-Moody Algebras

63 THE TRIANGULAR DECOMPOSITION OF G

The derived Lie group G associated with g has remarkable pair of decompo­
sition over the Weyl group W called the Bruhat and Birkhoff decompositions.
The latter gives us something resembling a triangular decomposition: G =
U ^^цrG_wHG+, and this decomposition is essential for proving the conju-
gacy theorems in Chapter 7. Both decompositions follow readily from the fact
that G can be realized as a Tits system or (B, A/^)-pair. This immediately
associates it with a large class of groups that are of this type and for which an
abstract chanpiber geometry exists. The theory of buildings, as these chamber
geometries are called, and (B, A/^)-pairs was developed by J. Tits [Til, Ti7] in
order to provide an abstract geometric scheme in which one could study all
the simple (finite-dimensional) Lie groups in a uniform way and over arbi­
trary fields. In this section we prove that G can be realized as a Tits system
and exhibit the triangular decomposition. The exposition follows [PK]. The
first results of this type for the infinite case are in [MoT].
Throughout this section (g, denotes a Kac-Moody Lie algebra over K.
We write ^ = (if, g+, Q+, o r ) as usual. Let P be a weight lattice of (g, ^ ) .
All the standard notation previously employed is in use.
Let G := G(P) be the derived Lie group of g with respect to P. We recall
some important subgroups of G that have already been defined

G“ := expg«, a
G + := <G“la
G_:= <G“k
^ = = ( « a ( 0 |« t e IK=^),
H--={h^{t)\a e"A , i e IK^).

Remark 1 The groups G “ are often called one-parameter groups in the


literature on Chevalley groups and are usually denoted by t/“.

Lemma 1
Let a, p e^'^A. Then G“ n G^ = {1} whenever a # jS.

Proof, Suppose that a p and that z e G" n G^, z ¥= 1, Then there exists
X e g“ and y ^ g^ such that exp x = z = exp y, x ,y ¥= 0. Thus

exp ad X = Ad exp x = Ad exp y = exp ad y .


Since

X = expad x (x ) = expad y (x ) = x + [y, x] +


6.3 The Triangular Decomposition of G 517

and (ad y T x g we conclude that [x, y] = 0. Thus [ad jc,ad y] =


ad[jc, y] = 0, and hence

1 = exp(ad jc)exp( —ad y) = exp(ad jc —ad y) = exp ad(x: —y).

Choose h with a(h) ^ ¡3(h), Then for all f g IK,

ad(x —y)^
h = exp(ad t( x —y ) ) h = /i + t[x — y, h] + ------r:------ h
2!
2iá(x-yY
+ ••• --------— h
nl

for some n independent of t. This can be true for all t if and only if
[x - y,h] = 0. Thus a(h) = contrary to the choice of h. □

Define (resp. 5 _ ) to be the subgroup of G generated by H and G+


(respectively H and G_).

Lemma 2

(i) H normalizes G"" for all a g In particular H normalizes G+ and


G_.
(ii) H • G^ and B_= H G_, H n G+= {!}, H n G _= {!}.
(iii) B ^ n N = H = B_ DN,

Proof. Let G g“, a g '^^A. Then for all p g '’^A and i g we have
[Proposition 6.1.14(iii)]

h p ( t ) e x p x ^ h p ( t y ^ = exp{Ad hp(t )x^)

This shows that H normalizes G", G+, and G_, and also that G+ is a
normal subgroup of A typical element of G +n / / is of the form

exp(xi) • • • exp(x„) = h y t i ) • • • / t J i J ,

where x¡ e g^', /3, and y, e''^A. Given any integrable module M and
any weight vector v e M'^, we have expixj) • • • exp(x^) • v = v + v', where
u' G On the Other hand (Proposition 6.1.8),

¿=1
518 Category ^ for Kac-Moody Algebras

Combining these two, we see that • • • hy(t^) acts like the identity
on M. From Section 6.1, Remark 1, we have = I q , proving
that G +n / / = {1}. Similarly G_D H = {1}.
Clearly H (zB+n N. Also G+n N = {1}. Indeed, if p e F+ is an integral
minimal regular weight and is a highest weight vector for L(p), then
G +*u^= u^, whereas n • +c L(p)^"^^ for n ^ N. But [n\p = p <=> [nj = 1
^ n ^ Thus G +n N = G + n H = {1}. Now x e B ^ D N ' =>x = /ix+e N,
B+n
where h ^ H and x , ^ G , , and hence x. N. Then jc+= 1 and x ^ H.
Similarly for B _n N. □

Often we denote B+ simply by B. This group is called the Borel group of


(g, Recall that if x and y are two elements of a group K then
(x, y) ’= xyx~^y~^ is called the commutator of x and y. If and A2 are
subgroups of K, then we will denote by (^1,^42) the subgroup of K
generated by all commutators (^i, a^) with e A^, i = 1,2.

Lemma 3
Let a, be such that (a, > 0. Then

(i) (Na + Np) n (A \ {0}) = {a, p ,a + p) n""A,

Proof, (i) Let y = m a + e A, m, n e I\l. Then

( r , P ^ ) = m( a , P ^y + 2n > 2n > n,

which shows that the p-ioot string through y contains y — np = ma


(see Section 6.2). Since a e^’^A, we have m < 1. Since (a^p"^) > 0 <=>
{ p , a ^ y > 0 (Proposition 5.2.8), we have similarly n < 1, and hence y e
{a, )8, a + p}. The above argument shows that 2y ^ A. Thus y e'^^A (corollary
to Theorem 5.8.6).
(ii) Let X e g", y e g^. Then by Proposition 6.1.14,

exp X exp y(exp jc)” \e x p y)~^

= exp((Adexp x )(y ))(ex p y) \

Now by part (i).

(A d ex p x )(y ) = y + [x ,y ]
6.3 The Triangular Decomposition of G 519

and
[y + [ x , y ] , y] = 0.
Thus

exp(y + [x ,y ])(e x p y ) ^ if a + j8 e'^^A,


(exp jc, exp y) =
expy(expy)"^ ifa + )8 ^ " " A ,

exp[x, y] if a + /3

1

For more on pairs of real roots see [BP2]


Remark 2 In the sequel we have many occasions to write down sets that
involve elements from the subgroup N, for instance, BnB, n ^ N. Often such
sets depend only on n mod H, that is, on [nj e W. For instance, if A e //,
then BnB = BnhB, so BnB depends only on nH = [nj. If [nj = iv, it is then
convenient, and standard, to replace « by w in the expression. Thus we write
BwB even though w ^ G . \ i both n and n~^ occur in this expression, then
the convention is to replace n h y w and n~^ by In other words, one
must understand that and stand for elements of N that are inverses
of one another. For instance, we will write rJ3^r~^ meaning
We will use these conventions without further comment throughout the rest
of the book.
For each a e II define a subgroup y of G+ by

7 “ := e G", y e G ^ p e^^AA{«}>.

We note that for p e '^^A+\{a}

(1) c Y“,
since

'■= r^cxp ^ = exp g'"«^

and r„(A+\{a)) = A+\{a).

Proposition 4
For all a ^ Yl we have

(i) SL^2^ normalizes 7 “,


(ii) G^= G^ Y ^,G ^ = {!},
(iii) Y“ = G + n r„ G + r;i.
520 Category ^ for Kac-Moody Algebras

Proof, (i) By Proposition 2.4.5, SL2(K) has a Bruhat decomposition. By


Proposition 6.1.7, SLjClK) = by a map with

expie^
"’( J ! ) -

5 ) = e x p t/„

where [e^, a /^} is some fixed, but otherwise arbitrary, ^ 12-triplet for
Combining these two results.

( 2) 5L<2“ ^ = U

where

5(“) := / / “G “, / / “ == {/i„(0k e IK''},

and = }n„(0J = [exp exp - i “ exp re„J [see Proposition 2.4.9].


Now Y “ is clearly normalized by H and G “, hence Thus we need
only show that
y a - 1 (- y a
ot ct

For this we need only show that for all x e G “, for all /3 e''®A \ {a},

r„x G ^x -V -i c y “.

If ( a , /3 '') > 0 then from Lemma 3,

(3) x G ^ x 'i c G“+^G^ c 7 “.

In fact from (1) and (2) we may deduce that (3) holds for all x e SL^2^ (still
assuming that <a, ^ 0).
If <a,j8 '^> < 0, then

= x 'G ''« ^ (x ')'‘ c 7 “

from (3) since x' g 5L^2^^ and <a, = —(a, j 8 ^ ) > 0 .


This proves (i).
(ii) y " is normalized by G“, and hence G ^y" is a subgroup of G^.. Since
it contains the generators of G+, it is G+. We show that G" n y “ = {!}.
Elements of G^ act trivially on the highest vector of any integrable
6.3 The Triangular Decomposition of G 521

representation L(A). However, if A e is chosen with <A, a'^ ) > 0, then


[r„](expie„)fr„]~^ = exp tf^ e G ““ for some r e K, and this does not fix
v+ if s # 0. Thus e G “ \{1} => i G+=> €
7 “ ^ X ^ 7 “.
(iii) Let g e G+, and write g = xy when x e G“, y e y". Then using (1)
\fa]g\ra]_ ^ ''al3'r''al ' ^ SinCC G “ € G +,
^ if and only if X = 1. Thus <x+nfr„]G +fr„]~‘ c 7 “. The
reverse inclusion is obvious from (1). □

Let 5 := { r j a e II}.

Proposition 5
(G, B, N, S) is a Tits system (see [Ti3] and [Soi]). By definition this means
that

TSl: G is generated by B and N and B N < N,


TS2: The elements o f S are of order 2 and S generates the group W ==
N /B n N,
TS3: For all s ^ S and w ^ W,

sBw c BwB U BswB,

TS4: For all s ^ S, sBs <t B.


Remark 3 According to TS2, W is defined 2ls N /B n N, We already have
a definition of W but, by Lemma 3, B n N = H so the two definitions
coincide. The group py of a Tits system is always a Coxeter group. Tits
systems are also called (B, AT)-pairs.

Proof of Proposition 5.

TSl, TS2: For each a G “ c G+. Also for each w ^ W,


fw]G“fH'l“ ^ c G*^“. Since we have G“ c for all
a e"^A; hence G = <B, N ). Also B n Af = / / <l and N /H = W.
TS3; Let s = S, and let w ^ W. There are two cases:

Case 1. w e A+: Then

sBw = sHG“Y “w
= sY “G“Hw
= sY^ssG^^ww-^Hw c Y ‘‘sG°‘wH

= Y “swG”'~'“H c BswB.
522 Category ^ for Kac-Moody Algebras

Case 2. e A_: Then (sw)~^a = w~^r^a = -w ~ ^a e

Now

sBs = sHs-^sG^s-^sY^s-^

c B u BsB

by (2). Thus

sBw = 5B55IV c (B U BsB)sw


c Bsw U BsBsw
c BswB U BsswB (by case 1)
= BwB U BswB.

TS4: sG°"s = G “ B; hence sBs € B. □

In the main result we go back to the original notation and use B+ instead
of B.

Proposition 6 (The triangular decomposition of G)

G= U B_wB^ {disjoint union).

In particular G = G_NGj^.

Proof. Evidently the right-hand side is closed under right multiplication by


elements of B+. We show that it is also closed by right multiplication by
elements of N. Then, since G = <B+, A/'), the right-hand side is indeed all
of G. It suffices to show that B_wB^s c B_wB^U B_wsB^ for each s =
a e n . There are two cases:

Case 1. wa e A_: Then

B_wB^s = B _wHssG^ssY°'s
(zB_w sG -^Y^
c B_G~^^^wsBB_wsB^.
6.3 The Triangular Decomposition of G 523

Case 2. wa ^ A^_: Then wsa e A_, and

B_wB+s = B_wssB^s c B_ws{B^yj B ^sB ^) (byTS3)


c B_wsB^\J B_wsB^sB^
c B_wsB^\J B_wB+ (by Case 1).

It remains to prove that the union is disjoint. Now any two (B_,B+)-dou-
ble cosets are either equal or disjoint. Consider B_wB^. Let L(p) be the
representation appearing in the proof of Lemma 2(iii), and let be a
highest weight vector of Lip). Then B _ w B B _ i n v ^ ) , where [n\ = w.
We have nv +^Lip)"^^ and B_nv +(z K^nu +-\- since this last
set is stable under G “, a e'^^A_, and H. Evidently, if B_wB^= B_w'B^, we
must have wp = w'p, and hence w = w'. □

This decomposition is also called the Birkhoff decomposition. We leave it


as an exercise to prove the following analogous result.

Proposition 7 (Bruhat decomposition)

G = U BwB {disjoint union).

In particular G = G+A^G^

Recall the group G — G(AT, P) of Section 6.1. We finish this section by


taking a closer look at the centers Z(G) and Z(G) of G and G, respectively.

Proposition 8
Let (g, ^ ) be a Kac-Moody Lie algebra and

Ad: G GL(g)

the adjoint representation o f its derived group. Then

(i) k e rA d c Z(G)
(ii) ker Ad = Z{G) if Q is radical free.

Proof. Let g ^ G, and let 5 — {jc e g“|a e'^^A}. Then

g e Z (G ) <=>g exp xg~^ = exp x for all x e 5


<=> exp(Ad g (x )) = exp x for all x e 5 [Proposition 6.1.14(iii)]
Ad g (x ) = X for all X e 5 [Proposition 6.1.14(ii)].
524 Categoiy á for Kac-Moody Algebras

This establishes part (i). It also shows that if g e Z(G), then Ad g


pointwise fixes Dg. Part (ii) will follow once we establish that

(4) (rad(g) = (0) and g e Z(G)} => Ad g (h ) = /i for all

We begin with two preliminary steps.

S t e p 1. Let A e and let v e L(A) be such that exp v(e)v = v for all
e e {ej]j^j. Then v g

We may assume v ^ 0 and write y = Uj + • • • + % , where v¡ e


L(A)'^“ '*''\{0} and y, < jj ^ i < j. From exp 7t( x ) í; = d we conclude that

Tr(e)Vi
Tr(e)vi H —------ h • • • +v{e)Vfj + = 0.
2!

Now v(e)v¡^ G L(A)^~‘>''^'^“ for some a g n , and no other term in the finite
sum above contributes to the weight space in question. Thus Tr(e)vj^ = 0 and
hence g (Ku^ since L(A) is irreducible. It also follows that y^^f = 0 and
hence that N = 1.
S t e p 2. Let A G P_^. I f g e Z(G), then g acts on (L(A), ir) as a nonzero
scalar.

Indeed let e G [e¡\j^y Then

'»»■(g)y+= Tr(g)exp'ir(e)i;+= ■rr(g)'ir exp(e)t;+


= Tr(g exp(e))ü+= ^ (e x p íc )^ )^ ^
= exp'n-(c)'ir(g)í;+.

By Step 1 ir(g)i;^e The rest is clear.


We now move to the proof of (4) above. Let /1 g 1^. By (6.14) for all
V G L(A),

■Tr(Ad g{h ))v = Tr(g)Tr(h)tr(g~^)v.

By Step 2, if A G P^ and u g L(A)^, then

ir(g )T r(/i)ir(g “ ‘)i; = ( p , h ) v .

It follows that Ad g(h) - h annihilates L(A) for all A g P^. Thus Ad


gih) = h [Proposition 2.7.5(vi)]. □
6.3 The Triangular Decomposition of G 525

Proposition 9
Let J c J', and let A c be a subCartan matrix o f the Cartan matrix
A e . Let R = (1^, П, П ^), and R' = (1^', П', П' be realizations of A
and A respectively, and let т/ = (17^, 7]¡f): R ^ R be an embedding o f realiza­
tions. Consider the Kac-Moody algebras g = д(Л, R) and g' = q'(A \ R ) with
their canonical triangular decompositions. Let P' c be a weight lattice and
let 17^: ]^'* ^ be the transpose map o f 77^. Then

(i) P ’= ri%iP') c ]^* has the structure o f a weight lattice in a natural way,
(ii) there exists a canonical homomorphism t|: Gp -> Gp, and furthermore
t|((Gp)+) c (Gp/)+ (We will see in Proposition 6.3.13 that t| is an
injection).

Proof We maintain the notation of the definition of embedding in Sec­


tion 4.3. Let < * , ‘ ):]^ * X ]^ ^ IK b e the natural pairing.
(i) P is a subgroup of Í)*, We check the axioms of weight lattices:

WLl: Let i e J. For ft e we have

= {a¡,h},

and hence a, = r,%U'i). Since Q' c P', we conclude that Q <z P.


WL2:

= {P’,r)¡,Q^) cz{P', Q:^ ) czZ.

WL3: If p is a minimal regular weight of P', then for all i e J,

<77*р',а,"> = < р ',аГ > = 1.

Thus rjj^p is a minimal regular weight of P.


WL4: Let Í1 = {(o'j\j e j'} be a fundamental set of weights in P'. For
У e J define o)j == Then for all i e J,

= {<о),г)цаУ) = = Sj¿.
526 Categoiy ^ for Kac-Moody Algebras

(ii) Let A and A' be the root systems of g and g', respectively:

G = * g'“

G — * q“
A

K p f ^ C \ ker TT, tt' a representation of g' in


Kp = C\ ker TT, 'fr a representation of g in ^ { P ) .

By definition Gpf = G/Kp, and Gp = G/Kp. By Proposition 4.3.2 we have a


natural homomorphism

tj: G —> Gpr,

If M is a g'-module and we view M as a g-module via the injection g ^ g',


then for all A e 1^'*, v e M^, and h we have h ' v '= r]p(hXv) =
(\,7]p(h))v = (7]%(XXh)v, This shows that the representations of g' in
restrict to representations of g in cX(P). It is clear therefore that
ker Kp and hence that there exists an induced homomorphism

Ti: Gp Gpr.
Finally it is clear that ti(G p)^c (Gpd+- □

Proposition 10
Let (g, be a Kac-Moody Lie algebra and let Ad: G GL(g) be the
adjoint representation of its derived group. Then the restriction o f Ad to G^ is
injective.

Proof. First assume that J is a finite set. Let ( tt, M ) be any integrable
g-representation, and let v e be any weight vector in M. Let

F := ir ( U ( g ^ ) ) i;= 0
/i, e A + ( ( 2 + { 0 } )

and let

L k = 0 ,1 ,2 ,... .
«ee+u{o)
ht(o:)>;t

Thus K == Kg D Fi 3 F2 D • • • is a filtration of V by g^.-submodules.


Let V ) := V/Vj^. Then fs a finite-dimensional Q-gi&d&A g+-module
with Jc e g^ acting as the endomorphism m + F* ■rr(,x)u + V^. Fur­
thermore the natural maps 9^: y W are g+-module maps.
6.3 The Triangular Decomposition of G 527

The endomorphisms of determined by U(g^) form an associative


algebra = K1 e where is a nilpotent ideal of (since
elements of increase the Z-degree each time they operate on
Now 7T determines a representation tt of G and hence by restriction a
representation ir of G_^. We fix some element g^= exp • • • exp g G+,
where Xj e /3^ ; = 1 ,..., r. We are going to show that if Ad g+ =
1, then ^г(g_^_)v = V as well, and hence 'rr(g+) = 1 for all integrable repre­
sentations ( tt, M), whence 1.
First, acts on as exp x^^^ • • *exp and since g we
may use the Campbell-Baker-Hausdorif formula to write this as exp(y^) for
some G of the form

(5) yk = -^1^^ + • *• +x^^^ -h 51 [ + higher commutators


i< j

(see Section 7.2 for a more detailed explanation of this). The important facts
are these:

1. Equation (5) has only finitely many terms that do not vanish.
2. The structure of the terms depends only on the Campbell-Baker-
Hausdorff formula (applied to r exponentials).
3. Consequently ^Ar+iiy^t+i) = yk for all k.
4. There exist elements z . g E,0<;<*9i such that z i ’‘^ = y ^ , k =
0 ,1 ,2 ,..., and such that = z*.+i modEhtaaA:9i- In fact we may
define z ^ ~ x ^ + ■■• +x^ + £, </;»:„ Xj] + • • • higher commutators
(using the formal expansion of the Campbell-Baker-Hausdorif formula)
where we truncate the series as soon as the Z-degree is greater than or
equal to k.

Now let us consider the case of the adjoint representation. Let y e A, and
let q e g^, V ■■=ir(U(g+)) • q, and so on. Suppose that Ad g+= 1. Then
exp y*. • g = 1 in and hence

(4) + ^ U k , z , , , q ] + ■■ ^ q m o d V , .

Let us suppose that some ^ 0 and that

^k ^n,k ^n-\-l yk ^n+s/cy ^ ’

where has degree ; in the Z-grading of and 0. (Of course


n > 0.) Then we observe that n < k and =■ is independent of k (for
k > n). From (4) we see that

[a„,q] = 0.
528 Category á for Kac-Moody Algebras

But this is true for all q ^ and hence lies in the centre Zg. But
Zg c gQ, whereas ^ n > 0. Thus = 0, and this contradiction
proves that = 0.
Now fact 4 above shows that = 0 hence that Tr(g+)u = u, which is
what we wanted. If J is infinite, we denote J, g, and G by J', g', and G'
consistent with previous notation. Let = exp . . . , exp x„ e G'+; x¿ e
g'^' be such that Ad = 1. Find a finite subset J of J' so that ...,
belong to the root system of the subalgebra

g= e © g“,
a^Q

where Q •= the last proposition we have a group homomor­


phism ti: G ^ G', where G is the group attached to g. It is clear that
has a pre-image g+ in G+. By assumption Adg. t|g+ acts trivially on g' and
a fortiori on T7g. Thus Adg g+= 1 e Ad(G). By the finite case g+= 1, and
hence g+ = 1. □

Let g ^ G, and let (M, tt) be an integrable representation of g. Then

Tr(g) e G L( M ) c E n d (M ).

From Section 1.3 we recall that irCg) is homogeneous (of degree 0) if

for all A e 1^*. We say that g is homogeneous (of degree 0) if Tr(g) is


homogeneous of degree 0 for every integrable representation (M, tt) of g.
Examples of homogeneous elements are those of H and X,

Lemma 11
Let (g, 0) e Z(G). Then (g, 6) is homogeneous.

Proof Let (M, tt) be an integrable representation of g, and let v e M^. For
all AT
A r(A )T r(g ,0 )i; = Tr(g,0)Ar(A)¿;

= Tr(g,0)(Tr(l,Ar)i^)
= Tr((g,0)(l,Ar))i^
= ir(l,A :)(T r(g,0)i;).

Thus, if w := '7r(g, d)v, then w satisfies the equation

= Af(A)iv for all x


6.3 The Triangular Decomposition of G 529

This forces w e M^. Indeed write w = + • ♦• where e M^‘ and


•••j are distinct. Then
At(A)w = ir{l,x)w = Ar(Ai)Wi + • • • + x {^n )^n -

Thus ;^(A) = = • • • = x (^ n ^ for x^ ^ind hence N = 1 and Aj = A


(since Z is a separating group of characters). □

Proposition 12
Let p be a minimal regular integral weight in P. For g ^ G the following
conditions are equivalent:

(0 g ^ H .
(ii) g acts homogeneously (o f degree 0) on (g, Ad) and L(p).
(iii) g is homogeneous.

Proof (0 => (iii) and (iii) => (ii) are obvious.


(ii) => (i) Let g ^ G, and using the Bruhat decomposition, write g = binb2
with b2 ^ B, n ^ N. Let p be a minimal regular integral weight, and let
(p, i^+) be a highest weight pair for L(p). Then

g-v^^K ^b,n-v^^L {py'"'’ + i:

since nL (p y (Z and e B. Furthermore g - v ^ has a nonzero


component in L(p)^"^^. Thus, if g acts homogeneously on L(p), then
g • +e L ( py , and hence [n\p = p. Since Stab^(p) = 1 (see Corollary to
Proposition 4.2.5), [n\ = 1 and we have g ^ B = G+H.
Write g = g^h with g + e G+, h ^ H, and apply Ad to obtain Ad g =
Ad g+ Ad h. Since Ad h is homogeneous, and we further assume that Ad g
is homogeneous, then Ad g+ is homogeneous. But Ad g+ acts on g as a
unipotent transformation, and hence Ad g^= 1. By Proposition 10, g+= 1,
and hence g = h ^ H. □

Corollary
G C ) X c : H , a n d G / G ^ X / X n H.

Proof The elements of X are homogeneous. Use now Proposition 6.1.15.


Proposition 13
Let the notation be as in Proposition 9. The natural homomorphism
T|: Gp
is injective.
530 Category ^ for Kac-Moody Algebras

Proof. Let g e Gp be such that t|g = 1. The explicit construction of P, il,


p, and r\ in Proposition 9 show that g acts trivially on (ad, g) and L(p). [Note
that the highest weight vector of L(p') generates an integrable module of
highest weight p for g. Both g and Gp act on its irreducible quotient, which
is Lip), and g acts trivially on it.]
By the last proposition g ^ H, and hence by Corollary 6.1.8(ii)

Since r\g acts trivially on L(a>}) the same argument as above shows that g
acts trivially on L((Oj). Thus tj = 1 for all j ^ J {ibid.). Thus g = 1. □

Theorem 14
Let {%, ^ ) be a Kac-Moody Lie algebra.

(i) I f P is a restricted weight lattice, then the center Z{G) o f G is given by

Z ( G ) = X q : ={ x ^ X \ x (Q) = 1}.

GO If g is radical free and P is any weight lattice, then

Z (G ) = ( n /ia X i,)| = 1 for all] e j \ c ii.

Proof, (i) It is clear from (6.1.12) that X q c Z(G). Conversely, assume that
(g, d) e Z(G). By Lemma 11, (g, 6) = (g, 1)(1,6) = g$ is homogeneous, and
hence g is homogeneous. Thus g ^ H (Proposition 12), and we can write

S = Y lhaiti)

By Remark 6.1.7, in G

and it is easy to see that (1, x) ^ Z(G) x ^ ^ q-


(ii) If g = with = 1> then A d g fixes 1^, all fj,
and all Cj, j e J, and hence Ad g = 1. Thus g e Z(G) [Proposition 8(0].
Conversely, if g e Z(G), then g acts like a scalar on Lip) and (g,Ad)
[Proposition 8(ii) and Step 2 of its proof]. By Proposition 12 we conclude
that g G i/. If we write g = n , ej/ia(t,), then Ad g = 1 clearly forces
n , eJ “<''> = 1 for all y G J. ' □
6.4 The Formulas of Weyl-MacDonald-Kac 531

Corollary
Let be a minimally realized finitely displayed Kac-Moody Lie algebra
with structure matrix A, Let G and G be its group and derived group. Assume
det(A) ^ 0. Then

(i) G = G,
(ii) P /Q Is a finite abelian group o f order |det(^)|,
(iii) Z(G) is a finite group and |Z (G )| | ldet(^)l,
(iv) Z(G) P /Q if IK contains all \dti{A)Vroots unity.

Proposition 15
Suppose that (g, ¿s radical free. Let P^ and P2 be any two weight lattices,
and let Gj and G2 be the corresponding groups. Then G1 — G2 by the
canonical isomorphism that sends exp x exp x /o r a//x e g", a

Proof Suppose that P^ d P2. Then by the definition of G,, i = 1,2, we have
a surjective homomorphism il/: G^ G2, exp x ^ exp x for all jc e g“,
a e'^^A. Furthermore we clearly have e/i ° Ad = Ad° (/r.
Now, if g e ker ij/, then Ad(g) = Ad(il/(g)) = 1, and hence by Propo­
sition 8(ii) and Theorem 14(ii), g e Z(G) c H and g = Uh^it^) uniquely.
Let j e J, and let (Oj e P2 be a fundamental weight. Then L((Oj) e ^ ( P 2\
so il/(g) acts trivially on L(o)j). But (/r(n/z^(/)) acts as / on L{o)jYK Thus
tj = 1, and finally g = 1. Now taking P' to be the standard weight lattice, we
obtain in general Gp — Gp> — Gp , all isomorphisms being the obvious ones.

6.4 THE FORMULAS OF WEYL-MACDONALD-KAC

The rules are tricky, but they are a much more efficient way of getting the answer
than by counting beans.
— R. Feynman, QED.

Throughout this section g will denote an invariant Kac-Moody algebra with


triangular decomposition and standard invariant bilinear form (• I *). Our
object is to prove the beautiful formula for the character of L(A), A e P^.
This formula, which expresses ch L(A) as the ratio of two WP^-skew-invariant
expressions, immediately gives an auxilliary formula called the denominator
formula, which relates W and the root multiplicities of g. As a bonus we
obtain the important fact that L ^ J ^ h ) = L(A) for all A e P^\ in other
words, L^J^K ) is already irreducible.
Recall the ring Z[P I Q which we defined in Section 2.5. In the present
context P is a weight lattice (see Section 6.1) of (g, ^ ) , and ¡2+ is the usual
positive cone in the root lattice. A typical element of I{P iQ+] is a formal
532 Category ^ for Kac-Moody Algebras

sum / = Em^eCa) whose support lies in a finite number of (2+-fans. Given


w E:W [the Weyl group of (g, we have wP = P. Therefore we can
define

wf = Y^m^e{wa) e Z [P]^ .

Unfortunately, even if / ^ Z[P i 0+], iv/ need not. However, as we will see,
the set

is nontrivial and is clearly a subring of Z[P i 2+].


Evidently W acts as a group of automorphisms on Z ^[F i 2+] and hence
also on its field of fractions i 2+)-

An element / of Qw/(P i 2 + ) is PT-invariant if

^ii w ^ W ,

An element of / of J. 2 + ) is W^-skew-invariant if

h/ = fo rall* v e» F

(recall that the map W ^ {± \], w ^ ( —1)'^*^^ is a homomorphism).

The simplest example of a IT-skew-invariant is the sum

(1) D := £ ( - l ) ^ ' ^ ’e(wp),


w^W

where p ^ is given by WL4 of Section 6.1. [Equation (1) depends on p


although this dependence wilt not be indicated]. By Proposition 5.2.5

wp = p - <5^> e p i 2 + ,

so indeed the sum (1) lies in Z[p 4 2+1 ^ Z[P i 2+1* Note that all the formal
exponentials occurring as summands in D are distinct; see Corollary to
Proposition 5.2. D appears as the denominator of the character formula.

Theorem 1 (Weyl-Kac character formula; [Wy2], [Ka2], [Ka3])


Let g be an invariant Kac-Moody algebra, let P be a weight lattice o f (g,
and let A e Let L be any integrable highest weight module with highest
6.4 The Formulas of Weyl-MacDonald-Kac 533

weight A. Then

I (-l)«*^>e(H;(A + p))
w^W___________________
ch L
e{p) n (l-ei-a))**“"«“
ae A+

This formula expresses ch L(A) as an element of i Q^.]. Both of these


are actually W^-skew-invariant expressions.

Corollary 1 (Weyl-Macdonald denominator formula, [Wy2], [Mac])

Z)== E i-lf'^^e{wp)=e{p) n ( l - e ( - a ) f " ’«“

Proof of Corollary. It is obvious that L(0) is a one-dimensional module [use


Proposition 6.1.5(iii)] whose character is e(0). Since c(0) is the identity
element of Z[P J, Q+], the result follows at once from the theorem. □

Theorem 2
Let g be an invariant Kac-Moody Lie algebra. Then for all A e
Lmax(^) = LUO. In particular there is, up to isomorphism, only one integrable
highest weight module with highest weight A.

Proof. The character formula is identical for L ^^{A ) and L(A); that is,
dim L„,3x(A)^ = dim L { K Y for all weights p. But L(A) is a homomorphic
image of L ^ ^ { K \ and under such a homomorphism L ^ J ^ K Y L(AY- It
follows that the homomorphism is injective; hence L^^J^A) = L(A). □

Theorem 3
Let g be an invariant Kac-Moody Lie algebra over C. For all A e the
hermitian Shapovalov form < • 1• > on L(A) is positive definite. In fact, given
A e ]^*, the hermitian Shapovalov form on L(A) is positive definite if and only
ifA ^P ,.

Proof. Using the extension trick of Section 6.2, we embed g in an invariant


Lie algebra g by adjoining new generators Cf and // (and by making some
appropriate extension § of 1^) so that under the adjoint representation of g
534 Category ^ for Kac-Moody Algebras

restricted to g, // generates an integrable highest weight module with highest


weight A. By Theorem 2 this is none other than L(A).
Now we know that § carries a positive definite hermitian contragredient
form < • I * > that is nondegenerate on all root spaces a ¥= 0. Since
L(A) = ©g“, summed over those a e A of the form -a ¡ + •**,<* I • )
restricted to L(A) is positive definite hermitian and contragredient. It follows
that it is a multiple of the hermitian Shapovalov form, say, c< • 1• >sh- But
then c = ifi\fi) > 0 [assuming that we use f¡ as the generator of L(A) in
defining the Shapovalov form, so </J/^>sh = 1], and we are done.
Now suppose that A © f)* and that the hermitian Shapovalov form < • I • )
on L(A) is positive definite. Let be a highest vector of L(A). Then for all
; e J and for all n © Z+,

Now ' fp • n«A , a ^ ) — n l)/y" (see Section 2.4.), and hence

<u+|ej' • / / • v ^ ) = n! n (<A,a;/> > 0.


J=0

This result is possible for all n only if <A, a / > = k for some A: > 0. Thus
A ejP + . □
Remark 1 The result of Theorem 3 is usually stated by saying that for all
A e P^, L(A) is unitarizable.

We now set about proving Theorem 1.

Lemma 4
Let XyfjL © P, and suppose that A, /x + p e and that p, © A —(2+. Then

(A|A + 2p) > (p ip + 2p).

Proof. Write p = A —a, Then

(A|A + 2p) - (p ip + 2p) = (A + p|A + p) - (p + pip + p)


= (A + p|A + p) — (A + p —a|A + p —a)
= (A + p |a) + ( a |p + p) > 0 □
6.4 The Formulas of Weyl-MacDonald-Kac 535

Lemma 5
/:= “ e( — is W-skew-inuariant.

Proof, It suffices to show that for each reflection r¿ = , i e J, we have


n f = - / . Now

r if = e{r¿p) n (1

= e(p - a,)(l - e (a ,)) O (1 - e(-a))'*""® ”


a e A+\{a/}

by Lemma 5.2.1 and the definition of p. Since e(p —a^)(l —e(a^)) =


e(pX(e(-a¿) - 1), we obtain r j = - / . □

Lemma 6
Let A e P_^. r/ie« the stabilizer o f K p in W is trivial.

Proof Let w E: W. Then

wA c P ( A) c A i ¡2+ [Proposition 6.1.10(i)]


wp = p — (S ^ ) (Proposition 5.2.5)

But <5^> ^ Q+ if w # 1 (see Corollary 5.2.5). The lemma follows. □

Proof o f Theorem 1. We are going to assume that J is finite. We leave it to


the reader to fill in the details when J is arbitrary. L is integrable with
highest weight A. Consider the Verma module M(A). Recall that the
Casimir-Kac operator F^(A) acts on this as scalar multiplication by (A +
2p|A). If p e A 1 ¡2+ and [M(A): L(p)] > 0, then (A + 2p|A) = (p + 2p|p)
(see Proposition 4.5.3).
Now L ^ M {K )/N for some submodule N, and hence by Propositions
2.6.2 and 2.6.12

(2) ch M( A) = ch L + ch N
= ch L + E [ ^ : ¿ (/i)]c h L (p ).

The sum runs over p. e A j, (2+. We may assume that fi ¥= A, since =


M(A)^. Therefore = (0), and hence [N: L(A)] = 0. Also note that

[N: L (p )] = [M: L (p )] > 0 ^ ( p + 2p|p) = (A + 2p|A).

Let Aq = A, Aj, A2,... be the elements p of A i (2+ satisfying (p + 2p|p) =


(A + 2p|A). We partially order these elements by increasing depth. For
536 Category ^ for Kac-Moody Algebras

convenience of notation we write L(A q) for L, even though we are not


assuming that L is necessarily irreducible. Then (2) reads

ch M (A o )= E [M (A o ):L (A ,)]ch L (A ,)
¿>0

Now in the same way Proposition 2.6.12 gives

(3) c h M ( A ,) = E [M (A ,):L (A ,)]c h L (A ,), k = 0 ,1 ,2 ,. ..


i>k

(since only weights of depth larger or equal than that of occur in

Taken together, the equations (2) and (3) form a system of linear equa­
tions with coefficient matrix

1 [M (Ao) :L(Ai )] [M (A o):L (A2)]


0 1 [M (A ,):L (A 2)]
0 0 1

where we have used the obvious fact that [M(A^): L(A*)] = 1 for all k. Now
this matrix is upper unipotent, and hence the system of equations is formally
invertible. Consequently we can express the ch L(Aj.) as formal sums of the
ch M(A,.);

c h L (A * )= E e* ,ch M (A ,.).
i>k

Here the e Z and = 1 for all k, since the above matrix has diagonal
consisting of Ts. In particular

chL(Ao) = E«/ ChM( A, )


¿>0

for some e Z, where = 1.


Now we use Proposition 2.5.4(ii) and obtain
-dim g®
c h L ( A o ) = Ea¿c(A,) П (l~e(-a))
¿>0

Thus multiplying by e(p) and setting

/:=e(p) П
«gA.
6.4 The Formulas of Weyl-MacDonald-Kac 537

we obtain

(4) /c h L(Ao) = L a,e(A,. + p ), an = 1.


/>0

Since ch L(A q) is Pf^-invariant and / is W^-skew-invariant, the left-hand side,


hence also the right-hand side of (4), is PF-skew-invariant. In particular W
simply permutes the elements (A^ + p) for which ^ 0.
Let Q in (4). Then PP^(A^ + p) c (A + p ) i <2^. has an element of
minimal depth, which is then dominant, say, >v(A^ + p) = A^ + p, and
furthermore Up = ( — Now suppose that Aq # A^. Then we have
Aq, A^ + p e and A^ e A q i 2+- Thus by Lemma 4, (A qIAq + 2p) >
(A^|A^ + 2p), contrary to the definition of the A/s. This contradiction
proves that A q = A^ and hence that

A^ + p = w ^(A q + p),

In view of Lemma 6, w w' w(Ao + p) =5^ >^'(Ao + p). We conclude then


that

/ c h L ( A o ) = E « , e ( A , + p ) = E ( - l ) ' ^ ’^ ^ w ( A o + p)).
¿>0 w^W

For another approach to these formulas, see [GLl] and [GL2]. The
following examples will give some idea of the appearance of the denominator
and character formulas in particular cases, as well as some indication of the
amazing amount of information that they contain. The Appendix includes an
extensive example of how to use these formulas.

Example 1 In the case of ^I2(K), A = {—a, 0, a}, W = {1, r^}, p = a/2.


The denominator formula reads

K p ) - e(''«p) = « (P )(l - < - « ) ) = « ( 2 ) “ ^ (“ 2 )’

which is certainly true (r„p = p - a)), though hardly enlightening.


538 Categoiy & for Kac-Moody Algebras

The dominant integral weights are na/2, n ^ N, and by Theorem 1,

e ( n a / 2 + a /2 ) — e( —( n a / 2 + a /2 ) )
chL
( r ) - e ( a /2 ) - e( - a / 2 )

_ e ( a / 2)"^^ - e ( a / 2)-^"*^^
e ( a /2 ) — e ( a / 2 ) ^

na \ Í (n — 2)a
)• - m

This result agrees with what we already know about the irreducible ^Í2(K)-
module of highest weight na/2. Namely its weights are n a /2 , ( n a /2 ) -
a, (na /2 ) — 2a , . . . , —n a / 2 and all of its weight spaces are one dimen­
sional.

Example 2 In the case of §I/+i(IR) we use the description of A i given in


Section 3.1 for the lattice Ai and the description of its roots given in Sections
3.2 and 3.3. Thus ..., is an orthonormal basis for and the roots
are

^ = K -
with simple roots

ai Si ^/+1? ^ 1? • • • ?

and positive roots

A+= (e,- - Sj\i <;•}.

These span the subspace V '= {Ec^eJEc^ = 0}. The Weyl group W acts by
permuting the £¿5.
Let A e P_^_, and write A = EA,e^. Then

2(A|a,.)
(5) ^ í + 1? ^ l?***5^j
(a,la,)

and the condition that A e reads

(6) A , . ¿ = 1,...,/.
6.4 The Formulas of Weyl-MacDonald-Kac 539

Since P + c F, we have the additional condition

(7) L a, = 0.

Together (6) and (7) characterize P^, The fundamental weights , (s>i,
defined by (co^\aj) = were worked out in Section 3.1:

(Oi = Si -\- • • • — («1 + ■■■ +«/+i)


/ +1

where
/ + 1 —/
if j <
/+ 1
i
if j > i.
I+ 1

Obviously for all A = e the coefficients A, e (/ + 1)“ ^Z.


If we translate each A^ by some real number k, then we obtain a point
A' = E(A^ + /:)£• G whose functional value on each a, is the same as
before (though in general A ^ F). It is convenient to make use of this to
“normalize” the A/s so that A/+i = 0 thus obtaining a 1 — 1 correspondence:

{(Ai .A2,---,A,)|A, e N, I = 1 , . . . , / , A, > A2 > ••• >A,}.

In particular for p, which has the defining property (p|a^) = 1 for each /,

p ^ (/,/ -

We now introduce the lattice L '= ( l / ( / + DXZej + *• • and its


group ring
±1
Z[L] = Z
( i J r l " ....... - ( S i l

Define
X, = e(£;), / = 1, . . . , / + 1,

and write for e(e,/(/ + 1)). We have an embedding

Z [P ]^ Z [L ]
e(ct>,) M -^/+1
c(a,) ^ x ,x ,+ V
540 Category & for Kac-Moody Algebras

W acts on Z[L] by permuting the variables. As we will now see, the


denominator and character formulas have very lovely interpretations in Z[L].
Consider the denominator formula. We have

«(p) = (-^1 ••• -^/+1) ^ xi.

where k{p) is the translation of the normalization of the coefficients of p,


namely - ( / + l)fc(p) + = 0. This only enters peripherally into the
picture. Observe that is fixed by W. Thus the denominator
formula reads

••• ^/ + 1) " E Sgn(o-)jct(l)Jf^(2)


a(Y)^<TÍ2) *** ^<TÍl+l)
cre5/+i

(^1 ••• ^/+1) X[X^2, * ••• ^ / 0 ( 1 -JC,


KJ

= (^1 ••• ^/+1) -^ y )-


i< j

Cancelling off we obtain the well-known formula from the


theory of symmetric functions for the skew-invariant polynomial of minimal
degree. It is even more familiar as the Vandemonde determinant formula:

X2
det = n ( ^ i -^ y )-
i<j

*^/+1 -^/+1

Having recognized skew-symmetric sums over as determinants, it is


natural to interpret the character formula as the ratio of two determinants.
Suppose that A e A <-» (A^, A2, . . . , A^) with translation A:(A) for normal­
ization. Then

£ + p))
w^W
-k (A )-k (p )
= (^1 • • X/+1
o-eS/.i

(^1 ••• ^/+1)


6.4 The Formulas of Weyl-MacDonald-Kac 541

SO

chL (A) = (ДГ1 •••


det(x-

Here is a specific example. Take §1з([К) which is of type A 2. Consider


Л= + й>2 ^ (2,1,0), so k(A) = 1 [see (6) and (7)]:

1
( ^ 1X2X3) xt xl 1
X3 x¡ 1
chL(A) =
-Хз)(Х2 - x ¡ )

A few row operations allow one to divide out the factors of the denominator,
and we end up with

•{xiX2 + xfx^ + X^xl + x^xj + Jc|x3 + X 2 XI + 2 x ^X2X^]


ХлХ^Х
l^2^3

= X^x^^ + X 1 JC2 ^ + -^2^3 ^ ^2^1 ^ ^1^^3 + 2

= e(a^ + «2) ^(^1) ^(^2) “*■^ ( “ ^2)


+ e ( —ai —0:2) + 2^(0).

This is exactly as it should be: L(A) is the adjoint representation of ^Í3(R)


(the highest weight is the highest root + «2) ±(0:1 + «2), ±
±«2? 0 are the roots, with the multiplicity of 0 being 2 since dim = 2.

Proposition 7 (Racah formula)


Let g be an invariant Kac-Moody algebra. Let A e F^_, fx ^ P. Then

w^W

^ j (— if fjL p = v(A + p) for some и ^ W,


,0 otherwise.
542 Category ^ for Kac-Moody Algebras

Proof, From the character formula

£ (-l)«-^>e(^(A + p ) ) = £ (-lf'^^e(w p) £ dimL(A)V(A)


wœW wœW

= L L dim L(A)^e(tvA)
w^W Ae]^*
Aeb*
v / ( » v ) J - * ___ r / A \ A .
= £ £ ( - 1) dimL(A) e ( w ( A + p ) ) ,
A s ii* w ^W

where we have used the fact that dim L(A)’^"' = dim L(A)"' for all w ^ W .
Now dim L(A)^ = 1, and hence after cancelling,

0= £ £ ( - l ) “’^Mim L(A) c ( w ( A + p ) ) .
Aei|*\{A} weir

Suppose that p. e P, p i W(A + p). Group together all the terms


e(w(A + p)) = c(p + p), namely A = w ' K p + p) - p. By assumption we
never have A = A. Thus

w^W

Finally, dim L( A) “'’ = dim = dim which is


what we wanted to prove.
Now suppose that fjL -\- p = u(A + p). Then for all iv e PF, p, + <5^> =
p + p - wp = u(A + p) - wp = ¿;(A + Thus dim =
dim Since e 2+ if and dim = 0 for
all a ^ Q^, only one term survives in the sum of our proposition, namely
( - iy<"> dim U A Y ^ = ( - iy<">. □

This formula can be used to compute the weight multiplicities dim L (A Y


by induction on depth. More commonly the Freudenthal formula is used. See
the exercises of this chapter for more on this formula.

6.5 COMPLETE REDUCIBILITY

In Section 2.4 we proved a complete reducibility theorem for the finite­


dimensional representations of ^I2(1K). In this section we generalize this
result to arbitrary invariant Kac-Moody Lie algebras. The proof is essentially
the same.
6.5 Complete Reducíbilíty 543

Theorem 1
[iCPl] Let g be an invariant Kac-Moody Lie algebra over K, and let M be an
integrable %-module in category 0 . Then M is completely reducible. In fact

M= e [M:L(m)]L(m)-
Remark 1 The multiplicities [M: L{¡i)] are defined in Section 2.6 as the
number of times L(/i) occurs in any composition series of M relative to p.
Remark 2 At the time of writing it is not known whether it is necessary to
assume that g is invariant.

Proof Let M be an integrable module in category 0 . Recall that the


Casimir-Kac operator acts on M as a degree 0 linear operator. It follows
that stabilizes the weight spaces of M, which are finite dimensional. Thus,
by extending IK if necessary, we can decompose M into a (direct) sum of
generalized eigenspaces

Mj := |i; G M |(r^ —t)^v = 0 for some k e Z + |,

where t runs over IK (see 7.1).


Each space is a g-submodule since commutes with the action of g
on M, Since e 0 , if ^ (0) it has a highest weight vector u e M f for
some (p e ]^*. Recall from Section 4.5 that Tm ^
= (q>\(p + 2p)u. Thus t = (<p\(p + 2p). Since M originally decomposed into
weight spaces, the weights occurring in are weights of M (whether or not
we made a field extension). Thus t = {<p\(p + 2p) e IK. Thus, just as in
Section 2.4, no extension of IK is necessary.
Let 5, be the sum of all the irreducible submodules of M, for some i,
where ^ (0) (5, is the socle of Mf), Since the irreducible modules in 0
are of the form L(p), fi e 1^*, (see Proposition 2.6.6) and L(p) is integrable
if and only if p e

S: 0 n,L(p,,.),

where f c P_^, ( p j p , + 2p) = t for all and the n¿ e Z^,


We claim that 5, = M^. Indeed, suppose that for some i e IK, S^,
Then Mf/Sf has a highest weight vector v e for some p e 1^*.
544 Category á for Kac-Moody Algebras

The highest weight module U(g) * v c is integrable. Hence ii e


Let be a preimage of ü in M¡^. From the definition of it follows that

+ 2p)i; mod 5,.

Since V e M,, (p |p + 2p) = t. Now, if ^ 0 for some generator


of then ej • u ^ S ^ \ {0}, and hence fi ^ v - a for some v e and
some a ^ Q+. However, from Lemma 6.4.4,

0 = Í - Í = (i'll' H- 2p) - ( mIm + 2p) > 0 since jjl, v ^ P+.

This contradiction shows that ej • v = 0, so g + - = 0 and t; is a highest


weight vector. Thus U(g) • z; is an integrable highest weight module, hence
isomorphic to L(/i) by Theorem 6.4.2. But then U(g) • i; c 5^ contrary to
V ^ Sf. This completes the proof that = S^, and hence M is completely
reducible.
The uniqueness statement of the theorem is a consequence of the results
on composition series in Section 2.6 (or see Proposition 2.6.11). □

Corollary 1
Let g be an invariant Kac-Moody Lie algebra, and let X , ijl ^ P+. Then
L(A) 0||^L(p) is completely reducible. □

Corollary 2
If g is an invariant Kac-Moody Lie algebra over C, then any integrable module
in category 0 is unitarizable.

Proof Use Theorems 1 and 6.4.3. □

Applying the theorem to the adjoint representation we obtain the follow­


ing proposition:

Proposition 2
Let g^ := R ) be defined by the Cartan matrix A = (yl,y), у
realization R . Let A = (^¿y)^ y ^ j be a symmetrizable submatrix o f A , and
let g := R) be a subalgebra o f g^ obtained by restriction from A to A
{see Remark 4.3.1, and Proposition 4.3.2). Set M д(у4')±\а(>4)±91*

(i) gi = M _0(g + f|) e M+.


(ii) M_ {resp. M+) is a completely reducible q-module [in category 0 ( q)
(resp. 0 4 q))1
6.6 The Shapovalov Determinant Formula for Kac-Moody Algebras 545

Proof. Part (0 is obvious. For part ii we use the facts that M_ is an


integrable g_ module in category and that g is iiwariant. Similar consider­
ations apply to M+. □

6.6 THE SHAPOVALOV DETERM INANT FORMULA FOR


KAC-MOODY ALGEBRAS

Recall that the Shapovalov form on the Verma module M(A) induces a
bilinear form
M(A) XM(A) K.

This is defined as follows: If v+ is a highest weight vector of M(A) and


X= y = «2 ■ where «2 ^ U(g_), then

where
^:U(9) ^ U ( 5 )

is the projection onto the first factor of the decomposition

U(g) := U (^) e(g_U(9) +U(9)9+).

In this section we prove the Shapovalov-Kac-Kazhdan determinant formula


for invariant Kac-Moody algebras. This is an explicit formula for the determi­
nant detSh.y, -y e g^u{0} of Section 2.8. The formula involves the Kostant
partition function K defined in Section 2.5 by

ch(U(g_))= L K(p)e(-p).
/3ee+u{0}

An essential ingredient in this section is the notion of Jantzen filtrations


discussed in Section 2.9.

Theorem 1
Let 9 be an invariant Kac-Moody Lie algebra with triangular decomposition
(^,Q+,Q+, o’)- Let K be the Kostant partition function and Sh^ the
Shapovalov form with value Sh^(A) on the weight space MiX-Y ^ for each
A e fi*, y e <2+u{0}. Then
K(y-na)
“ / («1«)
detSh^ = C - n n U “ + P (« ) - " — ^
aeA+n=l\ ^
546 Category û for Kac-Moody Algebras

where

. the determinant is to be understood as follows: M(A) =: U(g)//(A), where


/(A) is the left Ui^ymodule defined in Section 2.3. is an ordered
basis o f U(g_)”'^ then {w, mod/(A)} is a basis o f under the
identification with U(9)//(A), and detSh^(A) == det((F|^(«¿, My))(A)) [see
(2.8.3)]. Then detSh^ is a polynomial function on ij* whose value at each
A is det Sh^(A).
. is defined by (4.4.8),
. the product is over A+ with roots repeated according to their multiplicities,
. p G ]^* is chosen so that <p, = 1 for all i e J (it is not necessarily
unique),
. C is a nonzero constant [which depends on the triangular decomposition of
g, the choice o/ (*| * ), r, and the basis chosen for
Remark 1 Note that scaling (1 • ) by a constant r also scales by r and
hence makes only an overall scaling of r in each factor of the product.
Remark 2 The proof here follows the paper [KK] of Kac and Kazhdan.
The original version of this theorem, for the case where g is finite-dimen­
sional semisimple, is due to Shapovalov [Sh].

Since any y e involves at most finitely many simple roots a¿, i e J, it


is not difficult to see that we may assume that both J and dim are finite.
As we know, detSh^ e U(i|) ^ 5(£|) = 05"(i|). When we refer to the
degree of an element of 5(i|), we are referring to this grading. In particular
the leading term of a nonzero element of 5()&) is its component of highest
degree.
Define A+c A+X by
A+= {(a,j )\a e A+, 1 < ; < dim g“}.
We think of Â+ as A+ in which the roots are repeated according to
multiplici^. For p = (a, ;) ^ A+we definep = a. If y ^ Q+ and y =
_l_ . . . where p^,...,p,^ aresome elements of A+, we will often simply
write y = + ** +i^A: notation down. Similarly for p e Â+, p^
means (p)^- These and otherways in which elements of A+ are treated as
roots should not cause any confusion. The determinant formula itself is really
a product over A+ and Z+ rather thanA+ and Z+. We have written it in the
way that it appears in the literature, in which it is understood that multiplici­
ties are taken into account. In the proof we try to keep the distinction visible.

Lemma 2
Let y e g +u{0}. The leading term in detSh^ is

nonzero constant X Y\
i8eÁ+ n =l
6.6 The Shapovalov Determinant Formula for Kac-Moody Algebras 547

The proof of lemma requires some preliminary discussion.


Let Pi < P2 ^ ^ total ordering of Â+ made in such a way as to
respect increasing height. Usmg Proposition 4.7.4, we can find a basis
of 9+ SO that ep e g^, (cre^lcp,) = 8^^,. In order to apply Proposi­
tion 4.7.4, we need our field to be ordered. We can overcome this_by taking
the basis from some rational form of g; see Section 4.3. Thus for /3 = /3' we
have [a-ep,ep,] = -8 ^^,^^ (4.4.10).
Let y e 0+u{0}, and let iK y) denote the set of all partitions of y into
elements of Â+ (note the convention mentioned above). Then card (ii(y)) =
iC(y). For each partition w = {/3,^ < /3,^ < • • • < /3,} of y, set ■=
o'(efl.)cr(e«. ) • • • a-fcg.) G U fg .)“"^.' Then {Ar„,|w g 0(y)} is a basis for
U(q_)“'^ (see the proof of Proposition 2.5.3). The length of a partition w is
the number of parts in it. By definition ft(0) consists of the empty partition 0
and ~ 1•

Lemma 3
Let w, w' e fKy), y e ¡2 + u{0}. Then the leading term o f qiX^X^r) has degree
less than or equal to 5(iv, w') •= \{length{w) + lengthiw')}. Furthermore equal­
ity occurs if and only ifw = w \ and in this case if w = [ô^ < 82 < *** < 5^},
then the leading term o f q (X ^X ^) is C n[=i5? for some positive integer C.

Proof It is useful to begin with the following observation which is easily


established by induction onr:If jL Ci, .. .,/ x^ eA+ , then for every permuta­
tion 77 of {1, . . . , r} (in particular for 77 so that m^(1) < • ** < M-Trcr)) have

e •**e =e ***e +i?,

where R isa linear combination of products e^^ • • • e^^ and ..., e A+


and s < r.
We prove the result for all y at once by using induction on 5(w, w'). If
s(w,w') = 0, then w = 0 = w' and q(X^X^r) = 1, so the result is true in
this case. In general we write

w = {5i ^ • •• ^ S J ,
w' = { e ^ < € 2 < ■■■ <e,) ,

where Z 8¡ = y = Ee,-. Since q(X^X^,) = q(X^>X^) (Proposition 2.8.1), we


may assume that < 8i. We write e_^_ instead of o-(e^), j = 1,..., 5. We
have

(t ( X J X ^ , = ■■■ Cs

= «3^ • • • es^[ese_,}e. . . . g

+ e_,_ =-A + B.
548 Category á for Kac-Moody Algebras

We consider the two terms separately. If then [e^e_^^] =


and

A = 6s^ •••
= e,

- ( £ 2 + ••• +e.)(5?)(es, ••• •••

Set Z := e,8r . By the induction hypothesis q(Z) has


leading term of degree less than or equal to s(w, w') — 1, and equality occurs
only if r = s, Si = 8i for all i. In this case the leading term of q{Z) is
C e Z+, and hence the leading term of q{Z8^) is
Thus A will contribute precisely C 'n[= i8i to our proposed lending term if
w = w', and otherwise nothing with degree this high.
If ¥= then w ^ w', and either [e^^, e_^^] = 0 and ^4 = 0 or [e^^, e_^^]
is a linear combination LCj-e^, (positive roots because Sj > e^). Thus by the
observation at the beginning of the proof A expands as a linear combination
of terms o-(X^)X^,, where s((p,cp') < s(w,w'). By the induction hypothesis
we may dismiss these terms from our consideration.
Now consider B. If r = 1, then ^(B) = 0. If r > 2, then

B = e s ^ - - esj_es^e_,^]es6_,^ ■■■
=-A' + B'.

We deal with A' as with A above. If we obtain a contribution of


^8/ ” ^82^81^-62 *** ^-Ss^2 terms that cannot contribute to a leading
term of degree s(w,w'). The partitions

< ^3 < ••• < 5^} and {£2^ ^^5}

can be equal only if £2 ^ ^1 ^ ^2 ^ hence w = w'. In that case we


obtain by induction a contribution of a positive integral multiple of IT-=i5f
again. If £5^ we obtain [e^^, e_^^] = 0 or [e^^, e_^^] is a linear combina­
tion and, as before, we are able to dismiss A' by the induction
assumption! The term B' is treated just like being moved leftward
again. The process repeats until reaches the leftmost position, where­
upon q kills the corresponding term. The lemma now is clear. □

Proof of Lemma 2. For 7 = 0 the result is trivial, so we suppose that


7 0. According to Lemma 3, the leading term of det(Sh^) =
0(7)) comes from the diagonal and is a multiple of

(1) n U 8?.
{ S J s i K y ) (S,)
6.6 The Shapovalov Determinant Formula for Kac-Moody Algebras 549

We have to count how many times a particular 8^ will occur. If 8 appears as


a part, say, exactly n times, of some partition w e il(y), then what remains
of w after removing these n occurrences of 5 is a partition w' of y - n8 with
no occurrences of 8. Now we have a bijection

ilr. (w'}

from the partitions of n (y - n 8) to the set of those partitions of fi(y) with


at least n8’s, and similarly a bijection

[w") ^ {w",8}

from partitions of iKy - (n + 1)8) to the partitions of fl(y - n8) with at


least one 8. Thus if/ maps the K(y — n8) — K(y - (n + 1)8) partitions of
il(y - n8) with no occurrences of 8 onto the set of partitions of fi(y) with
exactly n occurrences of 8. The number of occurrences of 8 in (1) is

E n { K ( y - n8) - K { y - (n + 1)8))
n>0

= ( K ( y - 8) - K ( y - 28)) + 2{ K ( y - 28) - K{ y - 38)) + ■■■


= Y . K ( y - n 8).
n>0

This concludes the proof of Lemma 2. □

Define Q+A+= {aa\a e Q^, a e A+}. Elements of (Q+A^.) U are


called quasi-roots (see Kac-Kazhdan).

Lemma 4
Up to a nonzero scalar factor in IK, detSh^, y ^ Q+, is a product o f linear
factors o f the form

where ^ is a quasi-root.

Proof For the purposes of the lemma we can assume K is algebraically


clo^d, for otherwise we can extend IKto an algebraic closure K, replacing 1^*
by IK®!,^]^*, and so forth, to obtain the result.
We view det(Sh^) as a polynomial function on i}*, and let its zero set be
denoted by V. For any A e 1^*, (det(Sh^))(A) = det(Sh^(A)) vanishes only if
the unique maximal submodule MA) of the Verma module M(A) meets the
weight space nontrivially (Corollary to Proposition 2.8.3). For this
to happen, MA) must contain a highest weight vector with weight A - j8 for
550 Category ^ for Kac-Moody Algebras

some P ^ Q+ with A - y e (A -- Using Proposition 4.5.3(ii), we


have
( A + 2 p |A ) = (A - ^ + 2p |A - j8),

which simplifies to

(A+p|/3) = 103|i3),
or equivalently

(A +p)()80)-i(/3|/3)=0.

Let ~ + p(^°) “ e 5(1^). As a polynomial function on


it acts as

^p(A) = (A +p)()8°) - A(j8|)8).

What we have proved is that for a fixed y, det(Sh^) vanishes only on the
union U of the hyperplanes of f)* defined by the zeros of the £^,/3 0
!2 +’^{0}, A - y e A - j S i Q^, i.e., y - jS e (2 +U{0}. The set of ¡3 satisfying
these conditions is finite and independent of A. Thus V<zU, and the ideals
of polynomials vanishing on V and U, say I(V) and I(U), respectively, satisfy
I(V) Z) I{U) By the Nullstellensatz (Section 7.11 [JA2], IT^£^^ lies in the
nilradical y^(det(Sh^)) of the ideal (det(Sh^)). Thus det(Sh^)|(n^£:^)^ for
some k, showing that det(Sh^) is a product of linear factors (with some
appropriate multiplicities k^).
The leading term of this product is a product Comparing with
Lemma 2 and using the uniqueness of factorization in 5(1^), we obtain
aa^ for some a e e A+. Given that p e 0 ^ , we see that
a Q+ and the lemma is proved.

Lemma 5

(i) Let A e Then M(A) is irreducible if for every quasi-root p we have


E^ik) ^ 0.
(ii) If p is a quasi-root and A e ]^* satisfies

. E^(k) = 0,
. for every 8 g Q+\{p}, E^ik) ^ 0,
. for every quasi-root 8, E^(k — p) ^ 0,

then M(k — p) is irreducible and N{k) is a direct sum of finitely many


copies of M (k - P),
6.6 The Shapovalov Determinant Formula for Kac-Moody Algebras 551

Proof. (0 By Lemma 4, M(A) is reducible <=> det Sh^(A) = 0 for some y e


Q+=> E^(X) = 0 for some quasi-root p.
(ii) By part (i), Af(A —p) is irreducible. The sum N of all the irreducible
submodules of M(A) isomorphic to M(A —/3) [possibly N = (0)] is a finite
direct sum of such modules. To prove that N = MA), we prove that L —
M { \ ) / N is irreducible. If it is not, then by Lemma 2.6.3(ii) applied to
L, M(A) has a primitive vector v of weight k — 8,8 ^ Q^, with v ^ N and
c AT. By Proposition 4.5.3(ii), (A + 2p|A) = (A —5 + 2p|A —5), from
which we have = 0. Thus 8 = p, and v e Now U(g+) • v
contains a highest weight vector v'. If v' e K^v, then v itself is a highest
weight vector and generates a copy of M(A — p). Then contrary to assump­
tion, u ^ N. If u' ^ K^v, then u' ^ N by the definition of u. But N has no
weight spaces above X - p, and we again have a contradiction. Thus L is
irreducible as claimed. □

Lemma 6
For fixed a e A+ the functions Q + Z defined by K¿J,y) •= K(y — ¿a),
/ = 0,1,2,..., are linearly independent {over Z).

Proof. Suppose that La¿K¿^ = 0 for some a¿ e Z. With y = 0 we obtain

0 = L a¡K{- i a ) = aoü:(0) = üq.


¿^0
Then with y = a we obtain

0 = E a¡K(<^ - ia) = a,K(0) = a„


i > l

and so on.

Lemma 7
Fixz so that for all s e A+, z(e^) 0. For each A e 1^* yei A — A + zt,
where t is an indeterminant. Following the construction o f Section 2.9 {applied
to our Lie algebra g), form the Verma module M(A) for IK[f ] ^i^g, and let

(JFl) M( X) = M q Z>M^z>
and
(JF2) M(A) =Mor>Mi =)M2 =)
be the corresponding Jantzen filtrations [see (JFl) and (JF2) o f Section 2.9].
Then for all y e Q^,
Ky,S)
detShr(A) = constant X El
S=quasi-root
552 Category ^ for Kac-Moody Algebras

where

detSh^(A) = constant X
5 = quasi-root

Remark 3 We already know that det Sh^(A) has the form indicated here.
However, we do not know the exponents /(y, 8). At the end of the proof of
the Theorem 1 we will have
00

detShy(A) = constant X
aeÁ+"=1

Proof. Let 7 G (2-b) consider the Shapovalov matrix Sh^(A) for the
weight space of M(A) (M(A) being treated as a lK[i]-module). Let

M(A) ^ M ( A )

be the usual map determined by the specialization t ^ a, a E: K. We have


(Lemma 2.9.3)
E^(Shr(A)) = Sh^(A + za)
and
E^(detShy(A)) = detSh^(A H-za).
But
/(y,5)
det Shy = constant X (5^ + P ( ^ ° ) “
8 = quasi-root

Thus

detShy(A + za) = constant X

This being true for all a e K, we have

det(Shy(A)j = constant X n ( ^ 5 ('^) +

Proof o f Theorem 1. It is useful to assume that K is uncountable for the


purpose of choosing suitable A e ]^*. We can always replace IK by an
uncountable extension, if necessary, to allow us to make this assumption. Let
y ^ Q+. We know two things: first, that the leading term of det Shy is

n n
aeA+"=1
6.6 The Shapovalov Determinant Formula for Kac-Moody Algebras 553

and, second, that det Sh^ is a product of factors

where jS® runs over various quasi-roots. Our object is to show that the factors
in the leading term are precisely explained by corresponding factors in
det Sh^:

(n, a^) ^ na^ + p(na^) — \{na^\na^)

= n |a ° + p (a °) -

The conceptualization of what is going on is complicated by the fact that if


a then na so + p(a^) - n{a^\a^)/2 occurs for several
‘‘reasons” in the factorization of detSh^. The proof avoids analyzing this in
detail. We introduce the notion of a root ray. Let a be a quasi-root. The root
ray through a is Q+a n 0 +j (the set of all quasi-roots p that lies in the
same ray of 1^* as a). For each root ray R the leading term of +
must correspond to We
deal with the problem one root ray at a time.
Now one part of the product of Theorem 1 is easily disposed of. If R is the
quasi-ray through a and (ala) = 0, then for all P ^ R we have ip\p) = 0
and p^ + p(p^) — n(p\p)/2 = H- p(p^) = c(a® + p(a®)) for some c =
c(p) e Since the factors of detSh^ that can contribute factors a® to the
leading term are precisely those of the form p^ + p(P^), P^ ^ R, we must
have for each a e A+ and n e Z+ a factor (a® + p(a^) —n (a |a )/2 ) in
det(Sh^), and this accounts for all factors of det(Sh^) related to the ray R.
Now suppose that R is a quasi-ray through a with (a |a ) 0. Let be a
quasi-root on R, so (p\p) ^ 0. Suppose that we prove that for all y e 0+,

( 2) P^ + p(P^) - HP\P) occurs n ^ K ( y - p) times

as a factor of detSh^, where is a positive integer independent of y.


Then up to a scalar factor, the contribution of this quasi-ray R to det Sh^
is
rifíK(y-p)
n - k m )

Comparing with the leading terms and using Lemma 2, we find that up to a
scalar factor,

n = constant X n n
á^R n= l
554 Category € for Kac-Moody Algebras

that is,

'L n J K (y -l3 )= L j:K (y-n a ) fo ra lly e e ^ .


ae A+, a^R = l
n

By the independence result (Lemma 6) equals card{(a, n)\a e A+, na =


p}, and hence the factors a® + p(a^) - n(a\a)/2, na = p occur with exactly
the correct multiplicity in detSh^.
It remains to prove (2). By definition

^^(A) := A(5") + p(8^) - ^{8\8), 8 e 2^u{0}, A e r •

Since we have (p\p) # 0, we can easily check that none of the hyperplanes
£g(A) = 0, 8 p, and £g(A - /3) = 0, of 1^*, is the hyperplane E^iX) = 0.
Then, since K is uncountable, we can choose A e 1^* satisfying the three
hypotheses of Lemma 5 (ii).
For this choice of A, M(A) has maximal submodule MA) = 0M(A - p)
(finite sum, conceivably empty), and M(A - p) is irreducible. Fix z g i)* so
that for all s e A+, z(s^) ^ 0, and set A = A + zt, where t is an indetermi­
nant. Following Section 2.9, we obtain a filtration

M(A) = M d M i d A?2 •••

and a corresponding Jantzen filtration defined by evaluation of t at 0:

M(A) = M d M j 3 M2 => •••

Since Ml = N(X) (Theorem 2.9.4), each Af^, A: > 1, is a direct sum of, say, n/^
copies of M(X — p) (which explains why the series is finite).
Applying Lemma 7, we see that the power of t dividing det Sh^(A) is pre­
cisely the number of quasi-roots 8 occurring in the product [taken Ky,8)
times] for which ^^^(A) = 0 (note the choice of z). By the choice of A the
only quasi-root with this property is p. Thus the number of factors of t in
detSh^(A) is equal to the number of factors of (p^ + p(P^) ~ j(P\P)) in
detSh^.
However, by Theorem 2.9.4(i) we know that the power of t dividing
det Sh^(A) is

E d im M r" = L n ,\K {y -p ).
k=i \k=i I

This proves (2) and concludes the proof of the theorem.


6.6 The Shapovalov Determinant Formula for Kac-Moody Algebras 555

Remark 4 Note that we obtain as a result of the proof that /(y, 5) =


K{y - 81

Given A e define
A+(A) == {(a, n)\a e n e Z + , 2(A + p )(a ° ) = n(a°|a«)}
(don’t forget that a itself is already an ordered pair).

The next result is the basis for Section 6.7.

Corollary
Under the hypotheses of Theorem 1, and with
M(A) D =) M2 =5 • • •
the filtration appearing in the proof, we have
00
£chM ^= Y. c h M (A -n a ).
(a ,n )e A + (A )

Proof. We see from Lemma 7 and Remark 3 that ord,(detSh (Á)) is pre­
cisely the number of factors (a° -I- p(a^) - for which
= 0, where a e n e Z^.. Now = 0 -«=> (A -l- pX«°) ~
n(a°la°)/2 = 0, and we conclude that

ord,(detSh^(A)) = £ K (y-na).
(or, / i ) s A^_(A)

Using Theorem 2.9.4(ii), we have


00

£chM *= £ ord,(detSh^(A))e(A - y)
k = l reG+UtO)
= £ £ K { y - n a ) e { \ - y)
y s Q ^ U iO ) (a ,„ )G Á + (A )

= L L K (y-na)e(A -y).
(a ,n )e A + (A ) r e 0 ^ u {O }

Now K(y — na) = 0 unless y = /3 + na for some ¡3 e Q+G{0}. Replacing y


by /3 + no: in the last sum,

£chM , = £ £ K i P ) e { \ - na -
^= 1 ( a , / z ) e A + (A) IB ^ Q + u {0 }

= X) ch M { \ - n a ) ^
(a, «)e A+(A)
556 Category ^ for Kac-Moody Algebras

6J THE B G G THEOREM A N D GENERALIZATION

In this section we derive a characterization of the weights e i)* for which


L(/x) is a subquotient of M(A). The original theorem of this type is known as
the Bernstein-Gerfand-Gerfand (BGG) theorem and was proved in the
context of finite-dimensional semisimple Lie algebras [BGGl]. The general­
ization to the invariant Kac-Moody case occurs in two steps over the papers
[KK] and [DGK]. The notation here continues that of Section 6.6. In
particular (g, is a finitely displayed invariant Kac-Moody algebra:

A+ (A) := [{a,n)\a e Á+ , n e Z + ,2(A + p ) ( a ° ) =

= [{a,n)\a E: e Z + , 2(A + p\a) = n{a\a)],

and A+ is simply A+ expanded so as to take account of multiplicities. P is


standard weight lattice.
Let A,p, e 1^*. We write A^— ¡i if A = p, or if there is a pair (a, n) e
A+(A) so that 11 = \ - na. We extend this relation transitively so that
A>— fi means that there exists a sequence of elements ^ P^
so that

Proposition 1
Let g be an invariant Kac-Moody algebra with triangular decomposition
(Í), 0-). Let A E r . Then for e

M(A) ^ M ( f i ) <=> [M(A): L(fi)] > 0 A > -/i.

Proof The first equivalence is proved in Theorem 2.11.1. For the second we
use induction on ht(A —¡i) (simultaneously for all \ , f i E i j * for which
\ - fi E (2+U{0}). The result is true if \ = fi. From Proposition 2.6.12 and
the corollary of Theorem 6.6.1 we have

E E [M,: L(,x)]ch L ( n )
fc=i IJ.SÍ)*
= E L [ M ( A - n a ) :L ( M ) ]c h L ( /i),
(л,«)еД^.(А)/*е6*

where M(A) э Afi э Л/2 э • • • is a Jantzen filtration of M(A). Since the


6.7 The BGG Theorem and Generalization 557

characters chL(^t), e 1^*, are linearly independent (Proposition 2.6.11),


the coefficients of ch L(^t) on both sides of this equation are equal. In
particular

1. [Af^: L(jLt)] > 0 <=> [M(A - na): L(/x)] > 0 for some (a, n) e A+(A).
2. Using the corollary to Proposition 2.6.13, since [M^:
UiJi)] > 0 <=> [M,: Ufi)] > 0.
3. Since = MA) and we have the exact sequence
0 -^M (A ) ^ L ( A ) ^ 0,
we have for \ fi, [M(A): L( ijl)] > 0 <=> [M^: L(^l)] > 0.

Now using our equation and the induction assumption, we have for A /jl,
[M(A): L(/x)] > 0 [M(A - na): L(fjL)] > 0
for some ( a , n) g A+ (A)
<=> A - na>— IX for some ( a , n) e A+(A)
<=» A>— IX, □
The preceding result suggests an interesting equivalence relation on 1^*:
A /X <=> there exists a sequence A = Aq, A^,. . . , A^_j, A^ = /x
so that for each / = 1 ,..., A:,
^/-1 -A, or А,-
For each equivalence class A we introduce the full subcategory of ^
whose objects are those modules of & all of whose composition factors L(^t)
have /X G A. Evidently contains all sums, submodules, and quotients of
its modules. □

Proposition 2
For all A G ]^* have M(A) g

Consider the case when dim q < oo. Every nonzero root is real, and hence
for a G A^_, w G Z^,
2(A + p\a)
( a ,n ) G A+(A) =n
(a la )
<A + p ,a ^ > = n.
If (a, w) G A+(A), we see from the definition of (see Proposition 4.4.7)
that X — na = rJ^X + p) —p. Because of the frequency of expressions like
w(X + p) - p in this context it is useful to introduce the dot-action of the
558 Category ü for Kac-Moody Algebras

Weyl group: For \ ^ t)*, w ^ JV,


w ' X '•= vv(A + p) —p.
It is trivial to check that this is a group action (although it is not a linear
action):
{ ^ 1^ 2) *A = Wj • W2 • A for all Wi, vv^2 ^ W, X e 1^*.
The expression above can be thought of as the action resulting from shifting
the origin to the point —p.
Define for each A e ]^* a set c'^^A and a.subgroup of W:
A^ := [a G'’^A|<A,a^> e Z},
:= <r^|a e A"^>.

Proposition 3

(i) A'^ is a subroot system o f A, and is its Weyl group.


(ii) For all w/Xf = A’^^ = A^^ and wW^w~^ =
(iii) I f w ^ then A^-^ = A^ and = W^.

Proof (i) If a, j8 e A^, then


( X ,r ^ P ^ ) = ( X , p - - ( a , p - } a - ) ^ Z .
This proves that A^ is a subroot system of A. Then W^ is its Weyl group by
definition (see Section 5.6)
(ii) Trivial verification (note that wp — p ^ P).
(iii) Follows from part (ii), since W^A^ = A^. □

In this case that dim g is finite we obtain a precise description of the


-equivalence classes.

Proposition 4
Suppose that dim g < 00.

(i) The equivalence class A 0/ A ш 1^* is • A.


(ii) jy A e ]^* and w El W, then the equivalence class ofw • A is (wW^) • A.
(iii) I f К = U and F c ]^* is a chamber o f the Tits cone, then the map

Лs —p -\-F

wW>^ wW^ • A
is a bijection (here F is the closure o f F).
6.7 The BGG Theorem and Generalization 559

Proof, (i) Let A e 1^*, and let a e A^. Then n — <A H- p, a e Z, *A =


A - na, and (A —« a + p, a ^ ) = —n. Thus, depending on whether na or
- n a is in ¡2+, we have
A or -A.
In either case • A A. Using Proposition 3(iii), we now see at once that
w • A A for all w e W^.
Conversely, suppose that /x e A. It will evidently suffice to show that if
— A or A>— jjL, then p = • A for some a e A'^. But in either case
fi = Á — na,
where a e A+, 2(A + p\a) = n{a\a), and n e Z. Since a this last
equation can be written <A + p , = n, and we have a e A'^, • A = p,.
(ii) The equivalence class of w • A is
• A = wW^w-^w • A = {wW^) • A.
(iii) Consider the map
U ^ f)* /'-
A s —p + F

wW^ •-> wW^ *A


We need to prove that ^ is 1-1 and onto.

onto: Let p , p e 1^*, be an arbitrary equivalence class. Then by Proposi­


tion 5.6.9 there exists a (unique) w ^ W with wifi + p) e F. Set
A + p = w(fjL + p), and observe that • A = p. Thus p e
• A, and -p+ F.
1-1: Let A,p e —p -h F. Suppose that wW^ • A = w'W^ • p. Then
wW^ • A = w'W^ • p => p e IF • A
=> p -h p e W ( \ + p)
= > p + p = A+ p = > p = A
(Proposition 5.5.9). Thus vw • A = (w'u) • A for some u e W^, and we
have
w • A = (w'u) • A => w~^w'u • A = A
=> w~^w'u(\ + p) = A + p
=> w~^w'u e Stabp^(A + p)
=> w~^w'u e <r^|<A + p, a = 0> c
where we have used Proposition 5.6.11 to describe Stabpj^(A + p). Thus
ylr(y^;W^) = ^(w 'W ^) => A = p and wW^ = w'W^. □
560 Category á for Kac-Moody Algebras

To generalize Proposition 4 to the infinite-dimensional case Deodhar-


Gabber-Kac restrict the class of allowable elements of i)* and Introduce a
correspondingly restricted class of g-modules. We assume IK = R (the case
K = C is a minor variation, see the Exercises).

Let F be the fundamental chamber for i)* relative to

n = {aj\j e j}

and set

Fo--=F\ U iV ,

where 5 == {/ c J]3a e A+, supp(a) c I, (a\a) = 0}.


Thus A e F q if and only if

(0 <A, a / > > 0 for all i e J, and


(ii) {(A|a) > 0 whenever a e A+ and (a |a ) = 0}.

Define

K = \ J wFo c X,
w^W
:= - p + K,
F := - p ^ Fo.

Observe that K' is closed under the dot-action of W: \ ^ K* => \ + p ^ K


=> w(\ p) ^ K => w ' X ^ K \ Certainly / g 5 implies that the subroot
system A^ of roots supported by I is infinite, since finite root systems do not
contain isotropic positive roots. (The converse is not true, however, as any
rank 2 hyperbolic root system shows). Thus / e 5 implies that the points of
Fj have infinite stabilizers in W (see Proposition 5.6.14) and hence are
boundary points of the Tits cone 3£. Thus is 3£ with some part of its
boundary removed.

is the full subcategory of ^ whose objects are those modules of Û all


of whose composition factors L(p,) have p, e K \

This “good” category has a theory similar to that of Û in Propositions 2


and 4. We begin by introducing an equivalence relation « on K \ For
A, p, e K \ write A>—p, if there exists a sequence of elements p,j,. . . , p^ = p
e K' so that

-P i^ P 2 -
6.7 The BGG Theorem and Generalization 561

and define \ ^ /jl if there is a sequence


A = Aq,A j ,...,A ^ =
of elements of K' so that for each i =
A^_i > A, or A^> A^_i.
Evidently A - /i, => A However, the converse is false.

Proposition 5

(i) A G K \ A^ -p. => ¡JL = ' k for some a and p e K \


(ii) k ^ f l <=>¡1 = W ^ -k .

Proof (i) By assumption p = k - na for some (a,n) e A+(A). Now we


prove that a e''^A. Since A e K \ there exists w ^ W with w(k + p) e F q (or
IV• A e F). If (ala) < 0, then by the definition of A+(A), (A + pla) < 0 [with
equality if and only if (a |a ) = 0]. Since a (see Proposition 4.4.8), we
have wa e^'^A^ and (w(A + p)\wa) = (A + p|a) < 0. This contradicts the
definition of Fq. Thus (a |a ) > 0 and a
Since a g '^^A, g W . From the definition of A+(A), g and
rj ^k + p) = p + p, that is, k = fjL, Since K ' is closed under the dot-ac­
tion, p G K \ This proves part (i), and part (ii) follows in the direction (=>)
immediately. We prove (<=) in Proposition 4' below. □
Beware k ^— p and p g FT’ do not imply that A g K \
For each equivalence class A^ , A g K “, we may introduce the full subcate­
gory of whose objects are modules of ^ with composition factors
L(p) that have fi ^ k ’^ ,

Proposition 2'
For all k ^ K' we have M(k) g , In particular M(A) g and K' is a
characteristic subset o/

Proof [M(A): L(p)] > 0 => k ^ p p G and p ~ A by Proposition 5.


Proposition 4'
(i) Every equivalence class o f ~ on K ‘ has the form • A, A g
(ii) The equivalence classes o f ~ are in bijective correspondence with the
disjoint union o f the cosets
U W /W ^ by wW’^ ^ (wW ^) • A.
AgF*
562 Category ^ for Kac-Moody Algebras

Proo^ Use the proof of Proposition 4 for replacing all occurrences of by


« , F by F q, and ii* by K \ □

Proposition 6
Let g be an invariant Kac-Moody algebra and let \ ^ F+. Then for all p e ]^*
and for all w

{M{w ' A): L(^l)] > 0 <=> /i = H'' • A,

where w' and w < w'.

Proof Note first that since A e F^., we have

A+ p G f
\^ F a K\
= W.
Also note that the map W ^ defined by m m ♦A is injective.
By Proposition 4' the « equivalence class of A is

x-= w ^\ =w \,
and by Proposition 2',

[M{w ' k) \ F (p )] > 0 = > p e A * = > p , = w'*A

for some unique w' e W. Thus we have to prove that for Wy w' e W,

(1) [M{w • A): L(w' • A)] > 0 w < w'.

Suppose that [M(w • A): L(w' • A)] > 0 and w w’. Then w • A>—w' • A by
Proposition 1, and by Proposition 5,
w' - \ = r, r^^-wX

for some )3i,. . . , /3^ where, if we set

A,- == r^. ■ ■ ■ K p ^ - w X , i = 0 ,...,k ,

we have
A;1-1

It will suffice then to prove that for « e IF, )8

(2) u • X^rpU • A =» M < r^u.


6.7 The BGG Theorem and Generalization 563

By definition of the Bruhat order, the latter condition is equivalent to


l(u) < Kr^u). Now using the strong exchange condition (Proposition 5.3.2),
we have

l{u) < l(rpu) <=> w e A+

<=» (Л + p\u~^p) > 0

(3) « (w(A + p)|/3) > 0


<=> <m(A + p ),)3^> e Z +
и *X ^ r ^ u ' A.

This completes the proof of (1) in one direction.


Since [M{u • A): L{r^u • A)] > 0 <=>M{r^u • A) ^ M{u • A), by Proposi­
tion 1, it suffices to prove that

и < •и [ M ( w • A): L[r^u - A)] > 0

for u, p as in (2). But by (3),

и < 'и и ' X ^ r ^ u • A,

and by Proposition 1 this is all that we need. □

It is a fact [R-C,W] that for A e w, w' g W,

dim Horng(M(]v' • A), M(w - A)) < 1.

In other words, there is up to at most one scalar embedding of M(w' • A) in


M{w • A). (Exercise 6.12). We combine this with Proposition 6 as follows:

Proposition 7
Let A G Then there is an order reversing isomorphism of partially ordered
sets between W under the Bruhat order and the set of Verma submodules of
M(A) under inclusion:

w ^ M (w - A). □

Corollary
Let Q be a Kac-Moody algebra of finite type. Let A e P_^. Then M(A) has only
finitely many Verma submodules. Furthermore, ifw^^^ is the opposite involution
in W, then M(Wqpp • A) c M{w • A) for all w ^ W , and • A) is irre­
ducible.
564 Category ^ for Kac-Moody Algebras

6.8 TRANSLATION FUNCTORS A N D THE GENERALIZED


CHARACTER FORM ULA

The Weyl-Kac character formula has a generalization due to Kac and


Wakimoto [KW] that allows “fractional weights” to appear. Although we
have no need of the formula here, its proof is a very interesting use of the
results of Sections 2.12, 5.6, and 6.7 and gives us an opportunity to introduce
the translation functors of Jantzen. In its general plan the proof of the
generalized formula parallels that of the Weyl-Kac character formula. The
novelty is that to keep track of the different terms and to prove that the
coefficients are just alternating + l ’s, we need to invoke a considerable
amount of the machinery of subroot systems and embeddings of Verma
modules. Since chamber geometry is an essential ingredient in our exposition,
we will assume throughout that we work over R. The transition to C is not
difficult and is left as an exercise.
The assumptions are

1. (K = R,
2. g is a finitely displayed invariant Kac-Moody algebra with triangular
decomposition ^ = (g+, Q+, <r), root system A, and standard weight
lattice P.

Proposition 1
Let M be a highest weight module with highest weight pair (i;+. A), and let L be
another highest weight module with highest weight A. Suppose that ^^
basis of L with Vi g L^', 6^ g I^*, and suppose the ordering is chosen so that
depths Si > depths => i > j. Then M L has a highest weight series
(0) = /?o ^1 ^ ^2 ^ ***. = M <S>L,
where each Rj/Rj_^ is a highest weight module o f highest weight A H- dj. I f Mis
a Verma module, then each R^/R^_^ is a Verma module, namely Rj/Rj_^ ==
Mix + Oj).

Proof Let Rj be the U(g)-module generated by [v+(S> f j , . . . , f Vj}, Clearly


Kv^<S> L c \JJ=iRj- Let U„(g) be the subspace of U(g) generated by all
products of at most n elements of g. Suppose that
(C/„_i( 9 ) i; ^ ) ® L c \J r ..

Then for u e U„(g), y ^ L,

u ■ y) e (m• y+® y + (U„_i(9)y^) ®L) n ( U R j ) ,


and so by induction u • v+® y ^ U R j and U„(9)y^® L c U Rj- Thus M ®
L = U Rj.
6.8 Translation Functors and the Generalized Character Formula 565

Now g+-(y+® D,) c v+® g+t), c R i-\- Thus u, + i?,_i


is a highest weight vector for RJRi_^. Thus is a highest weight
module. By induction on i we observe also that .R, = Ey^,U(g_) ■ Vj.
Now suppose that Af is a Verma module. We show that R, is a free
U(g_)-module with basis {y+® Indeed suppose that Mj, . . . , u, e
U(g_), not all 0. Consider the filtration {U„(g_)}“^o> p e Z+ be the
smallest integer for which Mj, . . . , « , e Up(g_). Let I <J q < i be such that
Then

W : = Wj • + ® i^ i) + * * * + U ¿ • (v ^ (S > U¿)

00

+ E w* • +®
k = \

where each lies in Then == Vj + u'jJ o V V¡^ =h 0,


since M is a free C/(g_)-module. Since M ^ L = 0 IK¿;^ and Wq is
the component of w in M <8) Jo' we see that w ¥= 0. This proves freeness.
The result is now clear. □

Proposition 2
Let fi e K \ V e P^. Then L{p) lies in

Proof. Form a filtration (0) = c c —• of L ip) <S>L(v) using a basis


{i;y} of L(z/) as in Proposition 1. We have Rj/Rj_^ - M ( ¡ + SjX j = 1 ,2 ,..., jl

where dj is the weight of Vj.


Now Oj e P(i/) and L(v) integrable (since z/ e imply that Oj is in the
Tits cone 3£ and hence that (0y|a) > 0 for all a e"'”A+. Since also jx + p ^ K,
we have for any a with (a\a) = 0, (p -\- p\a) > 0, and hence
(fi + p 6j\a) > 0. This shows that p 6j p ^ K \ and hence /x + 6>y e
K \ By Proposition 6.7.2', Mip, + 6j) e
Now we can prove that L(p) If [L(p) ® L(v): L((p)] > 0,
then [Rj/Rj_^: L((p)] > 0 for some j, and hence [M(p + 6j): L(<p)] > 0. By
the definition of this means that <p ^ K \ and the proposition is proved.

Proposition 3
If M e and A P+, then M <8>L(A)

Proof Suppose that [M <8>L(A): L(A)] > 0. We have to show that A g K..
The weights of M <8>L(A) lie in the fans (cp + A) i Q+, cp e P(M). In
particular X ^ (<p A) i Q+ for some such <p, and hence X - A ^ cp i Q +
for some weight <p of M.
566 Category ^ for Kac-Moody Algebras

Let M = *** 3 D Mq = (0) be a local composition se­


ries for M at A - A. Each that involves weights f i > \ - A is
isomorphic to L(Ay) for some Ay e
We have

M 0 L( A) = M, 0 L(A) D 0 L( A) d

and for some j\

[Mj 0 L(A)/My_i 0 L(A): L(A)] > 0.

Then \ — A ^ (p i for some weight cp of Mj, so > A —A. Thus,


successively,

Mj 0 L(A)/M,._i 0 L(A) = L{Xj) 0 L(A),

[L(A^) 0 L ( A ) ; L(A)] > 0.

Since [Mj-. L(Xj)] > 0 => [M: L(Ay)] > 0 Xj ^ K ' (Proposition 6.7.2'), we
have by Proposition 2, L(Xj) 0 L(A) e and hence A e A'. □

We are now able to define the Jantzen translation functors. First we recall
Section 2.12. According to Proposition 6.7.2', K' is a characteristic subset of
1^* and the equivalence relation ~ is what is denoted as « in Section
6.7. By Theorem 2.12.4 for any M e ^ ' = M = ®M^, where e
and ft runs through the equivalence classes of K' under = . For simplicity
we will often denote by if A e ft. We denote by 17^ or the
projection functor

M ^M ^.

Let A, /X e K ‘. We assume that A, /i satisfy the condition

( 1) ¡X - X ^ di n p.

Then W(/x - A ) n F n P # ( ^ and W(fi - A) n P+= {A} for some unique


Ae (Theorem 5.9.2). Furthermore (/x - A, a'^> g Z for all a g ''*A, and
hence = A**, (see Section 6.7 for the definitions).
6.8 Translation Functors and the Generalized Character Formula 567

Let M e Then we define (see [Jz, Section 2.10])

(2) m M )= 7гJ^M ® L{^)).

According to Proposition 3, M ® L(A) g so (2) makes sense. It obviously


defines TJ¡^ as a functor

Tr- ^
Furthermore it is exact: If
0 M' M ^ M" ^ 0
is exact, then so is

0 ^ M' ®rL(A ) M ®r L(A ) ^ M" ®rL(A ) ^ 0,

which we rewrite as

0 ®(M' ® L( A )) n ^ ®(M ® L ( \ ) ) a ^ ®(M" ® L ( \ ) ) a ^ 0.

Since g-maps preserve the types, we have exactness of

0 ^ (M ' ® L (A ))^ ^ (M ® L (A ))^ ^ (M " ® L (A ))^ ^ 0.

Proposition 4

(i) Let M, M' e M ' c M. Then

and
c h T t i M ) = chT>t{M') + chT/^CAf/M').

(ii) If M G and
M = ■■■ D M,_y z>M, = (0)

is a sequence o f submodules of M, then

c h T ^ M ) = E chTJ^{Mj/Mj_,).
; =i

(iii) For all M e

c h 7 7 (M )= E [M :L (c.,)]ch r,^(L (a,)).


0)^P
568 Category ^ for Kac-Moody Algebras

Proof. The proof of part (i) follows from exactness, and the proof of part (ii)
follows from part (i). For part (iii) we choose any cp and let

M = Mo d M i D = (0)

be a local composition series at cp. Now either Mj_^/Mj - L{(o) for some
(o > (p or has no weights o) > <p. Then by the definition of the
functor Tf, either Tf(Mj_-^/Mj) ^ Tf(L((o)) or has no
weights (o > cp A. Thus

c h M = L ch(My_i/My)
y=i
= i ; [ M : L(o))]chL(a>) (mod IK[f)* iG+](<p))

(see the proof of Proposition 2.6.11), and

ch r / ( M ) = E chTJt{Mj_,/Mj)

= Z [ M : Lico)]chTi^{Lia>)) (mod K [ r i C+li^+A))-

Since this is true for all (p e fi*, we obtain part (iii). □

Proposition 5
Let A, ju ^ K' with pt —A e 3£ Pi P. Suppose that for some chamber C of the
root system A^, we have

A+p^C, p+p^C.
Then

T f{ M { w • A)) = M{w • p) for all w e W \

Proof Let W{p - A) n {A}. For each w g W^, M { • A) e w (Pro­


position 6.7.2'), and Tf{M{w • A)) = 7t^(M(w • A)) 0 L(A)).
Let {z;y} be a basis of L(A) as in Proposition 1, Vj e L(A)^>. Then
M{w • A) 0 L(A) has a HWS with factors Miw • A + 0y), ; = 1 ,2 ,... . Thus
we will be done if we can prove that for exactly one Sj e P(A), w • A + 0y lies
in the equivalence class /i* of p, and that for this dj, w • A + 0y = w • p.
Since p* = *p = • p (Proposition 6.7.5), we have to solve

(3) w • A + 0y = w ' • p
6.8 Translation Functors and the Generalized Charactek form ula 569

for fixed w € and for the variables w' e W^, 6¡ e P(A). Set A' — A + p,
ff ■= p! ■•=p + p. Then (3) reads

(4) A' + 0' = vpi,

where ff e P(A) and v e are variables. If 0' and v solve (4), then

w f = \ \ n ' f - 2 i v p : \ x ) + "^'»^

Use C to define a base for (Exercise 5.12). By assumption A' and


are dominant integral for 11^. In particular the weights of the irreducible
g^^^-module lie in jj! 1 ( í2 + ^ X s o v / i ' = ¡j ! - e N, and
{vi¿\X) < (/x'|A') with equality if and only if vjj! = ¡j! (A is strictly dominant).
Thus

l|0'll" > Wp’ f - 2(p’\X) + IIA'II" = Wp' - A'll"

with equality if and only if u^j! =


However, 6' e P(A) and by Proposition 6.2.9, and the fact that )l¿' - A' =
jLi - A e WK, we have

11011" < IIMI" = \\p! - All"

with equality if and only if 6' e WA. Comparing the two inequalities, we see
that there is only one possibility:

0' e WA and Vfj! = ij!.

Thus (4) has at most one solution:

= w-^Sj = - A' = /X - A,
6j = w (p - A).

Since this is indeed a solution of (3) (with w' = w), we are done. □

The proof of the main theorem that follows depends on the clever choice
of translation functors. The next result is used in making this choice.

Lemma 6
Let A G K \ and assume that (A + p, ^) > 0 for all e n A^.. Assume
that A^ ^ 0 a n d that I I is a base for A^ that lies in A+ (Exercise 5.12). Fix
any a e n^. Then there exists p e 1^* such that

(i) <p + p, o:^> = 0, <p + p, '^> > 0 /or all p {a},


570 Category ^ for Kac-Moody Algebras

GO W ill - x ) n p ^ i ^ 0 .

Proof. It is easy to see that P n is a group that spans Fix any y o ^ P


with <7o> ~ Z), and set ¡jlq ■= X yQ — p. Then
^0 + P ^
liQ -X ^P .

Let be the Tits cone defined by the subroot system of A (see


Section 5.6), and let C be the chamber of 3E^ defined by the base 11^. By the
definition of 3E^, C n x 0 , where 3£ is the Tits cone of A. Let be the
face of C corresponding to the root a. If B is an open ball in C n x, then
the convex hull of B U r^B contains an open subset of lying in
n X. The union of the rays from 0 through produces an open cone in
lying in C^.
Let := i4 ^ P- This is a semigroup. It is nonempty since D P is 2l
group that spans and contains open balls of of arbitrarily large
size. Let m e M+. For all N ^ Nm e n X, and hence for large
enough N,
/¿0 + p + ^ ^4’
o
Po P ~ p) ^
Set jLi := )Lto + Nm. Then p satisfies part (i), and p - X e F n i . Some
IT-translate w(p — X) of p — X lies in the fundamental chamber F of A that
defines II. Thus w(p —A) e and part (ii) holds. □

Theorem
Let q be a finitely displayed invariant Kac~Moody algebra over IR. Let X e K\
and suppose that

< A + p ,a^)> 0 for all a ^ C\


Then
ch L(A) = det(iv)ch M(w • A)

— \ dim g®

Proof. Let A denote the equivalence class A of A, and order the


elements of A n (A J, 0 + ) according to increasing depth

A n ( A i Q + ) = (A = Aq, A j , A2, ...}.


6.8 Translation Functors and the Generalized Character Formula 571

Now we know that

chM (Ao)= E [M(Ao);L(M)]chL(M),

and since Ag e K \

[M(Ao):L(^t)l Ag
=» /i e {А,|г e 1^}.

Thus

(5) chM(Ag) = E [ A /( A g ) :L ( A ,) ] c h L ( A , ) .

and [M(Ag): L(Ag)] = 1.


Since A, e K \ we can repeat this process to obtain

chM(A,) = E[ M ( A ,) :L ( A , ) ] ch L ( A ,. ) ,

( 6) [M (A ,);L(A ,.)] = 1, i=

We may view (6.46) and (6.47) as a system of linear equations whose matrix is
of the form

1
* 1
* * 1

We may now formally invert this system to obtain

chL(Ao)= ^m ¿chM (A ¿), e Z, mo = 1.


/ =0

If = 0 , then of course = {!}, A = {A}, and (5) already gives M(A)


L(A). So there is nothing to prove. Suppose that ^ 0 . We rewrite the
preceding formula as

chL(Ao)= 12 ^(w^)ch M(iv • A), m(l) = 1,

There remains to prove that m(w) = ( — It is tempting to invoke


If"'^-invariance, but ch L ( A q) is not in general WT^-invariant (see comments at
the end of this section).
572 Category ^ for Kac-Moody Algebras

Fix a e and choose according to L ^ m a 6. Then A + p e C (C is


the fundamental chamber of H^), / ¿ + p e C , —A e X n P . Thus is
defined, and by Proposition 5,

T¡t{M{w • A)) = M(w • !i) for all w e WK

We have from Propositions 6.7.1 and 2.11.4,

• A => [M(A): L(r„A)] > 0


^ M ( r ^ A ) -^M(A),

SO we have an exact sequence

M(A)/M(r,A) ->L(A) ^ 0 .

Now

r/(M (A ))

and since Tj^ is exact, we obtain

T a L { \ ) ) = (0).

Thus

0 = chrr(L(A))
= Tj^M{w • A)

= 53 w(w)ch M(w • p,).

From the independence of characters we have in particular for each w e W^,

0= 52 f n{ u) .
VfJi=WfX

However, fi + p lies in the face of F, and hence Stab,^A()u. + p) = {1, r„}.


Thus v - p = w - p < ! ^ v ^ [w, wr^, and we obtain

0 = m { w ) + m{wr^).
6.8 Translation Functors and the Generalized Character Formula 573

Starting from m(l) = 1, we then have

m(H-)=
where /' is the length function for in terms of the generators {r^la e II''}.
However, (-1 )'^ ^ is the unique homomorphism of into {±1}, that is,
-1 on reflections. Since the length function 1: W ^ [ ± \ ) when restricted to
has the same property, we obtain
IM
m(w) = ( - 1 )
We have proved that

chL(A)= E ( - l ) ' ^ ’"^chM(w • A).

The rest of the theorem follows from


e(w • A)
ch M(w • A) = g“ *

If A e P, then = A, and the formula reduces to the usual character


formula (Section 6.4). An example for A ^ P appears in the example of
the Appendix. It is somewhat paradoxical that ch L(A) can fail to be
PT^-invariant, even though it is the ratio of two PT'^-skew-invariant expres­
sions. This happens in the example of Section 6.7. The following example
shows clearly what the problem is.
Let Q := Za be an abelian group of rank 1, and set Q +— Z+a. Then
K[G >1G+] is an associative (K-algebra and lK[j2] is a subalgebra (Section 2.5).
The latter has an involution r defined hy a —a, e(a) <-> e(—a), but this
clearly does not extend to K[Q iQ+], Consider
e{ —a) — e(a)
/== e ( —2a) — e(2a)

This is evidently the quotient of two r-skew-symmetric elements of K[Q]- But


e(a) — e(3a)
/ = 1 - e(4a)
= ( e ( a ) — e ( 3 a ) ) { l + e { 4 a ) + e ( 8 a ) H- *• • }
= e ( a ) — e ( 3 a ) + e ( 4 a ) — e ( 7 a ) + e ( 8 a ) *• •
on which r is not defined.
574 Category û for Kac-Moody Algebras

EXERCISES

6.1 Let A be an indecomposable Cartan matrix of finite type, and let (g,
be the corresponding split simple finite-dimensional Lie algebra. Let P
be the (unique) weight lattice of (g,
(a) Let A e P^. Using the characterization of P(L(A)) of Section
6.2, show that L(A) is finite dimensional. (One can also use the
character formula.)
(b) Let M e ^ ) . Show that the following are equivalent:
(i) M is finite dimensional.
(ii) M is integral.
(iii) M = 0/!^ i L(A¿) for some Aj,. .. , Ay^^ in P+.
(c) Let A e Show that L(A) has a unique lowest weight A* [i.e.,
A* - a¿ ^ P(L(A)) for all fundamental roots a¿] and that this is a
highest weight for g with its opposite triangular decomposition.
(d) Show that A* e WA. In fact A* = WqA, where Wq is the opposite
involution of W,
(e) Let A e Prove that 0 e P(L(A)) if and only if A e g . (Show
that A e g+U{0}.)
6.2 Let A = be an indecomposable Cartan matrix. Assume that
J is finite. Let P be a minimal realization of A, and consider
g := g (^ , R). Let P be the standard weight lattice of g, and let
A e P+\{0}. Show that the following are equivalent:
(a) g is finite dimensional.
(b) L(A) is finite dimensional.
6.3 Let (g, ^ ) be a split simple finite-dimensional Lie algebra over I Let
P be its (unique) weight lattice.
(a) Let E = and F = Show that there exists an ele­
ment such that ( P , 2 p E) is an § l2-triplet. Let § = KF
0 [K2p^0 [KP [called the principal §I2 subalgebra of (g, 3^)].
(b) Let A e P+. Define the depth of L(A) by dL(A) == max{d(p)|p
G P(L(A)}. Consider L(A) as an ^-module to show that dL(A)
= <A,2p^> + l.
6.4 Let (g, *^) be a finite dimensional split simple Lie algebra over IK:

g = Í) e © g“ ,
aeA

where ^ and A c 1^* is a finite root system. Consider the natural


Exercises 575

W^-invariant nondegenerate IK-bilinear form

k : ij X Í) ^ K,

K{x,y) ■■= ^ (x,a)(y,a).


ae A

(a) Let t : -> 1^* be the isomorphism defined by k. Let ( | • ):


1^* X 1^* ^ IK be given by (T(x)|T(y)) = k( x , y) for all ,y e i.
Show that (T(x)|T(y)) = E „ s 4(T(x)|a!XT(y)|a).
(b) Show that (;u.|iu.) = for all ¡x e b*.
(c) For each ¡x in the weight lattice P, define £>(/x) ==
e I[P]. Show that D(p) = (e(a/2 )
-e (-a /2 )).
(d) For each A e F define I[P] IK[[r]] by e(p) =
Show that for A,p e F,

f,D (n) = +■■■,

where AT = 1A+|, and that

= n ( f 1«) == d^,
OtS A^.

^N+1 “ 0,
d„

(e) Let L(A) be the finite-dimensional irreducible g-module of high­


est weight A e P_^, Apply to both sides of the character
formula D(A + p) = £>(p)ch(L(A)) to conclude that
(a) <¿A+p ^ dp dim L(A) (Weyl’s dimension formula),
(b) E c^ÍmIp )^ = (dim L(A)/24)(A|A + 2p), where ch(L(A)) =

(f) Let <l>be the highest root of A (the highest weight of the simple
adjoint g-module g). Use the Casimir-Kac operator to conclude
that (<f>\<f> + 2p) = 1.
(g) Show that (pip) = dim g/24 (Strange formula).
(h) The dimension formula

„ (A + p|a )
dimL(A) = n
asA+ (Pl^)
576 Category ^ for Kac-Moody Algebras

is independent of the choice of PF-invariant bilinear form (*| *) on


Í)*. Compute the dimensions of the modules L(A) for the simple
Lie algebra of type G2 for the following A:
A = (l,0):<A,ai^> = 1, <A,a2^> = 0,
A = (0,l):<A,ai^> = 0, <A,a2^> = l,
A = (l, l) :< A , ai ^ > = 1 = <A, a2^>.

6.5 Let (9, be a minimally realized Kac-Moody Lie algebra. Let


Ae Show that the weight system P(A) of L(A) lies in the interior
of the Tits cone 3£ iff 5 — {/ e J|<A, = 0} is a subset of finite
type (i.e., is finite).
6.6 Let (G, B, N, S) be a Tits system with associated group W, For each
w W let C(w) = BwB, Prove that for any r 1? • S and any
w W,
C(ri U C (r,

running over all sequences i^< • • • < in [1, p].


In the next two exercises we obtain a recursive formula (due to Dale
Peterson) for the root multiplicities mult(o:) = dim 9", a e for
any invariant Kac-Moody Lie algebra and then use it to obtain the
Freudenthal formula for the weight multiplicities mult^(A) =
dim L(A)^ for any integrable highest weight module L(A). The Peter­
son formula (unpublished) first appears as an exercise in [Ka5]. It
appears in more detail in [KMPS]. The Freudenthal formula for the
finite-dimensional case is very well known. It is a recursive formula (on
depth) and is the basis of most computer algorithms for weight
multiplicities. Our proof of it here fully uses the infinite-dimensional
setup, even for the finite-dimensional case. For the conventional proof
in the finite-dimensional case, the reader may consult [FdV].
6.7 Let (9, y ) be a minimally realized invariant Kac-Moody algebra with
structure matrix A. Begin with the denominator formula (corollary to
Theorem 6.4.1), and denote the common value of these formal series
by D.
Let {a^,{a') be dual bases for relative to (*| *). Then for all
a, e ij*, (a\p) = E(o:|«^)()3|a0. Define derivations d-, d' on
Q iP iG Jby
d^e'^ = ()a|a,)e^

and the “Laplace” operator by = 'Ld¡d‘.


Exercises 577

(a) From the sum version of D show that

d^D = (p\p)D.

(b) Compute d‘D /D using the product version of D, expand the


terms c(—aXl —e(—a ))“ *, apply д¡, and sum over i to obtain

- = E mult(a)(a|a) E
(c) Using d^id‘D / D ) = did‘D /D - a‘Da,.D/D^ obtain

^ ^ mult(a){A:(Q:la) — (a|2p)}e( —¿ a )
k>l asA+

= E E m u lt(a ')m u lt(a " )e (-p a '- ^a").


p,q>l a',a"eA +

(d) Set Cp := > i(l/A:)mult()8/ k ) [mult()8/ k ) is understood to be 0


if p / k ^ A+]. Show that (c) can be rewritten as

E C pim -2p)e(-l3)

= E c^,c^»(i3'l/3")e(-/3'-r)-

(e) Deduce that for P ^ Q+,

(p\p - 2p)c^ = E(i8'li8")c^^V

when the sum is over all pairs (p',p") ^ Q+>^ Q+ such that
P' P" = p. This is the Peterson multiplicity formula.
(f) Show that (P\P — 2p) < 0 if j8 is a root and hi p > 1 and hence
that the formula effectively determines root multiplicities recur­
sively beginning at mult(a,) = 1, / e J.
6.8 Maintain the notation of the previous exercise. Let A e and use
the extension trick outlined after Proposition 6.2.4 to form § based on
an extension ^ of >1 so that § has a generator / q that generates L(A)
as a g-module.
(a) Apply the Peterson formula to obtain for p = ^
G .,

(P\p - 2p)mult()3) = 2 ^ m ult(a) ^ (a\p - ka)mu\t(P — ka),


aeA+ k>\
578 Category ^ for Kac-Moody Algebras

(b) Now letting A e A i and writing P = X, show that

((A + p|A + p) —(A + p|A + p)}mult(A)


= 2 ^ muIt(o:) + fca)mult(A + ka).
aeA+ k > l

(Note that one first obtains this with p, so it is necessary to see


that replacing p by p makes no difference.) This is the Freuden-
thal multiplicity formula.
(c) Prove that (A + plA + p) - (A + plA + p) # 0 if A A, so the
formula does recursively define multiplicities.
6.9 Write a program for the Peterson multiplicity formula, and use it to
compute some root space and weight space multiplicities. Even for
small examples like ^ is quite instructive.
6.10 The theory of category was developed for K = IR. Show that this
theory can be generalized to K = C by redefining Fq as

A G Fq

(a) Re<A,a^^) > 0 for all i e J a n d I m ( X , a y ) > 0


ifRe<A,a,.^> = 0;
(b) (A|a) =5^ 0 if a e A^_ and ( a |a ) = 0.

6.11 Let IK = IR. Given an example of a pair A,p, e 1^* for which A^—p,
p e a:-, A ^ K \
Let g be an invariant Kac-Moody algebra over IR with triangular
decomposition R = (g+, ij, Q^, a), and let A e We wish to prove
that if w, iv' e PL with w <w' in the Bruhat order, then

(1) Honig(M(M'' • A), M ( h' • A)) = 1.

This will complete the proof of Proposition 6.7.7. This extended


exercise runs from 6.12 6.30 and except for the final part is based on
the paper of Enright [En]. Exercise 6.30 comes from [R-CW].
We fix once and for all j e J, and let ^ == = Cej + Chj + C/y c
C ® 55g. We will usually suppress the subscripts j. Many of the exer­
cises depend only on § (not on g) and hence should be considered as
part of § 12-theory. We will use the following notation for a e C, M(a)
is the Verma ^-module of highest weight a. Its C/z-weight spaces are
ie
If V is any ^-module then for all n e C, •= {x e V"\ex = 0}.
Exercises 579

The category consists of all g-modules V such that as an


^-module by restriction,

El: K is a weight module for Ch,


E2: / acts injectively on V,
E3: e acts locally nilpotently on V

together with all g-module maps between these spaces.


When g = ^, we write for Note that for all A e
Mix) e

A module V of is complete if induces a bijection of


onto for all n ^ N .
Let K be a module of ^ ( ^ ) . A C/z-weight module Q( K) is a
completion of V if the following axioms hold:

Cl: Q (F ) is a g-module and there exists an embedding of g-


modules

i:V --C ,iV )

by which we identify F as a submodule of Q (F )) such that


C2: For each n e N for which there is ^-module injection

M ( - n - 2) F,

there is a unique extension to an ^-module injection

M (n)^Q (F ).

C3: Q ( F ) / F is Ui^)-\oceA\y finite (i.e., each element of the quotient


space lies in a finite-dimensional ^-module).

6.12 (a) For a, b ^ C, a b, Mia) ^ Mib) as ^-modules iñ b ^ N, a =


- b - 2. For a e C, Mia) is complete iff « ^ {-2, - 3 , - 4 , . .
Any direct sum of complete modules is complete.
Let F be a g-module and W a g-submodule. Set

F ' := {í; e V\e^v = 0 for some m


F" = {i; e Fl/'^z; = 0 for some m ^ N } ,
V'" = (z; e F|(U(^)z; + W ) / W is finite dimensional}.

(b) Prove that F', F", F'" are g-submodules of F.


580 Category ^ for Kac-Moody Algebras

6.13 Let V be an §-module that is a C/i-weight module and on which e acts


locally nilpotently. Then the Casimir operator 2Y = + 2ef + 2fe =
(A + 1)^ — 1 + 4/e acts on V.
(a) Show that V = ®^^(V[c], where V[c] is the c-generalized
eigenspace of V (see Section 7.1).
(b) Show that if x is a highest weight vector in V of weight a and
y G 11(5) is a highest weight vector of weight b, then b ^ - a -
2.
(c) Suppose that X c F is a submodule of V and V / X is U(i)-lo-
cally finite. Prove that V is generated by X and [V"\n g N} as an
d-module.
6.14 Let V G We construct a completion of V. Let n y, and
suppose that z g 2 # 0. Define
L , ■■=U(g) - 2),

and let be a highest weight vector of M ( —n — 2).


(a) Show that there is a g-module map
L , ^ U ( g ) -z c F

defined by 1 ® z
(b) Let z vary over a basis Z of all the spaces « e N, say,
z e Let
L' == F© e L ,.
zeZ

Using (a) construct the obvious map


k = L'-^V , k\y=idy.
Let K ’ ■ = ker k so that K' n V = (0). For each z g Z let —
t/(g) ®u(g)M(m(z)).
(c) Show that there exist g-module embeddings

and
L' == F © © L^.
zeZ

(d) Set K q ■■=image (AT') c L, and define


K ~ {x ^ L \ f ”x G K q for some m

Set F = L / K . Show the following:


Exercises 581

(e) V is embedded in V under the maps

U ^ L

(f) V / V is U(ê)-finite.
(g) K is a completion of V,
(h) V is in ^ ( ^ ) .
(i) V is complete.
6.15 Show that if C is any completion of F, then there is an isomorphism
(p\ C of g-modules with (p\v= \dy. [Exercise 6.13(c) is useful
here.] Conclude that any completion of V is in and is actually
complete.
6.16 Using the uniqueness of completion given in Exercise 6.15, prove that
if F is a g-module and we treat it just as an ^-module Fg, then C /F^)
has exactly one g-module structure by which it becomes C^(F).
6.17 Show that if 0: F^ F2 is a g-module map, then 0 has a canonical
extension to a g-module map 0: €^(¥2).
6.18 Set A = {(¿;, d u ) \ u e F J = graph of 6. Let A = {(jc, y) e Q (F j) X
C^(V2)\(x, >0 generates a finite-dimensional i7(ê)-module modulo A},
Show that A is the graph of a function, thus defining 0, (Again
Exercise 6.13(c) is useful.)
6.19 This exercise is about § and the category <J^(ê). The main result is to
determine the indécomposables in Let n e N. In U(§) define

= U(§){/i + n + (2ru(,) - «2 _ 2nf } ,

= U(g){/i + n + 2,c'’-"2},
T(n) == U (i)/2W „,S(«) := U(ê)/5R„.

There is a natural surjective ^-module map S(n) Tin). For each


e C, let := a} + 2a. Let n ^ N.
(a) Let L == linear span {/'e^|0 < ; < n H- 1, /, j ^ N}. Prove that
U(§) = L 0 9in and Sin) is a free C[/]-module.
(b) For 0 < / < n + 1 define to be the C[/]-modulo generated by
{e\ . . . , Show that this leads to a filtration of Sin) = 5 q ^
5i D • • • D + i Z) 5„+2 = (0) whose factors are M i - n - 2 +
20, i = 0 , .. ., n + 1.
582 Category ^ for Kac-Moody Algebras

(c) Set c := c„ = «^ + 2n. Prove that the c-generalized eigenspaces


of the filtration of (b) lead to a filtration of S(nXc] in which

S(n)[c] 3 S„^i[c] D(0)

with S ( m)[ c ]/ S „ + i [c ] = M ( —n — 2) and S(n + iXc] = M(n).


(d) Prove that 5(/i)[c] T(n) and hence that there is an exact
sequence

0 M(n ) -> T(n) M( —n — 2) 0.

(e) Prove that T(n) is indecomposable [use the fact that M(n) and
M ( —n — 2) are indecomposable and T(n) is monogenic].
(f) Prove that T(n) e cX(§) and is complete.
(g) Show that if ^ e and A = A[c] (c = c„) and if v e
then the mapping x xu of Vi(ê) into A factors through T(n).
6.20 In this exercise we determine that the M(a), a ^ C, and the T(n),
n Ei Ny are the indecomposable objects of cX(ê). Let A e We
wish to decompose it into indécomposables. For this we can suppose
that A =A^ for some c = a'^ + 2a ^ C.
(a) If a ^ I or a = - 1 , then A^ and A~^~^ are composed of
highest weight vectors, whence y4 is a direct sum of modules of
the type M(aX M{—a — 2).
(b) Suppose that a ^ N . (This covers the remaining cases from (a)
since, if ÎÏ e Z_, we can replace it by -¿z - 2). Set L —
N := A^. I f N = (0), ^ is a direct sum of modules M { —a - 2).
Assume that N ^ (0). All elements of N are highest weight
vectors. Let and let N2 be a subspace so that
N = N 1 ^ N2- Let Li any subspace of L mapped isomorphically
onto N 1 by Let L 2 =

Prove that n L2 = (0) and that

L = Lj 0 L 2 ® L 3?

where L 2 0 L 3 is some subspace of L consisting entirely of


highest weight vectors.
(c) Prove that for any nonzero element x e L^, U(^)x ^ T(a), and
hence that U(§)Li is a direct sum of r ( a ) ’s.
(d) Prove that U(^)Li + U(^)N2 is a direct sum of ^-modules iso­
morphic to T(n) and M(a),
(e) Prove that U(^)L3 is a direct sum of modules M{ —a — 2).
Exercises 583

(f) Prove that

A = U(ê)Li e U(ê)N 2 ® U (ê)L 3 .

Conclude A is decomposable into summands that are of the types


Mia), M i - a - 2), and Tia) if a ^
6.21 Fix n ^ N, and let Lin) be the irreducible ^-module with weights
I := { - n , - n + 2 , .. ., Let V Œ C, and form F — Lin) ®^Miv).
In this exercise we prove that F e ^ ( ê ) and determine its indecom­
posable components.
(a) Set b ■= Ch Ce. Let be the one-dimensional b-module of
weight V. Show that

L(n) 0 c C ,).

(b) Set Li = 'L/ç^iLin)^, i e I, thereby giving rise to a filtration

where F^ — U(ê) i e I. Show that + 0.


(c) Prove that Lin) ®^Miv) e cX(§) and hence that Lin) ®^Miv)
is a direct sum of indécomposables, as in Exercise 6.20(f).
(d) Write F = ©^^^F[c] (generalized eigenspaces). Prove that if
1/ e Z, then r is injective and hence that F ^
+ r).
(e) Suppose that v ^ Z. Show that for r,s ^ I, => (after
possibly interchanging r and s)

V + r > —1 ,
s = ~iv r) — 2 — V.

(f) Set

/ ' := [r e I\v + r > 0, 5 — —(ï/ + r) —2 — e /},


/" := {5 e I\s occurs in /'} ,
/" = / \ ( / ' U / " ) .

(Note that if v r = —1, then r occurs in /"'.) Show that

© P[c^+r] ® ® +
re/' k ^ I'"

and show that each factor F[c^+i^] == M(/a. + /c), /: e


584 Category ^ for Kac-Moody Algebras

(g) Prove that for r e /', is indecomposable and hence that


T{v + r). [To show that is indecomposable, let v e
L(n)“'^ be a lowest weight vector of L{n) so that z; <8> gener­
ates F as an ^-module, and consider the projection tt: F ^

(h) For n e N, v e C, L(n) + r) ®


M(v + r).
6.22 Show the following:

© r(2i) if n is odd,
0 < i < ( n —l ) / 2
(a) L{n) ® M( —1) = M (-l)e © 7(1 + 2i)
0<i^(n-2)/2
if n is even.

(b) if 1/ e N, then

© M{v - n 2i) if V — n > 1,


0<i<n

© T(2i) ® © M(2j)
0<i<(fi-2)/2 -fji/2<j<(nu+n)/2
if fji '•= V — n is even and < -1 ,
L(n) <S>M{v) = <
M( - 1 ) ® 0
0siS (/x-3 )/2
T(2i + 1)

© M(2j + 1)

if /i ~ V — 1 is odd and < - 1 .


(c) Prove that in (a) and (b) above, L(n) ® M(v) is complete.
6.23 Let V G ^ ( § ) , and let L e ^ ( q, T ) be an integrable module. We
wish to prove that L ® F e ^ ( § ) and C(L ® F ) = L ® C(F).
(a) Show that L ® F e ^ ( ^ ) .
(b) Show that C(L ® F ) =) L ® C(F), since L ® C (F )/L ® F is
U( §)-locally-finite.
(c) Prove that L ® C (F) is complete. [By Exercises 6.16 and 6.20(f)
we may assume that g = § and that C (F) is indecomposable.
Now use Exercise 6.22.]
6.24 Let A G ]^*(]^ = g° ). Let § = Then prove that Mg(A) is complete if
\(hj) ^ - 2, —3,...} and that the completion of Mg(A) is MJitjX) if
X(hj) G { - 2, - 3 , . . . } . [Prove that as a ^-modules Mg(A) =
® c W (^ X ]
Exercises 585

6.25 We are now in a position to prove the Horn result (1). By Proposition
6.7.6 it is enough to prove that

Honig(M(w' • Л), M{ w • Л)) < 1.

Show that we can restrict to the case = 1, and reason by induction


on /(w'). Prove that

Z (ry) > /(w ') -A e { - 2 , - 3 , . . . } ,

using the completion function C^. to complete the induction step.


6.26 Establish the Bruhat decomposition of Proposition 6.3.7.
6.27 Prove the corollary to Theorem 6.3.14.
6.28 Let G be the derived group associated with a Kac-Moody algebra g.
Prove that Z(Ad(G)) = 1.
6.29 Let ^ be a Cartan matrix and R a minimal realization of A, Form the
Kac-Moody algebra g = g(^4, i?). Let G be the associated derived
group.
(a) Show that the restriction map res: Ad(G) ^ Aut(Dg) is injective.
[Write g = f © Dg, i c 1^. Prove that g e ker(res), k => gk
= k -\- a^^Kg), where a^^Kg) e Zg. Finally, argue that ker(res) is
central.]
(b) Suppose that in addition A is symmetrizable. Prove that there is
a natural mapping

<p: Ad(G) A ut(D g/Z g).

Prove that (p is injective.


6.30 Let ( tt, M) be an integrable representation of g, and let tt: G ^
GL(M) be the corresponding representation of G. Show that if kerCrr)
c i) then keri-rr) c ker(Ad) = ZG. In particular G is a central exten­
sion of Tr(G), and there is a natural surjection tt(G) Ad(G) whose
kernel is central.
6.31 Let g be a finitely displayed Kac-Moody Lie algebra. A g-module is
said to be weakly integrable if it satisfies condition IM2 of Ch 6.1.
Show that the following are equivalent
(i) g is perfect
(ii) Every weakly integrable g-module is integrable (That (i) => (ii)
was communicated to us by Jun Morita)
Chapter Seven

Conjugacy Theorems
T h e r e is in f in it e h o p e ; b u t n o t f o r u s.
—F. Kafka

One of the most useful concepts in the theory of finite-dimensional Lie


algebras is that of a Cartan subalgebra, a nilpotent subalgebra that equals its
own normalizer. A central result of the classical theory is that (in the
algebraically closed case) all Cartan subalgebras are conjugate.
The main objective of this chapter is to generalize this to the case of
invariant Kac-Moody Lie algebras. The discussion is based on Peterson and
Kac’s cryptic [PK]. Our treatment is self-contained save for some key parts of
Section 7.4. There we need to use the conjugacy theorem for finite-dimen­
sional Lie algebras (the necessary result may be shown by following Exercise
7.6) and some results from algebraic groups, particularly the Borel fixed point
theorem.

7.1 L O C A LLY FINITE ENDOM O RPH ISM S A N D


JORDAN-CH EVALLEY DECOM POSITIONS

An endomorphism jc of a vector space V over K is locally finite if each v


lies in a finite-dimensional x-invariant subspace. A considerable amount of
the finite-dimensional theory of linear transformations extends without dif­
ficulty to the locally finite setting. In this section we develop the locally finite
theory as far as we need it. Our main result is the locally finite version of the
Jordan-Chevalley decomposition theorem. We give a completely self-con­
tained proof of this theorem. This is preceded by a proof of the finite-dimen­
sional version and by a short development of some basic linear algebra, both
of which are undoubtedly familiar to many readers.
The importance of locally finite endomorphisms from our point of view is
that they arise completely naturally in the study of integrable modules of a
Kac-Moody algebra g. Locally nilpotent endomorphisms are locally finite.

586
7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 587

and so are their exponentials. Thus the generators of the derived Lie group
G of g are locally finite. In addition the diagonal algebra of g is locally
finite, in the sense that it consists of elements that act locally finitely on any
module of and hence all the generators of g are locally finite.
A few words about the nature of the base field K. For convenience and
consistency of notation, K will still denote a field of characteristic 0. How­
ever, in this section such an assumption on K is not necessary. The key to the
existence of a Jordan-Chevalley decomposition of a given endomorphism x is
that the roots of its minimal polynomial lie in a separable extension of K, and
this is the only assumption that we use in the proof.
Let F be a K-space, and let x be an endomorphism of V, Consider the
polynomial ring lK[i] in one variable t with coefficients in K. Evaluation at x

fit) --fix)

determines a ring homomorphism

^ E n d ^ (F ).

Notice that makes V into a K[i]-module via f(t)v = f( x X v ) for all v


If Endj^(F) is finite dimensional, our map must have nontrivial kernel. On
the other hand, K[t] is a principal ideal domain, and it follows that there
exists a unique nonconstant monic polynomial p^^yU) ^ K[t] called the
minimal polynomial of x such that

k e r ( e j = <p ^,k(0 >

where < ) stands for “the ideal generated by.” The subalgebra IK[a:] of
End^fF) generated by x is evidently

Thus IK[j:] is a field if and only if p^^yit) is irreducible.


Let X e End(F). For each polynomial /( i) e K[i] define

■■= (u e F |/ ( = 0 for some n e N}.

Evidently F^ is j:-invariant.
588 Coivjugacy Theorems

Proposition 1
Let V be a finite dimensional K-space. Let x g Endu^(K), and let Px^y(i) be its
minimal polynomial. Then

V=Y®
i=l

where p^^yO) = pft)^^ **• is the decomposition o f p^y{t) into monk


irreducible relatively prime factors. Each is x-inuariant and p¿(tY‘ is the
minimal polynomial o f x viewed as an endomorphism of

Proof. Suppose that / ( 0 = a{t)b{t\ where a (0 and b{t) are nonconstant


and relatively prime. We claim that ® V^.
In fact, write 1 = a{t)a{t) + p(t)b(t) for some a(t) and p(t) in K[t]. Let
V e V f Then V = a(x)a(x)v + p(x)b(x)v. Now

b ( x ) a { x ) a ( x ) v = a ( x ) f ( x ) v = 0,

so a{x)a{x)v e V^. Similarly p(x)b(x)v e V . Thus V = V V^. Further­


more V Ei C\ V ^ ya(t)aO)+p(tMt) ^ yid ^ showiug that the sum
is direct.
The first part of the proposition now follows easily by induction (note that
for any polynomial /, Finally, let q¿(t) be the minimum polyno­
mial of X acting in V^‘. Then qfOlPiitT for some n, and since p f t ) is
irreducible, q f t ) = p¿(tY‘ for some l¿ > 1. Since qf^x) • • • q„(x) annihilates
V, we have PxyiOlqft) • • • qJ^O, whence k ¿ < f . But both Px,y(0 and
p¿(ty^ annihilate hence also their greatest common divisor which is
pftY^. Thus k¿ = /¿. □

The decomposition of V given by Proposition 1 is called the primary


decomposition of V relative to x.
Let F be a K-space (not necessarily finite dimensional) and let x be an
endomorphism of F. x is said to be semisimple if the action of x in F is
completely reducible. That is

F=
yej

where each Vj is jc-stable and no nontrivial subspace of Vj is jc-stable (i.e.,


each Vj is an irreducible x-space). x is said to be diagonalizable if there
exists a basis of F and a family of scalars {A¿}¿g, such that

XV: = X:V: for all i e I.

In other words, there exists a basis of F consisting of eigenvectors of x.


7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 589

To say that x is semisimple is equivalent to saying that V is completely


reducible lK[jc]-module. Thus by Proposition 1.6.2 we have

Proposition 2
Let V be a K-space, and let x e End[j^(K). The following conditions are
equivalent:

(i) X is s e m i s i m p l e ,
(ii) V is the sum o f simple IK[jc]-modwfe5.
(iii) V is the direct sum o f simple ^x\-modules,
(iv) Every ^x^subm odule M o f V admits a supplement. That is, there
exists a submodule N o f V such that V = M ^ N. □

Proposition 3
Let V be a K-space, and let x be a semisimple endomorphism of V. I f W <zV is
an x-stable subspace, then xis a semisimple transformation o f W. □

Proposition 4
Let V be a K-space, and let x g End,,^(F). The following are equivalent:

(i) V is an irreducible x-space.


(ii) V is finite dimensional, monogenic (= cyclic), and the minimal polyno­
mial o f X in V is irreducible.)

Proof Suppose V is irreducible. If v ^ V , v i-Q, then [K[a:]í; = V, and


hence V is monogenic. Make V into an (irreducible) lK[i]-module by defining

f{t)w = f(x)(w ) for all f { t ) e K[t],w g V.

The map
a: K[i] V,
given by

fiO = f{x)(v).

is a surjective K[/]-module homomorphism. Since IK[i] is not irreducible as a


K[t]-module, ker a # (0), so ker a = (p (r)), where p(t) is monic and irre­
ducible. Also p(x)(V) = 0, since /?(a:)(í;) = 0 and i; generates V. Thus p(t)
is the minimal polynomial of x in V. Finally, V « K[t]/(p(t)}, and hence V
is a finite dimensional K-space.
590 Coi\jugacy Theorems

Conversely, choose ü e F such that IK[;t]i; = V, and define a: K[t] V


as above. Since the minimal polynomial pit) of x is irreducible by assump­
tion, we see from pit) e k era that k era = (pit)). Thus F « K[t]/(pit)},
which shows that V is an irreducible K[r]-module, that is, an irreducible
x-space. □

Proposition 5
Let V be a finite dimensional K-space. Let x e End^iF). Then for x to be
semisimple, it is necessary and sufficient that each o f the irreducible factors of
its minimal polynomial has multiplicity 1.

Proof. Using Proposition 1 we see that F is a semisimple x-space if and only


if each of its primary components is semisimple. Thus we reduce to the case
where the minimum polynomial has the form p i t Y , p i t ) irreducible. If
h = f, then F is a IK[t]/<p(t))-module, namely a vector space over the field
IK[t]/<p(t)>]. It follows at once that it is completely reducible.
If k > 1, then W ••= [v ^ V\pixY~'^v = 0} is a nonzero proper x-in-
variant subspace of F. If W were to admit an ;c-invariant supplement N,
then p i x ) N c ATn IF = (0) from which p ix Y ~ ^ V = p ix)'‘~ \W + N ) = (0).
Thus IF has no such supplement, and hence F is not completely reducible.

Let F be a K-space, and let F be a field containing K. Let Fp = F ®^V. If


X e End^CF), then there exists a unique F-linear extension Xp of x to an
endomorphism of the F-space Fp that satisfies
Xf-. a ® V a ®x(v) for all a e F, u e F.
When F is finite dimensional we let char^ p.(t) and p^ yit) denote the
characteristic and minimal polynomials of x on F.

Proposition 6
Let Vbe a finite-dimensional K-space. Identify K[i] inside F[i] on the natural
way. Then
(i) char_, p,(i) = char^^
(ii) Px.v^t) = p,,^yft).

Proof Since p^^yixj:) e EndpiFp) annihilates F, it annihilates Fp. Thus


p^p j/p(i)|P;c.F(i)- On the other hand, let {A,} be any basis for F over K, and
write p^p ,/f(0 = EA,/,(i), for some unique f f t ) e K[i]. We have
0=Px,.Vf(x^) = EA,./,.(xp) e Endp(Fp).
Restricting to F, we have 0 = EA,/,(x) from which f f x ) = 0 for each i.
7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 591

Thus leave part (i) as an easy


exercise. □

The next result is well known and easy to prove.

Lemma 7
Let V be a finite-dimensional K-space, and let x e End[j^(K). Let K be the
algebraic closure o f IK, and consider e End(IK<S>(|^i;). Then the roots of
p^ y{t) in IK are the eigenvalues o fx ^. □

The roots of Px,y(0 together with their multiplicities are called the
spectrum of x.

Proposition 8
Let V be a K-space, and let x e Endj^(F). The following conditions are
equivalent:

LFl: Every element o f V belongs to a finite-dimensional x-invariant subspace


ofV,
LF2: For all v the span o f the family is finite dimensional.
LF3: There exists a family {Vj)j ^ j o f finite-dimensional x-invariant subspaces
of V such that V = Ey^j^-

Proof Clear. □

An endomorphism x of a IK-space V is said to be locally finite in K if x


satisfies the equivalent conditions of Proposition 8.

Proposition 9
Let X be a semisimple endomorphism o f a K-space V. Then x is locally finite.

Proof. By definition K is a sum of x-invariant irreducible spaces and by


Proposition 4 each of these irreducible spaces is finite dimensional. Thus x
satisfies LF3. □

Lemma 10
Let X be a locally finite endomorphism o f a K-space V. I f x is injective, then the
restriction of X to every x-invariant subspace o f V is surjective. In particular x is
an automorphism o f V.

Proof. If W is an x-invariant subspace of V, then clearly x is locally finite


in W. Write W = E y ej^? where each Wj is finite dimensional and
x-invariant. Then xWj = Wj for all j, and hence xW = Lj^jxWj = W. □
592 Coi^ugacy Theorems

Proposition 11
Let ¥ be a field containing (K, and let x e End,,^(F). Consider Xp = 1 0 e
EndpiKp). Then x is locally finite if and only if x^ is locally finite.

Proof If Xis locally finite, then write V = (LF3). Then = E/eiF ®


Vi, and each F 0 is finite dimensional (over F) and x^ invariant. Thus x^ is
locally finite. Conversely, if x is not locally finite, then there exists v
such that the K-span of [x^v\n e l\l} is infinite dimensional. Let
be an infinite linearly independent set of this span. Then {1 0 ^
{( X ( p )''( l 0 u)\n e N } is an infinite linearly independent subset of over F.

Proposition 12
Let Vbe a K-space, and let ¥ be a field extension o f K. Let x End|,^(F), and
consider the endomorphism x^ ofV^ = ¥ 0| kI^* Then forx to be semisimple it is
necessary and sufficient that x^ be semisimple.

Proof If either a : or X p is semisimple, then either a : or X p is locally finite


(Proposition 9), and hence both x and jcp are locally finite (Proposition 11).
We can therefore assume that V = E/eil^, where each is A:-invariant and
finite dimensional. Thus Kp = E¿ei(^)F» (I^)p is JCp-invariant and
finite dimensional (as an F-space). Now
X is semisimple <=>x|i/ is semisimple for each i
<=>JCp|(K,)p is semisimple for each i
<=>jcp is semisimple.
The only nontrivial equivalence is the middle one. However, by Proposition 6,
x\y. and ^ pI(i/) f have the same minimal polynomial, and the equivalence now
follows from Proposition 5. (Here we need to know that irreducible polyno­
mials over IK do not have repeated factors in F. This is true as long as F is a
separable extension of IKand happens in the characteristic 0 case.) □

Proposition 13
Let Vbe a K-space and x an endomorphism o f V. The following are equivalent:

(i) X is diagonalizable.
(ii) X is semisimple, and all the eigenvalues of x lie in K.
(iii) X is semisimple, and its irreducible subspaces are one dimensional.

Moreover, if these conditions hold then x acts diagonally on every x-invariant


subspace o f V.
7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 593

Proof, (i) => (ii) Suppose that x is diagonalizable. Then there exists a basis
of V consisting of eigenvectors of x. The corresponding eigenvalues are all
the eigenvalues of x. Part (ii) then follows.
(ii) => (iii) Since X is semisimple, V = where each is an
irreducible jc-space. Let p f t ) denote the minimal polynomial of x restricted
to V¿. By Proposition 4, p f t ) is irreducible. By Lemma 7 and by assumption
p f t ) = t — a¿ for some e K. Thus x acts on like scalar multiplication
by so is one dimensional.
(iii) =» (0 Clear.
Finally, if PT c K is an jc-invariant subspace of V, then x is semisimple on
W, and all eigenvalues of x in PT are eigenvalues of x in V. Thus x acts
diagonally on W. □

Proposition 14
Let X be an endomorphism of a K-space, and let IK be the algebraic closure of
K. Then for X to be semisimple j t is necessary and sufficient that be a
diagonalizable endomorphism o f 1K<S)[|^K

Proof The result follows from the last two propositions. □

Proposition 15
Let V be a K-space, and let S c End|,^(F) be a set consisting o f semisimple
endomorphisms that commute with each other. Let E be the subspace o f
End[,^(K) spanned by 5, and suppose that E is finite dimensional. Let A be the
associative subalgebra o f End,,^(F) generated by E.

(i) Every element o f A is a semisimple endomorphism o f V.


(ii) I f E admits a basis consisting o f diagonalizable endomorphisms, then
there exists a basis o f V on which all elements o f A act diagonally
(iii) Let ^ = Lie(^) be the Lie algebra o f the associative algebra A. Then
is an abelian Lie algebra. I f the assumption made in (ii) above holds
then the ij-module V admits a weight space decomposition.

Proof, (ii) Let {xi,. . . , jc„} be a basis of E, and suppose that all x^ are
diagonalizable. It is clear that A is generated hy x ^ ,. .. , jc„ as an algebra. We
establish (ii) by induction on n. If n = 1, then A = !K[jcJ, and the result is
obvious. Assume that the result holds if E is of dimension n - 1. Let
S' = {^2, . . . , x^}, and let A' be the subalgebra of End„^(F) generated by S'.
By the case n = 1 we can write

AGIK
where
i; e P^ <=>x^v = \v.
594 Coixiugacy Theorems

Fix A Then A' stabilizes V^, since v E: and / ' e A' implies that
X if 'v = f'x^u = Xf'v. Since are diagonalizable, it follows
from the induction hypothesis that we can find a basis of in which A' acts
diagonally. Since each of these basis vectors are eigenvectors of and x^
and A' generate A, the result follows. _
(i) Let JCi,. . . , be a basis of Fi’jionsisting of elements of 5. Let IKbe the
^gebraic closure of K. We set_F = 1K0 F, and let = 1 0 jc- e E^jj^(F). If
A is the subalgebra of Endjj^(F) generated by the 3c/s, then >1 = 1K0^. By
Proposition 14_all the x /s are diagonalizable, and hence by part (ii) we
conclude that A consists of diagonalizable, hence semisimple, elements. Now
part (0 follows from Proposition 12.
(iii) f) is abelian, since is generated by commuting endomorphisms. If
assumption (ii) holds, we can find a basis {Vj}j^j of V such that hvj = \j(h)vj,
Xj(h) e IK, for all j e J. It is immediate that the mapping Ayi h Xj(h) is a
linear functional on and hence V = □

Locally nilpotent transformations were defined in Section 1.5. Evidently


they are locally finite.

Lemma 16
Let F be a field containing IK, and let x e Endj^CF). Then

(i) for X to be locally nilpotent it is necessary and sufficient that Xp be


locally nilpotent,
(ii) if X is both semisimple and locally nilpotent, then jc = 0.

Proof (i) The proof is straightforward.


(ii) Since X is semisimple, x ^ is diagonalizable (Proposition 14). By part
(i), x^ is locally nilpotent, so 0 is its only eigenvalue. Thus x ^ = 0, and hence
X = 0. □

Let F be a finite dimensional IK-space, and let x e End,,^(F). By a Jordan-


Chevalley decomposition of x we understand a pair (s, n) where

5 is a semisimple transformation of F,
n is a nilpotent transformation of F,
s and n commute,
X = s n.
Note Observe that in such a decomposition both s and n commute with jc.
7.1 Locally Finite Endom orphism s and Jordan-C hevalley D ecom positions 595

Theorem 17
Let V be a finite-dimensional K-space and x e Endu^(F). Then there exists a
unique Jordan-Chevalley decomposition ( 5 , n) o f x. Moreover

(i) s and n are polynomials in x with zero constant terms,


(ii) a subspace U of V is x-invariant if and only if it is s and n invariant,
(iii) every endomorphism o f V that commutes with x commutes with both s
and n.

Proof Let K be the algebraic closure of IK. Then

Px,viO = ( i - “ 1)*' ■■■{ ( -


where a^,. . . , a ^ e IK are all distinct.

Case 1. . . ., e IK: By the Chinese remainder theorem applied to


IK[i], we can find a polynomial f( t) e IK[i] such that
f ( t ) = a, mod(r — for all i,
f ( t ) = Omod t.
(The last congruence is redundant if = 0 for some /.) Define
V-=^vEi V\{x — ol^Y' v = 0 for some n e Klj.
Then by Proposition V= where each is jc-stable.
Define s = fix), and note that 5 is a polynomial in x with no constant
term, and hence each is 5-stable. To compute the action of s on
write / ( 0 in the form

/ ( 0 = « ,i +
Since g^ixXx - a^)^‘ annihilates (Proposition 1), it follows that
s\Vi is scalar multiplication by a,, so s is diagonalizable and hence
semisimple. Define n = x — s. Then n is a polynomial in x with no
constant term. If e then nv^ = (x - s)(Vi) = (x - a¿)(v¿) so that
n^‘ annihilates Thus n^ = 0 where k •= max{A:/,. . . , k j , and hence
n is nilpotent. We have shown that x admits a decomposition of the
required type. It remains to be shown that such a decomposition is
unique.
Suppose that
X = s' n'
is any other such decomposition. Note that s' and n' commute with 5
and n because the latter are polynomials in x and that s' and n'
commute with one another. We have s — s' = n' - n, and by Proposi­
tion 15, 5 — 5 ' is semisimple. It is obvious that n' — n is nilpotent. Then
by Lemma 16, s = s' and n = n'.
596 Coiyugacy Theorems

Case 2. The general case: Let F = , a„), then F is a finite Galois


extension of IK. Let Vp == F and consider e EndpCKp) = F
End|}^(F), By Case 1 we have a Jordan-Chevalley decomposition
Xp =

Let [e^,. . . , 6/} be a IK-basis of V, Then {1 0 , 1 ® c j is an F-basis


of Fp. Let X, S, and N denote, respectively, the matrices of Xp, 5p, and
Wp with respect to this basis. Observe that since jcp(l 0 F ) = 1 0 x(V)
c 1 0 F, the matrix X has all of its entries in IK. For all a e Gal(F: IK)
and for all matrices M with coefficients in F, let denote the matrix
obtained by applying cr to each of the entries of M. We have

X = X ^ = {S = S^ +

Since 5 is diagonalizable (see Case 1), it is clear that so also is 5"^.


Similarly is nilpotent, since = ( N ^ y . Finally,

= ( S N y = ( N S y = N^S^,

By the uniqueness of the decomposition of Xp into semisimple and


nilpotent parts, we have

S = and N = for all o- e Gal(F, K) ,

and therefore both S and N have all their entries in IK. This shows that
5p and «p determine endomorphisms of F, which we will denote by s
and n, respectively. Then 5p = 1 0 5, Wp = 1 0 n, jc = 5 + n, and sn =
ns. By Proposition 12 and Lemma 16 it follows that s is semisimple and
n is nilpotent. It remains to be shown that both s and n are polynomials
in X with no constant term. Choose a basis {Aq, A^, . . . , A^} of F over IK
such that Aq = 1. By Case 1 there exists / ( 0 e F[i] such that 5p =
jCp/(jCp). Write f{t) in the form

/(0 = ¿ A , ® / , ( 0 e F ® ^ K [ i ] = F[/]
;=o

for some (unique) fj{t) ^ IK[i]. We have

1 ® s - (1 ® a: ) ( 1 ® / o( ^ ) ) = (1 ® D A; ® / y ( j : ) j .

Now the left- and right-hand side of this equality are linearly indepen­
dent elements of F 0 IK[jc] over IK. Hence both sides vanish, and
therefore s = xf^ix). Similarly for n.
It is clear at this point that parts (i), (ii), and (iii) hold. □
7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 597

Remark 1 Let W (z V be an x-invariant subspace. Let {s, n) be the


Jordan-Chevalley decomposition of x in V, Then both s and n stabilize W,
We know that s acts semisimply and n nilpotently on W. Thus n\w) is
the Jordan-Chevalley decomposition of the restriction x \ w

Next we generalize the concept of Jordan-Chevalley decomposition to


locally finite endomorphisms.

Theorem 18
Let V be a K-space, and let x be a locally finite endomorphism o f V. Then x
decomposes uniquely as a sum
X =s n .

where

s is locally finite and semisimple,


n is locally nilpotent,
s and n commute.

(i) if W <zV is an x-inuariant subspace, then W is both s and n invariant


and {s\w^ n\w) is the Jordan-Chevalley decomposition o f x\w\
(ii) every endomorphism o f V that commutes with x commutes with both s
and n. In particular, both s and n commute with x.

This decomposition is called the Jordan-Chevalley decomposition of x.

Proof. Define s and n as follows: Given e F, consider any finite-dimen­


sional jc-invariant subspace U o i V such that v ^ U. Then the restriction x\u
decomposes into semisimple and nilpotent parts

x\u = (-^1/7)5 + (^li/)v


according to Theorem 17. Define s(v) and n(i;) by

s(i;) = (:i|£/)^(u) and n{v) = (Jilt/)N (f)-

If U' is any other such space to which v belongs, then v s U + IT and


U + U' is j:-invariant and finite dimensional. By Remark 1 we have

(^l£/)i = and (Jtlt/Oi =


Thus

This shows that s is well defined, and similarly for n.


598 Coi\|ugacy Theorems

Since X is locally finite, we can write V = in LF3. Because of


the way in which s and n are defined, it is easy to see that

s stabilizes and acts semisimply on each l^, and hence s is locally finite
and semisimple as an endomorphism of V;
n stabilizes and acts nilpotently on each and hence n is locally finite
and locally nilpotent as an endomorphism of K;
X = s n and s commutes with n, since this is the case in each K.

Next we prove parts (i) and (ii).


(i) Let W be an jc-invariant subspace of V. Let w ^ W, and consider a
finite dimensional ^-invariant subspace Í7 of F to which w belongs. Then in
U both s and n act like polynomials in x. Thus sw and nw belong to U <^W,
showing that W is 5-invariant and Az-invariant. In addition {s\w,n\w) is the
Jordan-Chevalley decomposition oí x\w because of the way in which s and w
were defined.
(ii) Let y e End|}^(F) be such that xy = yx. We have to show that y
commutes with both 5 and n. By LF3 it will suffice to show that

( 1) sy\u = ys\u and ny\u = yn\u

for all finite-dimensional jc-invariant subspaces U of V, Let U be such a


subspace. For convenience set x = x\u, and let x = s + n ho the Jordan-
Chevalley decomposition of x in U. Let x = x\yu. Note that x stabilizes yU
and that the diagram
y
u yU

‘1
u y^ yU

commutes. If f( t) e K[t], it follows that

(2) f ( x ) y u = yf { x ) u for all u ^ U,

Let gU) and h{t) in K[t] be such that s = g(x) and ñ = h(x). Define
s = g (i) andñ = h(x). We show that s is semisimple and ñ a nilpotent
transformation of yU, First observe that both s and ñ leave yU stable, since
they are polynomials in x. To show that s acts semisimply in yU, write
U= where U¿ is 5-stable and irreducible. Then yU = Eyt^, and 5 will
be semisimple if we can show that each yU¿ is 5-stable and irreducible. Let
u¿ e Then by (2)

(3 ) syu¿ = g(x)yUi = yg(x)u¿ = ysu¿ e yU¿,


7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 599

Furthermore, if N czyU¿ is an 5-stable subspace, then, using (3), we see


that M := {u¿ e U¿\yu¿ e Ñ} is an 5-stable subspace of U^. Hence either
N = ( 0 ) or N = yU¿, showing that yU¿ is an irreducible 5-space. This estab­
lishes that 5 is semisimple (Proposition 2). To see that ñ is nilpotent, choose
k ^ N such that n ^ U = 0 . Then

n'^yU = h { x Ÿ y U = y h ( x y u = yn*i/ = (0).


Having established our claim, we observe that (5, n) is the Jordan-Cheval-
ley decomposition of x i n y U . Indeed, if y u ^ y U , then
(5 + n )y u = g ( x ) y u + h ( x ) y u = y (g (jc ) + h ( x ) ) u

= yxu = Scyu.

Thus Jc = 5 + Я in y U .
We can now finish the proof of part (ii). For this we have to establish (1).
If MG [/, then
syu = {x\yu )^yu = syu = g ( x ) y u

= y g ( x )u = ysu = y {x \u )sU = ysu .

From this it follows that y n = n y . Uniqueness is established as in the


finite-dimensional setting of the theorem. □

Given a locally finite endomorphism x we will denote its semisimple and


nilpotent parts by x^ and Xj^ respectively. If no confusion is possible we may
simply denote this by 5 and n .

Proposition 19
L e t X b e a lo c a lly f in it e e n d o m o r p h is m o f a K -s p a c e K, a n d le t x = s + n b e it s
J o r d a n -C h e v a lle y d e c o m p o s it io n . I f F is an e x t e n s io n o f th e n = 5p + «Р
is t h e J o r d a n - C h e v a l l e y d e c o m p o s it io n o f x ^ in F

By Proposition 11, x ^ is locally finite, and hence the Jordan-Chevalley


P ro o f.
decomposition exists. Now use Proposition 12, together with Lemma 16 and
the uniqueness of the Jordan-Chevally decomposition, to complete the proof.

We briefiy recall some basic results about eigenspaces and generalized


eigenspaces. Our objective is Proposition 23 which says that, just as in the
finite dimensional case, a locally finite endomorphism of a vector space over
IK is triangularizable if and only if all its eigenvalues lie in the field IK. Let x
be an endomorphism of a IK-space V . For each A e IK define the A-eigen-
space
V \ x ) := {v e V \x v = \v )
600 Coiyugacy Theorems

and the generalized A-eigenspace

V^{x) := e V\{x - A)”i; = 0 for some n = n{v) ^ N]

of X in V, If X is understood from the context we sometimes write and


instead of V^(x) and V \ x ) , respectively.

Proposition 20
Let X be an endomorphism of a K-space V,and let IK be an extension of K. Let
K := IK K and jc = 1 <S>jc e End^iV). Then for all A e K,

(i) Vfix) and V \ x ) are stable under x,


GO V^(x) = k ^ V , ( x X
(iii) I f X is locally finite and x c is the semisimple component o f x then
= ^ "(^ 5)-

Proof (i) The proof is straightforward.


(ii) Let be a basis of the K-space K. Any u ^ V can be uniquely
written in the form
V= ®
/el
Thus for A e K,

(x ~ \)^u = (1 <S>X — 1 <8>A)”l5


= (1 0 (jc —A))”C
= E a , ® (ac -
/el
from which part (ii) follows.
(iii) V^ix) is x-stable, and hence jc^-stable. By Proposition 5 we see that
X5, which is locally finite, acts on K;^(x^) like scalar multiplication by A. Thus
Vjf^x) c K/X5) = ^^(x^). The reverse inclusion is clear since for v e ^^( a:^),

( x ~ X)^v = ((^5 - A) + Xj^)^u = 0


for k large enough, given that Xp^ acts locally nilpotently. □

Proposition 21

(i) The sums tmd direct.


(ii) if V - ^AeiK^A^-^) tmd U is an x-stable subspace o f V, then U =
7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 601

Proof, (i) Since V \ x ) c it will suffice to show that the sum


is direct. Let Aq, Aj , . . . , A„ be distinct elements of K, and let j;, e
i = 0 , . . . , r. Assume that
^0 = + + Vr-
We want to show that Vq = 0. Let « q. be chosen so that (x -
A,)"'f, = 0 for all 0 ^ i < r. Consider P(t) and Q(t) in K[i] given by

^ ( 0 = n ( i - A , ) " ',
1= 1

Q(t) = ( t - x , y \
Then P( x ) vq = Q( x ) vq = 0, and hence Vq = 0, since P(t) and Q(t) are
relatively prime.
(ii) Let u e U, and write u = where 5 c K is finite and e
F^(x). We show that each g U; the result is then clear. We choose N ^ N
so that (x — = 0 for all A g S. Choose A g 5. By the Chinese remain­
der theorem we can find a polynomial p (t) g K[i] so that

p ( t) = 0mod(r - for all ¿li e 5 \ {A}


and
p{ t) = lm o d (t - A)^.
Since p(x) stabilizes U, we have p(x)u = e U. □

Proposition 22
Let V and W be vector spaces over K, and let X, p. ^ K. Let x g End(F),
y e EndOF). Then

( F ® W)j, +^(x ® 1 -I- 1 ® y),


^"a(^ ) ® WJ^y) c
(F® ® y).
Proof. The proofs of both statements are similar. We prove the second. Let
u ® w G F;^(x) ® W^(y). Then
(x ® y) —X p ) ^ ( v ® w)
= [(ji ~A) ®y -l-Al ® (y —ju,)]^(i; ®w)
in \
= - A) ® y)^(Al ® (y - p ) ) ^ ~ \ v ® w)

N
N
= y J(JC - A)'A^ ^v<&yj(y - •'w,

and this latter expression is 0 for A » 0.


602 Coi\jugacy Theorems

Here is a typical application of this result. Suppose that 9 is a Lie algebra


of derivations of an algebra A. Suppose that V, W are g-submodules of A
under this action. Let дг e g. Then Vj(,x)Wj^x) c (V W \^ J ix ). In fact we
have the commutative diagram

V®W ------------------------- » V(»w

X ~ { \ + f x)

where m is the multiplication m: A A ^ A, and the result follows from


the proposition.

An endomorphism x of a vector space V over IK is called triangularizable if


there exists a totally ordered basis B of V such that

(4) XV ^ ^ K m for all v e B.


u^B
u<v

We then say that x is triangular with respect to B. We say that x is strictly


triangularizable if
XLJ e ^ Ku, for a\\ u ^ B
u^B
u<v

Proposition 23
Let X be a locally finite endomorphism o f a vector space V over IK. Then the
following are equivalent

Tl: X is triangularizable
T2: All the eigenvalues o fx {in some algebraic closure o f IK) belong to IK
T3: F=
T4: V=

Proof. Let K be an algebraic closure of K. The result being obvious for


V = (0), we assume that V (0).
Tl_=> T2: Let B be a basis with respect to which x is triangular. Let
y e K® F be a A-eigenvector for 1 ® jc, and write y = ® u for some
u<.v
V E: B, where c^. = 1, e IK for all w, almost all c,^ = 0. Then
Ay = (1 ® x ) y = L ® E ^wu^-
U < V W <U

Hence ACy = c^,a^J^ and A e IK.


7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 603

T2 =►T3: If X is semisimple, then by Proposition 13 x is diagonalizable,


and T3 obviously holds. In general, write x = x^ + Xj^. Then we have from
Proposition 20(iii),
K ( x ) = V,(xs),
and the result follows.
T3 => T4: Use Proposition 21.
T4 =* Tl: Let ^ be the set of all totally ordered linearly independent sets
B of vectors in V for which
(5) XV for all V ^ B.
u&B
u<v

Let us first see that ^ is not empty. Since x is locally finite, any nonzero
vector y e K lies in a finite dimensional subspace F that is stable under x.
By T4 and Proposition 21(ii) we have F = © F / jc). A s is well known there is
a basis of each F / jc), and hence F, for which x has a triangular form. Briefly
(x —A) is nilpotent on F/jc), and hence F /x ) has an ordered basis for
which (x —A)i; c E IKw for all v g B^. Stringing these bases together
ueBx
gives us the required basis B of F. Thus B satisfies (5), which shows that

Partially order ^ by an order-respecting inclusion. Then, in choosing a


maximal element B^^^^ in apply Zorn’s lemma to If B ^ ^ were not a
basis of V, we could choose y e K, which is not in the linear span ( B ^ ^ ) of
Fjnax- F containing y with basis B, be chosen as above. Choose a subset
C = {ci,. . . , c^} c F so that C U B^^^ is a basis of F + and for all
; = we have xcj e L^^jKcj + ( B ^ ^ ) , Then C U B ^ ^ may be to­
tally ordered by using the orderings of C and B^^^ and by decreeing that
w > y for all w © C, z; © It is easy to see that C U B^^^ © which
contradicts the choice of B ^^. Thus B ^ ^ is a basis of V and part (i) is
proved. □

In the sequel we will find it necessary to work with Lie algebras of linear
transformations. We are interested in extensions of the classical theorems of
Engel and Lie that tell us that certain Lie algebra of transformations are
simultaneously triangularizable.
Let t be a Lie algebra over IK and let (ir,V) be a representation of t.
Then t is ( tt, F)-triangularizable, or triangularizable on V if there is a
totally ordered basis F of K such that F satisfies (5) for all x = ttUX í ^ t.
It is convenient first to introduce a more restrictive concept.
Let F be a vector space. A flag in F is a sequence F =
subspaces of F such that
(Fl) K ^K -i^ ^ y i ^ y o = (0),
(F2) dim - j.
604 Coi^jugacy Theorems

By abuse of notation we will often identify F with its top space For
example, when we say that e F, we mean that v ^
Let t be a Lie algebra, and suppose that tt is a representation of t on V.
A flag F = {FJ}/Lo ^ said to be t-stable if
7r{i)VjdVj forallO <y < « .
We say that t is ( tt, K)-hoistable if every element of V belongs to a t-stable
flag in V, Observe that if this is the case, then the elements of t act on V as
locally finite endomorphisms. If t is a subalgebra of a Lie algebra g, then we
say that t is hoistable in g if it is (ad, g)-hoistable.
Let t be a Lie algebra, and let ( tt, F ) be a representation of t in which
every element 7r(x), jc e t, is locally finite. We say then that V is 7r(t)-locally
finite.

Proposition 24
Let ( tt, K) be a representation of the Lie algebra i. I f i is {7r,V)-hoistable,
then it is ( tt, V)-triangularizable.

Proof The proof is a simple adaptation of the proof of T4 => T1 in Proposi­


tion 23. □

Proposition 25
Let t be a finite-dimensional Lie algebra, and let ( tt, F ) be a locally finite
representation of t. Then

(i) 7r(t) is locally finite in the sense that for each v ^ V, v lies in a
finite-dimensional Tr{t)-stable space,
(ii) (EngeTs theorem) I f for all jc e t, 7t(jc) is locally nilpotent, then t is
( tt, F ) = hoistable. In particular t is ('Tr,V)-triangularizable, In fact
it is strictly triangularizable, and for all x g t, tt{ x ) has only the
eigenvalue 0.
(iii) {Lie's theorem) I f 7r(t) is solvable and for each x G t 7t(jc) is
triangularizable, then t is {Tr,V)-hoistable,

Proof, (i) Let V F, and let [t^,. . . , be a basis for t. Then by the PBW
theorem.
ir(U (t))(t;) = u (U (K O ) • • • 7t(U(K íi ) ) ( í^),
and this is clearly finite dimensional and 7r(t)-stable.
(ii) Let i; G F. Then U '= 7r(U(t))(i;) is finite dimensional and 7r(i)-stable.
By the finite-dimensional Engel theorem (Exercise 1.22), there is a flag
F= with U^ = U such that Tr(t)L^ c ; = 1 ,..., n, and part (ii)
follows from Proposition 24.
7.1 Locally Finite Endomorphisms and Jordan-Chevalley Decompositions 605

(iii) Since for all X e t, 7t( a:) is triangularizable, all the eigenvalues of all
the 7t( x ) lie in K. Let v ^ V, and let U ■= 7r(U(t))(i;). Then by the finite­
dimensional Lie theorem (Exercise 1.22) we can find a 7r(t)-stable flag F for
U, □

The notion of t-hoistable is far more restrictive than t-triangularizable.


For instance, let K be a vector space with basis {?;oj • • •)? define
endomorphisms i = - 1 , - 2 , . . . of K by

a^Vj = 0 ii j 0,

Then t := is an abelian Lie algebra of nilpotent locally finite transfor­


mations and is visibly triangularizable. But t is not locally finite and hence
not hoistable.

Proposition 26
Let A be an algebra over IK. Let x be a locally finite endomorphism of A (as a
K-space), and let
X=s +n

be its Jordan-Chevalley decomposition,

(i) I f X is a derivation o f A , then both s and n are derivations o f A,


(ii) I f X is an algebra automorphism o f A , then both s and 1 s~^n are
algebra automorphisms o f A.

Proof. Extending the field if necessary, we can assume that IK is algebraically


closed. Indeed, let F be a field containing (K, and let x^ and A^ be as above.
By Proposition 19,
X^ — 5|p /Ip

is the Jordan-Chevalley decomposition of x^ as an endomorphism. Evidently,


if 5p (resp. /ip) is a derivation of A^, then s (resp. n) is a derivation of A.
Likewise, if s^ is an automorphism of A^, then s is an injective homomor­
phism of A. Since s is locally finite (Proposition 9), it follows that s is
surjective (Lemma 10). Thus s is an automorphism of A.
Assume then that K is algebraically closed. Let U and V be finite-dimen­
sional x-stable subspaces of A. Let

UV= [Zuv\u G i/, z; e K}.

Then UV is finite dimensional, and it is j:-stable whenever x is a derivation


606 Coi^jugacy Theorems

or an algebra homomorphism of A. Let


m n P
(6) u= 0 c/,, v= © 1 / , UV= ® ( U V ) y „
¿=1 ' ;=1 k=l

be the generalized eigenspace decompositions of U, V, and UV with respect


to X. Suppose that ;c is a derivation of A. Using Proposition 22 (see in
particular the remarks following it), we obtain

Thus, if u ^ U and £; e F, we have

(7) s{uv) = (A + fi)uv = Xuu + ujjLV = s{u)u +

By (7) and LF3 of Proposition 8 we conclude that 5 is a derivation of A, and


therefore n = x - s, being the difference of two derivations, is also a deriva­
tion.
Next suppose that x is an automorphism of A. Once again by Proposition
22 we have c (C/F),^. It follows that for all u ^ and v e F^,

s(uv) = XjjLUU = s{u)s(u)

and hence that 5 is a homomorphism of A. To show that s is surjective, it


suffices to show that s is injective, and this is clear because the spectrum of s
is the same as that of x and 0 is not an eigenvalue of x. We have therefore
shown that s is an automorphism of A. Finally, 5(1 + s~^n) = jc, so 1 + s~^n
is also an automorphism of A. □

7.2 LOCALLY FINITE ELEMENTS IN KAC-MOODY ALGEBRAS

In this section we investigate the relationship between (1) ad-local finiteness


and (2) local finiteness on integrable highest and lowest weight modules for a
Kac-Moody algebra g. Remarkably (1) => (2), a result that is important for
the conjugacy theorem of Section 7.4. In the meantime we use it to establish
an abstract Jordan-Chevalley decomposition of locally finite elements of
finitely displayed Kac-Moody algebras.

Lemma I
Let Vbe a K-space, and let x ,y ^ End|,^(F) be locally nilpotent. Then

(i) if V ^ V, the linear span o f the set {exp tx(u)\t e K} is the smallest
x-invariant subspace o f V containing u;
(ii) if exp tx = exp ty for all t ^ K, then x = y.
7.2 Locally Finite Elements in Kac-Moody Algebras 607

Proof, (i) Let n e Z^. be such that x"v = 0. Then the smallest x-invariant
subspace of V containing v is spanned by {x'l; | i = 0 ,1 ,..., n — 1}. Let
/ q, t i , . • •, t„-i be any n distinct elements of IK^, and consider the n vectors

tfx^ tj
exp tjx(v) = v + tjxv + +

We then have the matrix equation

XV

1 tn .n-\ X^V
X^V
(1) (expi^x(i;)) = := M
Jl
n —\ ''n —\

(n -l)!

and since the Vandermonde matrix M is invertible, part (i) follows.


(ii) Given V ^ V, choose n e Z+ such that = 0 and let
io,... ,tn-i distinct elements of IK. Define M as above. Then (1) gives

x^v y^u
M =M
7T
so XV = yv because M is invertible.

Lemma 2
Let I be a Lie algebra and let t and a be subalgebras o f I such that
[t, a] c a. Let ( tt, V) be a representation o f I and suppose that

(i) ad Í : a ^ a is locally finite for all t


(ii) there exists a finite dimensional subspace S of V which is 7r{i)-invariant
and generates V as an a-module.

Then t is TT-locally finite.

Proof Let F be the subspace of V consisting of all elements that lie in some
finite-dimensional 7r(t)-invariant subspace of V. The lemma will follow if we
show that F = V.
Let V ^ F and x e a. We show that ( ) e F. Set tt x v

M = 7r{U(t))v(zV
608 Coivjugacy Theorems

and
A^:= ad(U (t))jc c a .
By assumption both M and N are finite dimensional. Thus the space
Tr{N\M) spanned by all elements of the form 7r(n)(m) with n ^ N and
m e M is also finite dimensional. If i e t and Tr(n)(m) e Tr(N)(M), then
Tr(t)Tr(n)(m) = 7t([í , n])(m) + 7r(n)Tr(t)(m) e Tr(N)(M), since [t, n] ^ N
and TrCtXm) e M. This shows that 7t(N)(M) c F, In particular irixXu) e F,
and hence F is 7r(a)-invariant. On the other hand, F contains S, and 5
generates V as an a-module. Thus V = F. □

In what follows we use the notation that was introduced for Kac-Moody
algebras and groups in chapters 4 and 6.

Lemma 3
Let q be a Kac-Moody Lie algebra, and let G be its derived Lie group. Then
U j^ j(A d G^)ej spans Q+. In particular is spanned by elements which are
locally nilpotent on every integrable representation o f g.

Proof Let 5 be the subspace of g spanned by all elements of the form


Ad g+{ef) such that g+E G+ and j e J. Clearly S is an Ad(G+)-invariant
subspace of g+. Let i e J and i; e 5. Then
exp(ad te¡){v) = Ad exp te¿{v) e 5 for all i e K.
By Lemma 1 we conclude that [e^,v] e S. On the other hand, ej e 5 for all
jE J, showing that g+c 5.
If ( tt, V) is an integrable representation of g, then by Proposition 6.1.4

7r(Ad(g)e¿) = 7r{g)Tr(e¿)Tr{g)~^ for all g ^ G.


Since the right-hand side of this identity is locally nilpotent in V (by
definition of integrable), the last assertion of the lemma follows from the
first. □

Remark 1 It is clear that in the above lemma G+ can be replaced by G_


(and the ef^ by /y’s).

Proposition 4
Let Q be a Kac-Moody algebra, and let m be a subalgebra o f o f finite
codimension. Let P be a weight lattice for g. Then for any A ^ P+,
ann^(^)(m) := [v E L(A ) | m • i; = 0}

is o f finite dimension.
7.2 Locally Finite Elements in Kac-Moody Algebras 609

Proof. Let < • I • > be the Shapovalov form on L(A) defined by our anti­
involution a (See Ch. 2.8). By Lemma 3 there exists jc^, . . . , e
U yej Ad(G+)ey such that m together with these x¿, span g+. Then ax¿ e
U Ad(G_)/y, and so using Proposition 6.1.13, a ( x f ) , ... ,a(xf^) act locally
nilpotently on L(A); in particular they are locally finite. Let V ■=
U(IKc7-(jc^)) • • • U(lKcr(xi)) • £;+ where v+ is a highest weight vector for
L(A). Then dimu^CF) < oo. By PBW we have

L ( A ) = U ( c r ( g ,) - : ; ,
= U(o-(m)) • U(J<o-( a:,)) . .. U(J<o-( a:i )) •
(2) = U (o -(m ))I/cU ^ (o -(m ))L (A ) + F.
Let z e L(A). Then
2Gann¿(A)(m) >m • z = (0)
« < U ^ (m ) - z |L (A ) ) = 0
« < z |U + ( a ( m ) ) •L (A )) = 0.
Thus by (2), each z e ann^^^^Cm) is determined as a function <z | • > on
L(A) by its values on V. Since ( • I • > is nondegenerate and dim F < oo,
dim(ann^(^)(m)) < oo. □

Proposition 5
Let q be a finitely displayed invariant Kac-Moody algebra with weight lattice P.
Let t be any finite dimensional ad-locally finite subalgebra o f g. Then for all
Ae t is 7T^ and Tr^-locally finite. In particular ad-locally finite elements
are 7T^ and Tr^-locally finite for all K E: P^ (for the definition of see
Proposition 2.3.5).

Proof. Fix an arbitrary A e P^, We show that t is Tr^-locally finite. The idea
is to construct a finite dimensional 7r^(i)-invariant subspace A of L(A)
containing v^. The result then follows by Lemma 2. We construct A as the
subspace of elements annihilated by an ad t-invariant subspace of finite
codimension in g+. Since dim t < oo, t c Eht(a>>-A/9“ some M ^ N. Set
b = ]^ + g+. Then [t, b] c Eht(a> > so b + [t, b] = b + a for some fi­
nite dimensional subspace a of g.
Let F := ad(U(t))b d b. Then dim F /b < oo. Indeed ad t(b + ad(U(t))a)
c b + a + (ad U(t))a = b + ad(U(t))a, and ad(U(t))a is finite dimensional.
Thus b c F c b + ad(U(t))a (in fact we see that F = b + ad U(t)a).
Consider F-^ relative to the proper invariant bilinear form assumed to
exist on g. Since g““ and g“ are nondegenerately paired for all a e A+, and
since F D b, we have F c b c b. Furthermore F c Eht(a)> some
610 Coiyugacy Theorems

N > 0 , so Eht(a)>N9^ showing that dim (b/K ^)


< 00.
V-^ is ad t-invariant, but we need an ad t-invariant subspace of g+. To
this end set M = V nD g, which is also t-invariant because dq is. Now
M c b PiDg c To prove the last inclusion, simply note that b and Dg
are both b-invariant, so it suffices to show that (b n Dg) n b -^ = (0). From
Section 4.4, IBF, we have b n Dg = for k e \ {0},
{k lb) = (A:®, b> ^ (0). Thus M c g_^, so M = F P i g ^ . Since the projection
K b obtained by annihilating the g+ component has kernel M and
dim b < Qo, M has finite codimension in hence in g^..
Let m denote the subalgebra of g+ generated by M. Then m is t-in­
variant and of finite codimension in g+. By Proposition 4, A — ann^^^^fm) is
of finite dimension. Also A is Tr(t)-invariant: For a y e m, t e t,

Tr{y)Tr{t)a = 7r([y,í])í2 + Tr(t)7r(y)a = 0.

Since v +^ A and L(A) is generated as a g-module by 7r(t) is locally


finite by Lemma 2 (applied with I = g = a, 5 = A).
To show that t is Tr^-locally finite, we apply what we have proved to the
locally finite algebra □

Corollary
Let % be a finitely displayed invariant Kac-Moody algebra with Cartan matrix
A. Let A be a sub-Cartan matrix of A , and let q be a subalgebra o f g obtained
by restricting A to A (Section 4.3, Remark 1, and Proposition 4.3.2 and 4.3.3).
If i is a finite-dimensional diá-locally finite subalgebra o f g, then t is an
did-locally finite subalgebra o f g.

Proof Let b be the diagonal subalgebra of g. It is clear^ that t is a


(finite-dimensional) subalgebra of g + b- Moreover, since [t,b ] c g, t acts
ad-locally finitely in g + b- By the reduction result of Proposition 6.2.4 and
Proposition 5 we conclude that t is ad-locally finite in g. □

We will eventually prove that locally finite elements of finitely displayed


invariant Kac-Moody algebras have Jordan-Chevalley decompositions. We
have here a weak form that only involves the adjoint representation.

Proposition 6
Let q be a finitely displayed invariant Kac-Moody algebra. Then every ad-
locally finite element jc g g decomposes as

(3) a: = a;5 +
7.2 Locally Finite Elements in Kac-Moody Algebras 611

where is 2id-semisimple, is locally 2id-nilpotent, and = 0. This


decomposition is unique up to addition of elements from the centre of g.
Furthermore

(i) For any subspace M of

[ a:, M] c M <=> [jc^, M] c M and [ x j ^ , M] ( z M,

GO g"" = n (see Section 1.1 for notation).

Proof Let A be the structure matrix of g, and suppose that A has dimen­
sion / X /. If ^ is singular, let ^ be a symmetrizable nonsingular Cartan
matrix of dimension (/ + corank X (/ + corank A ) in which A is embed­
ded as a sub-Cartan matrix (see Lemma 4.2.1). Then the minimally realized
invariant algebra g with structure matrix A has a Cartan subalgebra § of
dimension (/ + corank A). Use Proposition 4.3.2 to embed g in g. Because
of the dimension of the Cartan subalgebra of g must be all of If .^4 is
nonsingular, then we set g = g. In either case g is centerless, andhence all
derivations of g are inner (Proposition 4.1.15).
Since X is ad-locally finite, it is also ad^-locally finite (corollary to Proposi­
tion 5). Then by Theorem 7.1.18 and Proposition 7.1.26. ad x is uniquely
expressible as where X^ and Xjs^ are, respectively, semisimple and
locally nilpotent derivations of g and [X^,Xj^] = 0. Since they are inner
derivations
ad X = ad ^5 + ad Xj^

for some unique (g is centerless) x^, Xy^ e g and [x^, Xj^] = 0.


Let n be the normalizer of g in g. Then
^
n z> g Z) 6, and so n = ©Ofe A
But, ii a ^ Q and x e n “, then g z> [x, if] = (Kx. Thus n = g. Now, since x^
and Xy^ normalize g (as x does), we see that x^, Xj^ e g.
The remaining parts (i) and (ii) follow from Theorem 7.1.18(0 and (ii).

Proposition 7
Let Mj, M2 be integrable Q-modules in ^ ( P ) , and let S be a G-invariant subset
of Mj. Then M5 := {/ e Homj^(Mi, M2) I /(5 ) = (0)} is a q-submodule o f
H ohI kÍM i , M2).

Proof. It suffices to prove that c and that for all a e and


e^ e g", e^ • c M^. We use the subgroup 2Í of G to establish the first
and the subgroups G “ == exp(g") for the second.
Let / e M5 . Let F^ c Mj be a finite dimensional A^-invariant subspace.
Since /(F j), is finite dimensional and the action of X is locally finite, we can
612 Coixjugacy Theorems

embed / ( jPj) in an A^-invariant finite dimensional subspace F2 of M2. Then


we can consider f\f^ as an element of the A"-invariant finite-dimensional
space Homj^CFi, F2). By Remark 6.1.6, and F2, hence also Hom^(Fi, F^),
are ]^-modules. Thus for h ' f)\r^ = h • ( / I fj) ^ ^om^(Fi, F^)- Fur­
thermore M^l/Tj is an AT-invariant subspace of Homu^ (Fj, F^), and hence by
Remark 6.1.6) again, it is 1^-invariant. In particular h ' / \ fi ^
Now every element of Mj lies in some finite-dimensional ^-invariant
subspace Fj of Mp and hence any g e Homj^(Mp M2) is determined by its
restrictions to such spaces. Since h - f E: M^I fi spaces, h •f ^ M5.
This proves that • M^ c M5.
For the second part repeat the argument using the groups G“, but this
time invoking Lemma 7.2.1. □

7.3 THE KOSTANT CONE

Throughout this section (g, ^ ) will denote a finitely displayed invariant


Kac-Moody Lie algebra over K. As usual ^ = (i), g+, Q+, cr). Let P+ denote
the set of dominant integral weights of some weight system F. If A e F+, the
g-module L(A) is integrable (Proposition 6.1.5), and L(A) ® L(A) is com­
pletely reducible (Corollary to Theorem 6.5.1). Moreover, if (A,i;+) is a
highest weight pair for L(A), then (2A, i; + 0 1;+) is a highest weight pair for
L(A) 0 L(A), and we see that

(1) L(A ) ® L(A ) = L (2 A ) ® © n^L(yii)


<2A

for some nonnegative integers n^.


Given A e P+, we define the Kostant cone T(A) by

( 2) r ( A ) = {y e L ( A ) It; ® v g L(2A)} \{ 0 } .

In this definition it is understood that v ® v is decomposed according to (1).


5^(A) is a cone in the sense that V(A) c 3^(A).
Let G be the group associated to g (Section 6.1), and let G be its derived
group. Then G acts on L(A), hence on L(A) ® L(A), and stabilizes the
decomposition (1). Hence for g & G,v ^ T(.A) => gv ® gv = g(v ® u) g
|( L ( 2A)) = L(2A) =» gi; G T(A), so V{A) is stable by G. Define an action
of the group IK^X G on T(A) by

( a , g ) v = a(gv).
7.3 The Kostant Cone 613

Theorem 1 [PK]
K^X G is transitive on TiK).

The proof of this theorem is the focus of this section. We begin by


expressing X(A) as the solution space of a specific set of quadratic equations.
Let be dual bases (relative to the invariant form of g, see Section
4.4) for g“ and g““, a e A, 1 < / < dim g“. We can clearly take
if a ^ 0, and it is convenient to assume this. Set Q ^= 0 .^ and
as is customary.

Proposition 2
Let A c P_^, Then T{K) is the set of nonzero solutions in L(A) 0 L(A) of the
equation

(3) (A 1A)(í; ® u) = Y.
ae A i

Proof Let r = + T' be the Casimir-Kac operator of L(A) (Section 4.5).


If y G L(A), then

T (u u) = r®(£; 0 ¿;) + P(i^ v).

Now

F ( ¿ ; <S> V) = 2 Yi ®
aeA+ i

= r ( i ) ) ® i; + y ® r '( y ) + 2 Y ®
a e A+ i

+2 Y Y^X^aV ® yi'^V
a e A+ I

= r(u ) V + V ® r(v) + 2 Y ® yi'^i’


a G A\{0} i

(using Thus

(4) r ( ü ® ü) = r ( y ) ® Ü + y ® r ( u ) + 2 Y Y^a^^ ® ya^^


a G A\{0} i

+ r^ (i; 0 í;) — T ^ ( v ) <S> V — u <S> T ^ ( v ) .

W riting u= w h ere e L (A )^ , ^ P, and using the fact that


614 Coivjugacy Theorems

acts on L(A)^ as scalar multiplication by (/jl + 2p\ fi), we have

(5) r ^ ( v (8> u ) ~ r ^ ( i ^ ) (8> u — u (8> r ^ ( i ^ )

= H ( ( m + i' + 2 p I jll + i ' ) - ( p -h 2 p I p )


/1, i/eF
- ( u + 2 p l v))v'^ ® d "

= 2 ^ (/i, 1 ® u"

= 2 Y, ® yo \v'^ ®
fX.V^P i

= 2Y,X q^ ® y o \ ^ ® ^)-\

Combining this with (4), we obtain

(6) r ( i ; <S> z;) = r ( z ; ) <8) f + í; 0 T( i; ) + 2 51 ®


ae A /

Now z; G L(A), and therefore F(z;) = (A + 2p | A)z; [Proposition 4.5.3(i)].


Thus

r(z; 0 z;) = 2( A H -2p|A )z;0z; + 2 52 ®


ae A /

Thus (3) is equivalent to

(7) r ( v 0 z;) = (2(A | A) + 2(A + 2p | A))z; 0 z;


= 4(A + p I A)i; 0 V.

Using (1) decompose v ® v = Yiv 0 z;)^, {v 0 v)^ e L(p,), p e


(2A 4 Q+) n into components using (1). Then (7) holds if and only if for
each p with {v 0 i;)^ ^ 0 we have

( p + 2 p 1 p ) = 4( A + p I A )

[Proposition 4.5.30)]. However, for p = 2A - a, a e Q+U{0},

(2A —a + 2p |2A —a)


= (2A + 2p|2A ) - ( q:|2A ) - (2A - a + 2p | a ) < 4( A + p | A)

with equality if and only if a = 0 (since 2A,2A —a e P^). Thus (7), and
hence (3), is equivalent to z; 0 z; e L(2A). □
7.3 The Kostant Cone 615

Let V e and write v = where e L(A)^ for all e 1^*.


Recall that the support of u, supp(i;), is defined by

supp(i;) = {)Lt e I =?i=0} c A + 2-

The arguments that follow involve a detailed discussion of supp(z;).


Let A be the /-dimensional real affine space A = R with translation
group F = R = ©¿^jRa,. Identify A + F with A by identifying A + o'
with a e A. This allows us to think of supp(¿;) as a finite subset of the real
affine space A.
Recall that a mapping / : A -> A is called an affine linear transformation
if there is a linear mapping df :V ^ V such that

f ( x - \ - u ) = f { x ) + df {v) iox dX\ X ^ V.

If A" c A and / : A ^ A is an affine linear transformation of A, then ( f ( X ) )


= f i x ) , where < > stands for “convex hull of.” The Weyl group W acts on F
in the standard way. Let w ^ W . We can view w as an affine linear
transformation of A by writing

H^(A + a ) = A + (wA - A) + w{a) for all a e F.

(Note that the last two summands lie in F.) It is easy to verify that this gives a
group action of W on A.
Let V e y{A), and let S ( v ) c A be the convex hull of supp(i;). An
element ¡i e S ( v ) is said to be a vertex of 5(i;) if

VI M ^ S(v),
V2 there exists an (affine) hyperplane H of A through fjL such that
S ( v ) \ { f j i } lies entirely in one of the open half spaces separated
by H,

By an edge of S ( u ) we will understand a pair of distinct vertices {A, of


S(v) for which there exists an (affine) hyperplane H of A such that

E l the closed line segment [A, joining A and /x lies in H,


E2 S ( v ) \ [A, ] lies entirely in one of the open half spaces separated by
ijl

H,

In broad outline the proof of Theorem 1 runs as follows: Let be a


highest height vector of L(A). Evidently + e V(A). Thus we have to prove
that 1K®G • u += T(A). We take any element u of TiA) and show that the
vertices of S(u) lie in WA. We then select the “lowest” vertex and proceed to
strip edges off S{v) by repeatedly replacing v by elements g • z;, g e G for
which Sig ' v) has less edges than S{v). Ultimately we arrive at an element u
616 Coivjugacy Theorems

for which 5(¿;) has no edges. Then v ^ where N is the subgroup of G


defined in Section 6.1, and we are done.

Lemma 3
Let g = ^ ^ grading o f g by (R, + ) satisfying Qq = suppose
that L(A) = is an U-grading o f L(A) compatible with that of g:

g^ • L{A)t c L{A)s+t for alls, i e R.

(i) If V ^ L(A)^ n V(A) for some r e R, then v e L {A Y ^ for some


w ^W .
(ii) Let V e T{A), and write v = e ^(A)^ for all 5 e R. //
m = max{5 e R | =?t 0}, then e ViA).

Proof (i) We first show that if a e A, then g" c g^ for some 5 e R. It


suffices to show this for a = ±a¿, i e J. Clearly s ¥= 0 if a ^ 0. Write
e- = E jc^, jc^ e g^. Then for all /i e 1^,

E [/i, Jcj = [h,e¡] = a¡{h)e¿ = £ a¡{h)x,.


S&U S^U

But [h, jrJ e g^, since = gg. Thus [h, x j = af h)x^, and we conclude that
there is only one s with x^ 0, and this x^ lies in g"'. This shows that
g“' c g^. The argument for g““' is similar.
Suppose that u e L(A)^ c V(A). Then (Proposition 2)

( A | A ) ¿ ; 0 ¿ ; = Y,
ae A i

The left-hand side of this lies in L(A)^ ® L(A)^, whereas x*¿^v ® y^'^v e
L(A),+j ® L(A)^_^ for some s, and s 0 unless a = 0. Thus

(A I A)u ® i! = E^o^^ ®

Writing V = with e L(A)'^ for all f i e P(A), we have [see (5)]

( Al A) i i ®i ; = iiJi\v)v>^ ® v ’',
!X,vE.P

SO(A I A) = (/£ I v) whenever 0 =?^=0. By Proposition 6.2.9, we conclude


that V = for some p e B^A.
7.3 The Kostant Cone 617

(ii) The R-grading of the g-module L(A) induces a compatible grading in


L(A) 0 L(A) by setting

(L (A ) ® L (A )),= L L (A ),® L (A ),.

By restriction L(2A) inherits this grading, for the argument by which we


showed that g"' c for some s can be used to show that the highest weight
space L(A)^ of L(A) lies in L ( A \ for some s. Thus u + ^ u + ^ L(A 0 A) 255
and it follows that L(2A) is graded.
Let V be as in (ii). Then since v ® u ^ L(2A) and v v = HS, / € U^S ®
with deg(z;^ u^) = s t < 2m, we see that 0 is the homogeneous
component of degree 2m of u u in L(2A). In particular e T(A). □

Lemma 4
Let A" be the usual n-dimensional real affine space. Let S c. be a finite set,
and let (S ) be its convex hull. Let X (Z be a countable set.

(0 Suppose that there exists Sq ^ S and an {affine) hyperplane H o f A"


satisfying
(ia) Sq ^ H n X ,
(ib) ( 5 >\ {5q}lies entirely in one o f the open half spaces separated by H.

Then there exists an {affine) hyperplane H ' satisfying (ia) and (ib) and such
that in addition

(ic) H'nX={s^}.

(ii) Let c Ei {S), and suppose that there exists an {affine) hyperplane H of
A" such that < 5 )\{c} lies entirely in one of the open half spaces
separated by H. Then c e S.

Proof, (i) Let R" be the real vector space of translations of A". For each
V E R", <5> V is the convex hull of the set 5 + i;. If we now translate 5,
X, and H by - 5 q, we see that we can suppose that Sq = 0.
Choose a basis [u^, f 25 • • • 5 of such that {v2, . . . , is a basis for H
and • / / = 0, where • is the usual inner product on R". Since 5 \ {sq} is
finite and lies entirely in one of the open half spaces separated by H, we can
suppose that there exists e > 0 such that

^\ {*^} ^R"I^ >e}-


0
618 Coi\iugacy Theorems

For each s ^ S \ set

t/^ == {w e R" I 5 • M> e}.

Then is open, and

i;i € i/ := f l U,.
s^S\{so)

For each a e A"\ {0}, //^ •= {w e R" | w • a = 0} is a hyperplane. By the


Baire category theorem, we can choose a point w ^ U \ \J ae x\{0)^a' The
hyperplane H' ■= {x ^ W \ x ' w = 0} satisfies (ia), (ib), and (ic).
(ii) Suppose that c ^ S. Then by assumption 5 lies in one of the open half
spaces, say, separated by H. Since is convex, <5> c a contradic­
tion. □

Lemma 5
Let V e y{K). Then the following two sets are equal:

(i) The set of vertices o f S(u).


(ii) supp(i;) n Wh.,

Proof Let /X be a vertex of S{v), By Lemma 4, /x e suppiu), and we can find


an affine hyperplane of A through fx such that

/ / ^ n ( A + 0 ) = {m},

and S(v) \ {ju,} lies entirely in one of the open half spaces, say, separated
hyH^.
Let H be the unique hyperplane of V (translation group of A) satisfying
= H ti. Choose p e V so that p, + p ^ Then V = H ® Rp.
Define
d:V^R

by means of the projection of V onto Rp parallel to H:

d{h + tp) = t for all h g H, í g R.

If we consider the restriction oi d to Q c.V , it follows from our hypothesis


that d{a) = 0 if and only if a = 0. By means of the group morphism

d:Q ^ R
7.3 The Kostant Cone 619

we can define a grading of 9 by (IR, + ) by setting

9 , == E { 9 “ U ( a ) = t},
ol^Q

for all i e R. By construction Qq = i). Similarly we can give L(A) an (U, + )


grading by setting

L(A ),== E { L ( A r " “ |d ( « ) = r}


a^Q

(with defined as above). By our choice of p we see that

V e E M A ),
i<0

and that

L ( A ) '‘ = L ( A ) o .

Since the grading of L(A) is compatible with that of Q, we apply Lemma 3 to


conclude that if i; = ^ ^(A), then Uq = ^ T(A) and p e H^A.
Conversely, let p e w k n supp(¿;). Let w and « e N lie over w, i.e.
n = [w\. Let u e cp e 1^*. Then nu = # 0, and clearly
supp(«i;) = vv’CsuppCi;)) and S ( n u ) = w5(¿;). Thus to prove that /x is a vertex
of S ( u ) , we may assume that p = A . But it is obvious that if A e S i u ) , then
A is a vertex, since all other weights in supp(¿;) lie in A —(2+. □

Lemma 6
Let V e V(A), Then the edges o f S(u) are parallel to real roots o f g.

Proof Let V = e L { A Y , and let [ \ , p ] be an edge of S(u). Then


\ ¥= p, so (X \ p) ¥= ( A \ A). Look at the component for L i A Y L { A Y in

(A I A )v V = Y , ®

namely

(A I A)u^ = Y ®
(or, <p, (/r)eF i

where F = {(a, (p,il/) \ a ^ A, cp a = k, if/ — a = p}. If for all (a, cp,il/) ^ F


620 Coi\jugacy Theorems

we have a = 0, then

(A I A ) f ® 0
i

= (A I 0

which we just saw is impossible. Thus there is an a e A \ {0}, for which


A - a and + a lie in supp(i;). By assumption S{v) \ [A, /x] is in one of the
half open spaces defined by some hyperplane H containing If
either A - a or + a does not belong to [A, ¡ \ then the midpoint ^(A + fji)
jl

of the line segment [X - a, + a] lies in


jjl contrary to |(A + /x) e [A, /x].

jjL a

It follows that A —a and fx a must lie on the line through A and /x. Thus
a is parallel to [A, /x]. Writing X — fx = ka, k ^ U, and choosing w ^ W so
that wX = A, we have A —wjx = kwa. Thus kwa e Reversing the roles
of A and jjL, we find that —kw'a ^ Q+ for some other w' g W. Thus a is a
nonzero root whose IT-orbit meets both 0 + and —Q+. By Proposition
4.1.5(iv) we have a e*^®A. □

Lemma 7

(i) For a and X e WK,

(A + Ua) n W X = {A,r^A}.

(ii) Every edge ofS(v), v e is o f the form [A, r^X] for some real a.
7.3 The Kostant Cone 621

Proof, For A + A:a e WK = W \,

(A I A) = (A + A:a I A + fca) = (A I A) + 2 k { \ \ a) + k'^{a \ a)


2(A I a)
k = 0 or k = —
( a Ia )

hence we have part (i).


Now let [A, fjb] be an edge of S(v), v e By Lemma 6, e (A + Ra)
n WK for some real root a, and hence, by part (i), ¡i = r^A. □

Next we gather together a few definitions and simple results that are used
in the process of “stripping off edges.”

For A e P^, A e WK, define

i/(A) = Pi wG_w~'^ n G^,


w G i W , wA. =A

which is clearly a subgroup of G and

'®A+(A) = {a 1r„A > a).

Note that we use the usual convention that wG_w~'^ makes sense since H
normalizes G_. (Remark 6.3.2).

Lemma 8
Let A e and let A e WK. Then

(i) is finite,
(ii) a e*^®A+(A) <=> exp g“ c i/(A).

Proof Let A = wA, a Then

r^A>A <=i> <A,a^> < 0

<=> < A , < 0


^ w
—1a '
622 Coiyugacy Theorems

and is finite by Proposition 5.2.3. This proves part (i). Now using
Proposition 5.2.6,

9“ = “w ' c wg_w ^

Thus a e*^^A+(A) => exp g“ c wG_w~^ Pi G+. □

We now prove Theorem 1, namely that G • u^= V(A).

Proof of Theorem 1. It is clear that IK®* G • f +c T{k). As for the reverse


inclusion, let V e T{A). Partially order Q by the height function, and
transfer the order to A + Q- Let A be minimal in the set of vertices of S{v).
If A has no edges from it, then S(u) = {\} [5(z;) is the convex hull of its
vertices], and u e L(A)'^ = L(A)’^^ some w ^ W. Then u e K^nv+, where
(n ) = w.
Suppose then that S(u) has at least one edge, say, [A, r^A], a g A+. Write
r^A = A + ka, and let Uj be the component of v in L(A)'^'^-'“, ; = 0 ,1 ,..., A:
(k is positive by the minimality of A). Look at exp te^ • v and its component
in

tfc-iefc-i
-Vn + +
k\ ( A :- ! ) !

If iei/j \)v^_j = ajv^, Oj e then this reads (i*a. + k-1


+ *** -\-aQ)Uf^. In an algebraic closure (Kof IKwe can find i = io to make this
0. Then at least in ^ (A ), r^A ^ supp(exp t^e^u). Let / / be a hyperpl^e
through [A, r^A] so that 5(i;) \ [A, r^A] (zH ^. Then supp(exp c
and since r^A is not in the support, S(txp tQe^u)\{\} c / / ^ . In particular
A + a ^ supp(exp tQe^u), so io^a^o + = 0. Thus e IK.
We now replace u by expio^a^^) so S ( u ) \ { \ } If r^A,)3 e A+, is
another edge then [A, r^A] points “into” Thus supp(exp se^v) \ {A} c //+;
in other words, when we remove more vertices, we cannot add a back again.
It remains only to recall that from Lemma 8(i),

{a I r^A > a} is finite.

A finite number of repetitions of the process just described removes all the
edges from 5(z;). □
7.4 Coiyugacy of Split Caitan Subalgebras 623

7.4 CONJUGACY OF SPLIT CARTAN SUBALGEBRAS

Proposition 1
Let I be a finitely generated Lie algebra with center 5. Let t be a subalgebra of
I, hoistable in I. Then

(i) adji is finite dimensional and solvable,


(ii) t is solvable. It is also finite dimensional z/ t n 5 is finite dimensional,
(iii) if Í is an ideal o f some Lie algebra g then t is hoistable in g.

Proof Let {x^,. . . , be a set of generators of I. For each 1 < j < m, let
be an ad(t)-stable flag containing Xj. Consider the following se­
quence X of subspaces of I:

m m —\
■■■
j=i j=i
m —1 m —\

E E K ^P ^ ••••
;=i y=i

By removing redundant terms in X, we obtain an ad {(testable flag, also


denoted by X, which contains x^ , . . . , x^. Thus adi(t)|;^^ c End(A") is finite
dimensional and solvable. Since X generates I the map adjt ^ (adjt)l;^^ is
faithful, and hence adit is solvable. Finally, t is solvable, since

t / t n a - a d i(t).

If t n 3 is finite dimensional, then clearly t is also finite dimensional. This


proves parts (i) and (ii).
Let i; G g \ I. Then [t, i;] is a finite-dimensional subspace of I (by part i).
As above, we can see that [t, i;] lies inside an ad t-stable flag • d Fq
of I. If we set = Kv ® V^, then Vn +\ ^ 3 Vq is an ad t-stable flag
in g that contains v. □

Let M be a g-module. Recall that Af * has a g-module structure via

X ' f ( v ) = —f { x • v) for alt jc e g, / e M*, v ^ M.

We let T \ M ) and S \ M ) be the space of degree 2 elements of the tensor


and symmetric algebras of M, respectively. These we view as g-modules
624 Coi\jugacy Theorems

(Section 1.6). The canonical surjection

v : T ^ { M ) ^ 5^(M )

induces an injection of the dual spaces

^ T^{Mf
via

*^7(2) = / ( v ( z ) ) f o r a ll/G z € r^ M ).

One verifies easily that the g-module structure of S \ M Y is that induced


from T \ M ) * by restriction to v‘( S \ M) * ) . More precisely, if / g SHM)*,
a: G g, and z g T \ M ) , then

{x - f ) { v { z ) ) = {x ■ v ' / ) ( z ) .

Let Q(M) be the space of quadratic forms on M. By definition F e Q(M)


means that F : M K satisfies

Q1 = a^F(u) for all i; e M for all e IK,


Q2 the map B : M <S>M K defined by

B{u, w) = j [ F{ v + w) - F{u) —F()v)} for all v, w E: M,

is bilinear.

Now B defined in this way is symmetric and hence defines a unique element
B^ e S K M T by

B^(uw) '= B( v , w) for all v, w ^ M.

Conversely, each element B^ ^ S \ M ) * determines a symmetric bilinear


form B and then a quadratic form F by

F( v) = B ( v , v ) =

This bijection of Q(M) and SKM)* determines, by transport of structure,


a g-module structure on Q(M). Namely for y g M,

(1) x-F {v):=x-B ,iv^)= - B^i xv ^)

= - 2 S * ( ( a: • v)v)

= + x ■v f ) + B ^ ( v ^ ) -h • u)^)
= —F{v + X ■v) + F( v ) + F ( x • v).
7.4 Conjugacy of Split Cartan Subalgebras 625

Define a g-module map

M* ®M* S^{M)*
by
f® g^fg
with

fg(vw) ■■= + /(w ')^ (^ )} for all e Ai.

Evidently fg = gf.
Let {y,}, ei be a basis of M, and let c M* be the corresponding set
of dual elements. Then all i, j e I, v* vf satisfies

U
1 if i = ; = * = /,
vO otherwise.

If we fix a total ordering < in I, we have

S^(M)* = n iK u f y /.
i< j

A typical element of S \ M Y may be thus written as a formal sum


= Li<jbijU*uf, or more conveniently as a formal sum

5* = L ^ u v r ^ r ,
i j

where

\bij if i < j.
^ij ^ji
, i>ii if i = J-

Then defines a quadratic form F whose associated bilinear form B


satisfies
B(Vi,Vj) =B^(viVj) =a^j.

In the sequel we will write quadratic forms in the notation

F = ' LaijVfvf,
‘J

thus identifying F with its associated element in In this notation the


626 Coi\|ugacy Theorems

g-module structure of Q(M) is given by

(2) x - Z a,jv* v * = Z a,X( x ■vr)v* + v H x ■v*)).


1<J 1<J

Lemma 2
Let cr, T e Endo^CM), and write

= LbfkVj,
j

JlbJkVj.

Let V = e M, Then in T \ M ) we have

au ® TV = Y .F lp’''\v)V j ® Vp,
j,P

where

k,q

Proof.

av <S)rV=i ' Zb^k^kv] ® ( E


^j,k ' ^P,q

= E (E ®V p .
J,P^k,q '

On the other hand, we have

Y. bf kb; , {vt , v){v*, v)


k, q

^ ^jk^pq^k^q'
k,q

Corollary
In T^{M) one has for all f e M,

v ® v = Zvfv^^{v){v^^v^).
j,p
7.4 Coiúugacy of Split Caitan Subalgebras 627

Proof. Set O' = T = id in the last lemma. □

Henceforth we assume that g is a finitely displayed invariant Kac-Moody


algebra, that P is a weight lattice, and that M is an integrable g-module.
Then S \ M ) is integrable, and by Proposition 6.1.16 we see that ==
Q(M) is a G-module satisfying

g-F{v) =g-B^{v^)

= = F ( g - 'y ) f o r a l l g e G , f e Q ( M ) ,ü G M .

In terms of a fixed basis we have

g • Y.ai jvf vf = i;a ,.y (g - v f ) { g - v f ) .

Our next objective is to interpret the defining equations of the Kostant


cone y = c L(A), A e in terms of quadratic forms. We maintain all
the notation of Section 7.3 concerning y. The representation of g on L(A) is
denoted by
For each a e A_^ we write

and

We have, for v =

® y^J'^V = Y,Flp’' \ v) Vj ® Vp,


J, P

where

k,q

Furthermore by the last corollary

(A| A)i> ® Ü = ( Á \ Á ) ' Z v ^ v * ( v ) { v j ® Up).


628 Coi^jugacy Theorems

Hence by Proposition 7.3.2, the equations defining the Kostant cone become

(3) ((A I \ ) v*v * - E i V p }’P ^

In short, the Kostant cone is defined by a set of quadratic equations.


Let

/r ( A ) = = { /^ Q ( M A ) ) l/( y ) = 0 f o r a lli) e r ( A ) } .

Lemma 3
Let A e P T h e n

(i) = {y e L(A) I f(,v) = 0 for all f e


(ii) is (^-stable.

Proof, (i) Since contains all quadratic forms of type

it is clear that TiX) is the zero set of Iy(^xy


(ii) The G-invariance of TiX) and the identification of 2(L(A)) with
5^(L(A))* imply by Proposition 7.2.7 that is g-invariant. □

Fix A ^ P+, and let t be a 7T;^-hoistable subalgebra of g. Consider the


finite-dimensional t-module F = U(t)i;+, where is a highest weight vector
of L(A). V affords a representation tt of t by restriction of to V. We have
a natural homomorphism

< p:G L (V) ^ G L {S \V )* ) =GL

of algebraic groups defined as follows. The map g e G LiV) extends uniquely


to GL(5(F)) and by restriction to an element g* of GL(5^(K)). Now (p(g) is
the inverse transpose of g*. Thus for g SKV)*,

(4) <p(g)B^(vw) = B:^((g~^v)(g~^w)) for all e F.

Since ^ is a polynomial map we can consider its differential [Bor, Ch3]


7.4 Coivjugacy of Split Cartan Subalgebras 629

In what follows we fix once and for all a 7r(t)‘Stable flag

F = F „ d K „_, d ••• :d I/o = (0)

of V. Having done this, we define the corresponding Borel group

By ■■= { g ^ G L { V ) \ g V k ^ V ^ , k = 0 , . . . , n )

and its Lie algebra

hy-.= { X ^ q l { V ) \ X V ^ < z V ^ , k = 0, . . . , n } .

Observe that by definition

(5) ^(t)<zby.

By restriction we obtain a map of to functions on F c L(A), which we


will denote by f >-*f. We set

/==

The elements of / are quadratic forms on V and can therefore be thought of


as elements of S \V )* .
There are two natural t-module structures on /. Since F is a t-module,
the space Q (F) of quadratic forms on F admits a t-module structure whose
action is given by (1). Let / e / c Q ( F ) , and let ;c e t. Since is
g-invariant, it follows from (1) that

( 6) X - f = x •f

and hence that / is a submodule of the t-module Q(F).


On the other hand, t can be made to act on / c 5^(F)* via

(V) X ■f = [d(pTr(x))f for all a: e t .

Let us see that these two actions of t are the same.


Let e S^(F)* be the linear map defining / , and let “ : T^(F) -» S^(F)
be the canonical map (called v above). For all e F and g e GL(V) we
have from (4)

<p(8)B^(vw) = B^{g~^vg~^w)

= ® ® M'))-
(>30 Coi^ugacy Theorems

Thus for all X e t,

d(p7r{x)B^{vw ) = —7t ( jc) (g) id — id <8> 7t ( x ) ) ( i; ®

= - B ^ { tt{ x ) { v) w) - B^{ vtt{ x ){ w)).

In particular

d<pv{jc)f{v) = d(pTr{x)B:^{v^)
= - 2B^ ( ' n- ( x) { v) v)
= x-f{v)
[see (1)].
Let

L-.= ( g ^ G L { S \ V ) * ) \ g f < z f ) .

This is a closed subgroup of G L ( S \V ) * \ and hence its Lie algebra I is


algebraic in the sense of Chevalley [Ch3]. One has by [Bor, Lemma 7.3],

i = { Z e g I ( 5 " ( L ) * ) |X / c / } .

Let
L:= {g^GL( V) \ <p( g) ^L} .

The Lie algebra of this closed subgroup of GL(V) is [Ch3, ch. 2, §14,
Theorem 12].

1 = {2fG gI(F)|rf<p(Jlf) e i } .

Lemma 4
7 r(t) CIn

Proof, We know that 7r(t) <zhy (5). Let x e t. Then since the actions (6)
and (7) are identical, d ( p ( T r ( x ) ) ) f = x ^ f = X ' f ^ f for all / ^ /. Thus
dipiirix)) e I, and 7t( x ) e I. This proves that 7r(t) c l . □

Lemma 5
There exists v e T{ \ ) n U(i)i;+ such that 7rfit)v c IKi;.

Proof Let P ■= (L n By)Q (connected component of the identity of L n


By). Then P is a connected solvable K-split algebraic group. We show that P
acts on the set T ( \ ) n V. Indeed, if g ^ P and v e V(X) n F, then g • v
7.4 Coixiugacy of Split Cartan Subalgebras 631

since g G By. Now by Lemma 3(i),

g • i; e T ( \ ) <=>f { g ’ u) = 0 for all / e

But

f { g • v) = f ( g • u) = {(p{g) ~V) ( v) .

Since g “ ^ G L we see that (p(g)~^f g /, and hence that ((p(gy ^ f ) ( v ) = 0»


It follows that P acts (morphically) on the projective variety ^(A ) n K of
lines of V(X) n V. Since projective varieties are complete, and v +^ ^(A) Pi
V ^ 0, we may use the Borel fixed point theorem to conclude that there
exists Kv Ei y { \ ) r \ V such that

P • Kv = Ki;
[Bor, Thm. 15.2]. The Lie algebra of P is [Ch3, ch. 2, §14, Thm. 11]

P = I n b^.D7T;^(t).
Thus '7T;^(t)z; c pz; c Ki;. □

Recall that b+= 0 g+ and b_= n g_= o-(b^).

Proposition 6 [PK]
Let t be a subalgebra o f an invariant finitely displayed Kac-Moody Lie algebra
g. The following are equivalent:

(0 t is a hoistable subalgebra o f g.
GO t is TT^ and Trf-hoistable for all A ^ P+.
Gii) Ad(g)(t) c b + Pi wb_ for some g 0 G and w W.

Proof (i) => (ii) We know that t is finite dimensional and solvable (Proposi­
tion 1) and hence that t is and irf locally finite (Proposition 7.2.5). Let K
be th^algebraic closure of K. We use “ to denote the corr^ponding objects
over K obtained by extension of the base field from IK to K. Since ! c g is
solvable, we apply Lie’s theorem (Exercise 1.22) to conclude that each
V g L(A) ot L ( x y lies inside an ad f-stable flag. _
The following argument shows that the extension to IK is superfluous. We
work with (the case of irf being analogous). We fix x g t, and let
g= the decomposition of g into generalized eigenspaces for
ad X (Proposition 7.1.23). Let v be an element of the generalized eigenspace
L(A)^ for 1 0 7t/ x). It is clear that for a g g^.

TT),(a)v e L (A ) CO+fX
632 Coivjugacy Theorems

and hence that

'^(11(9))^’ E L ( ^ ) fJL+ÜÍ •
<y € IK

Since L(A) is irreducible, it follows that

L(A) = £ L(A)^+<,.
0>GlK

Now given y e L(A), we can find a finite-dimensional 7T;^(t)-stable subspa^


Y so that y e y. If 17/x) has eigenvalues /i, + Wi,. . . , /t + o)„ on Y,
< t ) j , e IK, then wju. + (wi + ■• • +<o„) = tr(ir;^(ji)|y) g IK so that /i e
IK. It follows that the eigenvalues of 7T;^(j:) e IK for every a: e t and hence
that Lie’s theorem already applies over K.
(ii) =» (hi) Let A e P++~ {fi ^ P \ {¡x, a ^ ) e Z+ for all i e J}, and let
V e T ( \ ) n U(t)y+ be such that c IKy (Lemma 5). Let gj e G be
such that g^v e IK®ti+ (Theorem 7.3.1). Then

-^AiAd g i(t))u + = ^ iir^ (t)g f IKy+.

We use this to show that Ad gi(t) c Indeed let

Z := { : e g_ I
a +

Since g““' n Z = (0) for all a¿ e n (since A e P++), we conclude that Z


is (ad b_^)-invariant. It follows that ad(U(g_))Z c g_ is an ideal of g, and
hence is (0) since g is radical free. Thus Z = (0). Now, if : e g satisfies a

7r^(x)u^<z 1Kl’+, then writing x = x_-\- x^, where jc_e g_ and jc+e b+, we
see that x _ ^ Z and hence that A :=jc+eb+. Hence A d g i( t) c b + , as
desired.
Similarly, since t is Tr^-hoistable, we can find that g G such that2

Ad g (t) (zb_. Write gig ^ = 8+ng_, where g + G + and n = wT e N


2 2

(see Proposition 6.3.6). Set g — g+^g^ = ng_g2* Then

Ad g (t) = A d(g+V i)t c b+,


Ad g ( t) = Ad(«g_g2)t c A d (n )b _ c wb_.

Thus Ad g (t) c b ^n wb_.


(iii) =* (i) It will suffice to show that b^.n wb. is a hoistable subalgebra of
g. Now b+n w b_c b ® Ece A+nwA_9“- Since A+n h'A_ c ^*A+ is finite (Pro­
position 5.2.3), we can write b+n wb_= b ® n, where n c ®„e^=A^.9“
finite dimensional and nilpotent. We show first that b + ti is ad-locally finite.
Let n e be an arbitrary root vector. Let {ati, be a basis of n
consisting of root vectors, say, Xi e g'»'- with y¡ e'®A+. Each ad a:, is locally
7.4 Conjugacy of Split Cartan Subalgebras 633

finite, and we have a sequence of finite-dimensional spaces


Fo == K d c Ki
:= (ad U(K a:i ))F o c
■■= (adU(lKA:2))Fi c ••• c F„
:= (adU(0<x„)F„_,).
By the PBW theorem
Vn = U(n)Ko = U(n)U(^)Fo = U(^ + n)Ko
so is ad(l^ + n)-invariant. Finally, + n is finite dimensional and solv­
able, and the argument of (i) => (ii) shows that admtits an ad(l^ + n)-stable
flag. □

Lemma 7
Let Q be a Kac-Moody Lie algebra. For w set

^ +^ © 9“*
ae A+nvFA_
Then

(0 §,, = f| ® n,„, n,, = and n,, = D§„,.


(ii) is its own normalizer in g.

Proof, (i) It is clear that ^® and that c n^. The reverse


inclusion follows from = g" for all a Finally, H^A_n
(ii) Let n be the normalizer of in g. Then c n, and hence n =
®aeA^“ with n “ = n n g“ (Proposition 2.1.2). Clearly [f),n“] c and
therefore n “ c Thus n c The reverse inclusion is obvious. □

Proposition 8
Let g be an invariant Kac-Moody Lie algebra, and l e t w ^ W . I f x ^ n ^ , then
X acts locally nilpotently in every integrable representation ( i t , F ) o f g, and
there exists exp jc e G [independent o f ( tt, V)\ such that tt exp x = exp tt( jc).

Proof. Let {jCj,. . . , be a basis of consisting of root vectors x^ e g^%


Pi The Xi act locally nilpotently on ( tt, V) (Proposition 6.1.3). Using
PBW with the above basis, we see that every v ^ V lies inside a finite­
dimensional 7r(n^)-subspace of V and hence that 7r(n^) c End(F) is strictly
triangularizable [Proposition 7.1.25 (ii)]. In particular exp7r(x) is a well-
defined automorphism of F for all a: e n^.
634 Coiijugacy Theorems

Let M be the free monoid on the symbols X and Y. In the ring we


have

exp A^expy = exp(A^ + y + [AT,y] + • • • )


= exp/i(AT, y ) .

Let be a nilpotent associative algebra of transformations of a finite­


dimensional vector space U. There exists a unique monoid morphism

End(t/>

satisfying

for any two given jc, y ^ N. Since any word of large enough length vanishes
under <p, there exists a unique extension of cp to an algebra homomorphism

(p:K^ ^ E n d ( U ) .

Under this map

exp X exp y = exp(x + y + |[ x , y] + • • • )

(both right-hand sides above are finite sums). Replacing x by x + y and y by


- y gives

exp(x + y) = exp(x + j [ x + y, - y ] + • • • )exp(y)

= exp(x - j [ x , y ] + ■■■ )exp(y)


= exp(jc + / ( x ,y ) ) e x p ( y ) ,

where /(x , y) is a linear combination of commutators in x and y of length at


least 2.
We now return to n^^ and reindex the x, if necessary so that ht > ht )32
> • • • > ht We claim that if x = E flia,x, e n„, then

(8) expTr(x) = exp'7r(bjX,) exp ir ( 62^2) ■■■ ®xp Tr(hjvXyv),

where {¿j,. . . , bj^} c IK depends only on {a,,. . . , a/^} and n^, but not on
( t7, V). It will suffice to establish (8) on every finite-dimensional 7r(n^)-stable
subspace U of V. Since ir(n^)lu is strictly triangularizable it generates a
7.4 Coiyugacy of Split Cartan Subalgebras 635

nilpotent (associative) subalgebra N of EndQJ). We use the above discussion


and induction on r, where x = 0. If r = 1, then exp irix) =
exp iriaiX^X as desired. Assume the result holds for r — 1 > 1 and that
jc = jc' + a^x^, where x' •= # 0. Then on U,

exp 7t( jc) = exp 7t( x ' + a^x^)


= expir(ji' + f { x , a , x,))eTop'jr{a,x,).

Now f i x , a,.xf) e 0 [r/K jr, because all roots involved in commutators of x


and a^jc^ are of height greater than ht /3^. Thus exp 7r(jc' + / ( jc , a^x^)) has
the desired form, and this completes the induction step.
Let exp JC — exp(fciJCi) • • • exp(feyy^Xf,) e G. By (8) this element satisfies

TTexp JC = exp 'Tr(jc)

for all integrable representations ( tt, F ) of g. □

Let I be a Lie algebra over K. A subalgebra i of I is called a Cartan


subalgebra of I if

CSl: f acts ad locally nilpotently on itself,


CS2: f is its own normalizer in I.

If in addition

SCS: f is hoistable in I.
We then say that f is a split Cartan subalgebra of 1.

Example If g = (g + ,^ ,!2 + ? ^ )isa Lie algebra with triangular decomposi­


tion, then ij is a split Cartan subalgebra of g.
Remark 1 Let f be a Cartan subalgebra of a Lie algebra I. If f if finite
dimensional then f is nilpotent by CSl and Engel’s theorem. Most of the
time we shall restrict ourselves to this case in what follows. The above
definition comes from a general theory of Cartan subalgebras and their
relations to regular elements. For more details see [BPl].
Remark 2 Cartan subalgebras need not exist for a given Lie algebra I.

Theorem 9 (Chevalley, Peterson-Kac [PK])


Let g be a finitely displayed invariant Kac-Moody Lie algebra, and let G be its
derived Lie group. Then Ad G acts transitively on the set o f split Cartan
subalgebras of g.
636 Coiyugacy Theorems

Proof. Let f be a split Cartan subalgebra of g. By Proposition 6 we may


assume that f c e n^. Thus f is a Cartan subalgebra of ^ in the
classical sense, and by a classical conjugacy result (Exercise 7.6) there exists
jc e such that expad x ( t ) = ij. Finally, by the last proposition we have
exp ad jc = Ad g for some g ^ G. □

Proposition 10
Let I be a finitely generated ideal o f a finitely displayed invariant Kac-Moody
algebra g. For a subalgebra a o f I the following are equivalent :

(i) a is simultaneously diá-diagonalizable on I.


(ii) There exists g ^ G such that Ad g(a) c n I.

In either case a is abelian and finite dimensional.

Proof (i) => (ii) If a d a c g l ( I ) consists of simultaneously diagonalizable


transformations then it is clear that a is hoistable in I and hence also in g
(Proposition 1). By Proposition 6 we may assume that a c for some
w E:W. We now follow [Bo4]. Let c be the centralizer of a in and let f
be a Cartan subalgebra of c. Since a lies inside the center 3 of c, we have
a c 5 c f. Let n be the normalizer of f in and let b be an ad a-stable
space so that n = Í + b. We have
[a,b] c [ a , n ] n b c [ f , n ] n b c f n b = (0).
It follows that b c c and hence that n is the normalizer of f in c. Thus ! = n
so that f is a Cartan subalgebra of Up to conjugation we may assume
that a (zi) and hence that a c n I.
(ii) => (i) Clear. □

Corollary
If a c Dg is ad-diagonalizable on Dg, then a can be conjugated by Ad(G)
into i).

Proposition 11 (Jordan-Chevalley decomposition)


Let q be a finitely displayed invariant Kac-Moody algebra^ and let x ^ ^ be an
ad-locally finite element. Then there exist unique elements x^ and m g such
that

(i) X=X5+X;^,
(ii) [X5, x^] = 0,
(iii) X5 {resp. Xff) acts semisimply (resp. locally nilpotently) on every
integrable representation o f g.
7.4 Conjugacy of Split Cartan Subalgebras 637

Proof. Assume first that IK is algebraically closed. Since x is ad-locally finite,


t := IKx is hoistable in g, and hence by Proposition 6, x can be conjugated
into for some w ^ W. This allows us to assume at the outset that x e
Let jc = JC5 + be the decomposition of x guaranteed by Proposition
7.2.6. Since [jc, c Proposition 7.2.6 shows that [x^, c and
c and hence by Lemma 7(h) that x^, x^^ Write Xf^ = h -\-
n, where h n ^ n^. Then for any 7 e A and any a e we have

(ad a = (ad(/t + «)) a = {y,h)^a + A: = 1 ,2 ,...,

where ¿ a: ^ ^ht5>htyS^- The local nilpotence of adxj^ then shows that


<7, h) = 0 and, this being true for all 7 e A, that h lies in the center a of g.
We write

X = (x^ + /1) + n

and observe that

[(xs + h), n] = 0,
n ^ acts locally nilpotently on every integrable representation (Pro­
position 8),
JC5 + /z acts semisimply on every integrable representation (for this we let
a ■= Kx^ + Kh and use Proposition 10).

This proves the existence part of Proposition 11 when IK is algebraically


closed. As for uniqueness, suppose that

X=X^+Xjs¡ = x'^ + x',^


''N

are two decompositions of x satisfying the properties (i)-(iii). Then x'^ - x^ =


Xn - x'j^. Also [x's, X5] = 0, since e g^ = g^^ p, c (see Proposi­
tion 7.2.6). Similarly [x^, = 0. Thus (jCy^ - x^) acts locally nilpotently,
and x's - Xs acts semisimply on every integrable representation of g. This
forces x's = Xs and x'^^ = x¡^. □

To prove the nonalgebraically closed case, we need the following:

Lemma 12
Assume that {Ay}y ^ P+ span 1^*. For each i let be the kernel o f the action
of g on L(Ay). Then X — fl = (0). {See Exercise l.A for the existence of
{Ay}.)
638 Coqjugacy Theorems

Proof. Let r be the skeletal graph of the Cartan matrix ^ of g, and let
be its connected components. These define submatrices of
A and subroot systems consisting of those roots whose supports lie in
j = 1 , r. Let be the subalgebra of g spanned by the root spaces
g“, a e A^^\ and ij. Then by Proposition 4.3.9, g = Ey=ig^-'^ By Proposition
4.3.10, any ideal of g^-'^ is either inside i) or contains Dg^^\ In particular
K n g^^^ (zi) oi K Dg^^\ The latter is impossible: If e A-' n II, then
there must be an / e / for which <A^, ^ 0. Then g““^ does not kill the
highest weight vector of L(A/^) (use § 12-theory), and so g““^ c X. Thus
K n g^^^ c 't) for all j. Since K is an 1^-module, it has a root space decomposi­
tion, and so K = IXK n g^^^). Thus K (zi). finally, since span !^*, we
see that K ni ) = (0). □

We now finish the proof of PropositionJ.1. Consider K arbitrary. Let IKbe


an algebraic closure of IK, and let g — IK<8>|,^g, x •= 1 ® x. For each r e
Gal(IK/[K) we have a IK-linear automorphism of g defined by (a (S>y Y =
a'^ <8>y, and g = 1 <8> g is the set of points of g fixed by all these automor­
phisms.
Let P+= {A G P |< A ,a /> ^ Z > q for all i e J} be as usual. For each
A e ]^* we have the irreducible representation of g on L(A). It is clear
that L(A):= IK0|,^L(A) is_the irreducible module of g under — 1_0
whose highest weight is A == 1 0 A e 1K0 1^* = ^*. Clearly r e Gal(IK/lK)
defines a IK-linear isomorphism (as a vector space) of L(A) : a u a"’ 0 y.
It is an elementary exercise to show that for y g g, D e L(A), (y • í;)^ =
y'" • u'".
Now consider x. By the above x = for some g g that
satisfy parts (i)-(iii) of Proposition 11. In particular x^ acts diagonally, and
Xyv acts locally nilpotently on L(A). If x^ • d = fiv, then x j • = (x^ • vY
= /I'v"', and xj acts diagonally on L(A); similarly Xn acts locally nilpotently.
Also [xj, xj^] = [x^, Xj^Y = 0. Thus x = x"* = xj -I- x^, and for each A g
^ /x ) = is a Jordan-Chevalley decomposition. But this is
unique, and hence ^ / x ¡ ) = tt/ x ]^) = Since {A | A g P_^j
span ]^*, we conclude from Lemma 12 that x j = x^, x]^ = Xj^ for all r g
Gal(lK/[K) and hence that x^, Xj^ g g.
If M is any integrable representation of g, then M •= .M is inte-
grable for g, and so x^ and x^ act semisimply and locally nilpotently on M,
as well as on M (see Proposition 7.1.16(i)). This concludes the existence part
of the proposition. Uniqueness is as before. □

The decomposition of the locally finite element x g g in Proposition 11 is


the Jordan-Chevalley decomposition of x. The following proposition is easy
Exercises 639

to prove (see also Proposition 7.2.6):

Proposition 13
Let Q be a finitely displayed invariant Kac-Moody algebra^ and let x be an
2id-locally finite element o f g. Let x=x^-\ -Xj ^ be its Jordan-Cheualley decom­
position in g. Then

(i) for any subspace K o f any integrable ^-module M ,


x - K c z K ^ X s ' K c i K and Xj^ • K <z K;
(ii) g"" = g""^ n g^^;
(iii) if IK is any extension of (K, then 1 <8>;c = 1 0 X5 + 1 <8) is the
Jordan-Chevalley decomposition o f 1 x in K

EXERCISES

7.1 Let t be a Lie algebra over IK and ( tt, K) a representation of t. Let L


be the set of functions from t into IK. For each A e L define

= (y e K 1i; g V^^yix) for all x g t},


F ^ (t) = {y G K Iy G for all X G t}.

(a) Show that the sums and are direct. (Use


Proposition 7.1.21).
(b) Let ( tt^, FO, i = 1,2,3 be representations of t. Let B ®
-> be a bilinear map satisfying TT^(x)B(vi, ^ 2^ =
B(v^(xXvi),V2) + B(Vi,TT2(x)v2), for all X G t. Show that the
induced linear map ^ :V^ 0 ^ is a t-module homomor­
phism. Use Proposition 7.1.22 to conclude that

for all Aj, À2 e L.

(c) Show that if a:, y g t are such that (ad 'n-(j:))''(ir(y)) = 0 for some
n G N, then Tr(y) stabilizes V^ix) for all k ^ K. [Use (b) applied
to End(F) ® V.]
7.2 Let t be a finite-dimensional solvable Lie algebra (Exercise 1.22). Let
V (0) be a finite-dimensional t-module. Assume that for all x in t
the eigenvalues of ir(x) in V all belong to K. In this Exercise we
establish Lie’s theorem. The first step is to prove (by induction on
dim t) that

(*) th ere exists A g t* such that F ' ' ( t ) ¥= ( 0 ) .


640 Coiijugacy Theorems

(a) Let Í} be an ideal of t of codimension 1 (Exercise 1.22). By


induction ^ (0) for some A e 1^*. Show that t stabilizes
VKijX [Show that A[a:, y] = 0 for all x e t, y g )^. For this, fix
VG \ {0}, and choose N maximal so that u, tt{x ) •
V , . . . , Tr(x)^ • V are linearly independent. Let W¿ =
0 J “¿IK7r(jc)^(i;). Show that stabilizes and that, if z g 1^, then
trp^^(7r(z)) = NX(z). Apply this to z = [x, y].]
(b) Write t = ]^ © IKjc. Find an eigenvector of v i x ) on Extend
A to t* and establish.(*)
(c) (Lie’s theorem) Show that if IK is algebraically closed every irre­
ducible representation of t is one dimensional. Using a composi­
tion series, conclude that V admits a 7r(t)-stable flag.
(d) Show that there exists a sequence of ideals 0 = t o C t i C ••• c
= t such that dim t • = i for all 1 < i < n.
(e) Show that for a finite-dimensional Lie algebra g to be solvable, it
is necessary and sufficient that be nilpotent.
7.3 Let t be a finite-dimensional nilpotent Lie algebra and ( tt, F ) a
finite-dimensional t-module. We maintain all the notation of Exercise
7.1.
(a) For all A G L show that F /t) is a submodule of F.
(b) Show that if 7t( jc) g End(F) is triangularizable for all jc g t, then
F = © ;^^^F/t) (Proposition 7.1.23).
( c ) If F;^(t) # (0), show that A g t* and that moreover Alot = 0.
[Assume that F = and show that A(x) = (dim V)~^ tiy 7t( jc).]
(d) Let f :V ^ M be a suijective t-module homomorphism. Show
that /( F /t) ) = M /t) for all A g L.
7.4 Let g be a finite-dimensional Lie algebra and t nilpotent subalgebra of
g. View g as a t module via the adjoint representation. We maintain
the notation of Exercises 7.1 and 7.3.
(a) For all A, /x g L show that [g;^(t), g^(t)] c g;^+^(t). Conclude that
go(t) is a subalgebra of g that stabilizes g /t) .
(b) Show that A/g(go(t)) = go(t), that is, that go(t) is its own normal-
izer.
7.5 Let g be a finite-dimensional Lie algebra. A subalgebra t of g is called
a Cartan subalgebra if

CSl: t is nilpotent,
CS2: A/g(t) = t (t is its own normalizer).

We maintain all of the notation of Exercises 7.1, 7.3 and 7.4.


Exercises 641

(a) Let t be a nilpotent subalgebra of g. Show that the following are


equivalent:
(i) A/g(t) = t (and hence t is a Cartan subalgebra).
(ii)
[If ^ consider the t-module ^ind use Engel’s
theorem.] Conclude that Cartan subalgebras are maximal nilpotent
subalgebras.
(b) Let f : g be a surjective Lie algebra homomorphism and t a
Cartan subalgebra of g. Show that /( t ) is a Cartan subalgebra of
f). [Use (a) and Exercise 7.3(d).]
(c) Let 5 be the centre of g and t a subspace of g. Show that the
following are equivalent:
(i) t is a Cartan subalgebra of g.
(ii) g e t and t/3 is a Cartan subalgebra of g/a.
[For (i) => (ii) use (b). The converse is straightforward.]
7.6 Let g be a finite-dimensional solvable Lie algebra. Recall the ideal
-^ (g ) (Exercise 1.23). The purpose of this exercise is to prove the
following central result of classical Lie theory.

Theorem (conjugacy of Cartan subalgebras: solvable case)


If t and t' are two Cartan subalgebras o f g, then there exists x e ^ ( g )
such that t' = exp(ad x)t.

We reason by induction on dim,,^ g.


(a) Let n # (0) be a minimal abelian ideal of g. Let ~ : g g /n be
the canonical map. Use Exercises 1.22 and 1.23 and the induction
hypothesis to show that
V = exp(ad jc)î

for some x e T^~(g). Conclude that, without loss of generality, we


may assume that Î = Î' and hence that
t + n = t ' -h n : = i .

(b) Show that t and t' are Cartan subalgebras of Î and conclude that
we may assume, without loss of generality, that
t + n = t' + n = g

There are two cases:

Case 1. [g, n] = (0),


Case 2. [g,n] = n.
642 Coiijugacy Theorems

In Case 1 show that t = t'. [Use Exercise 7.5(c).] Assume therefore


that [g,n] = n.
(c) Show that n c
(d) Show that n is an irreducible t-module under the adjoint action.
Use this to reduce to the case when t n n = (0).
(e) Assume henceforth that t © n = g = t ' 0 n . Given a e t, let ¿z' e
n be the unique element satisfying a - a' e t'. Show that for
a, h e t,
[ a , b j = [a,b' ] - [b,a' ].

Conclude that there exists jc e n such that a' = [jc, a] for all
a ^ t. [Make M — n 0 K into a t-module via a • (x, k) = (a • x
— ka',0). Consider the canonical module homomorphism M ^ K,
and use Exercise 7.3(d).] Show that exp(ad jc)t = t'.
7.7 Let g be a minimally realized invariant Kac-Moody algebra. In this
exercise we provide a description for the group Aut(g) of automor­
phisms of g.
(a) Let f be a supplement for in so that g = f 0 0
a\{0}9“ and g/D g = f. Show that each / e Hom^Cg/Dg, Zg)
determines an automorphism <pf of g that pointwise fixes Dg in
the following way: Pull back / to a mapping / : g ^ Zg and set
<Pf~ 1 + f. We use this to identify Hom|^(g/Dg, Zg) as a sub­
group of Aut(g).
(b) Prove that the kernel of the natural mapping Aut(g) Aut(Dg) is
Hom^j(g/Dg,Zg).
(c) For each 8 e A ut(r) show that there is a unique element 8 e
Aut(Dg) such that 8f^ = i e J. Show that 8 lifts
to an automorphism of g with the same properties. [Let k e f .
There is an element k' e ^ satisfying <a,, k> = k'> for all
i e J.] This leads to a subgroup D c Aut(g) consisting of auto­
morphisms that (i) permute the root vectors e,, i e J, and the root
vectors /„ i e J, (ii) stabilize f|, and (iii) satisfy

0 ^ HomK(g/Dg) D ^ A u t(r) ^ 0.

(d) Using the conjugacy theorems for Cartan subalgebras and for root
bases show that every element r e Aut(g) can be expressed as a
product

( - a - Y x • 5 • A d (g),
where g ^ G, 8 ^ D, x ^ Hom((2, K®), e e (0,1} [Here a is the
anti-involution that is part of the definition of (g, ^ } . ] In particu­
lar Aut G = < - o->Hom((2, IK®)£> Ad(G).
Exercises 643

7.8 In this exercise we establish a result of Kac and Peterson [KP2] on


bounded subgroups (even bounded semigroups) of the group G associ­
ated with a complex Kac-Moody Lie algebra (g, ^ ) . The original
results of this type (for a different class of groups) are due to F. Bruhat
and J. Tits [Groupes reductifs sur un corp local, IHES, no. 41, 1972].
The proof of Kac and Peterson is unusual because of its use of
techniques (due to Kempf) involving the theory of convex functions. In
fact most of the results we need from this theory are easy to prove as
they are included here as part of the exercises. One may consult A. W.
Roberts and D. E. Varberg, Convex Functions (New York: Academic
Press, 1973).
Fix once and for all A = a Cartan matrix where J is
finite, and (J) = /. Let R = (l^j^, II, II ^) be a minimal realization of A
over U, Let C ^,C , 3£^,3£, be the fundamental chambers and Tits
cones relative to 11^ and II in and respectively.
Form the Kac-Moody algebra = g(^4, R) over IR, and let g =
C <8>p5g be its complexification with diagonal subalgebra = C
We are interested in the associated group G of g.
Let be the full subcategory of (the integrable weight modules
of g) whose objects are those integrable g-weight modules V satisfying

V is U(g+)-locally finite [i.e., dim(U(g+) • u) < oo for every v VI


P(V) c Int(3E).

Let y be a set. A set ^ of subsets of y is a system of bounded subsets


for y if
M, M ' c ^ => M U M ' e
M ' c M, M G ^ M' e

If y is a group, then the system ^ is compatible with the group law


on y iff

(a) Prove that if (G, B, N, S) is Si Tits system with associated group W


and we define ^ to be all subsets of G that lie in finitely many
R-double cosets, then ^ is a system of bounded subsets. In the
sequel this is the system that we will always have in mind.
(b) Let [/ be a nonempty convex subset of W . A function f : U ^ U i s
convex on U if for all x, y ^ U and for all t g [ 0 , 1 ] ,

fitx + ( l - t ) y ) < t f ( x ) + ( l - t ) f ( y ) .
We say that / is strictly convex if the inequality is strict whenever
X ¥=y and t G (0,1).
644 Coi\jugacy Theorems

(c) Show that jc is strictly convex on R and z \e^\ is convex on


C.
(d) Let t/ c IR'* be a nonempty convex set, and suppose that {/^} is a
sequence of convex functions f ^ : U R such that L/^. converges
pointwise to a function on U, Show that / is convex and / is
strictly convex if at least one is strictly convex.
(e) Let / be a convex function defined on an open convex set U c R",
and suppose that / is differentiable at Xq e U. Let a g R" \ {0},
and let be the partial derivative in the direction a. Show that

/ ( ^ 0 + 5« ) -f(X o ) >s(dJ)(Xo)

for all 5 > 0 for which Xq sa ^ U and furthermore the inequal­


ity is strict if / is strictly convex.
(f) Show that there exists a function

/:Int(3E) ^ R^
satisfying

/ is H^-invariant,
/ is strictly convex.

Set A^o = maxiCardCW^^) IK c J, K is a subset of finite type}. Fix a


basis c Int(C ^) of Show that for all A e Int(3£) and
for each k.
,-r(mi+ ••• +m/)

where r (which depends on A) is some positive constant. Show


that / : A gives us the desired function.]
In what follows, / refers to the function / that we have just
constructed.
(g) By extending the definition of / in the obvious way to a function /
on Int(3£) 0 show that / is the restriction of a
complex analytic function and that hence it is real analytic. (This
requires some minor analysis about uniform convergence of convex
functions.)
(h) Show that for any IF-invariant subset D of Int(3£) that lies in the
union of finitely many (2+-fans, ¡jl i Q ^ ,

/ ( |( A + j/))
A e D , i/e D n C

converges.
Exercises 645

(i) For each subset T c ]^* let <T> denote its convex hull. Show that
if T is finite, then / assumes a minimum value on ( T ) at some
unique point of <T>.
(j) Let D be a PF-invariant subset of Int(3E) lying in the union of
finitely many <2+-fans. Show that for each M e there are, up
to H^-equivalence, only finitely many finite subsets T oi D for
which > M.
(k) Let V e Obj(5^). For v = eP(K)^ V, supp(i;) — {A I
0} c Int(3£). For any nonzero finite-dimensional vector subspace U
of K, define

supp(i/) := U supp(i;) c Int(3£).


v^U

Show that supp(i/) is finite thus enabling us to define

r = 7 t/:G Int(3E),
y ( g ) = M<supp(g • i/)>,

where • is the integrated action to G of g on F. Let B, G+, N, H,


be the standard subgroups of G (Section 6.2).
Prove that for all n e N and w = nH e W,

y(ng) = wy ( g) for all g e G.

(1) Let U # (0) be a finite-dimensional subspace of K e OhjiS^), and


let y = yu be defined as in (k). Let g ^ G, and suppose that
y(g) e C.
Show that for all a e f)g such that y(g) + a e. <supp(g • i/)> we
have (d^fXyig)) > 0. Use the IL-invariance of / to show that for
all I e J we have /(y (g )) > 0. Given b e B, conclude that any
element A of <supp{bg • U)) can be written as

^ =y(g) +a +P

where a e is such that y(g) -f- a <supp(g • t/)> and /3 e

(m) Prove that for g e G with y(g) C and b ^ B, /(y(bg)) >


fiyigJ).
(n) Let U # (0) be a finite-dimensional vector subspace of F e
O b j(^ ). Show that there exists a finite subset {A,} c C so that
m ( g ) - G ) c uA,ie^.
(o) Let y = yj; be as in (k). Prove that the function g /(y (g )) on
G assumes a maximum at some point gg e G for which yCgg) e C
646 Coi\jugacy Theorems

(p) Let V, U, y, gQ, be as in (o): Show that for g ^ G,

y{g) = y(go) ^g^PKgo^

where K = {/ e J 1(ji g^X « /> = 0} and is the parabolic sub­


group BWj^B of G. (Consider the element ggo^, and use the
Bruhat decomposition.)
(q) (Kac-Peterson) Let M be a subsemigroup of G. Prove the follow­
ing are equivalent:

Bl: M is contained in a parabolic subgroup where K c J is


of finite type. (Parabolic subgroups of this type are some­
times called Iwahori subgroups.)
B2: M is bounded.
B3: M acts locally finitely on every V e Obj(5^).
B4: M leaves invariant a nonzero finite-dimensional subspace
of some V e Obj(<9^).
A*l*-An Extended Example

APPENDIX TO CHAPTER 1

Let K[t,t~^] denote the ring of Laurent polynomials with coefficients in K:

K[t , t — I X) ^ ^
I j = tn

where addition and multiplication are defined in the usual way

{ L a j t ’) + {j:bjt^) = E i a j + bj )t \

+ {T,bjt^) = e ( e « A -;)^ * -
k ^ j ’

Let L§l2(IK) = §I2(1K[L i ” M), the set of 2 X 2 matrices with coefficients in


K[t,t~^] and trace 0. This is a Lie algebra, called the loop algebra of ^I2(IK),
under the usual bracket and contains ^l2(IK) as the subalgebra of matrices all
of whose entries are constant polynomials. We can think of g as an algebra
over the ring K[i, i “ ^], in which case it is of rank 3 as a module. However,
over K it is infinite dimensional and reveals an entirely different structure.
Consider the standard basis of ^l2(K):

‘ =(2 i). A /=(; II


Then the matrices t^e =
Z, form a basis for L^l2(IK) over K. An alternate description of L^l2(lK) is as
lK[i, 0ij^^l2(W, the isomorphism being obvious: t^x ® x for all
X G ^ l2 (tK ).

647
648 Extended Example

There are infinitely many integer gradings that can be placed on L§l2(IK)
(a fact that becomes obvious later on). For now, however, we settle on one,
called the principal grading, by defining

deg r*e = 2k + \,
d e g r* /= 2k - 1,
deg t'^h = 2k. A: e Z.

The grading is schematically shown below. Notice that dim = 1


or 2 according to whether k is even or odd.

degree

3 te
2 th
1 e tf
0 h
-1 t f
-2 t-^h
-3

Treating L§l2(IK) as a module for §I2(IK) (under the adjoint action) we can
see from this schematic picture how decomposes into a direct sum
of infinitely many §l2(IK)-submodules each isomorphic to itself (as an
§l2(K)-module).
Let K be the Killing form on §I2(IK). (See Example 1.20). Define a bilinear
form ( I •) on L§l2(IK) by setting

(A.1)
(i*ali'b) = QK{a,b) for all A:, / e Z , a , b e [ e , h , f )

(5 = Kronecker delta). The invariance of this form is an immediate conse­


quence of invariance of the Killing form [but (-| • ) is not the Killing form
of §l2(IK[r, r " '])]• Evidently Lêl2(IK)* J. L il2 (IK )' if A: -I-
l Q. On the other hand, L§l2(IK)* and L§l2(IK)“* are nondegenerately
paired, namely x e L§l2(IK)** and (x|L§l2(IK)^*) = 0 => x = 0. It follows
that (-| • ) is nondegenerate, symmetric, and invariant.
Since §I2(1K) is perfect (i.e., is equal to its own derived algebra) so is
L§l2(K). Thus by the theory of Section 1.9, Lil2(K) has a universal covering
Appendix to Chapter 1 649

algebra. In fact we can explicitly construct a covering algebra, and later we


will show that this algebra is the universal cover.
We let

(A.2) §I2 (K) := e IKC

be a one-dimensional vector space extension of § l2(IK[i, t ']) and define


multiplication on it by bilinear extension of

1
(A.3) [ra ,t'Z )] = t'‘*‘[a,b] + k8^+, Q-K{a,b)<t.,

[t*a,e] = [e,r*a] = [«:,c] = 0 for all k ,l ^ Z,a,b ^ {e,h ,f).

The bilinear mapping

e; L§l2(IK) XL§Í2(IK)

defined by

(A.4) e { r a , t ‘b) = k8k+i^Q-K{a,b)

is a 2-cocycle in the sense of Ch 3.8(3), and hence § I2(IK) with [•, • ] is a Lie
algebra and a eentral extension of L § l2(IK). Since [th,t~^h] = \K{h,h)<t =
2c, c e D §l2(IK), and it is very easy to see that §I2(IK) is perfect. Note that
although L612(lKHsa subspace of §I2(IK), obviously it is not a subalgebra of
§ I2(IK). In fact, §I2(IK) has no subalgebra m of codimension 1 isomorphic to
L § l2(IK). If there were, then since the centre Z (L g l2(IK)) = (0) and Kc is
clearly central, c i m, and we would have § I2(IK)= m -l- Kc. But then
D §l2(IK) = m ^ §I2(IK), which we have seen is not true. The fact that § I2(IK)
is a universal covering algebra is shown below.
The principal grading lifts to §I2(1K) by inheriting the grading on LSljOK)
(as a subspace) and by defining deg c = 0. Thus

_____^ if yfc e 2Z \{0 },


if A: e Z \ 2Z,
(IK/i -I- IKc if A: = 0.
^^0 Extended Example

a p p e n d ix t o c h a p t e r 2

Using the principal grading set

^>0

k<0
5 := Kh + K0,-
SO

^i2(K) = §i2(iK)_0 § e

We consider this as a potential triangular decomposition of ^ l2(IK).


Define an anti-involution a on §I2(IK) as follows: §I2((K) has a natural
anti-involution, transposition of matrices. Extend it to an anti-involution a of
L § l2(IK) by t^a t~^cr(aXk G Z, Observe that this is not
transposition on § l2(lK[i, Lift a to §I2(IK) by setting cr(c) = C. Then

or([i*a, t'b]) = o-(r*+'[afe] + k8^+, o\K(a,b)c)


= t~^‘‘'^‘\<Tb,cra] + kd^.+, QjK{a,b)<t
= t~^’^^‘\ab,(ra'\ — 1 8 Q^K^ab, (To)(t
= [a-(i'i)),<r(i*a)],

where we have used the easily established fact that transposition preserves
the Killing form of ^ l2(K).
Although all this bodes well for a triangular decomposition, there is a
problem with TD(1). Indeed we have

= 0,
i 2i*a I e,
= < 0 according as a = I h,
\f-

Thus, defining a ; : 5 ^ IK by « j : e 0 and ^ 2, we see that § I2(IK)^.


decomposes into three eigenspaces relative to

§I2(K)+ = g l2 (K )7 ' ® §l2(»<)+ ® ^i2(i<) +


Appendix to Chapter 2 651

and

§i2(»<)+=
fc> 0

The problem is that ad § does not create enough different eigenstates. We


remedy this by enlarging
We let d: gljilK) ^ § l2(IK) be the linear map defined by d\.— = scalar
^I2(IK)
multiplication by k. Since d e Der(§l2(IK)) we can form the Lie algebra
§I2(IK) = Kd © i l 2(IK) using the construction of Section 1.5:

[d,d] = 0

[d, x] = d ( x ) for all x e §I2(IK),

in which § I2(IK) appears as an ideal of codimension 1. Set

(A.4) S==IK d® §,

and extend <r to an anti-involution a on §I2(IK) by defining aid) = d. Then

(A.5) §I2(IK) = §I2(IK) _ © ]^ © §I2(IK) + = §I2(IK) _ © ft © §I2

(A.6) § Q k )_ , <r|5 = id,

where we define § I2(IK)^= § l2v-v+-

Define «o> ^ ft* by

ao(d) = 1, a^iit) = 0, ao(^) = “ 2,

" i(^ ) = 2,
and let

8 = Q!g -|- C f].

Then

= {2k -h l)t*e]
—— —— — Qri+Ac5 - ——o;n+(^+ 1)5
[e,i*e] = 0 c §I 2(IK) = §I 2(IK)
[/i, i*e] =
652 Extended Example

Similarly t'^f ^ = gÍ2(IK)“'>'"^*“ *^, i*/j e §I2(IK)**. Thus

(A.7)
kd -a^+kd
§Í2(í<) += © §Í2(K) © §l2(K )''"e © §I2(K)
^>0 A:>0 k>0

and all of these root spaces are one dimensional. With

(^•8) Q +'= {^«0 + ma^\k, m > 0} \ {0},


we have

§Í2([K)_,= © §Í2(IK)
«^í2+

(note that §Í2(D<)^ao+mai = (0) unless \k — m\ < 1).


Together (A.5), (A.6), (A.7), (A.8) show that § 12( 11^) has a regular triangu­
lar decomposition. The Lie algebra ^ Í2(IK) is called the affine Lie algebra of
type The root lattice is g = Za^ + Zctj, and the set A of roots is given
as

(A.9) A = [ka^ + ma^\k,m e Z,|fe — m\ < l}.

^ : = (§I2(K) Q+, (t ) is a regular triangular decomposition. We set

n := [ao,a{\

and

n -= = { a o ^ < } .

We introduce further notation that will be convenient later on.


Set Cj := e, « 1^:= h, / j ■=f , and define i Í 2(IKy^^ = Ke^ + Kai'^+ K/j.
Define

(A.10) eo '=tf, ao - = t - h , f o - = t ^e,


(0)
and observe that := K cq + IK«(^+ IK/q = §I2(IK), for example
[^o»/ol = W> t~^e] = - h + jKif, eH = <t - h = a^. Observe that

e, e § l 2 ( K ) ', e§l2(0<) d-, h ^ % , i = 0 ,l-

Let us see that the subalgebra <eo, gj) generated by Cq and is precisely
Appendix to Chapter 2 653

Evidently c However,

(ad Cj ad eo)*ad Cj(eo) =


[i*^, e j = 2i*e,

[i*;i,eo] = -2i*=^*/, A :>0,

from which the reverse inclusion follows. Similarly

§ l 2W - = </o,/l>-

Thus

^ I 2( I K ) = < /o ,/i> e i| 0 < eo,C i> .

Next we consider as a module for (defined above) under


the adjoint representation. This action is integrable (see Section 2.4). In fact

(ad c) annihilates d,0,t^e

(ad e)^ annihilates t^h

(ad e)^ annihilates t ^ f

for all k e Z, so (ad eY = 0. In the same way (ad f Y = 0. Thus by Proposi­


tion 2.4.4, § 12( 11^) is completely reducible into irreducible modules for
§I2(IK)^^\ Two of these are the one-dimensional modules and Wd —
The others are the three-dimensional submodules that we already saw in
Lél2ÍKX
The situation here is somewhat unusual because ad e and ad / are actually
nilpotent, not simply locally nilpotent. This reflects itself in the bounded
dimensions of the irreducible ^ 12-modules that we just obtained. More
typical of integrable representations is the discussion below of the basic
representation of §Í2( 0^) ¡n which e and / no longer act as nilpotent
operators.
For the same reasons as above, (ad e^Y = 0 = (ad / qY, s o the representa­
tion of on §12( 1^ ) determined by the adjoint representation is
integrable and hence completely reducible. Using the above description of
the roots

—cíQ + kS ~kS aQ+kd


^l2(K) = © (§I2(K)
k^Z ^
0 ^I2(1K) 0 ^12'
)■
654 Extended Example

For ^ 0 we are looking at irreducible § l2(IKy®^-submodules each isomor­


phic to the adjoint representation. For A: = 0,
- - ocq
§Í2(IK) 0 §I2(1K) + §Í2

= lK ( d -I----- 0 IK c 0 ^ I2 (1K )

which consists of two trivial representations and the adjoint representation


again.
The two ways that we have just described for decomposing § 12(11^) into
^ l2(IK) submodules is made schematically evident by looking at the root
system A.

+ 5 28 an + 25
«1 8 Uq + 8
- “0 0 «0
-a 0 ^ -8 -« 1
-Uq — 25 -2 5 — —8

The vertical (resp. horizontal) sections correspond to ^I2(1K)^^^ [resp. §I2(IK)^®^]


submodules. Since each root has several descriptions (« q? ^ linearly
dependent) the naming of the roots as shown here is best understood by
starting at 0 (corresponding to §) and successively applying operators ad Cq
and ad Cl to move right up and ad / q and ad / j and to move left and down).
Let us return to our integrable representation of §I2(IK)^^^ on §Í2(1K).
According to Proposition 2.4.4, there is a compatible action Ad [in the sense
of condition Ch 2.4(8)] of SL2(K) on ^I2(1K)‘ for all nilpotent X e ^ Í 2(KY^\
exp ad¡^)A^ = Ad(exp X ) .

Since ad^----- X is a nilpotent derivation, Ad(exp AT) is an elementary auto-


éÍ2(IK)
morphism (Proposition 1.5.3) of ^I2(1K). Since SL2QÍ) is generated by
{exp X \ X nilpotent} (see Proposition 2.4.8), we see that we have a subgroup
5L2(IK)^^^ - PSL2ÍK) in Autg(él2(lK)) determined by g l2(IKy^\ In the same
(0)
way we obtain SL2(KY^^ c A u t/^ l2(IK)) from ^ I2(IK) Together these
generate most of A ut(^l2(IK)) (see below).

Remark 1 Since § l2(ll^) decomposes into a sum of only odd dimensional


irreducible modules for §Í2(IK)^^^ under the adjoint representation, the
corresponding group SL2(KY^^ of autmorphisms is isomorphic to PSL2ÍK).
Appendix to Chapter 2 655

Let us turn to the internal ideal structure of §I2(IK). Taking account of


Propositions 2.7.5 and 2.7.2, we have

rad(§l2(lK)) =

for any triangular decomposition T of ^I2(1K) (in particular for the one we
have constructed) and

n(§l2(IK)) n $ = Z(§I;

These become fairly trivial-looking statements in this case, since we can


prove that

Z{§Í2(K)) = Ki,

the ideal lattice of §Í2( ^ ) is

__ §I2(1K)
I __
§ I2 (K ) = d (§Í2(IK))

{0}.

Indeed let a be a nonzero ideal of ^ I2(IK). Then a is an l^-module and hence


decomposes into root spaces: a = a “ ^ (0) for some
a # 0, then t^e, or t ^ f is in a for some k. Then one easily finds that
a D §I2(IK). If a" = 0 for all a # 0, then a c $ and from [t^e, a] =
0 = [t^f, a] V/& obtain a c Kc. Thus

^7^= (0),

r a d ( i r ; w ) = (0),

n(§l2(lK)) = K0.

The ideals of are precisely all the subspaces of the form


a • §I2(IK), where a is some ideal of IK[i, i “ ^]. In fact the correspondence
a a • §I2(1K) is an isomorphism of the lattice of ideals of K[t, t ~^] and
L^I2(1K). Every nonzero ideal of has a unique preimage in §I2(1K),
and this accounts for all ideals of ^12(11^) except (0) and K0. However, with
the exception of ^ l2(IK), Kc, and (0), none of these other ideals is <i-in-
variant. This is clear from what we have already said about ideals of ^I2(1K).
656 / lí‘*-An Extended Example

We can clarify this by an example. Consider the ideal <i - 1> of lK[i,
The corresponding ideal of L § i 2(K) is (t - l>Lgf^([K) = (fUXt - l)x |/(i)
e IK[i, r“ *], Ai e g l2(lK)}. The preimage of this in §I2(IK) is an ideal of
that contains c. As a vector space it is
KC + <r - 1>L§I2(1K).
However, this is not ¿-invariant; If x e § I2(IK)“ \ {0} then ct e §I \
{0} and > 2 \

d{t - l) x = (« (¿ ) -I- 2)x - a { d ) x = 2x ^ (t - 1>L§I2(IK).

In fact the smallest ideal of §I2(IK) containing x is S T J k ).

A PPE N D IX TO CH APTER 4

Our triangular decomposition (§ l2(IK)+,h,Q+,or) satisfies CT1-CT3,


and hence §I2(1K) is contragredient. Since ade, and a d /,, i = 0 ,1, are
nilpotent, and a fortiori locally nilpotent, §Í2(IK) is integrable; in other
words, it is a Kac-Moody algebra. The Cartan matrix A is determined from
'afh) = 2 if i = j = 1,
aj(C - /i) = - 2 ifi = 1 ,; = 0,
a^{h) = - 2 •fi = 0 ,; = 1,
-h) =2 i f i = 0 ,/ = 0.
Thus
2 -2
(A.11) A =
-2 2

with Coxeter-Dynkin diagram O O of type The Lie algebra § I2(IK)


is often referred to as A^i\
The Weyl group is given by W = ( tq, Tj ). Direct computation using the
structure matrix gives
rQ-.ao^ - a o , a , -> 2ao + ttj = «Q-I-8,
rj : «0 “i + “i -« 1
(here, as before, 5 == « q -I- « 1). Then
r, 5 = 8 , i = 0,l,

''o'*!: «0 “ 0 + 25, «1 a - 25,


Appendix to Chapter 4 657

Thus IV has infinite order and has

<rori> = (Z, + )

as a subgroup of index two. A presentation for W is

<''o.''ll''o = = 1>-

This is in accordance with Theorem 5.3.1.

W = [I, ro, rj, r^r^, r j r o , roz-irg,. . . } .

The roots are, as we have already seen,

(A .12) = W{aQ, a^} = [ka^ + m a j \k — m\ = 1}

= Z5.

The triple i? := {§, II, II is a realization of A (the natural realization).


One notes that it conforms exactly with Example 1 of Section 4.2. From
Proposition 4.2.4,

u(^, /?) = u_e ij ©u+,


where is freely generated by two generators eg, and u_ is freely
generated by two generators, /g, /i- According to Proposition 4.2.10 there are
surjective homomorphisms <p.

(A.13)

where g^(A, R) ■= u(A, R)/J(R), J(.R) is the ideal of u{A, R) generated by

(adCifej, ii^i,

d r - ( a d / ,) V y , i* i,

and q(A, R) = u(A, f?)/rad u(A, R)). But we have already seen that
rad(§l2(IK)) = (0), and hence (see Proposition 4.2.8)

§l2(K) = 9 ( ^ ,i? ) .
<p'
In fact the Gabber-Kac theorem (Theorem 4.6.4) says that Q^(A, R) =
g(A,i?). This means that § I2(IK) is defined by the generators d, /i,, e¡, /„
658 Extended Example

i = 0,1, and the relations

(A .14) [A o,A i ] = 0, [¿ ,A ,] = 0, / = 0,1

[A„e,] [d,ej\= ej

[hi, fj\ = - A j i f i [d,fj\ = - f j / , / = 0,1,

[^<>/;] ~ ^ij^i

(ad e^fcj = 0 = (ad f i f Cj, i Фj

(see Section 4.2, Remark 2), and furthermore

D ( i l ^ ) = [§I^0<),§I^(IK)] = D iTI(i<j“ + IKAo + IKAi

is defined by the generators h^, e^, f-, / = 0,1 and the same relations (A.14)
less those involving the generator d (Proposition 4.3.4).
One would suspect that in the relatively simple case here there would be
an elementary way to see that g^(^, R) = д(Л, R) without recourse to the
Gabber-Kac theorem. It is a worthwhile exercise to look at this and convince
oneself of how difficult it is. According to Proposition 4.1.14, 9^(Л, R) and
^ I2(IK) = д(Л, R) have the same root systems and in both cases the real root
spaces are one dimensional. It follows that ker <p = rad(g^(^, R)) c
£*e2\{0>9^(A,/?)*«.ThuS

[e„rad(g4A,/?))] c r a d ( g 4 ^ , / ? ) ) n E = (0),
fcsZ\{0}

and similarly [/¿,rad(g'^(yl,/?))] = (0). It follows that rad(g^(^, Л) central­


izes D(g^(^, Ю) (which is generated by e„ /), / = 0,1), and hence

D ( g t( ^ ,/? ) ) - ! ^ ii^ ^ 0

is a central extension. In fact one can show that §I2(IK) is centrally closed
[Ks], so this sequence splits. It follows at once that Dg^(^, i?) = ^ §I2(1K).
The Cartan matrix (A.11), ~ | 2 2 ) ’ ** symmetric and hence
symmetrized by ( sq, Cj) = (1,1). The map of Ch 4.4.(8) is then

(A.15) Х
Л.1
a- , / = 0,1.

In particular 5° = = <г. According to the prescription of Ch 4.4, we


Appendix to Chapter 5 659

define a symmetric bilinear form (*| • ) on ^ by using the matrix

(A.16)

relative to the basis a / , d). Here * = (d\d) is arbitrary. This extends


uniquely to all of i Í 2(IK)= g ( /l ,/?) by the use of Ch 4.4(10a), (10b), and
(10c)

[t"e, r " f ] = ( i'’e |i - 7 ) ( a , + n S f = ( t " e \ r ' ’f ) { a ¡ ' + n<f),

[ r h , r " h ] = ( t " h \ r " h ) ( n 8 f = (t'‘h \ r ’' h ) m .

However, we already have a bilinear form ( | • ) on L ^ l2(IK) [see (A.1)] and


L ^ Í 2ÍK) is a subspace of §Í2((K). The definition of multiplication (A.3)
together with /c(e, / ) = 4, /c(/z, h) = 8, show that these two forms coincide on
L^l2(K).
The definition of the contravariant form < • | • > in Section 4.7 gives us

= (í"e|í7-í"e) =
-(e\f) = h
(r hU^ h) = ( t^ h lr ^ h ) = (h\h) = 2,

from which < • | • > is clearly positive definite if K is a totally ordered field.
This is deceptively simple. In general Kac-Moody algebras root spaces are
not one dimensional, and there is no easy way, here, to make the computa­
tion of < • I • >.

APPENDIX TO CHAPTER 5

We have

$ = lKao''+ Kd,

where (a¡, d) = 1, i = 0,1, and

]^* = K ckq + [Ko!^ + H.d*.


We define

(d*,an= S,^o,
( d * ,d ) = * arbitrary in IK.
^i^^-An Extended Example

The matrix of ( • , • ) on $ x 1^ relative to these bases is

2 -2 11
-2 2 0
1 0 *J
[compare with (A.16)] and

»
where

n - { a o ,a ,} , n ''= = { a o " ,< }


is a set of root data.
The reflections /*q, defined above extend to by

rod* =d* - ( d * , a ^ ) a o = d* - a o ,
rid* = d* - { d * , a ^ ) a i = d*.

The group that they generate restricts faithfully to + Z aj according to


Corollary 2 to Proposition 5.2.3. But this is in any case obvious from the
presentation of W given above. Likewise we define reflections r^ , r / on $
by
= h - (ai,h}ay, i = 0,1.

These generate a group with presentation ( r ^ , r^^lr^'^ = = D . which is


isomorphic to W (Lemma 5.1.2). The Bruhat ordering on W is, by Proposi­
tion 5.4.2,

ro^iro '• i v l

rcJ
O'l r^r,
I'O
(A.17)

The Tits cone and fundamental chamber in are illustrated in Figures


5.3 and 5.4.
Appendix to Chapter 6 661

For w ^ IV,

5^ := {a e A + k “ ’a e A_}
and

(SJ^= L « (Section 5.2).

With
p := 2d* - jao,

we see that p is a minimal regular in weight:

<p,a,^> = 1, / = 0,1.

Note that for all /: e [K, p + A:5 is also a minimal regular in weight. We have
(Proposition 5.2.5)
<5^> = p - w p .

It is easy to see by induction on k that

(A.18) (''o''i) P ^ P ^ 2ka^ — k{2k + 1)5,


r^ijQ r^) p = p — {2k + l)ai — k{2k + 1)5 for all k

Thus
= - l k a , + k { 2 k + \ )8,
(A.19) = (2* + 1)«! + k(2k + 1)5,
, , i m ( m — 1)
{<5,,>k e IF} = m a, + ---- -8\m g Z

Notice that (A.19) establishes a 1-1 correspondence between W and Z under


which l(w) s m (mod 2). For further reference we also note

(A.20) (ro/'j)"d* = d* + noiy — n^8 for all n

APPENDIX TO CHAPTER 6

We define a restricted weight system P ■= Zd* + Z (aj/2 ) + Z5 with set of


fundamental weights il = {wg, w j, where Wq = d* and o)^ = d* +
<562 Extended Example

We next examine what the denominator formula of Corollary 1 to The-


orem 6.4.1 has to tell us in the case of §I2(IK). Using (A. 18), the fact that
dim §l2(IK) = 1 for all a e A^., and the explicit description of A^. as seen
in (A.9), we can write the denominator formula as

E c(p + Ika^ - k{2k + 1)5)


k = —OO

~ H ^(p — (2A: + I)«! —k(2k + 1)6)


k = -00
00

= e (p )n (l-e (-a o -A :5 ))
k= 0

X(1 - e ( - « j - it5))(l - e { - { k + 1)5)).

We cancel off e(p) from both sides, and set x = e(-a^), q = ei —8) [so
e (-a o ) = obtaining
j.-2k^k{2k-^\) ^ ^(2k +D^k(2k +\)
E
k^Z k^Z

= ‘9*-"‘) ( l- x ( ? * ) ( l - «*■"')
k= 0

or, finally, combining the two sums,


(A.21)

E (-l)"xV^"- D /2
k =0
This is a nontrivial fact called the Jacobi triple product identity. One may
consult [HW, An] for applications of this to number theory. Let us remark on
two of its specializations: replacing q by and x by q,

E (-i)V3"-'^/^= ñ íi- í" ) ,


which is Euler’s pentagonal number theorem. Since n ” =i(l —^") expands as
~ where p^ {k) [resp. p^ik)] is the number of distinct
partitions of k into an even number (resp. an odd number) of parts, we
obtain p^{ k) = Po(k), except when A: is a pentagonal number, namely of
the form ^nOn - 1) (See [HW] for an explanation of the terminology). The
right-hand side is usually denoted by (p(q):

n= i
Appendix to Chapter 6 663

It has already appeared in Section 2.5, Example 3. Euler used his pentagonal
number theorem to compute the partition function. Since

l.y ( ,) - V ( 9 ) - ip (k ),4 z (-ir« " » » — « ) ,


^=0 meZ /
we obtain, by looking at the coefficient of n > 0,

0= E (-irP (A :),

where I(n) •= {(A:, m)\k + |m(3m — 1) = n}. Thus

(A.22) E - iw (3m - 1)) = So „


meZ

[p(j) == 0 if j < 0], which gives a very effective recursive determination of


pin).
If we replace q h y and x by q, then the Jacobi triple product identity
gives the Gauss identity

E (-i)V = n ( 1 ( 1 - «^")
«eZ "=1

<
pW ■
Next we consider in detail the integrable highest weight module L(fi*) of
§I2([K) with highest weight d*. According to Theorem 6.4.2 there is, up to
isomorphism, only one such module (hence the use of the word “the” above),
and it is irreducible.
The weight system P{L{d^)) lies in the set d* - (Na^ + Na^X Let
II G P(L(6i*)). We claim that also ¡jl - 8 ^ P(L(d*X). Since W8 = 8 and
WjjL n P^¥^ 0 (see Proposition 6.1.11), it is enough to prove the claim for
jji ^ P ^ n P(L(d*)X By Proposition 6.2.5(i), ¡x is connected through d*. This
means that if we write jx = d* - a ^ d* i Q+, then either fx = d* or
(d*, a / > # 0 for some i e supp(a). In our case this means that jx = d* or
«0 ^ supp(a). In any case i e supp(a) => i e supp(a + 8 \ and since
/X —5 e P^r\(d* i Q^X we may use Proposition 6.2.5(ii) to conclude that
fx — 8 ^ P(L(d*X), (This argument works quite generally for affine Lie
algebras.)
If /X e and we write ¡x = d* - maQ — na^, m, n e N, then from

(fx, a ^ } = 1 —2m 2n ^ N,
( f x , a ^ ) = 2m — 2n ^ N,
664 Extended Example

we obtain 0 < 2(m — n) < \ from which m = n. Thus ¡jl = d* - m8. Putting
this together with the remark above, we see that

P (L (d * )) = Wd* - N5.

Since r^d* = d* and W = u (ro ri)ri, we obtain from (A.20)

(A.23) P(L{d*)) = [d* + na^ - { n^ + k)8\n e Z, A: G N) .

This is illustrated in Figure A.1 (see [FL]). The dotted curves represent
PT-orbits. Multiplicities are constant on orbits. It is a consequence of Proposi­
tion 6.2.8 that the multiplicities increase (not necessarily strictly) as we pass

Figure A.1. The weight system of L ( d * ) for §I2(1K). If IK = [R the shaded region
represents that part of d * + IKao + IKaj that lies inside the fundamental chamber F.
The dotted curves represent PT-orbits.
Appendix to Chapter 6 665

to successively lower orbits. In fact we will soon describe the multiplicities


exactly.
From our description (A.23) of the weight system of L(d*), we have

ch(L (d*)) = E E - kd)


j^Z k^ N

Since the weight multiplicities of L(d*) are IF-invariant, the inner sum can
be written as

b f, ■■= E d im L (d * )‘'’ ~**e(-A:S)


k^N

and

(A.24) ch L ( d * ) = b r . - L e { ( r o r , y d * ) .

The expression b^* is called a string function or branching function. (Of


course, as it stands, it is not a function at all. However, see below.) The
notation b^* anticipates the more general notation of branching functions.
Write

bj* = E p ( k ) e ( - k d ) = E
k=0 k=0

where q •= e ( - 8 ) as usual. We plan to prove that p(k) really is the partition


function. For this purpose we apply the Weyl-Kac character formula (Theo­
rem 6.4.1). Here we use (A.24) and replace the denominator in Theorem
6.4.1 by the skew-invariant sum of the denominator formula (corollary to
Theorem 6.4.1).

(A.25)

^\k=0 l ^ j^ Z '

= E ( - l ) ''" 'e ( w ( d * + p ) ) .

Now observe that on the right-hand side the only term of the form e{d* +
666 Extended Example

p - nS), fi >0, occurring is e(d* + p), since d* + p - nd ^ P^, but by


Proposition 6.1.11, W(d* + p) n P^= {d* + p}. Let us compute the coeffi­
cients of all the terms e(d* + p — nS) on the left-hand side, namely

(A.26) £ L ( - l ) « “'V(A:))c(d*-t-p-nS),
n = 0 \ (w, kJ)^B(n) /

where Bin) ■■=[iw,k,j)\wp ~ k8 + ir^r^Vd* = d* + p - nS}.


Using (A. 19) and (A.20), we see that

wp - k 8 + (rotiYd* =d* + p - n8
m (m — 1)
<=> —mai — 8 - k8 + ja^ —j^8 = —n8

[where w ^ m in (A.19), and we have used p — wp =

(m - j = 0,
(A.27) - m ( m — 1)
- k = -n,

f m - y = 0,
<=> ( m
I k + — {3m — 1) = «.

Since each m e Z and n I determines a unique A: e N satisfying (A.27),


we have from (A.26),

(A.28)

— \m{3m — l))e(<i* + p - n8) = e{d* + p)


« =0 m

[the right-hand side coming from the remark that we already made about the
right-hand side of (A.25)]. But this is precisely the formula (A.22) that we
have seen determines the ordinary partition function. Thus indeed p{k) is
the usual partition function, and
. -1
(A.29) bf.=cpiqy

Thus we have the remarkable fact that

(A.30) dim L ( d * ) ‘‘ = p(k).


Appendix to Chapter 6 667

We can now rewrite the character formula (A.24) as

ch
neZ

1
Y, e{d* + na^ - n^8)
<P(9) n

ejd*)
E x-^q""
< p (^ ) n

where x == e( —a^).
Specializing x to 1, we obtain

^(^*) -^(^*)I specialized with e(-ai)= 1


©((?)

where ©(^) — = 1 H- 2^ + 2q^ + • • • . This is the theta function


of the lattice Z. All of these things have suitable generalizations, which have
been the subject of considerable investigation (see [KPl] for more on this).
Suffice it to say we have arrived at the point where modular forms enter into
the theory. The transition into functions is made hy q where r
ranges over the upper half plane in C (i.e., {z e C|Im z > 0}, and
the modularity refers to the transformation properties of these functions
ia h \ ar-h b
under the action of SLSX) on defined by j\ = ------- 7-
\c a } CT d
The way in which the weight system of L(rf*) decomposes into strings
- k8, k = 0,1,2,... seems to suggest that L(d*) itself decom­
poses with respect to some suitable representation of a Lie algebra. This is in
fact the case. To see how this comes about we assume that K = C.
Consider the Lie algebra

a = Cc + E ^I2(C) = Cc + E

We have
Extended Example

Define a(.n) = t " h / y/l n ^ Z \ {o}. xhen

® I I Ca{n) + Ci,

[a(m ),e] = 0.

Note that by our choice of d, above

[^>a(m )] =2ma {m ).

The Lie algebra a is a Heisenberg algebra. It can be identified with the


Heisenberg algebra of Section 2.2., Example 2, by

Now we want to apply Proposition 2.8.4. We have

L(d*) is unitarizable (Theorem 6.4.3),


[d,a] c a,
cr(a) = a [<r(a(n)) = a(-n )],
<a, d) = htia) > 0 for all a e fl = {«(,, a j .

Thus Lid*) is completely reducible as an a-module. Note that with á ~ a +


Cd, a decomposition of Lid*) into irreducible S-modules is also a decompo­
sition of Lid*) into irreducible a-modules (see the proof of Proposition
2.8.4).
Now consider the space

- k S

*-0

formed from the weight string {ir^r^yd* - It is evidently an fi-mod-


ule. Now & has the triangular decomposition á = d_-f-(Ce -l- Cd) -I- fi+. Let
Vj G Lid*y'°'''^‘‘^* \ {0}. This spans L(á*)^''»''i^‘'* and is evidently a highest
weight vector for S (in the displayed triangular decomposition). Since ^v¡ ^ 0,
this vector generates a Verma module MUr^r^Vd*) c Vj [here ir^r^yd* is
considered as a linear functional on (Cc -t- Cd)], and by Section 2.13 it is an
Appendix to Chapter 6 669

irreducible a-module. By the remark above, it is an irreducible a-module. It


follows that its character is

chM ((ro r,)V * - k8) = e{(r^r,yd* - k8) O (1 - e ( - a ) y " ' \

= dim a “ (see Proposition 2.19).

In this case Q+= Z+5 and = 1 for all e Z+. Thus

chM ((ror,)'d* - k8) = e((rori)'d* - A:5) O (1 “ e { - k 8 ) y

= «{(roriYd* - k8) n (1 - 9*)


k= l

= e{ir^r,yd* - k8) E p(k)q>^.


k= 0

But we also have

chVj = e((rori)''d* - *8)chFo = e^iror^yd* - k8)bf.,

since dim is constant on the orbit Wk, Thus by (A.30),

(A.31) ch = ch Vj.

But now there is a nice trick that shows that — k8) = Vj. We
already know that M d r ^ r ^ d * — k8) ^ Vj. Hence for all weights a,
dim M d r ^ r y d ^ Y < dim Vf. But (A.31) then gives equality. Thus we have
proved that

Vj is an irreducible a-module,
L(d*) = ®j^jVj a direct sum decomposition of L(d*) into irreducible
a-modules.

We can now give a structural interpretation to (A.24). Each weight space


jg dimensional and is spanned by a vector that is
annihilated by a^_. These highest weight vector or “vacuum vectors” span the
entire space annihilated by a+. Equation (A.24) expresses a factorization of
L(d*) as an irreducible a-module times the space of vacuum vectors. This
was one of the insights that led to the Frenkel-Kac vertex operator construc­
tion of affine representations ([F-K, KMPS, MoPi]).
It is interesting to look at the meaning of Proposition 6.7.7. Corresponding
to the Bruhat ordering (A. 17), we have the following inclusions amongst
Extended Example

Verma modules for M(A), A e p^:

/ Ai(A) \

and furthermore there are no other submodules. Since tv >v • A is an


injective map and tv • A c A J. (3^, it is clear that for any e {w\ii e (w •
A)i Q+} is finite and hence ■A) = (0).
We now consider the character of the irreducible module L(A) =
M(A)/(M(rQ • A) + M(fi • A)). As a first estimate we have
ch M(A) - {ch M(ro • A) + ch M{r^ • A)}.
But A/(fo • A) + M(ri • A) is not a direct sum; indeed all the submodules
M(w • A), Z(tv) ^ 2 lie inside the intersection of these two modules. Thus we
have removed ch Mir^r^ • A) + ch Mir^r^ ■A) twice, and as our second esti­
mate we have
ch M(A) - ch M{r^ • A) - ch M (rj • A) -1- ch M{rQV^ • A)
-I- ch M{r^rQ ■ A).
Of course the third level has now to be corrected because we have added too
much, and so by the principal of exclusion and inclusion we obtain

c h L (A )= £ (-l)'^ '^ ^ c h M (w A ).
w^W
Since ch M(w • A) = • A)/n^^eA+d ~ ^-a)dimg“ Proposition 2.5.3),
we obtain in this way a simple explanation of the character formula (6.29).
For more on this type of approach to the character formula in general, see
[GLl] and [GL2].)

APPENDIX TO CHAPTER 7

Next we consider the Lie groups G and G of Section 6.1 that are attached to
§I2(IK). We can derive a considerable amount of information by constructing
a special integrable representation of ^ I2( 1K).
Appendix to Chapter 7 671

Let ( it, V), where V = be the natural 2-dimensional representation of


§I,(IK):

(A.32)
\c d)[v ) [cu + dv ) ’

and set L F ~ IK[r**] ® F. Then we have a representation of L § l2(IK) on


L F by (A.32), where now a , b , c , d , u , v e By letting (5 act trivially,
we can lift this to a representation

^ g l( L F ) .

Define a Z-grading on L V (as a vector space over K) by

deg| I = 2k,

deg =2 k-l.

and observe that this is compatible with the Z-grading of i l 2(IK), namely for
X e §i^([K)*, y G L F '”, X • y G LF*+'”. For example, deg t ^ f = 2p - 1,
deg| Q j = 2q, deg(i^/ ‘10 1) = deg| ^ = 2{p + q) - 1. Now we extend
our representation to a representation tt of ^12( 1^ ) on L V by defining

d ' X = kx
on the elements x e LK^, /: e Z.
Since the real root spaces of ^I2(1K) are spanned by the elements t^e and
t ^ f , k ^ Z, it follows that ( , LV) is integrable. Thus affords a representa­
t t t t

tion TT of G and G = G(P) on LV. Now we observe that

K exp5i*e)|“ j = 1
0 1 jU i’

K ex p si* /)(“ ] =
1
St*/ :i(;i
Thus ir(G) c S L 2(IK[r±‘]).
The study of SL2(R) for commutative rings is a whole world in its own
right. If f? is a euclidean ring, then 5L 2O?) is generated by the elementary
matrices ( j (j J ) . a G R. In particular we see that ir(G) =
5L2(lK[r±*])(see [HoM]).
672 ^i^^-An Extended Example

Now we invoke the result of Exercise 6.30. Since ker tt = c ]^, we have
ker TT c Z(G) and G —>5L2(lK[i -^]) is central.
According to Theorem 6.3.14

Given an integrable representation (i//,M), 9fr(h^Jis)h^f±s)) acts on as


^<A,ao >(+5)<A,ori >Yhich in general is not trivial.Vhus we see that Z(G) =
[K^ X {±1}.
For LV, aQ ai ^ c acts as 0, so <A, = “ (A,
and h^^s)h^(^±s) acts as ( ~ 1)<^’"i''>. Since | ¿ j ^ is a weight vector with
weight A where <A, a /> = 1, we see that — does not act trivially,
namely ^ ker tt. [For the adjoint representation, however, the
weight satisfy <A, ao^> = and furthermore <A,ai^> e 2Z. Thus
hc^s)h^(,±s) acts trivially for all s, as we would expect.] We conclude that
we have the central extension

{1} ^ G 5L2(IK[i±']) -> {1}.

According to Exercise 6.30 the mapping Ad induces an epimorphism

5L2(K [i±i]) A d(G ),

and Ad(G) - SL2(K[t - ^])/Z(SL2(K[t - It is not hard to see the centre is


± IP ^ j , and hence

Ad(G) -P5L2(lK[i=^^]) == 5L 2(K [i± ^])/{± l}

(projective special linear group).


We have already defined

/>= Zd* + Zh
( tr ) +

Now we wish to consider the larger group G = G(P) = GX, where X •


Hom(P,K^).
Appendix to Chapter 7 673

Consider an element e H. As an element of X it corre­


sponds to the homomorphism of P into K” satisfying

d* = to,
^ f^cc,/2,a^'y^(a,/2.c:n =

S ^ 1.
Clearly X = H X X q, where

jifo = e H o m (P ,K O k (d * ) = < p (y ) = 1

= Hom(Z5,IK^) = Hom(Z,[K=") == K^.

Also by the corollary to Proposition 6.3.12, G n X = H, G n X o = l , G =


G ■X q. Finally, recalling X q ~ {;t' e X\x(Q) = 1} (see Theorem 6.3.14),

Z (G ) = X q = H om (P/!2,IK'') = Hom(Z X (Z/2Z),IK ^)

/ Z
= Hom(Z,lK^) X Horn — ,IK^

= K"^x{±l} = Z ( G ) .

In fact f o i h ^ H and e X q, hxoiQ) = 1 => ;^o = 1> hence Z(G) =


Z(G). Thus G = G X IK^, Z(G) = Z(G) = IK^ x ( + l}.
Finally, we consider the group ¿¡/~ Aut(§l2(IK)). Though Ad we have
.QfD Ad G = G /Z G = G • Xo/ZiG) = P5L2(IK[t ±‘]) • X^. We already know
that P5L2(IK[i **]) embeds in as the group of inner automorphisms
generated by the elements

|expad se„|s e K-*, a e " A , e„ ^

or equally well by the elements

{exp ad sc,, exp ad s/,|s e K^, i = 0,1}

(Corollary to Proposition 6.1.9). If x ^ is defined by x(8) = a e K^,


Xid*) = x(<Xi/2) = 1 then a: acts on i l 2(IK)'"“»^”“' as o'".
Let 0 e S3/, then Of) is some split Cartan subalgebra and by Thm. 7.4.9
there is an element <p ^ G with (Ad(^))“ ‘01^ = 1^. Let = (Ad <p)~^6. It
follows by Proposition 2.1.3 that there exists a linear isomorphism /x* on
so that fi^g“ = g** *“ for all a e A. In particular the base II = {«q, « J is
mapped into the base fi* II. By Proposition 5.9.2 there is an element w e IT
with w ll = ±^*11.
674 Extended Example

Let n ^ G satisfy l^J = w. Then 1x 2 == (Ad n) ^Ad <p) ^0 sat-


isfies fjL2^ = = ± n . There are three possibilities:

Case 1. /Lt*«/ = oiiy i = Q? 1- In this case 112^^2 = ^I2(K)" for all a.


In particular /jl2[^12{^)^\ = M2' a / ) = But then
^/] = 2/i,2^z leads at once to 112^ 1^~ Then M2^z ~
and fi2fi = i = 0,1. Let x ^ HomCQ, K^) be chosen so
that A'(<^/) = ^ = 0? 1; Af induces an automorphism, also denoted by
X, of g by A la" = multiplication by ;^(g). Then x ~^t^2 [e^, a / ,
/¿\i = 0,1}, and so fjL;^ — at” V 2 as 1 on §I2(IK).
Now it is a straightforward exercise to show that ¡x^id) = d + ac for
some a e K. Conversely if « e K, then there is an automorphism cp of
^ I2(K) defined by <pl----- id, <p(d) ■- d a<^. The group C of these
¿I2OK)
automorphisms may be identified with H orndI¡0K) / D l I ¡ ^ ) ),
Z ^l2(IK)) (see Exercise 7.7 where this group appears in full generality).
Thus /X3 e C. Thus, finally, we have shown that d = Ad(g)Aff, where
^g C, geG, ATeHom(i2,K^).
Case 2. At*n^o = In this case we introduce the automor­
phism y of the Coxeter-Dynkin diagram T of A : y : Vq (see
Section 3.4). This determines an automorphism y of ^12( 11^) by Corol-
lary 4.4 which effects Cq <-> «o ^ /0 ^ f v nnd y ’* : «n ^ «1
This automorphism can be lifted (not uniquely, but up to a coset of C)
to an automorphism y of ^ I2( 1K). Then replacing fi2 by y^t2?
back in Case (1).
Case 3. /x*!! = - I I . In this case we use the automorphism -cr of g. We
have ((-o-)jLL2)*n = II, and hence we are in Case (1) or (2).

In summary every element of Aut(^l2(IK)) is expressible in the form

A d(g)Arly(-o-)%

where g g G, x ^ Hom((2, K“^), ^ ^ C, y ^ Aut(F), e ^ {0,1}.

Heaven and earth grow together with me, and the ten thousand things and I are
one. We are already one—what else is there to say? Yet I have just said that we
are one, so my words exist also. The one and what I said about the one make
two, and two and one make three. Thus it goes on and on. Even a skilled
mathematician cannot reach the end, much less an ordinary man... Enough.
Let us stop!
—Chuang tsu
Bibliography

[An] Andrews, G. E., The theory of partitions, E n c y c l o p e d i a o f M a t h e m a t i c s a n d


I t s A p p l i c a t i o n s . Vol. 2. North Reading, MA: Addison-Wesley, 1976.

[Ar] Artin, E. G e o m e t r i c A l g e b r a . New York: Wiley Interscience, 1957.


[Bg] Berge, C. T h e T h e o r y o f G r a p h s . London: Methuen, 1962.
[Br] Berman, A. Cones, matrices, and mathematical programming. Lect. Notes
in Econ. and Math. Systems 79. Berlin: Springer-Verlag, 1973.
[Be] Berman, S. On derivations of Lie algebras. C a ñ a d . M . M a t h . 28 (1976):
174-180.
[Be2] Berman, S. On generators and relations for certain involutory subalgebras
of Kac-Moody Lie algebras. C o m m . i n A l g . 17(12) (1989): 3165-3185.
[BM] Berman, S. and Moody, R. V., Multiplicities in Lie algebras. P r o c . A m e r .
M a t h . S o c . 76 (1979): 223-228.

[BMW] Berman, S., Moody, R. V., and Wonenburger, M., Cartan matrices with null
roots and finite Cartan matrices. I n d i a n a U n i u . M a t h . J . 21 (1971-72):
1091-1099.
[BGGl] Bernstein, I. N., GeFfand, I. M., and Gel’fand, S. I. The structure of
representations generated by a vector of highest weight. F u n k . A n a l , i P r i l o
5:1 (1971): 1-9 (in Russian); F u n c . A n a l . A p p l . 5 (1971): 1-8 (in English).
[BGG2] Bernstein, I. N., GeFfand, I. M. and GeFfand, S. I. Category of g-modules.
F u n k . A n a l , i P r i l o . 10 (1976): 1-6 (in Russian); F u n c . A n a l . A p p l . 10:2
(1976): 87-92 (in English).
[Bel] Borcherds, R. Vertex algebras, Kac-Moody algebras, and the monster. P r o c .
N a t . A c a d . , USA, 83 (1986): 3068-3071.

[Bc2] Borcherds, R. Generalized Kac-Moody algebras. J . A l g . 115 (1988):


501-512.
[Bor] Borel, A. L i n e a r A l g e b r a i c G r o u p s . 2d ed. New York: Springer-Verlag, 1991.
[Bol] Bourbaki, N. Algèbre. A c t u a l i t é s S c i e n t i f i q u e s e t I n d u s t r i e l l e s 1261. Paris:
Hermann, 1958, ch. 8.
[Bo2] Bourbaki, N. L i e G r o u p s a n d L i e A l g e b r a s . North Reading, MA: Addison-
Wesley, 1975, chs. 1-3.
[Bo3] Bourbaki, N. G r o u p e s e t a l g e b r e s d e L i e . Paris: Hermann, 1968, chs. 4, 5, 6.
[Bo4] Bourbaki, N. G r o u p e s e t a l g e b r e s d e L i e . Paris: Hermann, 1975, chs. 7, 8.

675
676 Bibliography

[BPI] Billig, Y., and Pianzola, A. On Cartan Subalgebras. J . A l g . (to appear).


[BP2] Billig, Y. and Pianzola, A. Pairs of real roots whose sum is a real root.
T o h o k u J . M a t h (to appear).

[Cp] Cassperson, D. Free products of commuting Lie algebras and strictly


imaginary roots. Ph.D. dissertation. University of Waterloo, 1989.
[Chi] Chevalley, C. T h e A l g e b r a i c T h e o r y o f S p i n o r s . New York: Columbia
University Press, 1954.
[Ch2] Chevalley, C. Sur Certeins groupes simples. T o h o k u M a t h . J . (2) 7 (1955):
14-66.
[Ch3] Chevalley, C. T h e o r i é d e s g r o u p e s d e L i e , vol. 2. Paris: Hermann, 1951.
[Ch4] Chevalley, C. T h e o r i e d e s g r o u p e s d e L i e , vol. 3. Paris: Hermann, 1955.
[CM] Coxeter, H. S. M., and Moser, W. O., G e n e r a t o r s a n d R e l a t i o n s f o r D i s c r e t e
G r o u p s . New York: Springer-Verlag, 1965.

[Cx] Coxeter, H. S. M. Discrete groups generated by reflections. A n n a l s o f M a t h .


35 (1934): 588-621.
[Cx] Coxeter, H. S. M. R e g u l a r P o l y t o p e s . New York: Dover, 1973.
[CDS] Cvetkovic, D., Doob, D. M., and Sachs, H. S p e c t r a o f G r a p h s . San Diego,
CA: Academic Press, 1980.
[De] Demazure, M. Identités de Macdonald. S é m i n a i r e B o u r b a k i , 2 8 e a n n é e
(1975-76), no. 483.
[Del] Deodhar, V. V. Some characterizations of Bruhat ordering on a Coxeter
group and determination of the relative Möbius function. I n v e n t . M a t h . 39
(1977): 187-198.
[De2] Deodhar, V. V. A note on subgroups generated by reflections in Coxeter
groups. A r c h . M a t h . 53 (1989): 543-546.
[DGK] Deodhar, V. V., Gabber, O., and Kac, V. G. Structure of some categories of
representations of infinite-dimensional Lie algebras. A d u . i n M a t h . 45
(1982): 92-116.
[Dy] Dyson, F. J. Missed opportunities. B u l l . A m e r . M a t h . S o c . 78 (1972):
635-653.
[En] Enright, T. On the fundamental series of a real semisimple Lie algebra:
their irreducibility, resolutions, and multiplicity formulae. A n n . o f M a t h . 2
(1979): 1-82.
[FL] Feingold, A. J., and Lepowsky, J. The Weyl-Kac character formula and
power series identities. A d v . i n M a t h . 29 (1978): 271-309.
[FK] Frenkel, I. B., and Kac, V. G. Basic representations of affine Lie algebras
and dual resonance models. I n v e n t . M a t h . 62 (1980): 23-66.
[FLM] Frenkel, I. B., Lepowsky, J., and Meurman, A. V e r t e x O p e r a t o r A l g e b r a s a n d
t h e M o n s t e r . San Diego, CA: Academic Press, 1988.

[FdV] Freundenthal, H., and de Vries, H. L i n e a r L i e G r o u p s . San Diego, CA:


Academic Press, 1969.
[GK] Gabber, O., and Kac, V. G. On defining relations of certain infinite­
dimensional Lie algebras. B u l l . A m e r . M a t h . S o c . 5 (1981): 185-189.
[Gr] Garland, H. The arithmetic theory of loop groups, I H E S , 52 (1980).
Bibliography 677

[GLl] Garland, H., and Lepowsky, J. Lie algebra homology and the Macdonald-Kac
formulas. I n v e n t . M a t h . 34 (1976): 37-76.
[GL2] Garland, H., and Lepowsky, J. The Macdonald-Kac formulas as a
consequence of the Euler-Poincaré principle. C o n t r i b u t i o n s t o A l g e b r a , A
C o l l e c t i o n o f P a p e r s D e d i c a t e d t o E l l i s K o l c h i n . Academic Press, pp. 165-173.

[GO] Goddard, P., and Olive, D. K a c - M o o d y a n d V i r a s o r o A l g e b r a s : A R e p r i n t


V o l u m e f o r P h y s i c i s t s . World Scientific, 1988.

[HO’M] Hahn, A. J., and O’Meara, T h e C l a s s i c a l G r o u p s a n d K - T h e o r y , New York:


Springer-Verlag, 1989.
[HW] Hardy, G. H., and Wright, E. M. A n I n t r o d u c t i o n t o t h e T h e o r y o f N u m b e r s .
Oxford: Oxford University Press, 1960.
[HS] Hilton, P. J., and Stammbach, U. A C o u r s e i n H o m o l o g i c a l A l g e b r a . Berlin:
Springer-Verlag, 1971.
[Hul] Humphreys, J. E. I n t r o d u c t i o n t o L i e A l g e b r a s a n d R e p r e s e n t a t i o n T h e o r y .
G r a d . T e x t i n M a t h 9, Berlin: Springer-Verlag, 1972.

[Hu2] Humphreys, J. E. R e f l e c t i o n G r o u p s a n d C o x e t e r G r o u p s . Cambridge:


Cambridge University Press, 1990.
[Iw] Iwahori, N. Private communication via unpublished notes. Northeastern
University, 1967.
[Jal] Jacobson, N. B a s i c A l g e b r a I . New York: Freeman, 1974.
[Ja2] Jacobson, N. B a s i c A l g e b r a I I . New York: Freeman, 1980.
[Ja3] Jacobson, N. L i e A l g e b r a s . New York: Dover, 1962.
[Jz] Jantzen, J. C. M o d u l n m i t e i n e n h ö c h s t e n G e w i c h t . L e c t u r e N o t e s i n M a t h .
750. Berlin: Springer-Verlag, 1979.
[Kal] Kac, V. G. Simple irreducible graded Lie algebras of finite growth. I z u .
A k a d . N a u k S S S R 32 (1968): 1323-1367 (in Russian); M a t h . U S S R — I z u 2
(1968): 1271-1311 (in English).
[Ka2] Kac, V. G. Infinite-dimensional Lie algebras and Dedekind’s 77-function.
F u n k . A n a l , i a g o P r i l o h . 8 (1974): 77-78 (in Russian); F u n c . A n a l . A p p l . 8
(1974): 68-70 (in English).
[Ka3] Kac, V. G. Infinite-dimensional algebras, Dedekind’s 17-function, classical
Möbius function and the very strange formula. A d v . M a t h . 30 (1978):
85-136.
[Ka4] Kac, V. G., Infinite-root systems, representations of graphs, and invariant
theory. J . A l g . 78 (1982): 141-162.
[Ka5] Kac, V. G. Infinite dimensional Lie Algebras. 2d ed. Cambridge: Cambridge
University Press, 1985.
[KK] Kac, V. G., and Kazhdan, D. A. The structure of representations with
highest weight of infinite-dimensional Lie algebras. A d v . M a t h . 34 (1979):
97-108.
[KPl] Kac, V. G., and Peterson, D. H. Infinite-dimensional Lie algebras, theta
functions, and modular forms. A d v . i n M a t h . 53 (1984): 125-264.
[KP2] Kac, V. G., and Peterson, D. H. Unitary structure in representations of
infinite dimensional groups and a convexity theorem. I n v e n t . M a t h . 76
(1984): 1-14.
678 Bibliography

[KP3] Kac, V. G., and Peterson, D. H. On geometric invariant theory for infinite
dimensional groups, in Algebraic Groups. L e c t u r e N o t e s i n M a t h . 1271.
Berlin: Springer-Verlag, 1987.
[KaWk] Kac, V. G., and Wakimoto, M. Modular invariant representations of infinite
dimensional Lie algebras and superalgebras. P r o c . N a t . A c a d . S e i . , USA,
85 (1988):
[KWg] Kac, V. G., and Wang, S. P. On automorphisms of Kac-Moody algebras and
groups. A d v . M a t h . 92 (1982): 129-195.
[KMPS] Kass, S., Moody, R. V., Patera, J., and Slansky, R. A f f i n e L i e A l g e b r a ,
W e ig h t M u l t i p l i c i t i e s , a n d B r a n c h i n g R u l e s , Boulder: University of Colorado
Press, 1990.
[Ke] Kempf, G. Instability in invariant theory. A n n . M a t h . 108 (1978): 299-316.
[Kol] Kostant, B. A formula for the multiplicity of a weight. T r a n s . A m e r . M a t h .
S o c . 93 (1959): 53-73.

[Ko2] Kostant, B. Theorem 1.1 of G. Lancaster and J. Towber, Representation-


functors and Flag-algebras, J. of Alg. 59 (1979), 16-38.
[Ks] Kassel, C. Kahler differentials and coverings of complex simple Lie algebras
extended over a commutative algebra. J. Pure and Applied Algebra 34
(1985): 265-275.
[Lp] Lepowsky, J. Generalized Verma modules, loop space cohomology and
Macdonald-type identities. A n n . S e i . d e V É c o l e N o r m , 1 2 (1979): 169-234.
[Lo] Looijenga, E. Invariant theory for generalized root systems. I n v e n t . M a t h . ,
61 (1980): 1-32.
[Lth] Lothaire, M. Combinatorics on words. Encyclopedia of Mathematics and
Applications, North Reading, MA: Addison-Wesley, 1983.
[Mac] Macdonald, I. G., Affine root systems and Dedekind’s 17-function. I n v e n t .
M a t h . , 15 (1972): 91-143.

[MKS] Magnus, W., Karass, A., and Solitar, D. C o m b i n a t o r i a l G r o u p T h e o r y . New


York: Wiley, 1966.
[Mr] Marcuson, R. Tits’ systems in generalized nonadjoint Chevalley groups. J .
A l g e b r a 34 (1975): 84-96.

[Mt] Matsumoto, H. Générateurs et relations des groupes de Weyl généralisés,


C . R . A c a d . S e i . , Paris 258 (1964): 3419-3422.

[Mx] Maxwell, G. Sphere packing and hyperbolic refiection groups, J . A l g . 79


(1982): 78-97.
[MPS] Michell, L., Patera, J., and Sharp, R. J. The Demazure-Tits subgroup of a
simple Lie group. J . M a t h . P h y s . 29 (4) (1988): 777-796.
[MH] Milnor, J. and Husemoller, D. Symmetric bilinear forms. E r g e b n i s s e d e r
M a t h e m a t i k , 73. Berlin: Springer-Verlag, 1973.

[Mol] Moody, R. V. Lie algebras associated with generalized Cartan matrices.


B u l l . A m e r . M a t h . S o c . 73 (1967): 217-221.

[Mo2] Moody, R. V. A new class of Lie algebras. J . A l g e b r a 10 (1968): 211-230.


[Mo3] Moody, R. V. Euclidean Lie algebras. C a ñ a d . J . M a t h . 21 (1969): 1432-1454.
[Mo4] Moody, R. V. Root systems of hyperbolic type. A d v . i n M a t h . 33 (1979):
144-160.
Bibliography 679

[MoPa] Moody, R. V., and Patera, J. Characters of elements of finite order in a Lie
group, S I A M J , A l g . D i s c . M a t h . 5 (1984): 359-383.
[MoPil] Moody, R. V., and Pianzola, A. Infinite dimensional Lie algebras (a unifying
overview). A l g e b r a s , G r o u p s a n d G e o m e t r i e s 4 (1987): 165-213.
[MoPi2] Moody, R. V. and Pianzola, A. On infinite root systems. T r a n s . A m e r .
M a t h . S o c . 315 (1989); 661-696.

[MoT] Moody, R. V., and Teo, K. L. Tits’ systems with crystallographic Weyl
groups, J . A l g e b r a 21 (1972): 178-190.
[MoY] Moody, R. V., and Yokonuma, T. Root systems and Cartan matrices.
C a ñ a d . J . M a t h . 34 (1982): 63-79.

[Ne] Neidhardt, W. L. Irreducible subquotients and imbeddings of Verma


modules. A l g e b r a s , G r o u p s , a n d G e o m e t r i e s 1 (1984): 127-136.
[PK] Peterson, D. H., and Kac, V. G. Infinite flag varieties and conjugacy
theorems, P r o c . N a t l . A c a d . S c L , USA, 80 (1983): 1778-1782.
[Pi] Pianzola, A. The arithmetic of the representation ring and elements of
finite order in Lie groups. J . A l g . 108 (1987): 1-33.
[R-CW] Rocha-Caridi, A., and Wallach, N. Projective modules over graded Lie
algebras I. M a t h . Z . 180 (1982): 151-177.
[Ro] Rotman, J. T h e T h e o r y o f G r o u p s . Boston: Allyn and Bacon, 1965.
[Sn] Seneta, E. N o n - n e g a t i u e M a t r i c e s . London: Allen & Unwin, 1973.
[Se] Serre, J. P. A l g e b r e s d e L i e s e m i - s i m p l e s c o m p l e x e s . New York: Benjamin,
1966.
[Sel] Serre, J. P. L i e A l g e b r a s a n d L i e G r o u p s . New York: Benjamin, 1965.
[Sh] Shapovalov, N. N. On a bilinear form on the universal enveloping algebra of
a complex semisimple Lie algebra. F u n k . A n a l , i P r i l o 6:4 (1972): 65-70 (in
Russian); F u n c . A n a l . A p p l . 6 (1972): 307-312 (in English).
[St] Steinberg, R. L e c t u r e s o n C h e v a l l e y G r o u p s . New Haven: Yale University
Press, 1967.
[Str] Strooker, J. R. I n t r o d u c t i o n t o C a t e g o r i e s , H o m o l o g i c a l A l g e b r a a n d S h e a f
C o h o m o l o g y . Cambridge: Cambridge University Press, 1978.

[Til] Tits, J. G r o u p e s e t g e o m e t r ie s d e C o x e t e r . I H E S N o t e s . Paris, 1961.


[Ti2] Tits, J. Theoréme de Bruhat et sous-groupes paraboliques. C . R . A c a d . S c i . ,
Paris, 254 (1962): 2910-2912.
[Ti3] Tits, J. Algebraic and abstract groups. Annals of Math 80 (1964) 313-379.
[Ti4] Tits, J. R é s u m é d e c o u r s . A n n u a i r e d u C o l l è g e d e F r a n c e (1980-81: 75-87.
[Ti5] Tits, J. R é s u m é d e c o u r s . A n n u a i r e d u C o l l è g e d e F r a n c e (1981-82): 91-106.
[Ti6] Tits, J. Uniqueness and presentation of Kac-Moody groups over fields. J .
A l g . 105 (1987): 542-573.

[vdW] van der Waerden, B. L. Die Klassication der einfachen Lieschen Gruppsen.
M a t h . Z . 37 (1933): 446-462.

[Wed] Wedderburn, M. J. On hypercomplex numbers. P r o c . L o n d o n M a t h . S o c .


(2) VI (1908): 77-118.
[Vel] Verma, D.-N. Structure of certain induced representations of complex
semisimple Lie algebras. Ph.D. dissertation. Yale University, 1966.
680 Bibliography

[Ve2] Verma, D. N. Möbius inversion for the Bruhat ordering on a Weyl group.
A n n . S e i . E c . N o r m . S u p . (1971): 393-399.

[Vi] Vinberg, E. B. Discrete linear groups generated by reflections. I s v e s t i j a A N


U S S R (ser. mat.) 35 (1971): 1072-1112; E n g l . T r a n s . M a t h . U S S R - I z u e s t i j e 5
(1971): 1083-1119.
[Wyl] Weyl, H. Theorie der Darstellung kontinuierlicher halb-einfacher Gruppen
durch linear Transformation. M a t h . Z . 23 (1925): 271-309.
[Wy2] Weyl, H. Theorie der Darstellung kontinierlicher halb-einfacher Gruppen
durch lineare Transformation. M a t h . Z . 24 (1926): 328-395.
[Wy3] Weyl, H. Theorie der Darstellung kontinierlicher halb-einfacher Gruppen
durch lineare Transformation. M a t h . Z . 24 (1926): 789-791.
[Wl] Wilson, R. L. Euclidean Lie algebras are universal central extensions. L i e
A l g e b r a s a n d R e l a t e d T o p i c s . D. J. Winter, ed. Berlin: L e c t u r e N o t e s i n M a t h .
933. Berlin: 1982 Springer-Verlag, pp. 210-213.
[Wi] Winter, D. J. A b s t r a c t L i e A l g e b r a s . Cambridge, MA: MIT Press, 1972.
[Wil] Witt, E. Spiegelungsgruppen und Aufzahlungen halbeinfacher Liescher
Ringe, Harris. A b h a n d l 14 (1941): 289-322.
Index

Absolutely irreducible representation, 35 Cartan matrix of an associative algebra, 185


Ad-nilpotent: Cartan or generalized Cartan matrix, 250
element of a Lie algebra, 28 Cartan subalgebra:
locally, 28 of a finite dimensional Lie algebra, 640
Adjoint group, 320 of a Kac-Moody algebra, 635
Adjoint representation, 30 Casimir-Kac operator, 372
Algebra, 1 Casimir operator, 119
associative, 2 Category 142
commutative, 2 Central extension, 50
identity element, 2 Central extension of abelian groups, 289
proper ideal, 3 Centralizer, 84
right, left, two-sided ideal, 3 Character, 130
structure constants, 3 Character map, 152
Associative algebra, 2 Characteristic set, 204
commutative, 13 Cocycle, 290
free, 13 Combinatorically symmetric matrix, 247, 353
modules and representations, 34 Completely reducible representation, 34
Composition series, 149
Base of a root system, 240, 472 equivalent, 149
Base ring, 9 local, 149
BGG duality, 198 proper and extraneous factors, 149
Bilinear form: Connected through (weights), 513
alternating, 291 Connecting homomorphism, 78
invariant, 35 Contragredient:
locally invariant, 355 bilinear form, 169
Birkhoff decomposition, 523 hermitian module, 169
Borel group, 518, 629 hermitian unity module, 169
Branching function, 665 Contragredient Lie algebra, 310
Bruhat decomposition, 126, 523 adjoin group, 320
Bruhat ordering, 426 Cartan matrix, 324
display, 311
Campbell-Baker-Hausdorff formula, 70 integrable, 323
Cartan matrix, 258 invariant, 359
affine, 258 minimally realized, 350
decomposable and indecomposable, 251 radical free of a pair 340
finite type, 276 rational form, 352
hyperbolic, 485, 482 real and imaginary ro<fts, 317
indefinite type, 276 structure matrix, 312
realization, 330 symmetrizable, 359
symmetrizable, 251 Weyl and dual Weyl group, 313

681
682 Index

Coroot, 321 root and coroot lattice, 303


Covering, 50 simple root, 240
morphism, 50 weight and coweight lattice, 303
universal, 51 Weyl group of, 232
Coxeter-Dynkin diagram, 250 Flag, 603
Coxeter group, 419 Forest, 249
Coxeter matrix, 418 Formal exponential, 130
Coxeter number, 477 Friedrichs’ theorem, 68
Coxeter transformation, 477 Fundamental chamber, 438
Crystallographic, 425 Fundamental region, 444

Generalized contragredient Lie algebra, 376


Demazure-Tits group, 302
Generalized eigenspace, 600
Denominator formula, 533
Generating function, 130
Depth function, 140, 510
Geometric lattice, 217
Derivation, 23
automorphism, 217
inner of a Lie algebra, 27
definite and indefinite, 218
inner of an associative algebra, 24
determinant, 218
Derived Lie group of a Kac-Moody Lie
dual, 219
algebra, 488
even and odd, 218
Display, 311
index of duality, 220
isomorphism, 217
Edge (of a support), 615 of type A , 223
Eigenspace, 599 of type D , 222
Elementary automorphism, 28 of type E , 226
Endomorphism: orthogonal sum of, 221
diagonalizable, 588 rationalization, 217
Jordan-Chevalley decomposition, 594, 597 realization, 217
locally finite, 591 signature, 218
minimal polynomial, 587 sublattice and full sublattice, 218
primary decomposition, 588 type I and II, 218
semisimple, 588 unimodular, 220
strictly triangularizable, 602 Graded:
triangularizable, 602 homomorphism, 17
Exchange condition, 424 ideal, 18
Ext, 72 module, 15
Extension of modules, 71 subspace, 18
Exterior algebra, 14 Gradings, 15
Graph, 248
Fan, 104, 130 incidence matrix, 248
source of, 130 Group algebra, 130
Finite root system, 232
ambient space, 232 Heisenberg algebra, 23
base, 240 Highest root, 270
Cartan matrix, 246 Highest weight, 103
coroot of a root, 303 module and representation, 103
decomposable and indecomposable, 239 series, 150
dual, 303 weight vector, 103
fundamental weights, 303 Hoistable representation of a Lie algebra, 604
highest root, 303 Homogeneous element (of a Kac-Moody
highest short root, 303 group), 528
isomorphism of, 233 Homomorphism:
natural bilinear form, 256 graded, 17
positive system of roots, 242 homogeneous, 16
reduced and nonreduced, 237 Hyperbolic plane, 224
Index 683

Ideal, 3 Locally invariant bilinear form, 355


residually nilpotent, 165 Locally nilpotent endomorphism, 26
Imaginary root, 317 exponential of, 26
Imaginary roots, of a set of root data, 464 Lowest weight module, 106
Indecomposable blocks of a matrix, 352
Index of connection, 271 Minimal regular weight, 367
Inferior matrix, 278 Module, see also Representation of a Lie
Integrable ^ 12-module, 124 algebra
Integrable module, 482 character of, 132
Integrable representation, 482 Multiplicity of a module, 150
Invariant bilinear form:
Normalizer, 84
proper, 359
Null root, 258
standard, 365
Isotypical component, 122 Ore domain, 85
Iwahori subgroup, 646
Partition, 135
Jordan-Chevalley decomposition, 594 Pentagonal number, 662
Perron-Frobenius eigenvalue, 272
Kac-Moody Lie algebra, 324 Poincare series, 128
finite, affine, and indefinite type, 324 Primitive matrbc, 272
Killing form, 37 Primitive vector, 119
radical of, 36 Principal grading, 648
Kostant cone, 612 Principal subalgebra, 574
Kostant partition function, 139
Quadratic form associated with a geometric
Lattice, see Geometric lattice lattice, 217
Laurent polynomial, 647 Quadratic form over field of two elements, 290
Lie algebra, 2 Quasi-roots, 549
abelian, 2 Quaternions, 7
algebraic, 630
antiautomorphism, 3 Racah formula, 541
bracket, 2 Radical of a Lie algebra with triangular
central simple, 83 decomposition, 162
centralizer, 3 Rational form or structure of a vector space,
centrally closed, 56 245
centre, 3 Real root, 317
centroid, 83 Realization, 330
derived, 4 dimension of, 330
free on a set, 61 dual, 333
free product of, 64 natural, 339, 343
given by generators, 64 universal algebra, 334
holomorph of, 82 Realization of a root system, 457
of an associative algebra, 5 Reflection, 230
perfect, 51 orthogonal, 232
presentation, 64 Representation of a Lie algebra, 28
quotient, 4 action, 29
semisimple, 5 annihilator, 29
simple, 4 as a module, 29
solvable, 89 character, 132
special orthogonal, 81 exterior representation, 32
special unitary, 82 faithful, 29
subalgebra, 3 simple or irreducible, 29
symplectic, 82 tensor product, 31
Lie group of a Kac-Moody Lie algebra, 502 tensor representation, 31
Lie ring, 280 trivial, nontrivial, 29
684 Index

Representation ring, 152 Tensor algebra, 13


Restricted dual, 17 Theorem:
Restricted dual in category 158 B ernstein-G erfand-G erfand (BGG), 556
Restricted weight lattice, 491 Chevalley, Peterson-Kac conjugacy of
Root: Cartan subalgebras, 635
imaginary, 317 conjugacy of bases of a finite root system,
indivisible, 496 244
long, short, 270 conjugacy of bases for a root data, 473
real, 317 Gabber-Kac, 377
strictly imaginary, 481 Jordan-Chevalley, 605, 636
support, 464 Kac imaginary root, 468
Root data, see also Finite root system, 396 Kac-Peterson complete reducibility, 543
base, 472 Poincare-Birkhoff-W itt, 41
chambers, 440 Shapovalov determinant formula, 545
coroots, 397 Weyl-Kac character formula, 532
diagram automorphism, 474 Weyl-Macdonald denominator formula, 533
dual, 397 Theta function, 667
faces, 441 Tits system, 521
facettes, 440 Tree, 249
fundamental chamber, 438 Triangular decomposition, 95
height, 404 diagonal subalgebra of, 96
imaginary roots, 464 extension of base ring, 99
morphisms, isomorphisms, and hermitian, 100
automorphisms, 431 opposite, 96
natural expression and height of roots, 404 regular, 95
obtained by extension of the base field, 434 root lattice of, 96
relative chambers, 457 root system of, 96
roots, 397 Triangular pair, 161
standard of a realization, 397 Triangularizable representation of a Lie
subroot system, 435 algebra, 603
Tits cone, 440 Twisted group algebra, 296
universal covering, 400
walls, 441 Unimodal sequence, 509
Weyl group, 408 Unitarizable (highest weight module), 534
Root string, 327, 464 Universal algebra of a realization, 334
Universal enveloping algebra, 38
Saturated set, 514 graded algebra of, 49
Schur’s lemma, 33 grading inherited from g, 48
Semiprimitive matrix, 272
Semisimple representation, 34 Vacuum vector, 103
Separating group of characters, 499 Verma composition series, 155
Serre’s theorem, 375 Verma module, 107
Shapovalov form, 169 Vertex (of a support), 615
Skeletal graph, 352 Virasoro algebra, 20
Skew invariant, 532
Special linear group, 122 Weight, 92, 367, 491
Split simple and semisimple finite dimensional dominant, 367, 491
Lie algebras, 354 fundamental, 491
Standard invariant bilinear form, 365 integral, 367, 491
Strange formula, 575 minimal regular, 367
String function, 665 Weight lattice, 491
Strong exchange condition, 424 Weight space, 92
Supplement of a module, 34 Weight space decomposition, 92
Symmetric algebra, 13 Weight string, 508
Symmetrizable matrix, 360 Weight vector, 91
Index 685

Weyl group, of a finite root system, 232 WIP, 401


Weyl group, of a set of root data, 408 Witt algebra, 20
length of a word, 408 WUP, 401
parabolic subgroups, 415
reduced words, 408
words, 408 Z-lemma, 428
'V '

■■-J --,
"' ’ •;/-^> Ï%' '
'¿’»V/.,-
' '■ ' :-t Ï-
' j) Í' Ч '’'
h.

~ -,'aÍ' ■•■.:. ^V'." I

íA á M ’ ^'■•*

!i¿6ir^- ' >,v *


>hifl-‘^.-
- i.í *^ 1
. .u ,
_ -

- 4 ííf. '■^'i'^'' C . '^ ; '

f ;*^Vv

f. ‘
4-¿s , | í 4 '- ': í ^ .ís - л ‘
••

¿r-'-i. •^*':flit;-(i;’' . ' ■ .'v '- , 4 . ’’ ' '


ь - ц , . , , , . .................., .

. 1 *

ж Е ''- ; " Ж " ' •

You might also like