You are on page 1of 212

An Introduction

to Symplectic
Geometry

Rolf Berndt

Graduate Studies
in Mathematics
Volume 26

American Mathematical Society


Selected Titles in This Series
26 Rolf Berndt, An introduction to symplectic geometry, 2001
25 Thomas Friedrich, Dirac operators in Riemannian geometry, 20(10
24 Helmut Koch, Number theory: Algebraic numbers and functions, 2000
23 Alberto Candel and Lawrence Conlon, Foliations 1, 2000
22 Gfinter R. Krause and Thomas H. Lenagan, Growth of algebras and Gelfand-Kirillov
dimension, 2000
21 John B. Conway, A course in operator theory, 2000
20 Robert E. Gompf and Andrds I. Stipsicz, 4-manifolds and Kirby calculus, 1999
19 Lawrence C. Evans, Partial differential equations, 1998
18 Winfried Just and Martin Weese, Discovering modern set theory. 11: Set-theoretic
tools for every mathematician, 1997
17 Henryk Iwaniec, Topics in classical automorphic forms. 1997
16 Richard V. Kadison and John R. Ringrose, Fundamentals of the theory of operator
algebras. Volume II: Advanced theory, 1997
15 Richard V. Kadison and John R. Ringrose, Fundamentals of the theory of operator
algebras. Volume 1: Elementary theory, 1997
14 Elliott H. Lieb and Michael Loss, Analysis, 1997
13 Paul C. Shields, The ergodic theory of discrete sample paths, 1996
12 N. V. Krylov, Lectures on elliptic and parabolic equations in Holder spaces, 1996
11 Jacques Dixmier, Enveloping algebras, 1996 Printing
10 Barry Simon, Representations of finite and compact groups, 1996
9 Dino Lorenzini, An invitation to arithmetic geometry, 1996
8 Winfried Just and Martin Weese, Discovering modern set theory. I: The basics. 1996
7 Gerald J. Janusz, Algebraic number fields, second edition, 1996
6 Jens Carsten Jantzen, Lectures on quantum groups, 1996
5 Rick Miranda, Algebraic curves and Riemann surfaces, 1995
4 Russell A. Gordon, The integrals of Lebesgue, Denjoy, Perron, and Henstock, 1994
3 William W. Adams and Philippe Loustaunau, An introduction to Grobner bases,
1994
2 Jack Graver, Brigitte Servatius, and Herman Servatius, Combinatorial rigidity,
1993
1 Ethan Akin, The general topology of dynamical systems, 1993
An Introduction
to Symplectic
Geometry
An Introduction
to Symplectic
Geometry

Rolf Berndt
Translated by
Michael Klucznik

Graduate Studies
in Mathematics
Volume 26

American Mathematical Society


0 ""1
Providence, Rhode Island
Editorial Board
James Humphreys (Chair)
David Saltman David Sattinger Ronald Stern

2000 Mathematics Subject Classification. Primary 53C15, 53Dxx, 20G20, 81S10.

Originally published in the German language by Ftiedr. Vieweg & Sohn Verlagsge-
sellschaft mbH, D-65189 Wiesbaden, Germany, as "Rolf Berndt: Einfnhruug in die Sym-
plektische Geometrie. 1. Auflage (1st edition)" © by Ftiedr. Vieweg & Sohn Verlagsge-
sellschaft mbH, Braunschweig/Wiesbaden, 1998.
Translated from the German by Michael Klucznik
ABSTRACT. The notions of symplectic form, symplectic manifold and symplectic group appear
in many different contexts in analysis, geometry, function theory and dynamical systems. This
book assembles tools from different mathematical regions necessary to define these notions and to
introduce their application. Among the topics treated here are
symplectic and Kiihler vector spaces,
the symplectic group and Siegel's half space,
symplectic and contact manifolds, the theorem of Darboux,
methods of constructing symplectic manifolds: Kiihler manifolds, coadjoint orbits and sym-
plectic reduction,
Hamiltonian systems,
the moment map,
and a glimpse into geometric quantization (in particular the theorem of Groenewold and
van Hove) leading to some rudiments of the representation theory of the Heisenberg and
the Jacobi group.
The goal of the book is to provide an entrance into a fascinating area linking several mathematical
disciplines and parts of theoretical physics.

Library of Congress Cataloging-In-Publication Data


Berndt, Rolf.
]Einfiihrung in die symplektische Geometric. English]
An introduction to sympletic geometry / Rolf Berndt ; translated by Michael Klucznik.
p. cm. - (Graduate studies in mathematics, ISSN 1065-7339 ; v. 26)
Includes bibliographical references and index.
ISBN 0-8218-2056-7 (alk. paper)
1. Symplectic manifolds. 2. Geometry, Differential I. Title. II. Series.
QA649.B47 2000
516.3'6- dc2l 00-033139

Copying and reprinting. Individual readers of this publication, and nonprofit libraries
acting for them, are permitted to make fair use of the material, such as to copy a chapter for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Requests for such
permission should be addressed to the Assistant to the Publisher, American Mathematical Society,
P. 0. Box 6248, Providence, Rhode Island 02940-6248. Requests can also be made by e-mail to

2001 by the American Mathematical Society. All rights reserved.


The American Mathematical Society retains all rights
except those granted to the United States Government.
Printed in the United States of America.
® The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at URL: http://wv.ame.org/
10987654321 060504030201
Contents

Preface ix

Chapter 0. Some Aspects of Theoretical Mechanics


§0.1. The Lagrange equations
§0.2. Hamilton's equations
§0.3. The Hamilton-Jacobi equation
§0.4. A symplectic interpretation
§0.5. Hamilton's equations via the Poisson bracket
§0.6. Towards quantization

Chapter 1. Symplectic Algebra


§1.1. Symplectic vector spaces
§1.2. Symplectic morphisms and symplectic groups
§1.3. Subspaces of symplectic vector spaces
§1.4. Complex structures of real symplectic spaces

Chapter 2. Symplectic Manifolds


§2.1. Symplectic manifolds and their morphisms
§2.2. Darboux's theorem
§2.3. The cotangent bundle
§2.4. Kiihler manifolds
§2.5. Coadjoint orbits
§2.6. Complex projective space
§2.7. Symplectic invariants (a quick view)

vii
viii Contents

Chapter 3. Hamiltonian Vector Fields and the Poisson Bracket 71


§3.1. Preliminaries 71
§3.2. Hamiltonian systems 74
§3.3. Poisson brackets 79
§3.4. Contact manifolds 85

Chapter 4. The Moment Map 93


§4.1. Definitions 93
§4.2. Constructions and examples 97
§4.3. Reduction of phase spaces by the consideration of symmetry 104

Chapter 5. Quantization 111


§5.1. Homogeneous quadratic polynomials and 912 111
§5.2. Polynomials of degree 1 and the Heisenberg group 114
§5.3. Polynomials of degree 2 and the Jacobi group 120
§5.4. The Groenewold-van Hove theorem 124
§5.5. Towards the general case 128

Appendix A. Differentiable Manifolds and Vector Bundles 135


§A.1. Differentiable manifolds and their tangent spaces 135
§A.2. Vector bundles and their sections 144
§A.3. The tangent and the cotangent bundles 146
§A.4. Tensors and differential forms 150
§A.5. Connections 158

Appendix B. Lie Groups and Lie Algebras 163


§B.1. Lie algebras and vector fields 163
§B.2. Lie groups and invariant vector fields 165
§B.3. One-parameter subgroups and the exponent map 167

Appendix C. A Little Cohomology Theory 171


§C.1. Cohomology of groups 171
§C.2. Cohomology of Lie algebras 173
§C.3. Cohomology of manifolds 174

Appendix D. Representations of Groups 177


§D.1. Linear representations 177
§D.2. Continuous and unitary representations 179
§D.3. On the construction of representations 180
Contents ix

Bibliography 185

Index 189

Symbols 193
Preface

I.e caractere propre des mcthodes de ''Analyse et de


Is Gt ometrie modernes consiste daps 1'emploi d'un
petit nombre de principes genttraux. independants
de Is situation respective des differentes parties on
des valeurs relatives des differents symboles; et les
consequences sont d'autant plus tttendues que les
principes eux-memes ont plus de generalite.
from G. DARBOUX: Prancipes de GEometrie Analy-
tique

This text is written for the graduate student who has previous training
in analysis and linear algebra, as for instance S. Fang's Analysis I and Lin-
ear Algebra. It is meant as an introduction to what is today an intensive
area of research linking several disciplines of mathematics and physics in
the sense of the Greek word ouµrrMecet.v (which means to interconnect, or
to interrelate in English).' The difficulty (but also the fascination) of the
area is the wide variety of mathematical machinery required. In order to
introduce this interrelation, this text includes extensive appendices which
include definitions and developments not usually covered in the basic train-
ing of students but which lay the groundwork for the specific constructions

11 want to thank P. Slodowy for pointing out to me that the name symptectic group, which
eventually gave rise to the term sympkche geometry, was proposed by If. WEYL, [W], 1938, in
his book, The Classical Groups (see footnote on p. 165). The symplectic group was also called
the comples group or an Abelion linear group, this last to honor ABEL, who was the first to study
them.

xi
xii Preface

needed in symplectic geometry. Furthermore, more advanced topics will


continue to rely heavily on other disciplines, in particular on results from
the study of differential equations.
Specifically, the text tries to reach the following two goals:
To present the idea of the formalism of symplectic forms, to in-
troduce the symplectic group, and especially to describe the sym-
plectic manifolds. This will be accompanied by the presentation of
many examples of how they come to arise; in particular the quotient
manifolds of group actions will be described,
and

To demonstrate the connections and interworking between math-


ematical objects and the formalism of theoretical mechanics; in
particular, the Hamiltonian formalism, and that of the quantum
formalism, namely the process of quantization.
The pursuit of these goals proceeds according to the following plan. We
begin in Chapter 0 with a brief introduction of a few topics from theoretical
mechanics needed later in the text. The material of this chapter will already
be familiar to physics students; however, for the majority of mathematics
students, who have not learned the connections of their subject to physics,
this material will perhaps be new.
We are constrained, in the first chapter, to consider symplectic (and a
little later Kdhler) vector spaces. This is followed by the introduction of the
associated notion of a symplectic group Sp(V) along with its generation. We
continue with the introduction of several specific and theoretically important
subspaces, the isotropic, coisotropic and Lagrangian subspaces, as well as
the hyperbolic planes and spaces and the radical of a symplectic space.
Our first result will be to show that the symplectic subspaces of a given
dimension and rank are fixed up to symplectic isomorphism. A consequence
is then that the Lagrangian subspaces form a homogeneous space £(V) for
the action of the group Sp(V). The greatest effort will be devoted to the
description of the spaces of positive complex structures compatible with the
given symplectic structure. The second major result will be that this space
is a homogeneous space, and is, for dim V = 2n, isomorphic to the Siegel
half space S) = Spn(R)/U(n).
The second chapter is dedicated to the central object of the book, namely
symplectic manifolds. Here the consideration of differential forms is unavoid-
able. In Appendix A their calculus will be given. The first result of this
chapter is then the derivation of a theorem by Darboux that says that the
symplectic manifolds are all locally equivalent. This is in sharp contrast
to the situation with Riemannian manifolds, whose definition is otherwise
Preface xiii

somewhat parallel to that of the symplectic manifolds. The chapter will


then take a glance at new research by considering the assignment of invari-
ants to symplectic manifolds; in particular. the symplectic capacities and
the pseudoholomorphic curves will be given.
In the course of the second chapter. we will present several examples of
symplectic manifolds:
First, the example which forms the origin of the theory and remains
the primary application to physics is the cotangent bundle T'Q of
a given manifold Q.
Second. the general Kahler manifold.
Third, the coadjoint orbits. This description of symplectic mani-
folds with the operation of a Lie group G can be taken as the sec-
ond major result of this chapter. We describe a theorem of Kost.ant
and Souriau that says that for a given Lie group G with Lie alge-
bra g satisfying the condition that the first two cohomology groups
vanish, that is H'(g) = H2(9) = 0. there is. up to covering. a
one-to-one correspondence between the symplectic manifolds with
transitive G-action and the G-orbits in the dual space g' of g.
Here we will need several facts from the theory of Lie algebras and
systems of differential equations. and we will at least, cover some of
the rudiments we require. This will then offer yet another means
for introducing one of the central concepts of the field. namely the
moment map. This will, however. be somewhat postponed so that
In the fourth and last example. complex projective space can be
presented as a symplectic manifold: this will be seen as a specific
example of the third example. as well as, the second: that is. as a
coadjoint orbit as well as as a Kahler manifold.

As preparation for the higher level construction of symplectic manifolds.


Chapter 3 will introduce the standard concepts of a Hamiltonian vector
field and a Poisson bracket. With the aid of these ideas, we can give the
Hamiltonian formulation of classical mechanics and establish the following
fundamental short exact sequence:
0 -. R - .F(M) Ham Af -0,
where.F(M) is the space of smooth functions f defined on the symplectic
manifold and given the structure of Lie algebra via the Poisson bracket, and
Ham M is the Lie algebra of Hamiltonian vector fields on the manifold.
The third chapter continues with a brief introduction to contact man-
ifolds. A theory for these manifolds in odd dimension can be developed
xiv Preface

which corresponds precisely to that of the symplectic manifolds. On the


other hand, both may be viewed as pre-symplectic manifolds. Here the
connection will be given through the example of a contact manifold as the
surface of constant energy of a Hamiltonian system.
The fourth and fifth chapters will be a mix of further mathematical
constructions and their physical interpretations. This will begin with the
description of the moment map attached to the situation of a Lie group G
acting symplectically on a symplectic manifold such that every Hamiltonian
vector field is global Hamiltonian. This is a certain function
4i:M-+g", g=LieG.
The most important examples of the moment maps are the Ad`_equivariant
ones, that is, those that satisfy a compatibility condition with respect to the
coadjoint representation Ad. The first result of Chapter 4 is that for a sym-
plectic form w = -dt9 and a G invariant 1-form t9 such an Ad`-equivariant
map can be constructed. This will then be applied to the cotangent bundle
T"Q, as well as to the tangent bundle TQ, where it will turn out that for
a regular Lagrangian function L E F(Q) the associated moment map is an
integral for the Lagrangian equation associated to L. As examples, we will
discuss the linear and angular momenta in the context of the formalism of
the moment map, and so make clear the reason for this choice of terminology.
Next, we describe symplectic reduction. Here, we are given a symplectic
C-operation on M and an Ad'-equivariant moment map 4i; under some
relatively easy-to-check conditions, for p E g', the quotient
Mµ = i-1(µ)lG,
is again a symplectic manifold. This central result of Chapter 4 has many
applications, including the construction of further examples of symplectic
manifolds (in particular, we obtain other proofs that the projective space
PI(C) as well as the coadjoint orbits are symplectic). Another application is
the result of classical mechanics on the reduction of the number of variables
by the application of symmetry, leading to the appearance of some integrals
of the motion.
In the fifth and last chapter, we consider quantization; that is, the tran-
sition from classical mechanics to quantum mechanics, which leads to many
interesting mathematical questions. The first case to be considered is the
simplest: M = R2" = 7'R^. In this case the important tools are the groups
SL2(R), the Heisenberg group Heis2n(R), the Jacobi group Ga (R)
(as a semidirect product of the Heisenberg and symplectic groups) and their
associated Lie algebras. It will follow that quantization assigns to the poly-
nomials of degree less than or equal to 2 in the variables p and q of R2.,
an operator on L2(R) with the help of the Schrodinger representation of
Preface xv

the Heisenberg group and the Weil representation of the symplectic group
(more precisely, its metaplectic covering). The theorem of Groenewold and
van Hove then says that this quantization is maximal; that is, it cannot be
extended to polynomials of higher degree.
The remainder of the fifth chapter consists in laying the groundwork
for the general situation, which essentially follows KIRILLOV [Ki]. Here a
subalgebra p, the primary quantities, comes into play, which for the case
of M = T*Q turns out to be the arbitrary functions in q and the linear
functions in p. Here yet more functional analysis and topology are required
in order to demonstrate the result of Kirillov that for a symplectic manifold,
with an algebra p in F(M) of primary quantities relative to the Poisson
bracket, a quantization is possible. That is, there is a map which assigns
to each f E p a self-adjoint operator f on Hilbert space N satisfying the
conditions
(1) the function 1 corresponds to the identity idN,

(2) the Poisson bracket of the two functions corresponds to the Lie
bracket of operators, and

(3) the algebra of operators operates irreducibly.


There is a one-to-one correspondence between the set of equivalence classes
of such representations of p and the cohomology group HI(M,C').
In the first two appendices, manifolds, vector bundles, Lie groups and
algebras, vector fields, tensors, differential forms and their basic handling
are covered. In particular, the various derivation processes are covered so
that one may follow the proofs in the cited literature. A quick reading of
this synopsis is perhaps recommended as an entrance to the second chapter.
In Chapter 2 some material about cohomology groups will also be required.
The third appendix presents some of the rudiments of cohomology theory.
In the final appendix, the central concept of coadjoint orbits is prepared by
a consideration of the fundamental concepts and constructions of represen-
tation theory.
As already mentioned, somewhat more from the theory of differential
equations than is usually presented in a beginner's course on the topic, in
particular Frobenius' theorem, is required to fully follow the treatment of
symplectic geometry given here. Since in these cases the difficulty is not
in grasping the statements, this material is left out of the appendices and
simply used in the text as needed, though again without proof.
It is not the intention of this text to compete with the treatment of the
classical and current literature over the research in the various subtopics
of symplectic geometry as can be found, for example, in the books by
xvi Preface

ABRAHAM-MARSDEN [AM], AEBISCHER of al. [Ae], GUILLEMJN-STERN-


BERG [GS], HOFER-ZEHNDER [HZ], MCDUFF-SALAMON [MS], SIEGEL
[Sill, SOURIAU [so], VAISMAN [V], WALLACH [W] and WOODHOUSE [WO].
Instead we have tried to introduce the reader to the material in these sources
and, moreover, to follow the work contained in, for example, GROMOV [Gr]
and KIRILLOV [Ki]. In the hope that this will provide each reader with a
starting point into this fascinating area a few parts of chapters 1, 2, and 4
may be skipped by those whose interests lie in physics, and one may begin
directly with the sections on Hamiltonian vectorfields, moment maps and
quantization.
This text is, with minor changes, a translation of the book "Einfiihrung
in die Symplektische Geometrie" (Vieweg, 1998). The production of this text
has only been possible through the help of many. U. Schmickler-Hirzebruch
and G. Fischer, on the staff of Vieweg-Verlag, have made many valuable
suggestions, as has E. Dunne from the American Mathematical Society.
My colleagues J. Michalilek, 0. Riemenschneider and P. Slodowy, from the
Mathematische Seminar of the Universitiit Hamburg, were always, as ever,
willing to discuss these topics. A. G6nther prepared one draft of this text,
and I. Kowing did a newer draft and also showed great patience for my eter-
nal desire to have something or other changed. I also had very successful
technical consultation with F. Berndt, D. Nitschke and R. Schmidt. The last
of these went through the German text with great attention and smoothed
out at least some of what was rough in the text. I would also thank T.
Wurzbacher, W. Foerg-Rob and P. Wagner for carefully reading (parts of)
the German text and finding some misprints, wrong signs and other mis-
takes. The translation was done by M. Klucznik, who had an enormous task
in producing very fluent English (at least in my opinion) and a fine layout
of my often rather involved German style. It is a great joy for me to thank
each of these.

R. Berndt
Chapter 0

Some Aspects of
Theoretical Mechanics

Symplectic structures arise in a natural way in theoretical mechanics, in


particular during the process of quantizat.ion. that is. in the passage from
classical to quantum mechanics. In order to motivate the study of symplectic
geometry. we will begin with a rough sketch of the relevant physics. although
we will not cover all the concepts of this field nor give all of the relevant
definitions. As references. one may consult the first chapter of VAISAIA` [V].
A more complete description of the principles of classical mechanics can be
found in WOODHOUSE [Wo]. in the third chapter of ARNOLD [A] and in the
third and fifth chapters of ABRAHAM-XIARSDEN [AM]. A further highly
recommendable classical source is the first chapter of SIEGEL-MIOSER [SM].
For the process of quantization. we refer the reader to §15.4 of KIRILLOV
[Ki]. It is the goal of this text to later return and cover the topics of this
chapter in greater detail.

0.1. The Lagrange equations


The purpose of theoretical mechanics is the discovery of principles which
allow one to describe the time development of the state of a physical system.
In classical mechanics such a state is given as a point P on an n dimensional
real manifold Q (see Section A.1). Q is called the configuration space. and
P is described b y the local coordinates q1, ... , q.. called position variables.
The time development of the system is then described by the curve 7 = P(t),

t .-- P(t) with P(to) = P°.

I
2 0. Some Aspects of Theoretical Mechanics

or in the local coordinates as


t'-'qi(t) with gi(to)=q0, i=1,...,n.
Here physical principles must be found which allow one to give the curve as a
solution to a differential equation. The starting point for this determination
is the classical mechanical principle of least action. For this it is assumed
that the system has a Lagrange function L of the form
L = L(q, 4, t),
which is gotten as the difference of the kinetic and the potential energies,
L=Eye - E,
which is also written as
L=T-V
The principle of least action now says that the change in the system
proceeds so that the curve -y minimizes the path integral
eJLdt.
l

to

The variational calculus now says (am Coua4w -HILBERT [CII], p. 170)
that for the minimum curve y = q(t) the system satisfies the Euler-Lagrange
equations
d 8L 8L
(1)
dt 8qi 8qi 0,

This can be seen as a system of ordinary differential equations in a


2n-dimensional space TQ with local coordinates
qn,
(which can be understood as the tangent bundle over the configuration space
Q (see Section A.3)). The desired curve y on Q is the projection of the
solution curve ry of (1) onto TQ.

0.2. Hamilton's equations


Classical mechanics now takes the following formulation: for a given La-
grange function L the coordinates position and velocity, (q, 4), are replaced
by the coordinates position and momentum (q, p) made possible by the
transformation
8L
pi = aqt, i = 1,...,n.
0.2. Hamilton's equations 3

The basis of this concept is the Legendre transformation (see ARNOLD [A].
p. 61 f.) between tangent and cotangent bundles (see Section A.3)
TQ -* T'Q,
(q, 4) '--' (q, p)
Then the time development described on TQ by the Lagrange function L =
L(q, 4, t) (which we can and will assume to be convex in the second argument:
see, for example. ARNOLD ([A], p. 65)) is replaced by the Hamiltonian
function H on phase space T'Q defined by

H(P, q, t) :=P4-L(q,4,t) with p= aQ


where we have used the usual abbreviated symbols for the n-tuple
8L 8L 8L
etc.
aq = \aql ..., 57q-
P=(Pi,...,P-),

The Lagrange equations (1) are here translated into Hamilton's equations
8H 8H
(2) q = , p = - .

ap aq

Because the total differential of H = H(p, q, t) (see Section A.4) gives

dH = dp + aq dq + dt

and by the definition H = p4 - L(q, 4. t). we also get

dH = 4dp - -q dq - dt.
8L
Comparing()
1 and p = a4 , we get

OH OH
q=-;, OL
aq = - aq
-p. W=--
OH 8L
8t
.

Hamilton's equations (2) are now (when H is independent of t) a system


of ordinary differential equations, which, given a particular set of initial con-
ditions p0, q', gives a unique curve ry' in phase space T*Q whose projection
ry onto the configuration space Q solves the original problem.
The Hamiltonian function is also written in the form
H=H(p,q,t)=(T)+V,
where V is the potential energy of the system and T is the kinetic energy
given in terms of the variables q and p.
4 0. Some Aspects of Theoretical Mechanics

0.3. The Hamilton-Jacobi equation


Yet another formulation of the problem passes from the solution of a system
of ordinary differential equations to the solution of a partial differential
equation. The resulting partial differential equation is the Hamilton-Jacobi
equation

(3) H(q, 9S,t) + =0


5
for the action function S. Here, giving a solution which is dependent on t,
the n variables q, and the n initial parameters a,
S = S(q, t, a),
is equivalent to giving a solution q = q(t), p = p(t) of (2). Here we present
only the following consideration:

Let S = S(q, t, a) be a solution of (3) with


(a82S
det k) # 0.
Then the n equations
as
Sae
= be, I= 1,...,n,
in the q; are solvable in the q, = cpi (t, a, b) , i = 1, ... , n. This allows one to
write
PC= as
aqe
as a function of t, a. and b:
pt = 't(t, a, b).
These qi, p, satisfy Hamilton's equations (2), since differentiating

(+) H I q, Fq s (q, t, a), t) + =0

with respect to at gives


aH 02S + 025
= 0, t =...,
1, n.
apk Sae aqk 5a-1 -5i
k

And differentiating as = be with respect to t gives

Ys a2S =
qk + 0.
aqk sac at aac
0.3. The Hamilton-Jacobi equation 5

Taking the difference of the two equations yields

a2S OH
Oak Oat
- qk I = 0, t=1....,n,
OPk

and since
as
det { )oo
ag Oa

we arrive at half of Hamilton's equations. Next, differentiating (+) with


respect to qt gives
OH OH 02S 02S
o
Oqe + k OPk Oqk aqt + Oqe Ot =

and then, differentiating pe =


as with respect to t, we get
aqt
025 O2S
of = aqk aqe
qk +
at aqt

Taking the difference of these two equations yields, considering the satisfac-
tion of the relation = q,
W OH
Pe=-aqt
There is yet another way (which at first glance looks different) to derive
the Hamilton-Jacobi equation (see ARNIOLD [A], pp. 253-5). Here the path
integral

SgO,tt (q, t) = Ldt


J
ti

is taken along the curve y from (qe. te) to (q. t) that minimizes the integral,
and it is shown that
dS = pdq - Hdt.
Then it is immediately clear that for S the equations
as _ and
q, t) as = P
9

hold, and therefore also (3).


6 0. Some Aspects of Theoretical Mechanics

0.4. A symplectic interpretation


Here we continue from Section 0.2. The Hamiltonian function H defines
a Hamiltonian vector field XH on phase space Q. Relative to the usual
coordinates (q, p), this is defined by (see Section A.4)
8H 8 8H 8
XH
"PiNi -
:

NiBpi
Given a vector field X, the question immediately arises as to the exis-
tence of integral curves 'y; that is, curves whose tangent vectors ^ (t) at every
point of -y(t) are equal to the given vector of the vector field at that point,
which, in symbols then, is ry' (t) = XH ('y(t)). For
7(t) = (q(t), p(t))
this condition leads to Hamilton's equations (2)
8H . 8H
8p = q, 8q = -p-
With the help of a little bit more from the theory of differential forms (see
Section A.4), this can be reformulated as follows: there exist a 2-form
w = > dqi n dpi E 02 on T'Q
i
and an inner product i such that, from any vector field X, the 2-form w
gives us a 1-form i(X)w. Then Hamilton's equations (2) are equivalent to
(4) i(XH)w = d. H.

0.5. Hamilton's equations via the Poisson bracket


On any pair of (arbitrarily often) differentiable functions f, g on phase space
T'Q we may take their Poisson bracket { , }, 1 defined by the equation
r of 8g
{f,9}8qi Of
8g
Bpi 8pi8gi
This Poisson bracket endows the space of functions.F(T'Q) with the struc-
ture of a Lie algebra (see Section B.1). This will be discussed in more detail
later. Here we only note that Hamilton's equations (2) can be written with
the help of the Poisson bracket as
(5) ¢ = {q, H}, p = {p, H).
l Warning- in the literature (for example, in [KI[) one sometimes takes for if, g} the negative
of what is taken here)
0.6. Towards quantization

From this it does not take too much work to see that, generally, for the time
development of an observable given by f. the above system must satisfy the
condition
(5') f = (f, H).
0.6. Towards quantization
The term quantization will indicate the process by which a corresponding
quantum system is constructed from a given classical system. Thus we must
find a transition from the point in phase space T'Q which describes the state
of a classical system to an element v (more accurately. a 1-dimensional sub-
space vC) of a complex Hilbert space f with the aid of the probability
distribution for the state of a quantum mechanical system. This transition
should give, for the Hamilton function H and the classical observables f.
corresponding self-adjoint operators H and f in fl for the quantum situa-
tion. Thus one would expect that the equation (5')
f={f.H}.
giving the time course in T'Q in the classical case should pass to an analo-
gous equation on the operators
f =c[f,H]j
where [ , J is here the natural Lie bracket,
[A.B] =AB-BA.
where c = - 2* (c is. up to the factor i, a factor ensuring the symmetry of
the operators, a constant from physics, and h is called Planck's constant).
We will later examine for which f E .F(T*M) one may define a mapping
f i-, f whose images are self-adjoint operators and which satisfies
1'--.1=id7
and
{f1f } = c [fl, f2] .

With this process considerable problems will appear, but we will finally
see that, at least for the so-called primary quantities (that is. f either a
polynomial of degree 2 in q and p or a linear function of the pl, ... , p and
an arbitrary function of the q , . . . . , solutions can be found.
Chapter 1

Symplectic Algebra

The soon to be introduced symplectic manifolds can be thought of locally as


symplectic vector spaces. It is therefore necessary to define and study vector
spaces with additional structure. This additional structure is given by
a) a scalar product,
b) a symplectic form,
c) a complex structure.
To begin, we rework some old and trusted linear and inultilinear algebra. As
a final result, we will describe all spaces with a given symplectic structure
compatible with a complex structure as a Siegel half space. This space is
important not only for geometry, but also for function theory, although as
to the latter, we will only be able indicate a small part of this importance.
The second chapter of VAISMAN: Symplectic Geometry and Secondary Char-
acteristic Classes [V], as well as the third chapter of ABRAHAM--MARSDEN
(AM] serve as good guides to this chapter. For additional background, E.
ARTIN's Geometric Algebra [Ar) can be recommended.

1.1. Symplectic vector spaces


Now we let K be an arbitrary field of characteristic 0. Later, we will restrict
to K = R. Also let V be a finite-dimensional K vector space (with dim V =
p). Then the basic underlying concept of symplectic space is given by the
following definition.
DEFINITION 1.1. A symplectie form
w:VXV,K
9
10 1. Symplectic Algebra

is an antisymmetric and nondegenerate bilinear form; that is, it satisfies


w(v,v)=0 forallvEV,
and if
w(v,w)=0 forallvEV,
then w = 0. A vector space V is called a symplectic vector space if it is
equipped with a symplectic form.
Remark 1.2.
(1) ABRAHAM-MARSDEN [AMJ consider infinite-dimensional symplec-
tic vector spaces, as well.
(2) In the case of K = R this definition is parallel to the definition of a
Euclidean vector space; that is, to a space having a scalar product, a
symmetric positive definite bilinear form which is usually denoted
bysorby(, ).
Let
e = (el,...,ey)
be a basis of V. Then the bilinear form w on V in terms of a can be given
in matrix form by
wc=(wj)EMp(K) with wj=w(e:,ei)
For K = R there is a nice classification of the normal forms of symmetric
and skew-symmetric bilinear forms.
PROPOSITION 1.3. Let V be a p-dimensional R vector space.
i) In the case that s is a symmetric bilinear form of rank r, then is
a basis a of V relative to which

3g = with ei = ±1, i = 1, ... , r.

ii) In the case that w is an antisymmetric bilinear form of rank r, then


r = 2n and then is a basis e of V relative to which
0 En 0
wr = -E, 0 0 with En E Mn(R) the unit matrix.
0 0 0
1.1. Symplectic vector spaces 11

Proof. i) By a process analogous to the Gram-Schmidt orthonormaliza-


tion, it can be assumed that since s is symmetric, it satisfies the polarization
identity
s(v, w) = (1/4)(s(v+w,v+w) -s(v-w,v-w)).
Thus for s # 0, there is an ei E V with s (ei, ei) # 0. ei can be multiplied
by a scalar, giving el with el := s (el, el) = f1. Let
Vi := Rej and V2 := {v E V, s (e, e1) = 0}.
It is then clear that VlnV2={0}, and so V1+V2=V; thenforvEV,
v - Els (v, el) el E V2.
One may now continue by induction. So long as s 0 on V2 x V21 we may
find an e2 # 0 in V2 with s (e2, e2) = E2 = ±1, and so on.
ii) For w:# 0, there must be e1, e,,+l E V with w (el, e,,+1) # 0. After
multiplying el by a scalar, it can be assumed that w (el, 1. Since w
is skew-symmetric, we have
w (ei, el) = w (en+i, 0,
and the matrix for w' in the plane El spanned by el and is
01 .

0 }
Now let V2 be the w--orthogonal complement of El, so
V2:_{vEVjw(v,vl)=0 forallvlEEl}.
We have El nV2={0}and V=El GV2,and for vEV
v - w(v, el) e,,+1 E V2.
Given that w # 0 on V21 one may repeat the above procedure for V2 and
obtain e2 as well as so that w (e2, 1. Inductively we get the
claimed matrix we. 0
The statement ii) clearly holds as well for more general fields K # R.
Let V` denote the dual space of V and e` the dual basis to e on V',
which satisfies
ei (ei) = (ei, a=) = 8 j.
One of the basic results of (multi-)linear algebra (see also Section A.4) is
that the space A9(V, K) of skew-symmetric q-linear functions from V9 to K
is isomorphic to the qth exterior product A9 V' of V*. AQV ` has as K-basis
the set
e;, A ... A e,*, with it < ... < iq.
12 1. Symplectic Algebra

In particular, an antisymmetric bilinear form w with the matrix wg = (wig)


relative to e can also be written as
w = E wi.7 ei A e!.
iv
Written this way, w works as a function by sending (v, w) E V x V to
w (v, w) = E wi) (e; (v)et (w) - e; (w)ef(v)).
i<j
The statement ii) can be reformulated as
COROLLARY 1.4. By an appropriate choice of basis e, the antisymmetric
bilinear form w can be written as
n
w=Ee; Ae; .

i=1

Such a representation will be called the canonical form of w, and e a sym-


plectic basis of V. Then
n
w (v, u) = Dxiyi - xiyi),
i=1

when for v E V the components relative tog are given by


n n p-2n
v= xiei + E yiei+n + : zie2n+i
i=1 i=1 i=1

Of particular importance for symplectic geometry is naturally the non-


degenerate skew-symmetric bilinear form w; in this case p = r = 2n, so
that V must have even dimension. A criterion for w E A2V' to be non-
degenerate is that the nth exterior power wn = w A ... A w of w must be a
nonzero multiple of the volume form
T=e;A...Ae,EAPV*.
For r, we have
P
r (v1, ... , v.) =det(ail) with vi = E ail ei, i = 1, ... , p.
=1

For w = Eel A en+i, we have, in the usual notation, wn = n!(-1)(n/21r,


where [x] is the function that returns the largest integer < x for all x E R.
In general,
(_1) (n/21
w
n!
defines an orientation on V (see ABRAHAM-MARSDEN [AM], pp. 165-166).
1.1. Symplectic vector spaces 13

A symplectic form w makes possible the identification of


wb : V V.
r wb(v).
with
wb(v)(v') = w (v, v') for v, v' E V.
We will use the letter i for a general inner product
i : V x AqV - Aq-1V'.
(v, d) .-, i(v)d,
where i(v)t7 is the (q - 1)-linear function given by
i (v)99(VI..... Vq-I) = 19 (v. VI...., vq_I),
and so
wb(v) = i(v)w; E V.
It follows easily for w. as in Corollary 1.4. that
i(e3)w = wb(ej) = ej+n j = 1....,n,
i(e1+n)w = wb(ej+n) = -e* J.

Although it is almost trivial, we recommend


EXERCISE 1.5. Given a bilinear form w on V with dine V = m. show
that the following are equivalent.:

a) w is non-degenerate.
b) wb is an isomorphism.

and. in the case that m = 2n.

C) w"= )A...Aw#0.
Although these statements are left unproved, we will not hesitate to use
them later in the text.
In parallel to the situation in Euclidian geometry, we may form the
following definition in symplectic geometry.
DEFINITION 1.6. Two vectors v, w from the symplectic vector space
(L; w) are called w-orthogonal, skew-orthogonal or - when there is no doubt
about which w is intended - simply orthogonal, whenever
w (v, w) = 0.
This is also indicated by v 1 w.
14 1. Symplectic Algebra

EXAMPLE 1.7. Ken with the symplectic form w given by w (v, v') _
(xi2 - xyi) for
v = xlel + ... + xnen + glen+1 + ... +yne2n,
v' = x11e1 +... + y'e2

with the canonical basis e = (el,... , is the standard symplectic space.

Given Proposition 1.3, every 2n-dimensional symplectic space can be de-


scribed this way.
EXAMPLE 1.8. Let W be an n-dimensional K vector space and W* its
dual space. Then V = W ® W* is a symplectic space with
w:VxV -+K
given by

w (t1 + 71, t2 +7'2) = n(t2) - r (ti) for t1i t2 E W, 71,1 E W*.

Remark 1.9. Relative to this last example, note that a symplectic space
V has many decompositions V = W ® W*. Let e be a symplectic basis for
V, such that W is the span of el,... , en. Then W is isomorphic, via wb,
to the subspace spanned by en+1, ... , e2n; so V = W ®W * with the form
defined in the previous example.
EXAMPLE 1.10. There is for every k E N a p = (2n + k)-dimensional
K vector space U with a skew-symmetric bilinear form w: U x U K of
rank 2n. Then the annihilator of w,
Uo:={uEU; j(u,w)=0 for all wEU},
has dimension k, and i induces a symplectic form w on V := U/Uo. We will
say that the symplectic space (V, w) is given by the reduction of (U, w).

1.2. Symplectic morphisms and symplectic groups


Just as in the case of a Euclidian vector space, where the scalar product per-
mits one to define orthogonal morphisms, we have, in symplectic geometry,
a natural definition of morphism:
DEFINITION 1.11. Let (V,wi) be two symplectic vector spaces and 0 :
V1 -+ V2 a linear map. Then we call 0 symplectic whenever
(*) w2 (0 (v), 0 (w)) = w1(v, w) for all v, w E V1.
1.2. Symplectic morphisms and symplectic groups 15

Remark 1.12. A symplectic morphism is necessarily injective, since if


0(v) = 0, then (*) forces v = 0, since w, is non-degenerate. For dim Vi =
dim V2 < oo, 0 must therefore be an isomorphism. Symplectic isomorphisms
will be called symplectomorphisms.

When (VI1 w1) = (V2, w2) = (V, w), any symplectic map 0 must be an
automorphism of (V; w). The collection of symplectic automorphisms forms
a group under the usual composition, called the symplectic group of (V; w)
and denoted Sp(V). In particular, for V = Ken with the standard form,
this group will be written as (unfortunately it is also written as
in the literature!).
The elements M E are naturally matrices from G" (K). They
can be written in the following way. For the standard form, we have
=E , ,

which, with the use of the matrix

J=J 0 E
-E 0
and column vectors v, v' with tv = (xl,... , x, yl,... , can also be
written as
w (v, v') = tvJv'.
The matrix M leaves this form invariant, that is.
W (Mv, Mv') = w (v, v'),

exactly when
(+) tMJM = J.
Then from simple computation, we get
Remark 1.13. For A, B, C, D E M (K), the following are equivalent:

t) M = ` C D) E Spn(K),

ii) tAC ='CA, tBD = tDB, 'AD - 'CB = E, ,

iii) AtB = BVA, CD = D`C, AtD - BC = E,,.


16 1. Symplectic Algebra

Special symplectic matrices are J as well as

uv
= (0 V`) with V* = (IV)-t,

and

m= (0 E) withtS=S.

PROPOSITION 1.14. These matrices generate Spn(K).

Proof. See EICHLER [E], p. 47. 0


As a consequence of this proposition, we immediately get that
det M = 1 for M E Spn(K),
since this is clearly the case for the generators. This statement can also be
derived from the fact that a symplectic automorphism yfi relative to w must
also fix w", the volume form. Therefore, we have
e! A... AeL(0e1, (det0)ei A...Ae2n(e1,..., e2")
=ejA...Ae (eI,...,e2n),
and thus det 0 = 1.
PROPOSITION 1.15. Let M E Spn(K) and X an eigenvalue of M with
multiplicity k. Then 1/A is also an eigenvalue with multiplicity k.

Proof. Consider
P (t) = det(M - tE2"),
the characteristic polynomial of M. Then by using (+) and the fact that
det M = 1, we have
P (t) = det(1M - tE2n) = det (J-' (LM - tE2,=)J)
=
det(M_1
- tE2n) = det M-1 det(E2n - tM)
= t2n det(M - (1/t)Fj ).
0
Remark 1.16. For K = R, if M E Sp,t(K) as a complex matrix has the
eigenvalue A E C. then M also has the eigenvalues A, 1/A and 1/A.

These statements about eigenvalues are fundamental to the qualitative


theory, including that of the stability of Hamiltonian systems. Here we give
just a few comments on the topic of stability (see ARNOLD [A], p. 227):
1.3. Subspaces of symplectic vector spaces 17

DEFINITION 1.17. A morphism ¢, of V is called stable if for each e > 0


there is a i9 > 0 such that
11O vII<e, forallNEN, as soon as 11vJJ<19.
EXERCISE 1.18. If 0 E Sp(V) has an eigenvalue A with JAI # 1. then 0
is not stable.
EXERCISE 1.19. Whenever the eigenvalues A of 0 E End V are distinct
and have norm 1, the transformation is stable.
DEFINITION 1.20. 0 E Sp(V) is called strongly stable if every neighbor
41 E Sp (V) (that is, given a fixed basis for V. the matrix entries of 01 differ
from those of 0 by less than a fixed e) is stable.
EXERCISE 1.21. If for 0 E Sp(V) all 2n eigenvalules are distinct and
have norm 1, ¢ is strongly stable.
The symplectic group is a significant object in both function theory
and algebraic geometry. For clarification the reader may wish to consult
SIEGEL [Sil, Si2]. Spr(R) is a Lie group and has as associated Lie algebra
spn(R) = Lie (see Section B.2). It is easy to see that with the
standard form w we have
op. (R) = {M E W (Mv, w) + w (v, Mw) = 0 for all v, to E liter }

_ (M E tMJ + JM = 0}.
Here, as in the last proposition, we have. for Al E sp,,(R) that if A is an
eigenvalue then so is -A, and both have the same multiplicity.

Particular symplectic morphisms are the symplectic transvections r,,,,a.


For W E V and A E R, r,,,,a : V - V is defined by
v -* Aw (v, w) to.
EXERCISE 1.22. Show that
a) r,,,\ E Sp(V).
b) Sp(V) is generated by the symplectic transvectionss (see JACOBSON
[Ja], p. 373)-

1.3. Subspaces of symplectic vector spaces


Let W be a k-dimensional linear subspace of the 2n-dimensional symplectic
space (V, w). Then k = dimd(W) and 21 := rank wlm(w) remain unchanged
under any symplectic morphism from Sp (V). We will show that these two
integers, k and 21, classify the subspaces of V, in that they are the only two
independent symplectic invariants for subspaces.
18 1. Symplectic Algebra

As in the proof of Proposition 1.3, let


W1:={vEV; w(v,w)=0 for allwEW}.
W 1 is called the skew (or w-) orthogonal space of W (and often simply called
its orthogonal).
Remark 1.23. We have
dim W -L = dim V - dim W = 2n - k.
Then the dimension formula from linear algebra (see, for example, LANG
[Li], P. 95),
dim V = dim W + dim W °
for
W°:=IV (w)=0foralwEW},
can be here applied to show that (for an arbitrary non-degenerate bilinear
form w) W-L can be identified with W° via wb

Specific to the symplectic algebra is the appearance of the radical rad W


of W of the form
radW:=WnW1.
This is also given by
rad W= {w E W; i (w) w I w= 0}
and satisfies
dim rad W = dim W - rank w I W = k - 21.
Clearly, the structure of w-orthogonal spaces has the following properties:
Remark 1.24.
a) W-CW,1for W'CW,
b) (W-)1=W,
c) (W+W')1=WlnW1,
d) (WnW')1=W1+W1.
A third space fixed by W is defined by
°W:=W+Wl.
The usual dimension formula here says that
dim °W = dim W + dim W1- dim rad W = 2n - (k - 21).
From this space we can form yet more spaces, in particular
Wr`d := W/rad W,
1.3. Subspaces of symplectic vector spaces 19

the symplectic space associated to W. This satisfies dim R"'d = 21, and for
a subspace U C V with
Wr`d = rad W e U,
(U, w1 u) is symplectic and isomorphic to Wred. U is also w--orthogonal
to rad W. Therefore, we can also write W = rad W1U. Further, U'
with Wl = rad W ® U' is symplectic and isomorphic to the reduced space
(Wl)red = Wl/rad W.
The most important examples of subspaces of a symplectic space (V, w)
are the following.
DEFINITION 1.25.

1) A subspace Q of V with ";IQ = 0 is called an isotropic subspace of (V, w).

2) A subspace W C V with w'N, non-degenerate is called a symplectic sub-


space of V.

3) A subspace W C V with W1 isotropic is called coisotropic.


4) A subspace L C V which is both isotropic and coisotropic (thus with
Ll = L) is called a Lagrangian subspace.

Before we discuss Lagrangian subspaces further, we prove the earlier


statement that the dimension and rank are the only two independent sym-
plectic invariants of a subspace. This result will be a corollary of the follow-
ing otherwise useful discussion, which is here taken from AIrrIN's masterful
parallel presentations of orthogonal and symplectic geometry in his Geomet-
ric Algebra ((Ar], pp. 114 f).
A two-dimensional symplectic space P is, according to Proposition 1.3,
of the form
P=eK+e.K with w(e, e)=w(e., e.) =0, w(e, e.) = 1.
P is also called a hyperbolic plane and (e, e.) a hyperbolic pair. An orthog-
onal sum of hyperbolic planes is then called a hyperbolic space
H.,. = P11.... P,.
By the same proposition, a symplectic space is always an orthogonal sum of
hyperbolic planes and so is a hyperbolic space. An important step in proving
our claim will be the following theorem, which says that every subspace of a
symplectic space has a symplectic hull. On the way the notion of symplectic
morphism will be extended to a morphism of those subspaces U which are
compatible with the restriction of the symplectic form to U.
20 1. Symplectic Algebra

THEOREM 1.26. Let V be symplectic and U a subspace, written in the


forIn
U = rad U.LW
Let e1, ... , er be a basis of rad U. Then there are elements e1., ... , e,.. E V
such that
P, = e,K + e,.K is hyperbolic (i = 1, ... , r)
with
P,1Pj for i j, and P,1W for all i.
Then
V DU:=P11...1P,.1W
is symplectic with U D U.
An injeetive symplectic morphism 0 from U into a symplectic space V'
can be extended to a symplectic morphism : U - V'.

Proof. i) The proof of the first statement goes by induction. For r = 0


there is nothing to show. Now assume the claim is true for r' < r. Let
Uo :_ (e1,.. , er-1)1W
Then er is orthogonal to Uo, er 0 U0, and
rad Uo = rad Uo = (ell ...,er-1).

Consequently, we have er E Uo and er ¢ rad Uo , and so there is an a E Uo


f o r which w (er. a) # 0. The plane spanned by er and a is contained in Uo ,
and in this plane a can be changed to er. so that e, er. is a hyperbolic pair.
It follows that
Pr:=erK+er.K
is contained in Ua and is orthogonal to Uo, and thus Uo C P, .L.
Since dim rad Uo = r - 1, the induction hypothesis may be applied to
Uo = rad Uo1W C P,l,
giving the existence of hyperbolic planes in Uo
P:=(ei,ei.)C PT, i=1,..., r-1,
which are pairwise orthogonal and each orthogonal to W. Since all P; are
also orthogonal to Pr, we have that
U:= P11...1pr1W
is the desired symplectic space.
ii) Let 0 be a symplectic morphism from U to V' and e; (e,) so that
W': = 0 (W). Then
O (U) = (e, ... , e'r)1W'.
1.3. Subspaces of symplectic vector spaces 21

From i) there are elements e=. E V so that for P' = (e,, e;,) we have
PJP for i0 j, andP'1W', i=1,...,r.
The rule (e,.) := e;. then gives the desired extension of 0.

A consequence of this theorem is a special case of a theorem of Witt.


COROLLARY 1.27. Let V and V be isomorphic symplectic spaces, U C V
a subspace and 0 an injective symplectic morphism from U to V. 0 can be
extended to an isornorphisrn
¢:V-.V.
Proof. From the previous theorem, 0 can be extended to a morphism
U -+ V. Thus we may assume that U is symplectic, and that V = U.LU-.
Let U' := 0 (U) and Ul be such that V = U'J U'1. It only remains to
show that Ul and U'l are symplectically isomorphic. But this is clear, since
both spaces have the same dimension and no radical, and symplectic spaces
of the same dimension are isomorphic.

It is also now dear that dimension and rank of a subspace U C V are


the only symplectic invariants; given two subspaces U, with the same rank
and dimensional, we can as in the theorem decompose them into rad UU and
W;, pair the bases of the radicals and W; against each other, and then use
the corollary to extend this map to a svmplectic automorphism of V.
Besides the isotropic subspaces, the Lagrangian subspaces are the most
important. Let's collect here some of the theory which can now be stated,
given what has already been covered.
For a subspace W C V with dim W = k, we have

W isotropic WCW k< n,


W coisotropic aWDW1 k> n,
W Lagrange e' W = Wi k = n.

A Lagrangian subspace L can thus be described as maximal isotropic. We


then also have (with the identification V V' via wb as in Section 1.1)
V=LeL1=V'.
Remark 1.28. Let W be a coisotropic subspace of V. Then the image of
L n W under projection to W,Yd is a Lagrangian subspace.
EXERCISE 1.29. Prove this last remark. (See VAISMAN [V], p. 35.)
22 1. Symplectic Algebra

DEFINITION 1.30. Denote by G (V) the collection of Lagrangian sub-


spaces L C V.

From the previous corollary, Sp (V) acts transitively on L(V); that is,
for any pair LI, L2 E G (V), there is a ¢ E Sp (V) with
0 (LI) = L2
In such a case, it is a general fact that L (V) has the structure of a ho-
mogeneous space. The concept of homogeneous space will later play a very
important role (see Section 2.5). We will briefly make its acquaintance here
with a special example.
Denote by GL the isotropy group of L EC (V); that is,
GL :_ {O E Sp (V), 0 (L) = L}.
Let e = (e;, ej.)f=l,...,,, be an L-related symplectic basis of V; that is,
(el,... , e,) = L and (el., ... , e,,,.) = L1. Then, relative to this basis,
V K2R, Sp(V) ?d L Lo :_ (ei,...,en), where (et) are the
canonical unit vectors inK2ia, and

GL '-
GL. =I( A B \I ;BD ='DB,'AD =
This is because M(Lo) = Lo. and thus

\C D/\0/EL°
forces C = 0, so that from Remark 1.13 about the general form of symplectic
matrices we get the claimed transitive action.
From this we get a bijection between C (V) and the set of cosets of GL,.
in Sp (K); thus
Sp,,(K)/GLO 2L (V), M i--> M(Lo),
because, from
M(Lo) = M'(Lo),
it follows that
M-IM' E GLo.
In a similar vein, we can describe another family of interesting subspaces
of a given symplectic space. Fix T (L), for a given Lagrangian subspace
L C V, to be the collection of all Lagrangian subspaces L' transverse to L,
so that
T(L):= {L' EL(V), LPL'=V}.
The proofs already given show that GL operates transitively on T (L). Take
(e=) to be a basis of L and (e;.) a basis of L' (so that both together form
a symplectic basis of V). Then, relative to this basis, the isotropy group
1.3. Subspaces of symplectic vector spaces 23

GL,L' C GI C SpV that fixes L and L' can be identified with the group of
matrices \
A
0 iA-I/
0 f ' A E GL,,(K),
C
which is naturally isomorphic to GL"(K). We then get that
7 (L) GL/GL"(K),

In VAISMAN [V], pp. 36-38, the following statement and a few of its conse-
quences are discussed.
THEOREM 1.31. T (L) has a natural structure as an affine space over K
of dimension n (n + 1)/2.
An affine space is a triple (A, V, ir), where A is a set, V is a K vector
space and 7r : A x A -* V is a morphism, such that 1rIfor a fixed
point ao E A, is a bijection and for all a, b, c E A we have
ir(a, b) + a(b, c) = ir(a, c).
This definition can easily be brought into agreement with the definitions
which often appear in elementary textbooks. Of the proof which is contained
in VAISMAN [V], we give only a sketch:
By choosing the earlier required basis, the matrix of a morphism which
carries L' E T (L) to L" E T (L) has the form
(A B

which then modulo GL(En


L, = has the form

l with X ='X.
0 E,
This shows already thatT (L) rel/ative to a fixed basis of V is in bijection
to a set of symmetric n x n matrices, and therefore can be seen as an
(n (n + 1)/2) -dimensional affine subspace of K"2. This observation can be
made basis-independent, in that the symmetric matrix X can be defined
as a coordinate-independent quadratic form q on the dual space P. The
following remarks can then be established:
Remark 1.32. For any pair (L, L') of Lagrangian subspaces of (V, w),
there exists a mutually transverse Lagrangian space L".
Remark 1.33. Let L, L', L" be Lagrangian subspaces of (V, w), and let
L fl L' = L fl L". Then there exists a (not unique) symplectic transformation
of V which fixes every vector of L and carries L' to L".
EXERCISE 1.34. The reader would benefit from supplying proofs for
these remarks (which in any case can be found in VAISMAN [V]).
24 1. Symplectic Algebra

1.4. Complex structures of real symplectic spaces


Up to now, we have considered only R2n equipped with
i) the canonical Euclidian structure with the form
2n
s (v, w) := (v, w) = tvw = > vow' for V, W E R2n (as columns),
j=1

ii) the canonical symplectic structure with the form


n
w (v, w) = =vJnw = E(Vi wn+i - Vn+i wi)
i=1
Now, as R vector spaces, R21 Cn.. With the identification

v= 1 X

the operation z .+ iz corresponds to the operation v '-. -Jnv =


This operation,
Jn : R2n -- R2n with Jn = - id,
supplies R2n with a complex structure. It is then natural ask in how many
ways R2n or, indeed, an arbitrary R vector space, can be supplied with a
complex structure. This question will be considered further below. For now,
we give a brief survey.
The invertible linear morphisms of R2n, that is, the invertible matrices
M, that
i) preserve the canonical scalar product, s, form the orthogonal group
O(2n);
ii) preserve the symplectic standard form, w, form the symplectic group
Spn(R);
iii) preserve the complex structure J, that is, with MJ = JM, are of
the form
M- (X -Y
YX
and form the general linear group GLn(C) via M i-. X + iY.
It can be shown that
O(2n) n Spn(R) = Spn(R) fl GLn(C) = GLn(C) n O(2n) = U(n).
The unitary group U(n) then preserves the hermitian scalar product
h (v, w) = s (v, w) - i w (v, w).
1.4. Complex structures of real symplectic spaces 25

As a consequence of earlier material, we have, for the space of all Lagrangian


subspaces of R2n.
L (R 2") U(n)/O(n),
and the space of all positively compatible complex structures J on R2n is in
bijection with
s7n = {Z E Ill"(C): tZ = Z. Im Z > 0} = Sp"(R)/U(n).
All of this will now be discussed in greater generality (following VAISMMAN
((V], pp. 40 ff.).
DEFINITION 1.35. Let. V be an 1EY vector space. J E Aut V is called a
complex structure on V if and only if
J2 = -id%,.

In the case that V is symplectic with the form w, then we call the complex
structure J compatible with w if
w(Jv, Ju') = w(v. w) for all v. w E V.
Slightly changing the notation as in the previous example. (V, J) with an
arbitrary J can be made into a C vector space via
V"--l V := Jv.
Further. J can be extended linearily to the complexification
6', :=VSRC.
Then J has the eigenvalue fv=l, and from Jw = Aw we deduce that
-w=J2u+_AJu,_.A2u'.
The eigenspace of \1.2 = f is n-dimensional, and is given by
V+:={v-%/---I JvvEV}, V :={v+v/ I Jv. vEV}.
We then have that
V, =VC+ gV,,
and
c- V - y-IJV
defines a C vector space isomorphism between (l: J) and (1' . ).
In the case that J is a complex structure compatible with the symplectic
form w. we have
g(v, w) := w(v. Jw) for v, w E V.
From the compatibility of J and w we have also
g(Jv, w) = w(v, w)!
26 1. Symplectic Algebra

and from J2 = -1 and the skew-symmetry of w we also get


g(v, w) = g(w, v)
and
g(Jv. Jw) = g(v, w).
Therefore g is a symmetric bilinear form, and, like w, is also non-degenerate.
g will be called an w-compatible pseudohermitian metric. When g(v, v) > 0
for all v E V, we call g a hermitian 1 metric, J a positive compatible complex
structure and the triple (V, u;, J) a Knhler vector space.
THEOREM 1.36. Every real symplectic vector space (V, w) can be given
a compatible positive complex structure J and a hermitian structure g. Any
two such structures Jo and Jl are homotopic in the following sense: there
is a differentiable family Jt, 0 < t < 1, of positive compatible complex
structures on V defining a path from Jo with Jl.
Remark 1.37. An analogous statement holds for Hilbert spaces having
a skew-symmetric weakly nondegenerate bilinear form w. The following
reasoning for the general case reduces the proof of ABRAHAM-MARSDEN
((AK, p. 173) to our simpler case.
Proof. Let 7 be a Euclidean scalar product on V. and let A : V V be
defined by
y(Av, w) = w(v, w) for all v, w E V.
Since w is skew-symmetric, we have
y(Av, w) = -w(w, v) = -y(v, Aw),
and so
y(A2v, w) = --y(Av, Aw) = y(v, A2w);
this is to say that A2 is self-adjoint and, since
y(A2v, v) = -y(Av, Av) < 0
is negative, must have negative, but not necessarily distinct, eigenvalues
-A j2, Ai > 0 (j = 1,... , 2n). V then has a y-orthonormal--basis a of eigen-
vectors for A2, al, ... , a2,,. Let B E Aut V with

Ma(B) =
A2n
Then B is the unique self--adjoint positive operator with
B2=-A2.
'This notation will later be clarified: It will be shown that g can be extended to a hermitian
metric on the complexificatlon of V.
1.4. Complex structures of real symplectic spaces 27

For J := AB-i we have, when we consider the eigenspaces of A and B,


J = B-1A,
and thus
J2 = AB-i. B-1A = -E.
This J is compatible with w, and so
w(Jv, Jw) = w(AB-iv, AB-'w)
--y(AB-ivy A2B-'w)
= y(A2B-lv, AB-1w) =
= -w(B-iv, A2B-iw) = w(B-iv, Bw)
= w(BB-iv, w) = w(v, w).
for g with
g (v, w) := w (v, Jw),
and we have
g(v, v) = w(v, Jv) = y(Av, Jv) = -y(v, AJv)
= y(v, Bv) > 0
by the construction of B. This J satisfies the requirements of the theorem.
Since it is dependent on the chosen scalar product y, we write J = J.. It
can be seen that every positive compatible complex structure J arises in this
manner from some such y (one need just note that J = J. with g from (*)).
The last statement of the theorem is now easy. Let Jo and Ji be given;
then they are of the form J.,o and Jy, , and can be carried from one to the
other via the family Jy with
yt:=tya+(1-t)7i (O<t<1).
0
In the standard case,
V = RZ",
with w (v, w) _ `rvJnw,

Jn = ` -1 q),
and (v, w) _ 'vw, /
we clearly have A = -Jn. So A2 = -E, B = E, and so J = -Jn gives the
positive complex structure, and g (v, w) = 'vw.
28 1. Symplectic Algebra

The scalar product described above is called the hermitian metric, since
g induces a hermitian bilinear form gc on the complexification Vc of V by
9c(vCiv,w)=-9c(v, f Tw)=vi9(v,w) for v,WEV.
Then there is the restriction of gc to V.
gc(v - v/--l Jv, w - Jw) = 2 [9 (v, w) - viw(v, w)],
and then
h (v, w) := g (v, w) - %/'--I w(v, w) for v, to E V
is a usual hermitian metric on (V, J) as C vector space. This immediately
gives us
Remark 1.38. The map
V D v ,-+ -(v - vr1 Jv) E V+
is an isomorphism of the hermitian spaces (V, J, h) and (V+, 9c).
From linear algebra (V, J, h) has an orthonormal basis relative to h,
thus a C-basis (ej ), j = 1, ... , n, with
h (ej, ek) = bjk.
Since h = 9 - vl'--l w and g (v, w) = 9 (Jv, Jw) = w (v, Jw), this is equiva-
lent to
g(ej, ek) = 9(Jej, Jek) = bjk, g(ej, Jek) = 0.
From this we may conclude that (ej, Jej; j = 1, ... , n) is a real g orthonor-
mal basis of V and, with ej. = Jej, the family (ej, ej.), j = 1, ... , n,
is a symplectic basis of V.
In the other direction, if (ej, Jej) is a real unitary basis of (V, w, J),
that is, a g orthonormal basis and at the same time a symplectic basis, then
(ej) is also an h-basis, and we have a corresponding C-basis of V+ given by

This is called a complex unitary basis.


The automorphisms of the structure (V, w, J) which leave the symplectic
form w, as well as the metric g (or, equivalently, those which commute with
J), fixed are called the unitary transformations and generate the unitary
group U(V, J). These automorphisms are also characterized by carrying
unitary bases into unitary bases. Other equivalent descriptions are given
as the complex-linear transformations of (V, J) which fix the metric h, or
(see the last remark) as the group of complex linear transformations of V+
which fix the metric gc. By fixing a basis one may derive the relationships
between matrix groups as described at the beginning of this section: The
choice of a unitary basis (ej; j = 1, ... , n) allows one to identify (V, J, h) as
1.4. Complex structures of real symplectic spaces 29

well as (V+, gj with C", where h in the complex coordinates (zj) relative
to (ej) is the canonical hermitian metric

h(z,z')zizi.
1=1
Here we then have that
U(V, J) c U (n) = {U E GLn(C); UtU = En}.
On the other hand, for the Euclidian vector space (V, g), O(V, g) is the
orthogonal group of linear isomorphisms which fix g. With the help of a g
orthonormal basis this group is identified with O(2n). The above discussion
can be summarized in the following proposition.
THEOREM 1.39. Let (V, w) be a real symplectic space and J a positive
compatible complex structure. Then
U(V, J) = Sp (V) fl O(V, g).

In particular, for V = R2n with

v=
the canonical basis e = (ei, e;,) and the coordinates ( Vi.

the symplectic standard form


n
w(v, w) = tvJw = E(vi wi. - vi. wi),
i=1

the canonical scalar product

9 (v,w) = tvw = Dviwi + vi. wi=),


i=1

the complex structure defined by Jei = ei., Jei. = -ej,

then the complex coordinates zi = vi + vi. (i = 1.... , n) satisfy


U(n) = Spn(R) fl O(2n).
This leads to the following alternative formulation of the description
given in Section 1.3 of the space G (V) of Lagrangian subspaces L C V in
the case that V has a compatible positive complex struture J. This arises
from the fact that in the above introduced notation, any g orthonormal basis
(ei)i=1,...,n of L (ei, Jej) gives rise to a real unitary basis of V. Each such
basis is called an L-related unitary basis. For L' E L (V) with L'-related
30 1. Symplectic Algebra

basis (e!, Je,) there is a unitary transformation which takes (ei, Jei) to
(e;, Je,), and so L to V. The unitary group U(V. J) operates transitively
on G (V), and the isotropy group which fixes an L is the group which carries
L-related bases to themselves and is thus isomorphic to the orthogonal group
O(L,g). Thus we have shown
THEOREM 1.40. We have that
,C(V) .,, U(V, J)/O(L, g) '=' U(n)/O(n).

Now we will fulfill the promise made at the beginning of this section;
that is, we describe the collection
9=J(V,w)
of all compatible positive complex structures J on real symplectic space
(V, w), and so the set of possible ways (V, w) can be made into a Kahler
vector space. From this it will follow that 3 can be identified with the Siegel
upper half plane
15n ^-' Spn(R)/U(n)
From Remark 1.37, for a fixed J E J, the space (V, J) can be identified
with a subspace V+ of the complexification of V. Here V+ is a Lagrangian
subspace of V, where the symplectic structure w of V is linearly extended
to V. This is because we have
w(v - Jv, W - Jw)
= w(v, w) - w(Jv, Jw) - (w(Jv, w) + w(v, Jw)) = 0.
Furthermore, since g(v, w) := w(v, Jw),
-/ w(v, `v) = 2g(a, a) > 0 for 0 96 v := a - Ja E V+,
and, additionally, Jv = -,./--l(vl - v2) for
V=V+$Vc 9v=vl+v2,
Then the Lagrangian subspace F of (V, w) is called positive if
forall0&vEF.
Remark 1.41. There is a natural bijection between 9 = 3(V, w) and
the collection L+ = L+(V,,) of positive Lagrangian subspaces of (Va, w).
Proof. i) The mapping 3 - G+ is defined from the map J - V+
ii) For F E G+ F is also a Lagrangian subspace of V, since w is real.
From the positivity of F, it follows that F fl F = {0} and so V = F (D F.
Now we can define J : V -+ V by
vl,v2EF
1.4. Complex structures of real symplectic spaces 31

Then J2 = -id, w (Jv, Jw) = w (v. w). for all v, w E V, and J (V) = V,
and so for v E Vc we have the bijection
vEVav=v.
Finally, we have
w(v, Jv) = -2vr--1w(vi, 1Y1) > 0 for 0 34v = vl + v1 E V,
and so J E J(V, w).
iii) The morphisms 3 -. G+ from i) and L+ - J from ii) are clearly
inverse to one another.

A Lagrangian subspace L. of a complex symplectic space VV is called


a real Lagrangian subspace, if it is the complexification of a Lagrangian
subspace L C V; that is, L, = L OR C. This is satisfied precisely when Lc
is carried to itself by complex conjugation. that is. when Lc = L. In V c.
all real Lagrangian subspaces Lc are transversal to every F E C+. This is
because for
00v=a+vr 1bELcnF witha,bEL
we have
0<- / u; (v, v)_-2w(a,b)=0.
and so also the converse.
The remainder of the treatment in VAISMAN [V] is based on the de-
scription of the space T (L) of Lagrangian subspaces transverse to a fixed
Lagrangian subspace as an affine n (n + 1)/2-dimension space, as given in
Theorem 1.31. Since this description was not fully given here, we can only
give a sketch of VAISMAN'S treatment, but enough to see the idea:
Let L, be a fixed real Lagrangian space. Then, as described above.
C+ = C+(ti') C T (Lc). That T (LJ is an affine space means that for each
pair F E C+ and L' E T (La) with L' real and transversal to L, (thus
Lc E3 L' = V) there is a symplectic transformation 0 which fixes every point
of L, and carries L' into F. If e = (ei, ei.) (e; E L. ei. E L') is a basis of
V, respectively V. appropriate to L. then ¢ (see Theorem 1.31). relative to
this basis, is written as a matrix

l ,, : I with ZEM,, (C), Z=1Z.

Then det Z 96 0, since F is /also transversal to L. The positivity of F sets


yet another condition on Z. And thus qi can be written, relative to e, as.

((ei), (ei.)) (0 Z ) = ((e,). (ei.) + (ei)Z),


32 I. Symplectic Algebra

and so
n
O(ei) = ei, 0(ei.) = ei. +EejZiji, (Z;j) = Z.
j=1

In fact, 0 (ej.) is a basis of F, and a small calculation shows that F is


positive, and so
forOOvEF
exactly when
Ire Z = 2 V--1
is a positive definite matrix, which can also be written simply as Ire Z > 0.
This shows that, with a choice of a basis e, the mapping F - Z gives a
mapping from C+ into the so-called Siegel upper half plane
J , = {Z E MM(C), tZ = Z, Im Z > 0}.
With a bit more care and the help of Remark 1.41, one can show
THEOREM 1.42. For a real 2n-dimensional symplectic space (V, w), the
set 3 (V, w) of positive compatible complex structures on V can be identified
with the Siegel upper half plane fin.

The identification of the theorem depends on the symplectic basis a of


V. Should the symplectic basis be changed to e, then, since

(*) (e,e.)=(e,e.)(C D) with ( C D ,

there is a matrix Z associated to F relative toe with the property that F


has the bases e. + eZ as well as e. + eZ. So there is a A E CLn(C) with
e. + eZ = (e. + eZ)A.
Applying the transformation formula (*), we get
eB+e`.D+(eA+e.C)Z=e,A+eZA,
and thus
A=CZ+D and Z=(AZ+B)(CZ+D)-1.
From complex function theory, the mapping

Z I. Z = (AZ + B) (CZ + D)-' =: (


C D) (z)
is recognizable as a complex-analytic automorphism of $,,.
1.4. Complex structures of real symplectic spaces 33

We have just seen that operates on gj,,, and previously we have


established the bijection t (V) = Therefore it is not so sur-
prising that b,,, and hence ,7(V, w), can be described as a homogeneous
space. It turns out that
.7(V, w) S7n ^-' Sp,(R)/U(n).
Thus the operation 0 E Sp(V) can be extended to an operation on V,; which
is also called 0. This 0 then also operates on C+, and gives a transitive
operation of Sp (V) on C+. This is because the equality
(*), h(v, w) :_ w) for v, w E F from C..
allows one to define a hermitian metric h on F, and for an or-
thonormal basis of F, we have that (A, -vi74);_1,,.. is then a symplectic
basis of V. Analogously, for F' from C+ with h orthonormal basis (f,),
(f j', -v -1 f i)i=1,..,,,, is a symplectic basis of V,. Then there is a d E Sp (Ve)
with
(f=) = A', f;.
This 0 maps F to F' and commutes with complex conjugation, and therefore
is in Sp (V ). Thus, the transitivity of the operation of Sp (V) on G+ is
demonstrated. The isotropy group of F in Sp(V) is, because of (*), the
unitary group U(F, h). And so we have shown
THEOREM 1.43.

.7 (V, w) = C+(V) ^' fn = Sp(V)/U(F, h) =' Sp.,(R)/U(n)

EXERCISE 1.44. The reader is recommended to show directly (that is,


independently of what has come before) that:
a) G = SL2(R) operates on bi = {r = x + iy E C, y > 0} by

(M, r) -. M(r) := ar
+d
for M=[ a d f E SL2(R) and r E 551,

b) The following isomorphism holds:


S51?` SL2(R)/SG(2) Sp1(R)/U(1)

c) The Poincare metric


dx2 +dye
ds2 =
y2
is SL2(R)-invariant.
34 1. Symplectic Algebra

The Siegel half space $ has many uses in the area of moduli problems
of Abelian varieties. It can be given the structure of a complex manifold of
dimension n(n + 1)/2 (see SATAKE [Sa], p. 78). The study of the geom-
etry of these manifolds and the holomorphic, as well as the meromorphic,
functions on 15 with known invariant or covariant properties under the op-
eration of the group or its subgroups was initiated by SIEGEL (see
his Symplectic Geometry [Sil] or Topics in Complex Function Theory [Si2]);
for some time it was exactly these topics introduced by Siegel that formed
the subject of symplectic geometry. One may find a new and particularly
nice treatment of these traditional topics in MUMFORD: Tata Lectures on
Theta, Vol. 2 ([Mul]). Nowadays, however, symplectic geometry refers to
a much broader range of topics, which we will consider in the next chapters.
Chapter 2

Symplectic Manifolds

Symplectic geometry arises from the globalization of the symplectic algebra


considered in the previous chapter. The central concept is that of a sym-
plectic manifold. We begin, in this chapter, by defining these, and continue
by studying some of their properties and by giving several examples. A good
reference for this material is the second chapter of AEBISCHER et al. ]Ae]
and Section 3.2 of ABRAHAM-MARSDEN [AM].

2.1. Symplectic manifolds and their morphisms


In what follows we will assume that M is a smooth manifold of dimension
p, that is, a C°°-manifold in the sense of Section A.1, and assume, as well,
that it is real, unless something is said to the contrary.
DEFINITION 2.1. M is called a symplectic manifold, if there is defined
on M a closed nondegenerate 2-form w; that is, an w E c12(M) such that
i)dw=0,
ii) on each tangent space T,,,M, m E Al, if
w(X, Y) = 0 for all Y E TmM,
then X = 0.

The assumptions on w say that its restriction to each m E M makes the


tangent space TmM into a symplectic vector space. Thus it is already clear
that the dimension p of M is even; thus p = 2n. In the next section, it will
be shown that all symplectic manifolds of the same dimension are locally the
same. This is in sharp contrast to the situation in Riemannian geometry,
and indicates that symplectic geometry is essentially a global theory.

35
36 2. Symplectic Manifolds

However, not every even-dimensional manifold has a symplectic struo-


ture. In AEBISCHER et al. ([Ae], p.3) is given the example of M = S4;
however, this uses a cohomological result (see Appendix C), which we should
not yet go into.
Given two symplectic manifolds (M, w) and (M', w'), let F : M M'
be a smooth map, that is, differentiable in the sense of Section A.I.
DEFINITION 2.2. The map F is called symplectic, or a morphism of
symplectic manifolds, so long as
rw' = w.
Given a symplectic diffeomorphism F, F-I is also symplectic, and F is called
a symplectomorphism. Sp (M) denotes the group of symplectomorphisms
from M to itself.

2.2. Darboux's theorem


Darboux's theorem has, in its simplest form, the following formulation. To
every point m of a symplectic manifold (M, w) of dimension 2n, there cor-
respond an open neighborhood U of m and a smooth map
with rwo=wIu,
where wo is the standard symplectic form on R. It follows immediately
that for an appropriate choice of symplectic coordinates x = (q, p), p =
(pl, ... , pn) and q = (q,, ... , qn ), w can be written on U in the form
n
w = dgAdpdq,Adp;.
i=1
In AEBISCHER et al. ([Ae], p. 17) a proof is given which follows that of
GUILLEMIN-STERNBERG ([GS], pp. 156ff). We will use this proof here
and supply somewhat more detail. It does go back (as the proof in [AM],
p. 175) to MOSER (1965) and WEINSTEIN (1977). As a preliminary we give
a standard result on the behavior of parameter-dependent differential forms
under transformations, which will be useful in other places. Although the
details are somewhat tedious, its proof will provide a good exercise in the
calculus of differential forms described in Section A.4.
LEMMA 2.3. Let M and M' be smooth manifolds and Ft : M M', t E
Ht, a smooth 1-parameter family of maps. Let Xt denote the tangent field
of M' along Ft; that is, Xt is the map
M - TM',
m '--i (m', Xt(m)), m' = Ft(m),
2.2. Darboux's theorem 37

where Xt(m) is the tangent vector at m' = Ft(m) E M' to the curve s
F,(m). Let (at)tER be a 1-parameter family of differential forms on M.
Then we have

3i (Ft at) = 1 t {( tt + i(Xt)dat) + d(FF (i(Xt)oe))

= Fi (4dt t + i(Xt)do,t + d(i(Xt)at))


so long as all the maps Ft are difeomorphisms from M to M' and, therefore,
Xt gives a vector field on M'.

The derivative gi applied to a one-parameter (t) differential simply


means the differentiation of the coefficients with respect to that parame-
ter. In any case, we use the usual notation for differential forms (see Sec-
tion A.4). An exception is the symbol Ft (i(Xt)at), whose meaning as a
differential form on M must be clarified in part b) of the proof, since with
the general preconditions Xt is not a vector field on M' and therefore the
symbol i(Xt)at does not give rise to an inner product in the usual sense.

Proof. a) The statement will first be proved for the special case M =
M' = N x I, where N is an n-manifold, I is an interval, Ft = vt with
r(it(x, a) = (x, 8 + t) for x E N, and 8 E I. A differential form at on N x I
of degree k, which depends on x, a and the parameter t, can be written as
at = ds A a (x, a, t) dxk + b (x, s, t) dxk+i
where the coefficients are given in the abbreviated form
a(x, a, t) dxk ai,...ik (x, 8, t) dxi, A ... A dxik.
i1G..<4
With this, we clearly have
ytat =dsAa(x, s + t, t)dxk+b(x, s+t, t)dxk+i
and so
d
(iP at) = ds A T (x, s + t, t)dxk + -(x, s + t, t) dxk+i
dt 8s as
(1)
+dsA (x,a+t,t)dxk+ 5i(x, a+t, t)dxk+i
We immediately deduce that

(2) ti ( j- J = ds A (x, s + t, t) dxk + (x, s + t, t) dxk+l


38 2. Symplectic Manifolds

The vector field Xt, in this special situation, can be written as Xt = $


And so we then have

i(Xt)crt = i ( a ) of = a(x, s, t) d?
and

d (i(X1)ot) = d (a(.)(xIs t ) dx, A ... A dii)


(')

_ (8am (x, s, t) ds A dxt, ... A dxtk


(i)

s)
+ (x, s, t) dxj Adxt, A... AdxiR},
ax 7-
jOl...ik)
which can be written more briefly as

d (i(Xt)ot) = as (x, s, t) ds A dxk + dsa (x, s, t) dxk+1.

Pulling back by ;t, we get

84
(3) t,I d (i(Xt) at) = (x, s + t, t) ds A dxk + 4a (x, s + t, t) dxk+1_

as
With the same notation,

8b
dot = -ds A 4a (x, s, t) dxk+1 + (x, s, t) ds A dxk+l + 4b (x, s, t) dxk+2
8s

is then

i (Xt) dot = -4a (x, s, t)dxk+1 + 8 (x, s, t) dxk+1

and

(4) rJlt i (Xt) dot = -dxa (x, s + t, t) dxk+1 + 8 (x, s + t, t) dxk+1

Adding (2), (3) and (4) and equating this with (1) gives the claim for this
special case.
2.2. Darboux :s theorem 39

b) The general case will now be derived from the case a). via the following
decomposition:
Ft=Foi,toj with j:A1-Al xI.
m,-+(m.0),
t;'"t:AMxl-Alxl.
(m, s) -. (m. S + t).
F:MxI - A1'.
(m, s) F,s(m)
given by the diagram

The image of the curve in Al x I given by s (m, s + t). which for s = 0


passing through (m, t) is the curve s'- F,,+t(m) passing through Fj(m) =
m' at s = 0. The tangent vector X1(m) from the tangent field along Ft at
the point m' = Fi(m) = F(m, t) is then the image of the tangent vector to
the curve {(m, s+ t), s E I}, which was previously given as a4; that is. we
have

(5) (F) m.t)


(
X1(m)
To prepare the key conclusion, we only need to trace a vector
17 = Xm ETmM
through the diagram which gives Ft as a composition. And so n is carried
by j. to
(6) (,1, 0) E T(m,o)(A1 x I),
by (0t).j to
(7) (77, 0) ET(m,t)(M x I),
40 2. Symplectic Manifolds

and finally by (Ft). = F.(0t).j. to


(8) (Ft)*+n(0) E TFe(m)M.
By Lemma 2.3, in its sharper statement that all the Ft are diffeomorphisms
from M to M', Xt can be understood as a vector field on M', and for a
given (k + 1)-form at on M' it gives meaning to i (Xt) at as a k-form on
M'. Under the general weaker statement, Ft (i (Xt)) at is also meaningful
and can be thought of in the f o l l o w i n g w a y. T
define
Ft (i(Xt)0t)m..,r/k)

:= (at)Ft(m) (Xt(m), (Ft)*m(th), ...,


and, with the presumption that m' = Ft(m) = F(m,t), this with (5) and
(8) gives
Ft (i(Xt)at),n(rll,-..,t1k)

(at)F(t,m ) (F*)(m.t)
8 (m,t)
I , (m,o),...,(F.)(m,t>(rlk,0)

Then pulling back by F to (m, t) gives


r, (i (Xt)at)m(tfl,-..,thk)

_ (rcl)(mt),...,(tik,0)
(Ll(m,t)(mO)
_ 8
: F*at ((rh, 0), ... 0))
88

and then pulling back by rot to (m, 0) gives, with (7),


Fi (i(Xt)at)m(m,...,nk)
l
= tyi (i (88) F*at) ((rh,0),...,(rlk,0)),
II (m.0)

and, pulling back by j to m,


Ft (i(Xt)at)m(th,...,r/k)

_ (j*tyi) (i (8s) (rh, ,tlk)


F'at) (m,0)
2.2. Darboux's theorem 41

This means that

(9) Fi (i (Xt) at) = j`C, (i (8s) F'at)


and so

(10) Fe (i(Xt)dat) =j''ipi (i (8s) d(F'at)) .

Now, since j' and F" are independent of t, we get


d d d
(Ft at) = (j* F'at) = je dt(tl'i F'at),
Wt Wt

and applying part a) of the proof to F'at we get. since d j* = j'd,


d . d(F'at)
dt
(Ft at) =j 't dt

+?otGt (: (as) d(F'ot)) +j'd F'at))


(i-)
and further, from (9) and (10),

dt(Fiat) =j'tP;F` t +j'tjiF'(i(X1)dat)


With this the claim is shown. Under the stronger condition that all the Ft
are diffeomorphisms, this can also be written more easily as
(dol
at) = F1 + i (Xt) dat + d (i (Xt)at)) .
dt(Ft
0
Remark 2.4. This formula degenerates when M = M'. at is independ-
nent of t and Ft is the flow arising from a vector field X on M. Then the
formula for the Lie derivative of a along X is
Lxa = i(X )da + d(i(X )a).

EXERCISE 2.5. Prove the remark.

The above notion of a flow Ft to a given vector field X on M will be used


later. Here we will only reproduce Theorem 8.1 from STERNBERG ([St], p.
90), which says that for every mo E M there exist a neighborhood U of mo,
an e > 0 and a family of differentiable maps Ft : U M with
42 2. Symplectic Manifolds

i)

F: (-E,e)xU - M,
(t, m) Ft(m) is differentiable,

ii) for Iti, IsI, Is + tj < e and m E U with Ft(m) E U, we have


Fa+t(m) = FF(Ft(m)),

iii) form E U, Xm is a tangent vector at t = 0 to the curve t ,-+ Ft (m).

For the further properties we will later need we refer to Section A.4 as well
as to ABRAHAM-MARSDEN ([AM], pp. 61-67).
The result stated at the beginning of the section is then a direct conse-
quence of the general theorem:
THEOREM 2.6. (Darboux's Theorem) Let wo and wi be two nondegen-
erate and closed forms of degree 2 on a 2n-dimensional manifold M with
wok,,, = wi Im for some m E M. Then there exist a neighborhood U of m and
a difeomorphism F : U -. F (U) C M with F (m) = m and F`wl = wo.

Proof. The idea of the proof is to use a deformation argument to get


a neighborhood U of m and a family (F()tEj, I = [0,1], of diffeomorphisms
from U to Ft(U) such that
Fo = id,

F1 = F,
Ft (m) = m, and
Ftwt = wo, for all t E I with wt :_ (1- t)wo + twi,
and so, in particular, F'wl = wo. The Ft are realized as flows to the time
dependent vector field Y on Ft(U) with

(o) da t (m') = Yt(m`) for all m' E U.

The determination of Yt and thus of Ft will proceed in several steps.

a) F om the equation wo = 1 t wt it follows from Lemma 2.3 as a necessary


condition that

0=
dt
(Ft wt) = Ft (awl + i(Yt)dwt + d(i(Yt)wt)) .
2.2. Darboux's theorem 43

Since wo and wl are closed, so are all wt; that is, dwt = 0; the necessary
condition is equivalent to

(*) d(i(Y)wt) dtwt = "0 - wl =: a.


This will now be taken as the defining equation for Y.
b) Since a is closed, there are, by Poincare's lemma, a neighborhood
U1 of m and a 1-form a on U with do = a and a(m) = 0. The defining
equation (*) becomes
(**) i(}')Wt = a.

c) We have, for all t E I, that wt (m) = wo (rn) . and therefore that it


is nondegenerate at rtt. It follows that this also holds in a neighborhood
Up of m, contained in Ul. Therefore, there is in U0 a vector field 1 which
satisfies (**) and therefore also (*). Because of the normalization a(m) = 0,
Y(m) = 0, we get, from the existence and uniqueness theorem for systems of
ordinary differential equations. that the family (l' )tEi of vector fields can be
integrated to obtain a family (FF)tE/ of diffeomorphisms which on an open
neighborhood U of m satisfy Ft(U) C U0 for all t, and satisfy (o) with the
normalizations FO = id and F, (m) = m. By construction, it follows that for
tEI,
(Ft"wt) = 0,
dt
and then also the expected equality
Fiwt=Fpwo=wo.

d) The 1-form a from part b). whose existence was guaranteed by


Poincart's lemma, can be made more explicit with the application of a little
analysis. Namely, this says that there is a neighborhood Ui of in which has
a smooth retraction from Ul to {m}, that is, a family of maps
c"t : Ul ---> Ut with Vi = id, (pt(m) = m for all t and wo : U, --. {m}.
For each (Vt)tEj we have a tangent field along Oat in the sense of Lemma 2.3,
and with its help we have
i
a-Soa= f d (vio,)dt
0
and

f (c (d + i(X,)d,) + i(X! )a))dt.


0
44 2. Symplectic Manifolds

Then, since a(m) = 0, a is closed and independent of t. We have


I

a = da with a = J(i(Xt)u)dt.
0
Then by substitution into the given formulas we get from this a the desired
family (Y)tEj, respectively (Ft)tEI, as in c).
In GUILLEMIN-STERN BERG ([GS], p. 156) there is a further equivariant
sharpening of this theorem, which we will now describe, although this deals
with the situation of manifolds with group operations, which we will study
in detail later. We will consider the operation of a compact group G on the
symplectic manifold M. Let m E M be a fixed point under this group oper-
ation. Then wo, as well as wI, will be G-invariant symplectic forms on M.
Then there exist a G-invariant neighborhood U of m and a G-equivariant
diffeomorphism F from U into M with F (m) = m and F*wl = "70. The
proof of this statement requires adding just a little bit more additional work
to the proof just given.
The result stated at the beginning of this section is then a consequence
of Darboux's theorem:
COROLLARY 2.7. For each point m on the symplectic manifold (M, w)
them- exist an open neighborhood U of m and a symptectomorphism F of U
onto a subset F (U) of R2" equipped with the standard symplectic form wo.
Proof. Here we will require a little more external yet routine analysis.
This will assure the plausible existence of a diffeomorphism Fl : UI , U
of a neighborhood UI of the origin of the tangent spaces T,,,M H22n to a
neighborhood U of m in M. Then wI := Fi w is a symplectic form on U1.
Therefore, after a linear transformation, it can be assumed that w1lo = wolo.
Darboux's Theorem 2.6 now guarantees that there exist a neighborhood
Uo C UI of 0 and a diffeomorphism
Fo:Uo - Fo(Uo)CU1with Fo(0)=0and Fowl=wo.
F (FI o Fo)-I, which is clearly the desired symplectomorphism.
As we have already mentioned, the coordinates given by the corollary
will be called symplectic and will be written as (q, p).
The fact, just proved, that all symplectic manifolds of the same dimen-
sion are locally the same immediately raises questions about finding global
distinguishing features. We will take a glance at some results on these ques-
tions at the end of this chapter in Section 2.7, and explore a little of what is
today a very active area of research. But first we need to give a few examples
of symplectic manifolds.
2.4. Kahler manifolds 45

2.3. The cotangent bundle


The most important example of a symplectic manifold for physical applica-
tions is the cotangent bundle, M = T'Q, to an n-dimensional manifold Q
(see Section A.3). Q here plays the role of configuration space and M that
of phase space (see Section 0.2). Such an M is necessarily a 2n-dimensional
manifold. As coordinates of a neighborhood U of a point m E Al we will
take, as usual, (q, p) = (q,.... , q,,, pi , ... , (which we will sometimes
give in the natural order p, q; here we just want to make sure that it is
understood that the q parametrize the configuration space Q, and that it is
customary to indicate the coordinates of the base space first when speaking
of a bundle). A 1-form 79 is defined on M = T'Q and is given on U by

d = pdq = > R dqj


1=1

This form is called the Liouville form. The form can also be understood,
using the notation from Appendix A, as follows: t9 is defined as a 1-form on
M, given by
19-(11) :=
for in E M. (Thus m = (q, µ,,), where k is a 1-form on Q, 71 E T,,,M, 7r is the
canonical projection M = T'Q Q carrying m = (q, p,,) to q, and (7r=),
is the induced map T,,,M - TTQ.) For the negative of the inner derivative
w := -dd, we have, in terms of the (q, p)-coordinates,

-dd=dgAdp=Jdq,Adp;.
==1

This form is clearly closed and non-degenerate, and so defines a symplect.ie


structure on M = T'Q. Each diffeomorphism F : Q - Q naturally extends
to F := (F-')' = F'-1 a diffeomorphism of M = T'Q to itself, which is a
symplectic morphism.
EXERCISE 2.8. Prove this last comment.

2.4. Kiihler manifolds


A Kuhler manifold is, roughly speaking, a complex n-manifold (thus the
transformation functions between the charts are holomorphic), equipped
with a Kiihler metric, that is, a hernitian metric for which the associated
2 form w is closed. This metric was introduced by KAHLER in 1932 [K],
and taken up by WEIL [We] among others, and has, because of the pecu-
liar properties of Kahler manifolds, become particularly significant. These
manifolds form an important class of examples of symplectic manifolds. In
46 2. Symplectic Manifolds

order to introduce them, we need the material of Section 1.4. There, among
other things, we defined:
A complex structure on a 2n--dimension R vector space V is a J E
Aut V with J2 = -idv.
-- A symplectic R vector space (V, w) is called Kdhler, if it has an w-
cvmpatible complex structure J (with J E Sp(V)) which satisfies
w (v, Jv) > 0.

it follows from the discussion in [K] and [We], that for a C


vector space W the following sets are canonically isomorphic:
The set of hermitian forms h on W; that is, the set of h : W x W -
C which are sesquilinear in h(x, y) = h(y, x) for all x and y E W.
The set of symmetric R-bilinear forms g on W that are invariant
under multiplication with i; that is, those R-bilinear maps g : W x
W -+ R satisfying g(x, y) = g(y, x) = g(ix, iy) for all x and y E W.
The set of antisymmetric R-bilinear forms w on W that are in-
variant under multiplication by i; that is, those R-bilinear maps
g : W x W - R satisfying w(x, y) = -w(y, x) = w(ix, iy) for all x
andyEW.
The isomorphisms are given by
g=Reh, w=-Imh,
h(x, y) = g(x, y) + ig(x, iy) = w(ix, y) + iw(x, y).
Thus h is positive definite precisely when g is.
It is now natural to generalize the question asked in Section 1.4 of
whether and in how many ways a given R vector space can be supplied
with a complex structure for a given real manifold M. The answer to this
question depends on the answer to the question of whether every real tan-
gent space T,,, M R2n can be supplied with a complex structure Jm so
that these structures vary smoothly from point to point (more precisely:
they satisfy an integrability condition). We will not pursue this question
further here, but will assume that M comes as a complex n-manifold. Then
the tangent space TmM ^ C", as an R vector space, has in a natural sense
a complex structure J,,,. This will correspond to the choice of local coordi-
nates zj = xj + iyj (j = 1, ... , n) and the concurrent identification of the
basis
8 _ 1 8 8
m-2 9xjlm-t jlm =1,...,n,
2.4. Kahler manifolds 47

of T,mM as C vector space with the basis


a a _
n,

as R vector space. The multiplication by i = in T,,, M as C vector


space will be given by the map J,,,,
IM,
Jm (7Im) 55-1

8
Jn, (Im) i:i; =1,...,n.
The compatibility of these structures J,,, with the holomorphic coordinate
transformation functions allows one to give a formulation of the above-
mentioned integrability condition. We will not go into this here (see CHERN
[Ch], p. 14, for details). We continue with the definitions.
DEFINITION 2.9. A complex n-manifold M with a symplectic structure
(as real 2n-manifold) is called a Kahler manifold, if for every point m E Al
the l[t vector space (T,,,M, w,,,, Jm) is Kahler.
This is equivalent to the description given at the beginning of this sec-
tion:
DEFINITION 2.10. Let M be a complex n-manifold with a hermitian
metric g. Then M is a Kohler manifold if the skew symmetric bilinear form
w (. , ) := g (J., ) is as an exterior 2--form, a closed differential form on M.

To this we add a few points of clarification. By a hermitian metric g


we mean, as in AEBISCHER et al. ([Ae], p. 24), a Riemanian metric such
that for every point m E M, g,,, is a J,,,-invariant inner product on the
2n-dimensional R vector space TmM, and then J,n is compatible with gm
in the sense that
g,n(Jmv, Jmw) = g,n(v, w) for all v, w E TmM.
As already seen in Section 1.4, g,,, is, in the usual sense, the real part of a
hermitian scalar product on TmM as n--dimensional C vector space, namely
( ')h = h(. , .) = 9m(- , ) +
,

In the case when the differential form w associated to this g,n is closed, we
call it a Kahler metric. The findings from Section 1.4 on the connection
between skew-symmetric bilinear forms w and symmetric ones g apply here
immediately to verify the equivalence of Definitions 2.9 and 2.10.
In practice we find ourselves with the following procedure: we are given
a complex n-manifold M on which there is a Riemannian metric g, so that
g,n and J,n are compatible at every point m E M. This means a non-
degenerate 2-form w is also given. To see M as Kahler, and so also as
48 2. Symplectic Manifolds

symplectic, we must prove that dw = 0. A useful criterion in this situation


is due to MUMFORD (IMu], p. 87). Let 0 be a group of diffeomorphisms
acting on M, which under the operation
GxM -b M,
(g, m) O9(m) _: gm,
leave the complex structure and the metric h unchanged. For m E M denote
by
Gm= (9EG;gm=m)
the isotropy group of in. Then 09 induces, for each g E Gm, a map
((4g)=)m : TmM ~ TmM,
and so a representation Pm of Gm in TmM, thus a homomorphism
Pm : Gm - Auti(T,,,M).
We have
THEOREM 2.11. (Mwnford's criterion) If Jm E e n(Gm) for all m E M,
then dw = 0.

Proof. Since G leaves the complex structure and the metric fixed. G
also leaves w and hence dw unchanged. Therefore, for all g E Gm and
u, v, w E 7,n M..
dwm(p,n(9)u, 8m(g)v, gm(9)W) = dwm(u, V, w).
Here setting p (g) = .In, and applying the formula twice yields
dwm(u, v, w) _m(Jmu, Jmv, J,nu) = clw.n(Jmt+, Jmu, Jmw)
= -v, -w) _ -dwm(u, v, w) = 0.
0
This criterion can be used to show that complex projective space P (C"+I) _
CP" is a Kahler manifold (see Section 2.6).
AEBIsCHER et al. ((Ae], pp. 27 ff.) go on to prove the following crite-
rion. The condition d.o = 0 is equivalent to
VxJ = O for all X E r (TM).
By V, we here mean the connection corresponding to the R.iemannian metric
g (see Section A.4); by X a vector field, thus a global section of the tangent
bundle; and the complex structure J appears as a tensor field of type (1,1)1
thus as a global section of T(RI)M
2.4. Kahler manifolds 49

They ([Ae], pp. 28 ff.) then use this criterion to show that the unit ball
in C",
B" :_ {z E C", [[zj[ < 1},
with the aid of the Bergmann-metric with the kernel
n! 1
K,,(z,tP) a" (1-Zip)"+lz,iPE B",

is equipped with a Kahler metric.

Towards a formalism of complex differential forms: the Kffhler


form. In texts in which complex manifolds are the central theme (see
CHERN ([Ch], p. 53), WEIL ([We], p. 41) or KAHLER [K]), the description
of Kahler manifolds uses the standard formalism of the real differential forms
as described in Section A.4,
a = E b;,...,Qdx;, A...Adx;?
extended to the complexes. Following the Wirtinger calculus of function
theory (see FIsCHER-LIES [FL], pp. 22-23), we assign to the complex co-
ordinates zj = xj + iyj (j = 1, ... , n) the symbols
(*) dzj = dx; + idyj, dzj = dxj - idy;,
as well as the differential operators known from the Cauchy-Riemann dif-
ferential equations,
a 1 a a a 1 a a
,
at; 2 axi i ft, azj = 2 \ ax; + = ayj
and define differential forms, for example
f! = E cjkdzj A dz,t (Cjk are C-valued functions.).
j.k
This is called a form of type (1,1), since it is homogeneous of degree 1 in the
dzj and the dz-k. Such a form is called closed when dig = 0, where here
a&jk
d=a+$ with 0910=F azt dzt Adz; A dzk
j,k.l

and &I = ac dz1 A dzj A dzk.


j,kj
This will now be used in an application. A complex differential manifold Al
of complex dimension n is called hermitian if TM has a hermitian structure;
that is, for every point m E M the n-dimensional C vector space T,mM has
a hermitian scalar product (, )," assigned in a smooth way. For a chart
50 2. Symplectic Manifolds

(cp, U) with coordinates z = (z1, ... , z,.) a positive definite hermitian matrix
is defined, for each m E U, by

H = (Hjk) with Hjk = U' k = 1, .... n).


axj , azk
To this matrix can then be associated the (1,1)-form
S2 = (i/2) E Hjkdzj A dzk,
j,k
which CHERN called the Kahter form. 51 is clearly real in the sense that we
have

_ -(i/2) Hjkdzj A dzk = (i/2) 1: Hkjdzk A dzj = R.


j,k j,k
As an easy variant (see below) of Definition 2.10, we have
DEFINITION 2.12. M is called a Kahler manifold if M is hermitian and
the corresponding Kiihler form iZ is closed.
Of particular importance, especially for applications of Kiihler forms,
is the fact that f) is closed exactly when D locally has a potential; more
precisely,
THEOREM 2.13. Let M be hermitian with a Kahler form f1. Then M is
Kahler precisely when there is locally an R-valued differentiable function f
satisfying
n = i8f.

EXERCISE 2.14. Prove this. (A proof can be found in KAHI.ER [K], as


well as in CHERN ((ChJ, p. 56).

The connection of Definition 2.10 with the real theory described above
is that the Kiihler form via (*) can be transformed into a real differential
form.
EXERCISE 2.15. By a short calculation, show that
0=- 8jk(dxj A dxk + dyj A dyk) + ajkdxj A dyk.
j <k j,k

Here the (ajk) = A = ReH form a symmetric matrix, while the (Ojk) _
B = Im H form a skew-symmetric one. We have
H(z, z') = tzHz'
= txAx' + tyAy' + txBy' - tyBa'
+ i(txBx' + tyBy' + tyAx' - txAy')
2.5. Coadjoint orbits 51

This can then be likened to the earlier real theory. There we spoke of a
real compatible metric g on Al as a 2n -dimensional real manifold with a
complex structure J. thus with
g(e', w) = g(w, v) = g(Jv. Jw) for all v. u' E T..Al R2".
When the matrix of g is replaced by

9=(C D) and J=(-1


1 10

we arrive at the condition


g= (-B A
with A = to and B = -tB.
Then. from tv = (tx.ty),ty' _ (tx'.ty'). tz = tx+ity and tz' = tx'+i'y', we
have
H(z. z') = g(v, w) + ig(Jv, w).
The written formalism of the symmetric form g gives the assignment of the
skew symmetric form as
,O(v. U') = g(Jv. w).
After this is multiplied by 1/2 it has the associated differential form
ajk(dxj Adxk+dyj Ady)L) - >ajkdxj A dyk
j <k j.k
This is not exactly the above Kiihler form f2 for H. but rather that for H.
WEIL, by the way, associated to H the (1.1) form
i/2 hjkdij A dzk.
j.k
(WEIL ((We]. pp. 15 and 41)). and this is then in full agreement with the
real formalism.

2.5. Coadjoint orbits


This section will need to lean more heavily on the representation theory of
Lie groups and algebras (see Appendix D) and on the theory of systems of
differential equations on manifolds than either the preceding or the following
sections will. However, courage will soon be rewarded, for we will gain a
better description and the means for constructing many more manifolds; this
will be needed in the discussion of the moment map and of quantization.
Coadjoint orbits arise in a natural way for any given Lie group G. The
group operates via a coadjoint representation Ad' on the dual space g' of
the Lie algebra g of G. The orbits of this action are called coadjoint orbits
and can (under known conditions) be made into symplectic manifolds. More
52 2. Symplectic Manifolds

precisely, the goal of this section will be to discuss the following theorem of
Kostant and Souriau
Let G be a Lie group with H1(g) = H2(g) = {0} for g = Lie G. (Here
by Hk(g) we mean the k-th cohomology group of g (see Appendix Q.
Then there is (up to covering) a one-to-one correspondence between the
symplectic manifolds with transitive G-operation and G-orbits in g*.
The study of the coadjoint orbits was introduced by KIRILLOV, and
the reader may find in [Ki], pp. 226 if., a readable introduction to this
topic. Here though, we will follow the treatment of AEBISCHER et al. ([AeJ,
pp. 32-39), where one may find a summary of the detailed treatment of
GUILLEMIN-STERNBERG ([GS], pp. 172 ff.).

The coadjoint representation on f24(G) AQg*. For what follows,


we fix G to be a Lie group and g its Lie algebra, which we identify with
TeG or with the space VI(M) of all left-invariant vector fields X on C (for
preliminaries, see Appendix B). Analogously, g* with TeG is identified with
the space of left-invariant differential forms of degree 1, and from this it
then follows that the space f2q(G) of left-invariant q-forms can be identified
with AQg*, the space of alternating q-forms on g.
Here is a good time to make the following fundamental observations.

Remark 2.16. Via this identification of I (G) with AQg*, the operators
of exterior differentiation on 52l (G) are exactly the coboundary operators
6 on AQg* defined in Section C.2. Thus, in particular, the space Z2(g) of
2-cycles in A'g* can be identified with the space of left-invariant closed
differentials on G.

Proof. i) Here, as in much of what follows, the Lie derivative acts on


differential forms and vector fields of a manifold M. This is described in
Sections A.4 and B.2. These actions are in fact linked according to the
following formula. For a E 0Q(M) and X1, ..., XQ E )(M) we have (see
ABRAHAM-MARSDEN ([AK, p. 117))

(Lxc)(X1,...,Xq) = LX(a(Xi,...,Xq))
(*) q
a(Xi, ... , [X, Xd,... , Xq).
i=1

As described in Section A.4 and as a special case of Lemma 2.3, the Lie
derivative is related to inner multiplication through the formula

(**) LXa = d(i(X)a) +i(X)da.


2.5. Coadjoint orbits 53

ii) Now let a be an element of f2' (G) and X. Y elements of LA(M). Then
a(Y) is constant on G and, therefore.
Lx(a(Y)) = 0.
The formula (*) for q = 1 then says that
(Lxa)(Y) + a([X, Y]) = 0.
Since i(X)a = a(X) is also constant on G. it further follows from (**) that.
(i(X)da)(Y) = -a([X. Y]).
This can then be formulated as
da(X, Y) = -a([X. YJ).
and with this we have the formula for 5. when a is taken as an element of g.
iii) Let. w be an element of f22 (G) and X. Y, Z elements of Vj(G). Then
w(Y. Z) is constant, and therefore
Lxw(Y. Z) = 0.
The equation (*) for q = 2 then gives
(Lxw)(Y Z) + w([X. Y]. Z) + w(Y, [X. Z]) = 0.
Front (**) we arrive at
(Lxw)(Y, Z) = (i(X)dw)(Y. Z) + d(i(X)w)(Y. Z).
which then. using the result of part ii) for a = i(X)w. gives
(Lxw)(Y, Z) = dw(X. Y. Z) - w(X. [Y, ZJ).
Putting these two statements together. we get
dw(X. Y, Z) = -w([X, YJ, Z) +w([X. ZJ. Y) - w.'(IY. Z], X),
thus the formula for b for q = 2.
iv) Through further and analogous iteration we get the formula for the
general case.
EXERCISE 2.17. Verify the formula (*) for q = 1 and 2. and fill in the
details for step iv) of the proof.

The coadjoint representation on flq (G) can now be easily described:


conjugation with go E G acts as an inner automorphism of G:
Kg, G --. G.
9 '' 9og90
Kg induces a map of the tangent spaces: in particular. for the case TeG = g
Ad(go) := g g.
2. Symplectic Manifolds

This map gives, through a roughly similar mechanism, the homomorphism


Ad : G -+ Aut g,
go -Ad go,
which is called the adjoint representation of G. This gives rise (see Sec-
tion D.1) to the coadjoint representation
Ad' : G - Aut g`.
For q > 1, this induces a representation
Ad':G-+ AutAQg',
which is also often denoted by the symbol Ad#. In those cases where AQg' is
identified with the space of left invariant q--forms w on G, we get the formula

Ad'(go)w = (rigo)*w = (eyu)'w


with a the right translation on G.
Unlike KnRtLLOV [KI , we do not study the coadjoint orbits in g' directly,
that is, the G-orbits in the coadjoint representation in g'
G#ry :_ {Ad'(go)t9; go E G} for V E g',
but rather study the orbits in A29 .
We begin with some notation. Let (M, w) be a symplectic manifold on
which G operates on the left as differentiable maps
GxM M, (g, m) i-+ gm.
Then we may consider two maps
Og:M-+M with0g(m):=gm forallgEG
and
bm:G--+M withryn(g):=gm forallmEM.
In the situation where w is fixed by each 0g, g E G, we then have
09*w=w forallgEG.
We call this given operation of G on M a symplectic operation.
For the right translation p and the left translation A we clearly have
Og 0 10m = 10m o ag,
-Ogm = OM 0 Log.
With the help of 0,, forms can be pulled back from M to G; in particular,
the closed form w on M induces a dosed form on G. More precisely, we
have the following statement:
2.5. Coadjoint orbits 55

THEOREM 2.18.
a) A symplectic group operation G x M M defines a map
41 : M Z2(9),
in -' Winw,
with

(#) T(gm)= Ad(9)(`I'(m))

b) 41(M) is the union of G-orbits in Z2(g).


In the case that the operation of G is transitive, the image of 4(M)
consists of one orbit. Because of the commutation rule is also called
a C-morphism.
Proof. a) ty;,, w, as the image of a 2-form, is itself a 2-form on G. It is
left-invariant, and so we have

'k is a G-morphism, and so we have


vlpm w = (hm o Log)t w = Lo.*q , w = P9'I'(m) = Ad' (9)41(m).
From the fact that 1(m) = ipM w is closed, we have
d (rt'm w) = >Gmdw = 0.

b) The orbit through the point 41(m) has the form


G* (m) := {'P(gm) = Ad`(g)4' (m); g E G} (also =: G#T(m)).
Clearly,
$(M) = U G#(m),
mEM
and for a transitive operation also 11(M) = G#(m). O

Here we now immediately pose, for w E Z2(g),

QUESTION 2.19. Are there, for a given G-orbit G#w in Z2(g), a sym-
plectic manifold M and a map 4 : M Z2(g) as above with G#w = 41(M)?

This gives us cause to ask whether M can be constructed as a homoge-


neous space of the form M = G/H, where H is a closed Lie subgoup of G.
To find such an H, the difficulty will be in showing that a symplectic form
w on G/H can be defined, so that for the given closed w E Z2(g) we have
56 2. Symplectic Manifolds

pr'w = w, where pr: G - G/H is the natural projection. So let w E Z2(g)


be a closed form on G, and set
l) := {X E g; i (X) w = 0}.

Remark 2.20. hW is a subalgebra; 4w = {0} when w is non-degenerate.

Proof. The last statement follows immediately from the definition of an


inner product i (X) w. To prove the first statement we need to show that for
X, Y E h, we have [X, YJ E 4, After the application of the Lie derivative,
this has the same appearance as the proof of the previous remark. In detail,
for Z E g in terms of the formulas (*) used there, we have
0 = Ly(w(X, Z)) = (Lyw)(X, Z) + w([Y, XJ, Z) + w(X, [Y, Z)).
and in terms of (**)
Lyw = i(Y)dw + d(i(Y)w) = 0,
implying that w is closed. Both statements together then give, for X, Y E
w([Y.XJ,Z)=0,
and so [X, YJ E lj ,. D

One of the central theorems of the theory of Lie groups now says that,
given such an l,,,, there exists, up to covering, exactly one connected sub-
group H,,, C G with Lie H. ^_- 4, In this case, this delivers the desired
Hw.

THEOREM 2.21. Suppose that H., is closed. Then there is exactly one
symplcctic form ;a on X := G/H, with w = pr* , where
pr:G-G/H4,=Mw
is the canonical projection.

Proof. Here again we will need some external help. We begin by showing
that the condition that H is closed implies that G/H;,, = M,,, is really a
differentiable manifold. We note first that the orbits aH4J (which is an
abbreviation for ai(H,,,)), a E G, generate the points of M, At the least,
this gives a locally transparent description of M,,, in coordinates, in which
the orbit aH, of a E C in G can be seen as an integral manifold of a
differential system A (also called a distribution). Here we can use a few
concepts and the central statement from the theory of systems of (partial)
differential equations (as can be found in the text STERNBERG ([St), p. 130),
for example):
2.5. Coadjoint orbits 57

i) A c-dimensional differential system A on an n-dimensional differen-


tiable manifold G (which does not necessarily need to be, as in this special
case, a Lie group) is, for 1 < c < n, a map which assigns to each a E G a
c-dimensional subspace A (a) of TUG:
GE)a-A(a)CTUG and dimA(a)=c.

ii) Such a system A is called smooth if for each a E G, there exist (a)
a neighborhood U(a) of a in G, and (b) smooth vector fields Z1..... ZZ on
U(a), such that for every m E U(a) the vectors (Zl )m, ... , (Z,),,, form a
basis of A (p) (this condition takes on a much more elegant form in the
language of vector bundles).

iii) A differential system A is said to be involutive if and only if A is


smooth and
[X,YJEA for allX,YE.A,
where XEAmeans that XaEA(a)for all aEG.
iv) A submanifold N ' M of M is called an integral manifold of A, if
i.(TqN) = A (i(q)) for all q E N.
This thus means that the tangent space TqN of N is, at every point q E
N, isomorphic to the given subspace A (i (q)) of the tangent spaces T,igiG
given by the differential system A. (It would perhaps be better to call this
a maximal integral manifold, since then it would make sense to consider
objects N with i.(TTN) C A (i (q)).)
A criterion for the existence of these maximal integral manifolds is de-
livered by the following:
THEOREM 2.22. (Frobenius' Theorem) Let .A be a c-dimensional smooth
differential system on a differentiable n -manifold G. Then:
a) A is involutive exactly when every point of a E G lies in an integral
manifold N = Na.
b) Should a) be satisfied, there are local coordinates xl,... , x so that
the integral manifold has the form
N = {xlx; is constant, i = 1, ... , k}, k = n - c.
This form of Frobenius' theorem, so written with the help of vector fields.
has a mirror image in the world of differential forms (which historically came
first). A proof of Frobenius' theorem in this form can be found in KAHLER
[K1J.
58 2. Symplectic Manifolds

v) Now we return yet again to the special situation of the proof of The-
orem 2.21. In particular, we want to interpret the Lie algebras as subspaces
of tangent spaces in order to define a differential system D by
A (e) := f) for the identity element e E G
and

(a) :_ for a E G with the left translation Aa.


From Remark 2.20, it immediately follows that this 0 is involutive and,
from iv), that every point a E G has an integral manifold N.. These are
clearly given by Na = aH,,,, and further, from Frobenius' theorem (part b),
Na, at every a E G, can be locally written as
Na = {xlxi is constant , i = 1, ... , k}, k = n - dim fk,.
The tangent space TzNa then has as basis
a a
OXk+l,...,.
Now the given closed 2-form w on G has the form
n
W = E a=j(x) dxi A dxj.
iv
It follows from the construction of fj that i (X) w = 0 for X E h,,,
thus, in particular, for X = -, i = k + 1, ... , n. This and the fact that
dw = 0 is closed show that w, in these coordinates, can be dependent only
on xl,...,xk. Thus
k
w =1: a+S(xl,...,xk)dx;Adxf.

i<j
With this, we are essentially done. The integral manifolds Na of 0 are
the fibers of the projection pr : G M,,, = G/H,,, and M,,, is described in
the local coordinates (xi.... , xk). For
k
0:= aij(xl,...,xk)&, Adxj
s<j
we naturally have pr`w = w. It is now routine to see that in this manner
also a global form 0 on M,,, with the desired properties can be given. 0
In GUILLEMIN-STERNBERG ((GS], p. 174), Theorem 2.21 is given for
a somewhat more general case. In this generality it lays the groundwork
for the so-called symplectic reduction, which has the goal of projecting a
given manifold onto a (smaller) symplectic manifold by exploiting additional
2.5. Coadjoint orbits 59

symmetries. We will give a presentation of this material in Section 4.3. The


proof uses the same ideas and is only a little harder.
THEoR.EM 2.23. Let M be a differentiable manifold and w a closed 2-
form on M. Then the vector space of vector fields X on M with
(*) i(X)w-0
is closed with respect to the Lie bracket. In the case that the dimension of
the space of those X satisfying (*) is constant at every point m E M, w
defines an integrable differential system A. Should there, further, be given
a manifold MO and a submersion p : M MO (that is, p is a differentiable
surjection which induces surjective maps on the tangent spaces) so that the
integral manifolds of A are the fibers of p, then there is exactly one symplectic
form wo on Mn with p`wo = w.

The symplectic manifold of an orbit G. With this last theorem,


we can now answer Question 2.19. For a given orbit G#w in Z2(g), H,,,
and (since under the presumptions H, is closed) the manifold M, are fixed.
Then G#w is the image of M . under the map T. Thus
'I+(M')=G#w.

Proof. When mo = eH.. is interpreted as a point of M, we have, for


a E G. with the canonical projection pr : G -+ M . = G/H and in the
notation introduced at the beginning of the section,
pra=aH,, =amo=V,.w(a);
thus pr When we further define T : MW -+ Z2(g) as in Theorem 2.18
by

for w as in Theorem 2.21, we get


T (mo) = pr`w = w.
Since ' is a G-morphism, we conclude that
T (amo) = Ad* (a) +l'(m)) = Ad* (a) w
and so
4, (M')=G#w,
since C operates transitively on M,,,. O

Now that this construction of Mme; for a given G-orbit of w E Z2(g) for
closed Hw has positively answered Question 2.19, we naturally turn to the
question of its uniqueness
60 2. Symplectic Manifolds

QUESTION 2.24. Is there more than one homogeneous symplectic man-


ifold attached to a given orbit G*w?

To answer this question, we let a symplectic manifold M = G/H with


symplectic form fl be given, where H in G is a closed but not necessarily
connected subgroup, and with a symplectic transitive operation G x M
M. Then this operation induces, for all b E G, a map
bbH : G -+ M,
a ' - ObH (a) = abH.
By Theorem 2.18, there further exists a map
W:M-*Z2(9)
We then set
w:=T(eH)=i,*H11,
and because of the transitivity of the G-operation we have that
41 (M) = {Ad`(a) w; a E G} = G*w
consists of but a single G--orbit. The uniqueness can now be positively
answered, since, by the details of the proof of Theorem 2.21, we have that
M,,, is the same as the given M. To see this, as in the proof of Theorem 2.21,
for g = Lie G let
h,,:_ {X E g; i (X) w = 0}.
Thus
hw = {X E g; n (X, ') = 0},
and since Q is nondegenerate, we also have
h,o={X E9;
But this shows that h := Lie H = 44 since
(WeH).Io:g=TeG - TTHM=9/h,
X -X+h
means that X E his equivalent to (ikeH),X = 0. As things are now arranged,
H, is the connected subgroup of C associated to h,,, = h, thus the connected
component of unity in H. And so M,,, = C/H,,,, when not the same as
M = G/H, is then shown to be a covering of M. With this, we have now
proved

THEOREM 2.25. Homogeneous sympleetic manifolds G/H are, up to cov-


ering, parametrized by the C-orbits G*w in Z2(g), if a given w gives rise
to a closed subgroup H, in G.
From these results the statement made at the beginning of the section
quickly follows:
2.5. Coadjoint orbits 61

THEOREM 2.26. (Kostant-Souriau) For H'(g) = H2(g) = {0} there is,


up to covering, a one-to-one correspondence between the symplectic mani-
folds for G and the G-orbits in g'.
Proof. i) The cohomology groups Hk(g) are introduced in Section C.2.
H2(g) = {0} says that for every w E Z2(g) there is a,3 E g' with d,3 = w.
In a similar vein, H' (g) = {0} is equivalent to Z' (g) = B' (g). But B' (g) _
d(A°g') = {0} (since A°g' = R), and therefore d,3 = d8' implies /3'
and so 0 E g' with d/3 = w uniquely given by w.
ii) This says that there is a one-to-one correspondence between the
G-orbits in Z2(g)
G#w = {Ad'(a)w; a E G}
and the G-orbits in g'
C#(3 = {Ad* (a),3; a E G}.
Thus, a bijective map
Ad'(a),3"Ad'(a)w for all a E C
is given, and it follows that
d (Ad'(a)i3) = d (p.3) = en w = Ad'(a):.r.

iii) The isotropy group


G,3 := {a E G; Ad* (a) 0 = 3}
is per se closed. Then H,,, is also shown to be closed as soon as it is shown
that H,,, is the connected component of unity of C3. We have
gyLie G3= {X Eg; Lx,3=0},
and thus /3 = Ad'(a),3 = g,',3 is, by the definition of the Lie derivative, for
a = exp X equivalent to Lx(3 = 0. This is further equivalent to i (X) w = 0.
Then for X E g, Lemma 2.3 says that, in connection with the formula (#+)
in the proof of Remark 2.16,
Lx,3 = i(X)d/3+ d (i(X),3);
thus
Lx)3 =i(X)w,
since i (X) /3 is locally constant (because X and /3 both are left-invariant).
From this we now get that
9.1 ={XEg;
and so H, is contained in C,3 and therefore closed.
62 2. Symplectic Manifolds

A further criterion that M, = G/HW really be a manifold was discovered


by CHU. It says that H,,, is closed when G is simply connected (for a proof,
see GUILLEMIN-STERNBERG ([GS], p. 179)). Of greater practical signifi-
cance for giving explicit symplectic forms is the following theorem, where a
coadjoint orbit in g" also shows up.
THEOREM 2.27. Let G be a Lie group with H1(9) = Ha(g) = {0), 0 E
g', w = d,9, and
43 =G/H4J-G*w-G08
the associated symplectic manifold to the form D. Then it follows that
(X3, 1Y.8) = -fj ([X, Y])
with
X0.= pr.X, Y,9:= for X,Y E g.
Here again, pr: C - G/H;, is the canonical projection.
Proof. From Theorem 2.21, w as a symplectic form on G/H,,, is uniquely
determined by pr* 0 = w. On the one hand, we have
iv (Xq, Ys) = iv (pr.X, pr.Y) = (pr` w)(X, Y) = w (X, Y).
On the other hand, as given in step ii) of the proof of Remark 2.16,
-0 ([X, Y]) = dQ (X, Y) = w (X, Y),
which can here be recognized as the formula for the coboundary operator 6
in the cohomology theory in Section C.2. 0
We close this section with an example which should give a first feeling
for the power of Theorem 2.27: one can recover the volume form wo of the
two-dimensional sphere S2 in the following way.
EXERCISE 2.28. Take G = SO(3) and identify S2 with an orbit G*# for
0 E g' = so(3)*, and use Theorem 2.27 to define a symplectic form rv on
M,3 = SO(3)/SO(2) S2.

Hint. As indicated in B.2 and used more explicitly in Example 4.22, the
elements B E g = so(3) can be identified with b E R3 in such a way that the
Lie products [B, B'] correspond to the vector products b x Y. In this way,
we come up with a form
w = xl dxa A dx3 + x2dx3 A dx1 + x3dxj A dxa
which, if we put coordinates on S2 by
xI = coscicos
X2 = cos ti sin Q5,
X3 = sin e9,
2.6. Complex projective space 63

pulls back to
wo = cos t9d0 A dt9.

Remark 2.29. In the same way, one can construct symplectic manifolds
as orbits G#3 for the Galilei and the Poincare group which are of great
importance for physics. One is tempted to do it here, but it would take too
much space. The interested reader may consult GUILLEMIN-STERNBERG
[GS], pp. 122--130, 437--445, or SouRIAU [So], pp. 144-192, and also the
thesis of M.A. EL GRADECHI [EG].

2.6. Complex projective space


From the treatment of coadjoint orbits, one can build the moment map as
in GUILLEMIN-STERNBERG [GS]. But we will deal with this central topic
later. Now we turn our attention to another concrete example which will
serve to make the previous concepts and techniques somewhat more explicit.
Later, the study of the moment map will give us cause to consider further
examples.
Let P" = P" (C) = P (C "+') be complex projective space, thus the space
of all (complex!) lines passing through the origin in C"+1 and denote by ;r
the map
Tr : C"+1 \{0} - P",
z = (zo, , z") - Z--
As in Section A.1, z. will mean the equivalence class gotten by setting z - z'
whenever z; = \zi, i = 0, ... , n, for a,\ E C. P" is now an example of a real
symplectic 2n--manifold. This can be shown in a variety of ways; we shall
do so by demonstrating that P" is Kiihler (and therefore, from Section 2.4,
symplectic). This itself can be demonstrated in a variety of ways.
i) In MUMFORD ([Mu], pp. 86-87) the following metric is introduced on
IP". Let V = VHerm denote the space of (n + 1)-rowed hennitian matrices,
V = {AEM"+1(C); to=A},
thought of as an R vector space, which then has dimension (n + 1)2. A
differentiable map ¢ can be given by
P" V,
z- '- (Aj) = (ztzj) for z with IZ12 = 1.
The group U(n + 1) _ IS E GL"+I(C); £SS = E"+1} operates on V by
conjugation and on P" by
(S, z.,.) e--- S(z-) :_ (Sz)..
64 2. Symplectic Manifolds

0 is then equivariant for the operation:


0 (S (z-)) = SO(z,)S-1.
V has a U(n + l)-invariant positive definite symmetric standard bilinear
form q, given by
q (A, B) = tr(AB).
induces at every point z. = 7r (z) a map of the tangent spaces
(0.)(:) : TT(.)P" T .,-)V = V.
By pulling back, this gives a U(n + 1)-invariant Riemannian metric on P"
as real manifold. A rather brutal computation (as conceded by Mumford)
shows that the associated (see Section 2.4) hermitian metric on the complex
manifold, called the Ftebini-Study metric, in the coordinates
xi=z;/zo, i=1,...,n, for zo#0,
has the form
ds2 = EdXi - dx,)
1+F,E1xj12 (1+EIxjI2)2
Here we give a few steps of the calculation for the enlightment of the inter-
ested reader. To
C" P- (C) V,

x
. x )- =
z- '' 1 +IIxI2 (x
with x as column, is associated the map of tangent spaces

C" = TzC" io* T;a(z)P" "' T (=)v = V,


which takes the tangent vector u to the curve -y (t) = x + ut in C" to the
tangent vector to the image curve (4 y)(t) at t = 0 in the space V. This
vector can be transformed (with (x, u) = Exiu;) to
a (u) := ¢.io.u =

1 + Ixl2u0 ) (1' )+\x/(0'`)


1
(u'1+Ixla'u)(x)(i)).

'
Then as a bilinear form on T(=)P" we take the pullback of the bilinear form
living on V; thus
g (u, u) = tr(a (u) a (u)).
From this, after a little calculation involving substitutions of the form

tr((I )(1,r)()(1,ii))=(1+(x,u))2.
2.6. Complex projective space 65

we arrive at

9 (u, u') =
(u, u') + (u', u)
1+ Ix12
- (u, x) (x,(1u')+ (x12)2
+ (x, u) (u', x)

This can now be recognized as the two-fold real part of the hermitian metric
given above (after substitution of it = ei and u' = e1).

ii) In AEaISCHER et al. ([Ae], p. 40), this metric is carried over to the
definition of Pn as the space of one-dimensional subspaces of Cn+i. Then
the map
Cn+1\{0} * Pn D Uo = {z.,,, zo # 0} - C°,
X,

with xi = zi/zo for (i = 1, ... , n), induces the map of tangent spaces
Cn+1 = TZCn+1 T=lm TZCn = Cn,

Since any { 0 can be interpreted as a tangent vector to the curve


t i-- ry (t) = z + t. in C"+1
it can be shown that C is the tangent vector at t = 0 to the image curve
t ,-+ 7o(t) = (4P0l1`)(t) in Cn
via the transformation
C = (Cl, ... I W with (i = &/zo - Cozi/zo
For two such tangent vectors C and (' , we can, in the following "natural"
way, find a hermitian scalar product ((, )). On Cn+1 the usual hermitian
scalar product is given by the formula
n
(c e) = F, &i.
i=0
Denote by W the orthogonal complement in T2(Cn+1) = Cn+1 of the lines
passing through X. Then
(zC)1=:W25 C",
and W can be identified with TXCn = Cn since a.(C"+1) = ir.(W). So we
have = cz +' for a uniquely defined c = (C, z)/(z, z) E C and n E W with
(cPo r).W = W O-07) = C.
Clearly,
(z, z)£ - (C, z)z
66 2. Symplectic Manifolds

and it follows that


(z, z) - (C z)(z, f')
(rl, n)
(Z, Z)
Now as hermitian scalar product for two tangent vectors in Tf(:)P" with the
coordinate vectors C. respectively C', we take

(z, z)
Here we can substitute for , respectively i;', as well as z an (n + 1)-tuple
with (;poir).(l;) = ( as well as cpon (z) = x (that is, & = (;zo, Co = 0), and
then we get straightaway the form of the metric given above in i).
The formalism developed in Section 2.4 can now be applied to
h( , ):=Re((,))
This then says that this is a Riemannian metric on P" which is J-invariant;
that is, for the complex structure J for which JIT(z) operates on TT(.)P" as
multiplication by i = v1---1 on W, we have
hp(Jpv, Jw) = hp(v, w) for p = 7r (z).
To show that the exterior 2-form w associated to h is closed, one uses Mum-
ford's criterion from Section 2.4. To this end, let G be the group of diffeo-
morphisms of P" which leave the complex structure and h invariant; thus
G = SU(n + 1). (Here, we are exploiting the classical isomorphism
P" SU(n + 1)/U(n).)
It is then observed that
SU(n + z U(W)
and that the representation pp in Mumford's criterion, Theorem 2.11, is here
given by
B,r(z) : SU (n + U (W).
Then we clearly have JI,r(z) = iE,i E U(W), and the criterion says that
(- ,)=h
is closed, thus that P" is Kiibler. Mumford proves this directly in (Mu],
p. 87. In AEBiscHER et al. [Ae] this is crystalized to the criterion of
Theorem 2.11.
iii) Thus with a little calculation one not only arrives at the metric but
also gets a good feeling as to its origin. On the other hand, one may arrive
at the metric a little more quickly using the formalism of Definition 2.12, by
2.6. Complex projective space 67

giving a potential from which the Kahler form is gotten through differenti-
ation. And this is given on U; = {(z).,,, z; # 0} C P", with the coordinates
xEC",by
I
f = log K with K(x, T) = + > xJaJ.
J=1
Then, in the formalism of Section 2.4, this can be written as
r( ! f
,Oaf tlxjAdxj
K
Y n xkdk
K2
i=I J,k
and this can be interpreted as a Kahler form from a hermitian matrix, which
belongs to the Fubini--Study metrics given earlier.
iv) It is then pretty nice to demonstrate (AEBISCHER et al. ([Ae], pp.
41 ff.)) that P" can also be realized as a coadjoint orbit. Here again the
underlying group is naturally G = SU(n + 1). It can be shown that
9=su(n+1)={AEg((n+1,C); `A=-A, trA=O}
is the associated Lie algebra. On the vector space of these traceless skew
hermitian matrices, we can give a scalar product (,) by the formula
(A, B) = Re(tr AFB).
This makes possible an identification
9 - 8w,
A FGA with cpA(B) (A. B),
which we will consider fixed in what follows.
Now we consider the coadjoint orbit in g',
G#B = {Ad'(a) B; a E G} = {a-IBa; a E G},
for the special case B = B0 with

Bo_- aE_E_ _ ((1/n) E 0


0 -1 )
Then we have
G#Bo SU(n + 1)/U(n) - P",
and this is a-'E_a
t0
E_ precisely for

a=( T 0t
Theorem 2.27 now says that the symplectic form won P" is given by
w (X, Y) :_ (-B, IAx, AY))
68 2. Symplectic Manifolds

for AX, AY E g tangent vectors such that


pr. AX = X, pr. AY = Y for pr : SU (n + 1) - SU (n + 1)/U (n).

2.7. Symplectic invariants (a quick view)


On the basis of Darboux's Theorem 2.6, all symplectic manifolds of a fixed
dimension are locally isomorphic. Thus to differentiate between them we
must find attributes of the global object. Here we will give only a hint
of two such attributes. A more precise and fuller account of this active
area of research can be read in HOFER-ZEHNDER ([HZ], chapter 4) and in
AEBISCHER et al. ([Ae], chapter 6). Both treatments rely on the earlier
work of GROMOV, who during his study of symplectic embeddings [Gr]
formulated the following statement.

THEOREM 2.30. (Gromov's squeezing theorem) For r, R E Rio, let


B(r) :_ {(x, y) E R2"; Ix12 + Iy12 < r2}
be the open ball in R' and Z(R) :_ {(x, y) E R2"; x + yl < R2} an open
cylinder. Let 0 be a symplectic embedding of B(r) defined by
q(B(r)) C Z(R).
Then
R > r.

EXERCISE 2.31. Prove this for the case that 0 is linear.

Pseudoholomorphic curves. A proof of the above theorem is given in


GRoMov [Gr] and uses the concept of peeudoholomorphic curves. This
concept, along with some important applications and a proof of Gromov's
theorem, is given by AEBISCHER et al. ([Ae], chapter 6). Let M be an
almost-complex manifold (that is, there is a J E Aut(TM) with J2 = -1
(see also Section 2.4)), and let E be a Riemannian surface, thus a one-
dimensional complex manifold, which is an almost-complex manifold under
the map of tangent spaces induced by multiplication by i on E.
DEFINITION 2.32. A pseudoholomorphic curve is a smooth map f : E -
M for which the induced map
Tf:=f.:TE-.TM
is (i, J)-linear; that is, it satisfies f. o i = J 0 f..
2.7. Symplectic invariants (a quick view) 69

Here the pseudoholomorphic curves are particulary interesting when


there is a symplectic structure w on M which is compatible with the almost
complex structure. As an example we may pick out a statement appearing
in AEBISCHER et at. ([Ae]. Lemma 6.3.6 on p. 128). Let f : E -. Al be a
pseudoholomorphic curve, µ a metric on E in the conformal class of the coin-
plex structure i on E, and g a metric on M. The smooth map f : E Al
induces the map f.: TE - TM, and so, at each point, a homomorphism
TfZ = f.Z : TZE TJ(Z)M, ZEE,
between Euclidean vector spaces. and therefore a norm
II Tf 112= tr(T f )` o T f.
The metric u on E fixes a volume form dr,,. The energy of f is defined by

E(f) := f II Tf 112 d-r,,


E

and the area off by


a(f) = Jfi).
E
The quoted lemma then says that for compatible (w. J. g) we have
E(f) = 2a(f)
and that this is a topological invariant. Using the notation [w] for the de
Rham class of w, this can also be written as ([w], f. (E)). As a consequence.
for example, one may derive that a pseudoholomorphic curve is a minimal
surface relative to the metric.

Symplectic Capacities. In GROMOV [Gr] the following concept is also


introduced and studied.
DEFINITION 2.33. The symplectic radius, rad M, of a symplectic mani-
fold M is the supremum of all r for which there is a symplectic embedding
of the ball B(r) in M.
This motif was adopted and further expanded to the notion of the sym-
plectic capacity by EKELAND, HOFER, VITERBO, ZEHNDER and others. This
was a part of their study of the question of the existence of given closed
curves, called Hamiltonian trajectories (to be defined in the next chapter),
on the hypersurfaces of constant energy on symplectic manifolds. There is
for this situation the following axiomatization (see AEBISCHER et al. ([Ae].
pp. 73-74), or HOFER-ZEHNDER ([HZ]. p. 51)).
70 2. Symplectic Manifolds

DEFINITION 2.34. c is called a symplectic capacity of dimension 2n when


c is a map which assigns to every symplectic 2n-manifold (M, w) (eventually
with boundary) an element of [0, oo] satisfying the following properties:

Cl (Monotonicity): Whenever there exists a symplectic embedding


':(M,w)- (M', we have
c(M, w) < c(M', w').

C2 (Conformality): We have
c(M, aw) = Ialc(M, w) for all a E W.

C3 (Non-triviality): We have, for the standard form wo on R2",


c(B(1), wo) = c(Z(1), wo) = 7r-
The condition C3 can sometimes be weakened to
C3' (Weak non-triviality): We have
0 < c(B(1),wo) and c(Z(1),wo) < oo.
That a symplectic capacity actually is a symplectic invariant follows im-
mediately from condition Cl. These axioms, however, do not fix a uniquely
defined capacity function.
Remark 2.35. 1) It can be shown that the square of the symplectic radius
is a capacity.
2) For n = 1, thus in the case of a two-dimensional symplectic manifold,
the total volume,
c(M,w)=I
M
f U,

is an example of a capacity function, which for (M, w) c (R2, wo) agrees


with the Lebesgue measure. For n > 1, (vol M)1V" is not a capacity function,
because of C3 and the fact that the cylinder has infinite volume.
EXERCISE 2.36. Show that, for an open subset U C R2" and .1 54 0 with
the standard form wo, we have
c(AU, wo) = A2c(U, wo),
and calculate c(B(r), wj).
With the help of the symplectic capacity, Gromov's squeezing theorem
can be proven. For this and other applications, as well as for a proof of the
existence of capacities, the reader is referred to HOFER-ZEHNDER ([HZ], in
particular, Chapters 2-4), and to AEBISCRER et al. ([Ae], Chapter 4).
Chapter 3

Hamiltonian Vector
Fields and the Poisson
Bracket

Already in Sections 0.4 and 0.5 it was shown how the fundamental Hamil-
ton's equations of theoretical mechanics can be elegantly formulated (and
then also studied) with the help of the formalism of symplectic geometry.
This formulation will be further detailed in this chapter. Sources for this
material are, among others, KIRILLOV ([Kij, pp. 231-233), GUILLEMIN-
STERNBERG ([GSJ, pp. 88 ff.) as well as ABRAHAM-MAP.SDEN ([AM], pp.
187-208).

3.1. Preliminaries
For a given vector field X E V(M), there are, as we've already partly seen
in Chapter 2, several operations on the spaces of functions, vector fields and
differential forms on M. You may find a collected, systematical overview on
these topics in Sections A.4, A.5 and B.I. However, the information we now
need is also collected here.

i) Lie derivative of a function. For X E V (M) and f E Jr(M), we


define Lx f E F (M) by
Lx f (p) = dfp(Xp) for p E M;
thus
ax,
L x (p) a, (x) of (x), when X I U _ a;
Ox; 8

71
72 3. Hamiltonian Vector Fields and the Poisson Bracket

in the chart (U, cp) with the coordinates x. Then Lx is an element of


Der (.F (M)) = {D E End.F(M); D(f g) = f Dg+g Df; f,g E .F(M)}.

ii) Lie derivative of a differential form. For X E V (M) and 3 E


flk(M), we define LXw E Qk(M) by

Lx,3 = dt (FF,3) I tom'


where Ft denotes the associated flow of X.

iii) Lie derivative of a vector field. For X E V (M) and Y E V (M),


with Ft the flow associated to X, we define LxY E V (M) by LxY =
j (F_t.Y)jt=o, and it satisfies
LxY = [X, Y].
Here we are using the fact that V (M) is a Lie algebra with the Lie bracket
[ , ] as product.

iv) Inner multiplication and the fundamental duality. For X E


V (M) and w E 122(M), the inner product of X and w, i (X) w (also written
wb(X)) E 1l'(M), is defined by
(i (X) w) (Y) = w (X, Y), for Y E V (M).
If w is non-degenerate, this gives rise to an isomorphism
V (M) fll(M).
In the standard symplectic coordinates x = (q, p) of M we thus have w =
=t dqq A dpi, and it follows that for
n
d= E E fI
J=1
the vector field w* (t9) = X E V (M) corresponds to

(a' a -a= 8 ).
X=En

=1 , dqi aPi

v) A collection of formulas. A few of the following statements have


already been used previously. The reader may find proofs in ABRAHAM-
MARSDEN ([AM], pp. 109-120), but we also recommend that they be done
as exercises. Let a be in f1k, X, Xo, ... , Xk, Y elements of V (M) and f an
element of F (M). Then we have
3.1. Preliminaries 73

k
1. da (Xo, ... , Xk) = >(-1)'Lx, (a (Xo,... , Xi, ... , Xk))
i=0

+E(-1) +3a([Xi,Xi].Xo,.. ,J.Ci... , tj..... Xk).


i<i

2. i (X) a is R-bilinear in both X and a, and


i(fX)a= fi(X)a=i(X)fa, i(X)i(X)a=0,
i (X)(a A 0) = (i (X) a) A /3 + (-1)ka A (i (X )13) (for /3 E &I*) .

3. Lxa = d(i (X) a) + i (X) da.


4. Lxa is R-bilinear in both X and a, and
Lx (aA/3) = Lxa A/3+aALx/3.

5. (Lxa)(Xl,...,Xk) = Lx (a(Xi,...,Xk))
k
- Ea(X., ..., [X. Xi],..., Xk).
i=1

6. Lfxa = fLxa+df A (i(X)a).

7. i ([X, Y]) a = Lxi (Y) a - i (Y)Lxa.

8. Lxda = dLxa.

9. Lxi (X) a = i (X )Lxa.

10. Lix,yja = LxLya - LyLxa.


Remark 3.1. These formulas are dependent on the given formula for the
wedge product ofaiEQl(M)(i=1,...,k)andXjEV(M)(j=1,...,k).
which we have taken from ABRAHAM-MARSDEN. With this convention, we
have the formula (see ABRAHAM-MARSDEN ([AM], p.104))
(al A ... A ak)(X1, .... Xk) = > sgn (a)al(Xo(1)) ..... ak(XC(k)).
CESk

where the summation is taken over all the permutations o of 1, ... , k. How-
ever, in the normalization that appears in, for example, KIRILLOV [Ki], one
must take in the above formula (1) an additional factor of k + 1 on the left
side as well as applying an additional factor to the operator i(X).
74 3. Hamiltonian Vector Fields and the Poisson Bracket

3.2. Hamiltonian systems


The following concept is central to the whole theory.
DEFINITION 3.2. Let (M, w) be a symplectic manifold and H E F (M).
Then a vector field XH on M is called a Hamiltonian vector field with the
energy function H, if for XH we have
i (XH) w = dH.
(M, w, XH) is then called a Hamiltonian system.
The following important statements are, after the material of the previ-
ous section, but remarks.
Remark 3.3. If (q, p) are the canonical coordinates of w, then, by Sec-
tion 3.1 iv), in these coordinates

XH = 8H 8
8qi
8pi i
8H 8
-
8Qi api ,

since dH (Ldgj + 4dpi).


Remark 3.4. If (M, w, XH) is a Hamiltonian system, then there is an
integral curve y = y(t) = (q(t), p(t)), t E I, for the vector field XH, precisely
when for this curve the Hamiltonian equations hold in their classical form,
8H 8H
8Pi - 8q}
And then the theorem of the conservation of energy has the form that
H(y(t))is aconstant for all tEI.
Proof. y(t) = (q(t),p(t)) is an integral curve for XH precisely when
ti(t) _ (XH). (t) for all t.
With XH as in the previous remark, this can be brought, with the use of
Hamilton's equations, to the above form. And, when y = y(t) is an integral
curve to XH, we have

dtH y (t)) = dH.r(t)(ti(t)) =


w7(t) ((XH)7(t), (XH)n(t)) = 0.
0
As an example of the usefulness of this Hamiltonian formalism we will
show how to handle the example of the motion of a particle of mass m and
charge e in an electromagnetic field of electric field strength E = (E1, E2, E3)
and magnetic field strength B = (B1, B2, B3). Physics (for the appropriate
3.2. Hamiltonian systems 75

physics we refer the reader to LANDAU-LIFSHITZ ((LL], p. 45)) then gives


the equations of motion for these particles with velocity v = (vi, v2, v3) and
momentum p = (pi, p2, p,3) as

dp=eE+-vxB,
and so

dtt = e E, + e (vj Bk - vk Bj),


c

for each of the triplets (i, j, k) E {(1, 2, 3), (2, 3, 1)j (3, 1, 2)}. In this situa-
tion these quantities satisfy the relation

3
M
P V,
-V / j=1

which, in the classical case of v2 « c2, has the approximation p = mv.


These equations allow one to find integral curves -y of a Hamiltonian system
(M, U;, XH).

EXERCISE 3.5. It is recommended that the reader determine the integral


curves for the following systems (for the formulation of Maxwell's equations
with differential forms see, for example, SCHOTTENLOHER ((Sch], p. 191).

a) In the classical case, for a time independent field, we have

M = T'R3 with the coordinates (q, p) = (Q1, q2, q3, pl, p2, p3),
3
w = wp - (e/c)w8 with wo = E dqj A dpi,
j=1
WB = Bidgj A dqk,
(i,j,k}
3
H(q, P) = 2m p; + 0(q) with -gradO = E.
3_1

b) In the relativistic case, there is the Minkowski space R1.3 with the co-
ordinates q = (qo, Q1i 92i q3) and the metric ds2 = dqa - dqi - dq2 - dq3,
76 3. Hamiltonian Vector Fields and the Poisson Bracket

so

M = T'R1,3 with the coordinates (q, p),


w = CIDO + e WF with
3

uro=dgoAdpo-dq,Adpi,

wF = wB + WE,
3
WE=>E=dq,Adgo,
i=1

H(q, p) = H(q, p) (1/(2m))(po -pi -pa -p3).


Here one looks for curves ^y = -y(s) _ (q(s), p(s)) satisfying

a(s) _ (XH)ry(5)
Hint. Use the following formula from relativity:
dt 1

ds c 1-u/
c) A variant of b) arises for the case that
M = T'R"3,
w = curd, and
H(q, p) = H(q, p - (e/c)A(q)) with A = (Ao, Al, A2, A3),
3
so that da = wF for a = > A. dq;.
i=o
It is easy to see that b) and c) describe equivalent problems. In GUILLEMIN-
STERNBERG (EGS], pp. 143-144) a recommendation is given of how a particle
with spin would be described in this formalism.
Now we fill in more details of the general theory.
Remark 3.6. (Liouville's Theorem) Let Ft be the flow of XH. Then Ft
is symplectic; that is, we have
lw = w,
and consequently Ft preserves the volume form r,,.
Proof. From Lemma 2.3, we have
d (Fiw) =F (i(XH)dw+d(i(XH)w)).
3.2. Hamiltonian systems 77

Thus, since dw = 0 and i (XH) w = dH,


(Ftw)=Fi(ddH)=0.
From this, Ft *w is seen to be independent of t, and since FO = id. it must in
fact be w. From Section 1.1. the volume form of the given w is fixed by
(-1)!n/21w
rm = n!
Since Ft preserves the 2-form w, it must also preserve r,,,. 0
Since the Hamiltonian vector fields have such nice properties, we are
compelled to give a special symbolism for them. Thus we let Ham(M)
denote the vector space of all Hamiltonian vector fields and Ham°(M) denote
the local Hamiltonian vector fields X E V (M) that have the property that
to every m E M there is a neighborhood U of m with XIU E Ham(U).
The question of the characterization of Hamiltonian vector fields has the
following answer.
THEOREM 3.7. X is an element of Ham°(M) exactly when one of the
following equivalent conditions is satisfied:

i) i (X) w is closed, or
ii) Lxw = 0 holds, or
iii) the flow Ft of X consists of symplectic maps.

Proof. That a local Hamiltonian vector field satisfies the conditions i) -


iii) is clear from the previous remark. Poincar6's lemma says that a closed
form is locally exact, and so for i (X) w there is a local function H with
dHIU =i(X)wIU.
From the definition in 3.1 ii) of the Lie derivative of a differential form,
we see that LXw = 0 is equivalent to w being fixed by Ft. This, along
with the proof of the first equation in Remark 3.6, shows that i (X) w is
closed. 0
Because of relation (10) in 3.1,
LLX,i iw = LXLyw - LyLXw,
local Hamiltonian vector fields give rise to a Lie algebra in V(M). Hamil-
tonian vector fields are naturally locally Hamiltonian. For the reverse state-
ment, a topological condition, which assures that closed 1-forms are also
globally exact, is useful. This says, in effect (see section C.3), that the first
homology group, H'(M, R), is {0}, which is exactly the fact needed. In
particular,
78 3. Hamiltonian Vector Fields and the Poisson Bracket

Remark 3.8. For M with H'(M, R) = {0} we have that


Ham(M) = Ham°(M)
is a Lie subalgebra of the Lie algebra V (M).

In general the codimension of Ham in Ham° is exactly


bl(M) = dimH1(M, R),
the first Betti number of M.
An example of a locally Hamiltonian vector field which is not Hamil-
tonian will now be given. We begin with a vector field X, defined on the
2-dimensional torus T = R2/Z2, with the coordinates (z, y), by
X(x,y) _ (a, b), a, b E R with a2+0960.
Naturally, here we have
m=dandy.
Then
i (X) w = (i (X )dx) A dy - dx A (i (X )dy) = ady - bdx
is closed. In other words, this says that X is locally Hamiltonian. But a
locally Hamiltonian vector field with no zero points on a compact symplectic
manifold cannot be Hamiltonian. Thus if X = X H, the function H must
have a critical point (thus, a maximum or a minimum) on the compact
manifold; but then X would have a zero point.
Remark 3.9. Let i denote the identification of each real number c with
the constant function on M taking the value c at each point. Let j assign to
each element h E 1(M) the associated Hamiltonian vector field Xh. Then
there is a sequence of R vector spaces

0- R .F(M) '+Ham(M)-+0.
The fundamental duality w# now says that this sequence is exact; this follows
precisely because it is the constant functions which give rise to the trivial
vector field. Thus Imi = ker j. We will call this the fundamental exact
sequence .

It is of great significance that this sequence is not only an exact sequence


of vector spaces, but is also short exact with regard to the Lie algebra struc-
ture, which is defined on F (M) by the Poisson brackets (see the following
section).
3.3. Poisson brackets 79

3.3. Poisson brackets


There are many ways to introduce the Poisson brackets, which then naturally
lead to the same result for the Poisson bracket of two functions in canonical
coordinates (at least, up to sign. depending on whether the result is given
in terms of wo = dq A dp or wo = dp A dq). Here, we will continue to follow
ABRAHAM-MARSDEN ([AM], p. 191). where Poisson brackets for 1-forms
are first introduced. For this we will fix the symplectic manifold (Al, W).
The isomorphism given in Section 3.1 iv).
V (M) , Q'(M),
w
will be abbreviated to
wb(X) Xb for X E V (M),
respectively,
w#(19) t9# for d E S21(M).

DEFINITION 3.10. For a, 3 E Q1 (M), the Poisson bracket of a and 3 is


the 1-form
{a. 3} := -[a3#]b

From this we get the commutative diagram

Id
V(M) x V(M) V (M)

Wb

{.}
S21(M) x S21(M) S21(Af )

Since V (M) has a Lie algebra structure with respect to the Lie bracket [ , ],
so does 111(M) with respect to the Poisson bracket { , }.
THEOREM 3.11. For a, $ E 01 (M), we have
{a, Q} _-LQ#$+Lj.a+d(i(a#)i(/3#)W)

Proof. Here we will use the calculus of the Lie derivatives (see Sec-
tion 3.1). The starting point is formula (1) of Section 3.1 (this is the same
as material in Section C.2):
(dw)(X,Y, Z) = LX(w (Y, Z)) + Ly(w (Z, X)) + Lz(w (X, Y))
-w([X,Y],Z) -w([Y,Z],X) -w([Z,X],Y).
80 3. Hamiltonian Vector Fields and the Poisson Bracket

For X = a#, Y =,3*, it follows from the observation w(a#, Z) = a (Z) that
0 = LQ* (,d (Z)) - L,3* (a (Z)) - LZ (i (a#) i ((3#) w)
+{a, 01(Z) +a (LO, Z) - ,3 (LQ* Z),
and from this, with an application of the standard formula LxY = [X, Y],
as well as (1) and (5) from Section 3.1, we derive that
0 = (La*(3)(Z) - (Lq*a)(Z) - d (i (a#) i (f#) w)(Z) + {a, 3}(Z).
0
COROLLARY 3.12. If both or and,3 E 01(M) are closed, then {a, #} is
exact.

Proof. Since for a closed 1-form y we have, from (3) in Section 3.1,
Lxy = d(i(X )y),
Theorem 3.11 immediately gives the result. 0
The closed and the exact 1-forms generate Lie subalgebras of III(M).
Since every function f E F (M) brings along a 1-form d f and via w#
also a vector field X f? we may define Poisson brackets on functions.
DEFINITION 3.13. For f, g E F(M) we take as Poisson bracket the
function
If, 9} -i (X f) i (Xg) w,
thus
= w (X f, Xg).

In KIRILLOV ([Ki], p. 232), the statement of the following theorem is


taken for the definition of the Poisson bracket.
THEOREM 3.14. For f, g E .F (M) we have
if, g} = -Lxf9 = Lx9f

Proof. On the basis of the definition of X9, we have


i(X9)w=dg
and on the basis of the definition of the Lie derivative
Lx,g =dg(X1) = w (Xg, Xf) = i (Xf) i (X9) ,
= -w (Xf, Xg) = -Lxp f.
0
COROLLARY 3.15. For fo E F (M), g ,-* {fo, g} is a derivation.
3.3. Poisson brackets 81

Proof. For g, h E F (M) we clearly have


{fo, 9h} _ -Lx1o(gh) _ -d(9h)(Xf.) _ -((dg)h +gdh)(XJ0)
= -(Lxiog)h - (Lx1oh)9 = {fo, 9}h + { fo, h}g.
0
COROLLARY 3.16. The following statements are equivalent:

(1) f is constant on the curves of X9,


(2) g is constant on the curves of Xj, and
(3) {f, g} = 0.
Proof Let Ft be the flow of the vector field X1. Then on the basis of the
definition of the Lie derivative of a function (see Section 3.1 i), and Sections
A.4 and A.5) we get that
d
-j(goFt)=Ft Lxjg.

EXERCISE 3.17. Prove this last equality.

Now from Theorem 3.14 we also have

(goFt)=-{f,9}oFt.
Thus g is constant along Ft exactly when (f, g} = 0. Because of the anti-
symmetry of (f, g}, we also get the equivalence of ii).
Remark 3.18. The relation {H, H} = 0 then says that H remains con-
stant on the curves of the tangent vector field XH, and therefore leads to
the theorem of the conservation of energy.

Finally, we can bring the Poisson brackets from Section 0.5 into the
picture:
COROLLARY 3.19. In canonical coordinates, i.e., with a chart in whose
coordinates (q, p) the symplectic form w has the standard form wt = dq A dp,
we have, for functions f, g E F (M),
Ofa9of891l
aqi 8p; ap; aq8
82 3. Hamiltonian Vector Fields and the Poisson Bracket

Proof. From Theorem 3.14


{f, g} = Lxof = c (X9),
and from Section 3.1 iv)

X9
=w#(dg)=ag a - ag a
aP a9 eq aP'
and so
dq + !M l ag a Of agOf ag
q} aq ap aq ap Op 8q

0
COROLLARY 3.20. For a Hamiltonian vector field XH E Ham(M) with
}low F` we have

(foFF)={foFt,H} for all f E.F(M).

Proof. From Theorem 3.14 we have


If o Ft, H} = Lx,, (f o Ft),
and so
{foFt,H}=d(foFf)(XH)= (foFg).
0
The connection between the Poisson bracket of 1-forms and that of
functions is now at hand.
THEOREM 3.21. For f, g E.F(M), we have
d { f, g} = {ti, dg}.

Proof. From Theorem 3.11


{df, dg} = -Lxldg + Lx,df + d (i (X f) i (Xq) w)
= d(-Lxfg+Lxef +i(Xf)i(Xq)w),
and from Theorem 3.14 we have
{df,dg}=d(i(Xf)i(Xg)w)=d{f,g}.
0
Since ftl (M) with the Poisson bracket is a Lie algebra, we can easily
arrive at the following important fact.
3.3. Poisson brackets 83

THEOREM 3.22. F (M), as Ii vector space with the Poisson bracket, has
the structure of a Lie algebra.

Proof. Since d and w# are R-linear, the map f -, X f is also R-linear


and it follows that If. g} = -i (X f) i (X9) w is R-bilinear. It is then clear
that If, f } = 0. It remains to verify the Jacobi identity. We begin with
{f,{g, h}} = -Lxf{g,h} = Lxf(L. Bh),
{g,{h, f}} = Lx9(Lx,,f) _ -Lxg(Lxfh),
{h,{f,g}} = Lxlf,g}h.

But from Theorem 3.21, we have


X{f,9} = (d{ f, g})# = {4f, dg}* = -[(df)#, (dg)#],
and therefore
X{f,9} _ -[Xf, X9],
from which the claim follows. 0
The reader should pay particular attention to this last relationship.

COROLLARY 3.23. For f, g E F (M), we have


X{ f.9} = -[X f, X9],
and so the Hamiltonian vector fields Ham(M) generate a Lie algebra. The
fundamental exact sequence in Remark 3.9 is, with -j(f) = -X f, of the
form
0 - R ' .F (M) - Ham(M) - 0,
and, therefore, also an exact sequence of Lie algebras.

In the context of this formalism, we can give a useful criterion for when
a diffeomorphism of symplectic manifolds is symplectic.
THEOREM 3.24. (Jacobi, 1837) Let (M, w) and (M', w') be symplectic
manifolds and F : M - M' a diffeomorphism. Then F is symplectic exactly
when for all h E .F (M') we have
F.Xh.F = Xh
Proof. We begin with
XA =W '# (dh) and Xhop = w# (d (h o F)).
Thus, for all Y' E V (M'),
w (Xh, Y') = dh (Y').
84 3. Hamiltonian Vector Fields and the Poisson Bracket

i) When F is symplectic, we also have, for all Y E V (M),


w (F.XhoF, F.Y) = w (XhoF, Y) = d (h o F)(Y).
Then, with the usual transformation formalism,
w'(F.Xh.F, F.Y) = dh (F.Y) = w'(Xh, F.Y).
Since w' is nondegenerate, it follows that F.XhoF = Xh is closed.
ii) For h E F(M') we have
i(XhoF)w = d(h o F) = Fdh = F'(i(Xh) w).
The precondition F.Xh,,F = Xh, along with the general formula
F'(i(X')o) = i((F-1).X')F'o,
gives

i(XhoF)W = i(XhoF)F*w'
Since every X,,, can be taken locally to be of the form (Xh.F)m for an
h E F (M), we arrive at the claim F'w' = w. 0
Symplectic maps can be defined by the property that they preserve the
symplectic structure. They can also be characterized as preserving the Pois-
son bracket:
THEOREM 3.25. The difeomorphism F from M to M' is symplectic
precisely when F preserves the Poisson bracket of functions on M or the
Poisson bracket of 1-forms on M. In other words, exactly when
171:.F(MI) F(M),
f r-. f o F,
or, respectively,
F*: ill (M') 0I (M),
19 F't9,
+-+

are Lie algebra homomorphisms with relation to the Poisson bracket.


Proof. For f, g E F (M'), we have, from Theorem 3.14,
If, g} o F = L(F-t).x;(f o F),
and
if
as well. Thus, from Theorem 3.24, both are the same exactly when F is
symplectic. We leave the proof of the corresponding statements for 1-forms
as an exercise. 0
Now we are able to say whether the coordinates of a chart are canonical:
3.4. Contact manifolds 85

Remark 3.26. Let (U, gyp) be a chart with coordinates (q, p). Then this
is a symplectic chart, that is, with
wo=1: dg1Adpi,
we have w = V*wo exactly when
{q1, gi} _ {p:,Pj} = 0 and {q:, pi} = btj,
for i, j = 1, ... , n.
EXERCISE 3.27. Prove the remark.

3.4. Contact manifolds


We have now seen how specifying a 2-form with particular properties on a
manifold will endow this manifold with the structure of a symplectic man-
ifold, and that it will then necessarily have even dimension. In a precisely
analogous manner, specifying a 1-form will define a contact structure on a
manifold, which then forces the dimension to be odd. These contact mani-
folds can be discussed in a theory parallel to that of symplectic manifolds;
however, both can also be developed as special cases of the general the-
ory of presymplectic manifolds. In any case, it will soon be clear that the
physically most important examples of contact manifolds arise as hypersur-
faces of constant energy in Hamiltonian systems or from the treatment of
time-dependent Hamiltonian functions. For this reason, we will give a few
highlights from the theory of contact manifolds as a finish to this chapter
on Hamiltonian systems. We begin by defining the major concept of this
material.
DEFINITION 3.28. Let M be a differentiable manifold with dim M =
2n + k, k > 0, supplied with a 2--form w which has rank 2n everywhere.
Then w is called a presymptectic form and (M, w) a presymplectic manifold
For k = 0 we get the definition of symplectic manifold, and for k = 1 we
get that M is a so-called weak contact manifold. Darboux's Theorem 2.6 for
symplectic manifolds transfers without difficulty to the following statement
about the normal form of w.
THEOREM 3.29. Let M be a (2n+k) --dimensional differentiable manifold
and w a closed form of rank 2n. Then for every point m E M there is a
chart (U, gyp) containing m with coordinates
(ql, ... gn,P1, ... ,Pn, W1..... Wk),
so that in these coordinateswlu can be written as

wIU = dqj A dpj.


j=1
86 3. Hamiltonian Vector Fields and the Poisson Bracket

ABRAHAM-MARSDEN ([AM], p. 372) give a proof by falling back on


the proof of Darboux's theorem that we gave in Section 2.2. On the other
hand, STERNBERG ([St], pp. 137-140) gives a direct proof. The concept of
a so-called weak contact manifold can be made sharper as in the following
definition.

DEFINITION 3.30. A differentiable (2n + 1)-dimensional manifold M is


called a contact manifold, if there is on M a 1-form t9 with
t9 A (dt9)" # 0 everywhere on M.
L9 is then called a contact form.

This notation is in agreement with that found in AEBISCHER et at. [Ae],


VAISMAN [V] and BLAIR [BI]; this last is a good place to begin one's study
of this theory. In BLAIR'S text one may also find (pp. 11-12) some thoughts
as to the choice of vocabulary. ABRAHAM-MARSDEN [AM] call the above
defined object an exact contact manifold, and use the term contact manifold
for what is here called a weak contact manifold.
It is clear that with w = dt9 contact manifolds are weak contact mani-
folds. Corresponding to Darboux's Theorem 2.6, we have here a statement
about the normal form of a contact manifold.
THEOREM 3.31. Let M be a (2n+1)-dimensional differentiable manifold
with a contact form t9. Then there is, for every point m E M, a chart (U, gyp)
containing m with coordinates (q1, ... , qn, pl, ... , p,,, w) such that
n
r9lu=dw+>pj4.
i=I

Proof. Let w = -dt9. Then by Theorem 3.29 there are a chart (U, gyp)
and coordinates (q1, ... , qn, pi, ... , pn, w1) such that in U

d(t9->pidb) = 0,
i=1
and so locally for w = w(q1i... , w1)

dw.
j=1

Then, since 6 A (dt9)" 0, (ql, .... qn, p1, ... , pn, w) are also usable as co-
ordinates. 0
3.4. Contact manifolds 87

The reader may find a more direct proof in AEBISCHER et al. ([Ae]. pp.
168-171).
The study of contact manifolds will be enlightened by the introduction
of a particular differential system; that, is, by the introduction of a system
of smoothly varying subspaces of the tangent spaces of a fixed dimension, as
we have already seen in the discussion of FYobenius' theorem in Section 2.5.
Alternatively, this may be given as particular subbundles of the tangent
bundle TM.
DEFINITION 3.32. Let. w be an element, of S22(M). Then we call
R,,,.:_ {(m, Xm) E TMI i(Xm)wm = 0}
a characteristic bundle of w. X E V(M) is called a characteristic vector
field of w if
i(X)w=0.

Remark 3.33. When w E S12(M) has constant rank, then Rte, is a subbun-
dle of TM, and its sections generate a differential system (the differential
system of the characteristic vector fields). When w is closed. R,,, is involutive.

Proof. The proof is not hard (see ABRAHAM-MMARSDEN ([AM], p. 371)).


Since for two characteristic vector fields X. Y. [X, Y] is also characteristic,
we get the result from the formulas in Section 3.1.
Remark 3.34. In particular, when w defines a weak contact structure on
Al. then R,,, is a vector bundle of rank 1. thus the characteristic line bundle.

In analogy to the association of a characteristic bundle to a 2-form one


may also associate a bundle to a 1-form.
DEFINITION 3.35. Let t9 be an element of 1l'(M) and t9m 54 0 for all
m E M. Then we call
Rfl :_ {(m, Xm) E TM: t9m(Xm) = 0}
the characteristic bundle of V.
Remark 3.36. (Al, t9) is a contact manifold precisely when dt9 is non-
degenerate on all the fibers of R,t.

Proof. Since R,9 is a vector bundle of rank 2n, t9 is thus non-degenerate


on Rd precisely when the n-th exterior power (dt9)" 34 0. And this is
equivalent to saying that t9 A (dt9)" 54 0 on all of Al.
88 3. Hamiltonian Vector Fields and the Poisson Bracket

One should be careful about the word usage, which wee have here taken
from ABRAHAM-MARSDEN [AM]. In BLAIR [B1], the differential system
D of a section of R, is called a contact distribution and an element X of
a one -dimensional complement of D in V(M) a characteristic vector field
with the contact structure given by d. Such an X is then fixed by d(X) = 1
and dd(X, Y) = 0 for all Y E V(M). The following examples should make
clear what may be confusing at this first glance.
EXAMPLE 3.37. Let M = lR2,+1 come with a contact structure via
n
d=dw - EPidh.
i=1

Then X = 8,, is, in the above sense, a characteristic vector field of d, and,
for w=dr9dq,Adp,, we have
d(X) = 1 and i(X)u) = 0;
that is, X also spans a one-dimensional space of characteristic vector fields
to w in the sense of the first definition. Further, the contact distribution D
is here spanned by
Xi:=0q,+p,8,, and i=1,...,n,
and then, clearly, for i = 1, ... , n,
d(Xi) = 19(Xn+i) = 0.
EXAMPLE 3.38. Let Se be a regular energy surface for a Hamiltonian
system (M, w, H), that is. a connected component of H-1 (e) for a regular
value e of H: in other words, for those in which dH,,, 0 0 for all m E H-1(e).
S, is then a submanifold of M of codimension 1. And it is (Sei t'w) for
t : Se --+ M, a weak contact manifold. XH IS, is a characteristic vector field
of Ow and induces the characteristic line bundle of Ow . To see this is
not difficult. In particular, we have i(XHIs)t'w = 0 because of the relation,
defined on XH,
wm((XH)m, t) = (dH)m(t) = 0

EXAMPLE 3.39. In order to produce a true contact manifold, we sharpen


the last example, and arrive at the physically particularly interesting case
of
M=T'Q -Q
of the phase space of the configuration space Q with w = -ddo (see Sec-
tion 2.3) and
H=K+VoaQ,
where V is a real potential function on Q and K is the kinetic energy asso-
ciated to the Riemannian metric. Then (Se, i'do) is a contact manifold, and
3.4. Contact manifolds 89

it can be easily shown that t90 A (dt3o)n on Se has no zeroes (see ABRAHAM-
MARSDEN ((AM, p. 373)).

This last example can be formulated in more abstract terms with the
help of a theorem.
THEOREM 3.40. Let i : S ' lR2i+2 be the immersion of a smooth hy-
persurface, where no tangent space of S meets the origin of R2n+2. Then S
has a contact structure.
And it will be given by i' a for
a = xldx2 - x2dx1 + ... + x2 +Idx2n+2 - x2n+2dx2n+I,
when x1i ... , X2n+2 are the coordinates in R2n+2 (for a proof, see BLAIR
(Bi], pp. 9-10)).
The next example takes the form of a recommended exercise.
EXERCISE 3.41. Give the 3-dimensional torus T = R3/Z3 a contact
structure as well as an associated characteristic vector field so that the con-
tact distribution D can be made explicit.
EXAMPLE 3.42. Closely related to Example 3.39 is the example, partic-
ularly interesting for the physicist, of the construction of a contact manifold
arising from a time-dependent Hamiltonian function. Here, one is given a
symplectic manifold (M, w) and a time direction R on R x M. Then we let
7r2:RxM - M,
(t, m) '-- in,
C D:= axw, and t be the vector field on R x M given by
t(s,m} = (1,0) E TR x TmM = T(,,,,,) (R x M) for (s, m) E R x M.
Then the following statements are easy to see (see ABRAHAM--MARSDEN
([AM), p. 374)).
Remark 3.43.
i) (R x M, w) is a weak contact manifold.
ii) R;, is generated from the vector field t E V(R x M).
iii) if w = dt9, then for 0 = dt + 7r2W also D = dt9, and (R x M, z9) is a
contact manifold.
EXAMPLE 3.44. This last example can be varied by the inclusion of a
time-dependent Hamiltonian function H. For this the following formalism
is useful. Let
X.RxM-+TM
90 3. Hamiltonian Vector Fields and the Poisson Bracket

be a time-dependent vector field; that is, for every fixed t E R, a fixed vector
field on M is given. Then X represents, via
X:RxM -+ T(RxM),
(t, M) '-' ((t, tn), (1, X(t,m)),
a vector field X E V(R x M), called the suspension of X. Tb an integral
curve 7 of X through m E M (that is,
ry:I M,
t -Y (t),

with -y(t) = X(t,.,(t)) for all t E I and y(O) = m) there is associated the
integral curve 7 of X through (0, m) E R x M given by
ry:I RxM,
t '--. (t, 7(t)),
with 7(t) y(t)) = (1, X(t,. (t))) and 1-Y(0) = (0, m). If we now have
a symplectic manifold (M, w) along with a time-dependent Hamiltonian
function H E.F(R x M), then with
Ht:M R,
m --r Ht(m) = H(t, m),
XH, is, as in the above, a time-dependent Hamiltonian vector field. At
each 'M E M the vector is represented by (XH,) ,, and XH is the associated
suspension. If we now put
wH :=Ca +dHAdt,
we get the following statement as a variant of Remark 3.43.
Remark 3.45.
i) (R x M, wH) is a weak contact manifold.
ii) XH generates the characteristic bundle RAH which satisfies
t(XH)WH = 0 and i(XH)dt = I.
iii) For w = dig, tH := 1r2W + Hdt and we have wH = d19H. In the case
that H+(t901r2)(XH) vanishes nowhere, we have that (R x M, 19H)
is a contact manifold.
Thus here the contact manifold carries physical information, and the
conservation of energy, which in the case of time-independent Hamiltonian
functions H can be written as LXHH = 0, is here
LXHH=
3.4. Contact manifolds 91

The phenomena described in the examples can in large part be found in the
following influential viewpoint of WEINSTEIN which forms the starting point
for further study (see AEBISCHER et al. ([Ae], p. 174)). Let (M, w) be a
symplectic manifold and i : S - M a hypersurface given as the vanishing
set of f E F(M) with df Is # 0. Then all the multiples of the Hamiltonian
vector field X f are called the characteristic line fields on S. and from this
GS:={cXf;CER}.
S is said to be of contact type if there is 1-form fl on S with
dt9 = i'w and t9(X) 6 0 for all X E Cs\{O}.
In the context of the concepts presented at the beginning of this section,
every cX f is then a characteristic vector field to d in the sense of Bt.AiR as
well as a characteristic vector field to d O in the sense of Definition 3.35. The
contact structure on S is naturally not uniquely defined; given /3, a closed
1-form on S with (t9+13)(X) 0, then t9+,3 also defines a contact structure
on S.
We close the chapter with a nice result that shows just how tightly
symplectic and contact manifolds are interwoven.
THEOREM 3.46. A manifold with an orientable contact structure can be
realized as a hypersurface of contact type in a symplectic manifold.

A proof of this statement, as well as an entrance to this modern devel-


opment, can be found in AEBISCHER et at. ([Ae], pp. 167-218).
Chapter 4

The Moment Map

A very helpful technique used in classical mechanics for the solution of com-
plex problems consists of deriving integrals (that is, expressions which re-
main constant with the motion of the system) by analyzing the symmetry of
the given system. The conservation of momentum and angular momentum
in systems with invariance under respectively translations, rotation.-,, is the
most frequent example. In this, the following beautiful and at. first glance
rather abstract formalism has been crystalized from the influential work of
Souriau, Kostant, Smale and Marsden. This is linked to the discussion in
Section 2.5.

4.1. Definitions
For what follows, we fix (M, w) to be a symplectic manifold, on which the
Lie group G operates symplectically via 0; that is, for
GxM M,
(g, m) ' -' gm = Og(m),
where all the Qg, g E G, are symplectic diffeomorphisms with ¢e(m) = m
and Ogg,(m) = tg(og,(m)). We let g denote the Lie algebra of G and g"
its dual. In this situation, g can be realized as the tangent space TG to G
at the element e E G. and this is in turn identified with the left--invariant
vector fields VI(G) on G. In the same manner, 9' can be realized as the
cotangent space T, G, which in turn is the left. invariant 1-forms on G.
For X E g, we denote by XM the vector field on M which for all m E M
is given by the rule

(XAff)(m) at f(oexpix(m))It=o for f E F (M),

93
94 4. The Moment Map

or, equivalently, by

(XM)m
or, yet again, by
(XM)m (Vm)seX for u,,,, :G - M,
g'-- V4. (g) = gm.
EXERCISE 4.1. It is recommended that the reader, using the introductory
material of Appendix A, verify that these three definitions coincide with one
another, and also verify the formula
Lx.v w = 0.
X,M will be called the infinitesimal generator of the operation on M
associated to X.
DEFINITION 4.2. A map 4? : M - 9' is called a moment map for the
group operation 0 if for all Y E g we have
(ie) d (Y)=i(YM)w
with
6(Y):M -+ R,
m 4b (m)(Y).
Remark 4.3. Here we have what might at first sight appear to be con-
fusing notation:
6 (Y)(m) = 4 i (m)(Y) f o r m E M, Y E 9,
which says, in any case, that 4i is fixed by 4i. With the abbreviation of
X f for the (Hamiltonian) vector field (*) associated to f E F (M), this is
equivalent to
Xi(y) = Yu.

NOTATION 4.4. (M, w, 0, 4?) is also called a Hamiltonian G-space.


Remark 4.5. Not every symplectic group operation has an associated
moment map. The existence of a moment map 4' for a given operation 0
is guaranteed only when the local Hamiltonian vector fields for M are also
Hamiltonian. Then Theorem 3.7 says that the symplectic diffeomorphisms
0s produce local Hamiltonian vector fields Y;/, which then, under the pre-
condition Ham°(M) = Ham(M), are of the form YM = Xji(y).
Remark 4.6. Should 4i and 4" both be moment maps for the group
operation 0, then there is a u E g' with
4?(m)-4Y(m)=u forallmEM.
4.1. Definitions 95

The meaning of the moment maps will be made a little clearer in the
following statement about their behavior.
THEOREM 4.7. Let fi be a moment map for the operation ¢ of G on Al,
and let H E F (M) be invariant under this operation, that is,
H (m) = H (Og(m)) for all mEM and9EG.
Then 4' is an integral for the vector field XH associated to H; that is, for
the flow Ft, t E I associated to XH, we have
4) (Ft(m)) = 4, (m) for all »t E Af, t E I.
Proof. For all Y E g, we have
H (mexpty(m)) = H (m),
since H is invariant. Here, after differentiating with respect to t and evalu-
ating at t = 0, we get
(dH)m((Y.Yt)m) = 0,
and so
LYA, H = 0 for Yaf =
From Theorem 3.14, we then have
(Y)}=0,
and from Corollary 3.16, we have
`F (S')(Ft(m)) (Y) (m)
for every Y E g. Thus we arrive at the claim. O

In the most important examples, the moment map has yet another nice
equivariance property. In order to describe this we recall the definition of the
coadjoint representation from Section 2.5 or D.3. The adjoint representation
Ad is given by
G x TeG - TeG,
(g, Y) Ad(g)Y = Ad9Y with Ado = (p 1 \9
and the coadjoint representation Ad' by
G X (TeG)` (TeG)' = T, *G.,
(g, a) Ad`(g)a = (Ad(g''))'a =: Ady`._,a,1
from which we define
(Ad(g))*

: (TTG)` (TAG)',
a e {Y -a (Ad(g)Y)},
1Thia usage is common in the literature of the moment map, and we will use it here also.
96 4. The Moment Map

and so
(Ad(g)'a)(Y) = a (Ad(g)Y).
DEFINITION 4.8. A moment map 4' on the group operation 0 is called
Ad'-equivariant, when
(09 (m))=Ad; 4D (m) for all mEM andgEG;
that is, when, for all g E.G, the following diagram commutes:

m
M

Ad'(g)= Ad, w
g

In ABRAHAM-MARSDEN [AM] the following is introduced as the means


for weakening this equivariance.
DEFINITION 4.9. Let G be a Lie group and g its Lie algebra. Then we
call o a coadjoint cocycle when or is a map or : G -+ g' which satisfies the
cocycle identity
a (gh) = a (g) + Adg_, a (h) for all g, h E G.
Such a cocycle 6 is called a coboundary if there is a u E g' with
6(g)=u-Ad9_,µ foraUgEG.
The cocycles modulo the coboundaries give a cohomology group, which is to
be understood in connection with the general theory of Appendix C.

A Hamiltonian C-space (M, w, ¢, 4') can be represented by a coadjoint


cocycle a (respectively, a cohomology class [a]), in that for g E C and t; E g
we can take a(g)(l;) as the value of the map 1,bg,t, which is defined by
,g,E M -' R,
m '-' --i
and which can be shown to be constant (see ABRAHAM-MARSDEN ([AM],
p. 277)).
4.2. Constructions and examples 97

4.2. Constructions and examples


We will now discuss only the Ad* equivariant moment maps; we will try to
explain the reason for the term moment map, in that the classical momentum
and the angular momentum will both appear in the formalism. Before we
can manage this, however, we must make a few general observations.
We proceed by taking as fixed a symplectic operation
0:GxM- MMl;
that is, for
GxM3(9,m)igm=tp9(m)EM
all the ¢9, g E G, are symplectomorphisms. For a 1; E g = Lie G we denote
by CAt the associated infinitesimal generator, that is, the vector field on M
which, as defined at the beginning of Section 4.1, is given by

(CA!)rn :_ -dttlexpt'(()It=o'
We will later need several properties of this generator:
Remark 4.10. For E 9 and g E G we have
((Ad94)A1)m = im

in the notation of ABRAHAM-MARSDEN ([AM], p. 269).


Proof. It follows from the definition of infinitesimal generators that

dl OexptAdt;(m')It=O+
and from the definition of Ad9 that
d
dtO9(exPg)9-,(7n)It-o'
Then, since 0 is a group operation,

Wt- (0g o Oexptt;(g-Im)) It=O+

which finally can be interpreted as the displacement of by 0.9. at


the point g (g-Im) = m. Thus
((Ad9C)Af)rn = (O90g-1rn(CM)g-'m
11

Remark 4.11. Fore, g E g we have [Cm, i7m] _ -[C, rl]AI.


EXERCISE 4.12. Prove the remark. See ABRAHAM-MARSDEN ([AM],
p. 269).
98 4. The Moment Map

Remark 4.13. The notion of the infinitesimal generator is functorial.


That is, when we are given two manifolds M and N with G-operations 0,
respectively, ip and an equivariant map F : M - N (thus with Forty = e, 9oF
for all 9EG),then we have.,for tE9,
TFoCM=CNoF,
where CM and i N are the respective generators on M, respectively N. Thus
we have the following commutative diagram:

F
M

far

11
TF=F.
TM TN.

EXERCISE 4.14. Prove the remark. See ABRAHAM-MARSDEN ([AM,


p. 270).

We recommend that readers take another opportunity to acquaint them-


selves with the notion of infinitesimal generators by proving the following.
EXERCISE 4.15. Let Ad : G x TeG - TeG be the adjoint representation.
Then for l; E TeG g the associated infinitesimal generator is &T G =: add
with
adt : TeG TTG,
n [C nl
Many of the examples of moment maps stem from the following situation.
THEOREM 4.16. Let 0 be a symplectic operation of G on M. Let the
symplectic form w of M be exact (that is, w = -0) and let the 1 -form V
be G-invariant; thus Oyt9 = d for all g E G. Then
4i:M--+ g'
with

't (m) W = (i (CM)19)(m) = Vml(C.lf)m)


defines an Ad' -equivariant moment map for 0.

Proof. i) Since t9 is G-invariant, we have (compare with Section 3.1 ii))


LfMt9=0,
4.2. Constructions and examples 99

and because of (3) in Section 3.1 also


d(i(EM) t9) + 0.

Thus
d i(eM)w,
that is,
(t;):=i(£M)t9 foreE9,
thus satisfying the characterizing relation of the moment map b.
ii) To demonstrate the Ad'-equivariance, we must show that
ii (C)(Og(m)) = $ (Adg-i)(m) for all g E G, M E M and E 9.
But from i) this is equivalent to
(i(Cnf)t9)(O (m)) ((Adg-1j)af)19)(m),
and so to

and, because of Remark` 4.10, even to


tt
19grn(SM)gm =

t9 (that is, the property that 6g19 = i) says


that, for all Y, E TmG, we have
t9gm((og.)mYm) = t9m(Ym).
Here substituting Y,,, = )m gives, on account of
r.
the claimed equality
t9gm((SM)gm) = t9m (((AdgM)rn).
0
This theorem will now be used on the phase space, i.e. the cotangent
bundle M = T'Q of the configuration space. Q. In this situation G operates
via diffeomorphically on Q; that is, we have
GxQ Q,
(g, q) i--r gq = cpg(q), Mpg diffeomorphic for all g E G.
In light of Section 2.3 this G-operation on Q can be carried forward to a
symplectic G-operation cP (also called a canonical transformation) on M =
T*Q according to
G x T*Q T"Q,
(9, (q, aq)) '-' (9q = +pg(q), V*g-,aq) =: O(q, aq),
100 4. The Moment Map

where the point m E M = T'Q is written as the pair m = (q, %) with


gEQand agETQQ.
THEOREM 4.17. 0 has an Ad'--equivariant moment map
4?:M=T*Q_.g'.
For m = (q, aq) E M and C E g with infinitesimal generator q on Q,
thus,
[ d
(SQ)q = dtVexpq(q)It_e,

4i is given by
4' (q, aq) :=
This can also be written as
$ w = 1'
where, for a vector field X E V (Q),
P (X) : T*Q R,

(q, aq) '-' aq(Xq),


is defined as the momentum to X.
Proof. The G-operation c, on Q is extended to a C-operation 0 on
M = T'Q so that the projection 7r : T'Q -. Q is G-equivariant; thus
cpgo7r= 7ro0y.

The equivariance in the construction of the infinitesimal generators, as


demonstrated in Remark 4.13, implies that, for M = T*Q, N = Q and
F=7r,
(*) Qo7r=7r.oEAf
The definition of the canonical 1-form 79 on M (see Section 2.3), along with
d1rm = (7r.),,, : TmM T q Q f o r m = (q, aq) and X m E T says that
(q,ay)(X(q,ay)) = aq (7'(q,0q)X(q,av)) -

Employing this and (*), we get


(i (EM) 10) (q, aq) = ((W(q,Qa))
=
aq((4Q)q)
= P (eQ)(q, aq)
From this and Theorem 4.16, we can deduce that 4? is an Ad'-equivariant
map.
4.2. Constructions and examples 101

Were the usual coordinates (q, p) for a point m E T'Q used, then the
momentum corresponding to Xq = E (X, (q) i9q,) E TQQ would be

P (X) (q, p) = piX:(q)-

In the treatment of quantization in the next section the following rela-


tions will receive meaning. These will be relations on lifting, of functions
f E F (Q) as position functions to functions f E F (M) such that f = f o n
for the projection
rr:M=T`Q Q,

(q, aq) q.

Remark 4.18. For X, Y E V (Q) and f, g E F (Q) we have

i) {P (X), P (Y)} = -P ([X, Y]),

ii) {J}=0 ,
iii) {f,P(X)}=X(f).
The proofs follow by routine computations (which appear in ABRAHAM--
MARSDEN ([AM], p. 284)).
From Theorem 4.17, a moment map can be established on the tangent
bundle M = TQ whenever Q is a Riemannian manifold with a scalar prod-
uct ( , ) on the tangent spaces TqQ (it is actually sufficient that Q be
pseudorieinannian) and a group C operates by isometries V. on Q such that
this operation can be extended in a natural way to symplectomorphisms
;pg = Tcpg on TQ. Then from Theorem 4.17, the following statement can be
deduced (or it can be proven directly in exact analogy to the proof of that
theorem).
THEOREM 4.19. The moment map (P associated to cp is given, for
(q,vq)ETQ with qEQ, vgETgQ,
by

4 (q, vq) 4' (C) (q, vq) := (vg, for t; E g.

ABRAHAM-MARSDEN ([AM], p. 285) show that this is a special case


of a general statement connected to Noether's theorem in the context of
the Lagrangian formalism regarding TQ, which (pp. 208 ff.) is developed
via the Hamiltonian formalism on T'Q. On this topic we have only enough
102 4. The Moment Map

space to say a few words. A Lagrange function L E F (TQ) can be assigned


a fiber derivative FL, that is, a map
FL:TQ-.T'Q,
which is defined by
(q, v4)'-' (q, d"Lq),
where d"Lq E TTQ at a vq E TqQ has the value (This
corresponds to the relation p = in Section 0.2 in the transition from the
Lagrangian to the Hamiltonian formalism.) L is called regular when FL is
a local diffeomorphism. Precisely in this case the symplectic standard form
wo on TQ is pulled back via FL to a symplectic form WL on TQ:
wL = (FL)'wo.
Correspondingly, the Liouville form d carries over to
>9L :_ (FL)`t9.
ABRAHAM-MARSDEN ([AM), pp. 285-286) now prove
THEOREM 4.20. Let the regular Lagrangian function L E F(Q) be G-
invariant, that is,
Loo.q=L foraligEG.
Then we have

i) t9L is also G--invariant; that is, O;$L = tL for all g E G.


ii) For this G-operation, an Ad'-equivariant moment map 4
can be given by
6 vq) = for E E 9-

iii) The moment map it is an integral for the L-associated Lagrange


equation.

Now we offer several examples.


EXAMPLE 4.21. Let Q = R", and let C = R" operate on R" via trans-
lations
GxQ -+ Q,
(8, q) ,--+ s+q=,p.(9)
Then the infinitesimal generator associated to E R" = 9 is also t;
(and therefore independent of q). From Theorem 4.17 the associated moment
map lb on T*Q is given in the standard coordinates (q, p) of M by
4.2. Constructions and examples 103

p) _ Me

which is also
4, (q,P)=P.
the momentum.
This is to be understood in connection with the conservation statement
in Theorem 4.7. For every system with a Hamiltonian function invariant
under the action of G = lR", the momentum is a conserved quantity (an
observation which the physicist makes. however, without the need of the
whole apparatus here constructed.)
EXAMPLE 4.22. Let Q = W', and let G be a Lie subgroup of GL" (1R).
The elements q E Q will be thought of as columns, and G will operate as
usual through multiplication; thus
GxQ - Q.
(A, q) ' -' Aq = P.t(q)
The infinitesimal generator for B E g = Lie G C Af,,(R) is BQ with
(BQ)q = Bq. We have further. from Theorem 4.17. that there is then an
Ad'-equivariant moment map given by
4 (B)(q P) = P (Bq)
where p is meant to be a row vector.
In the special case of n = 3 and G = SO(3). we have (see the end of
Section B.2)
g = so (3) = {B E A13(1R). B = -1B} )R3

with
0 -b3 b2 b1
B= b3 0 -bt b= b2 E 1R3.
-b2 bl
0 b3
Were we now to realize the moment map 4 on TQ. as in Theorem 4.19, the
result would be. with the standard scalar product (.) on 1R3.
6 (B) (q, v) = (v, Bq)
= (b x q, v) = det(b. q, v)
= (g x v, b).
Thus, with the identification of so (3) with R3 and 1R3 with (1R3)". we have
4(qv)=qxv.
104 4. The Moment Map

which is the usual angular momentum. For instance, for the harmonic os-
cillator with the Hamiltonian function
H(q,4)=(1/2)(IIq 112+114112),
this moment map is an integral.
EXAMPLE 4.23. The Lie group G operates on itself by left translation
GxG -+ G,
(g, h) - gh _ Ag(h).
Then the infinitesimal generator associated to this operation &, for f E g =
Lie G is given by the right-invariant vector field that takes on the value
at e; thus (£G )g = (eg, )el;, where Bg is the right translation. From this we
get for the moment map on T'G
4' ( )(g, ag) _ 1(Pgs)e!O _ (pgag);
that is,
4' (g, ag) = B?ag(e)
ABRAHAM-MARSDEN [AM] discuss further examples in their exercises.
Also related to the material of this section, GUILLEM[N-STERNBERG [GS]
introduce the example of the operation of the Euclidean group E (3) _
SO(3) x 1R3 on 1R3.

4.3. Reduction of phase spaces by the


consideration of symmetry
A classical theorem, going all the way back to Jacobi and Liouville, says
that by giving k first integrals whose Poisson brackets vanish. Hamilton's
equations can be reduced to a system of equations in 2k fewer variables.
In a similar way, rotational invariance in an n-body problem allows the
elimination of four variables. These two processes indicate how, in a general
way, with the help of symplectic reduction, one can reduce from higher to
lower dimensional symplectic manifolds when a symmetry group operates on
the given manifold. Going back to Elie Cartan, these procedures allow one
to form quotient spaces and are a completely general central theme of later
constructions; they have, in fact, already been seen at the end of Section 2.5
in the discussion of construction procedures for symplectic manifolds.
Here we will follow the treatment of ABRAHAM-MARSDEN ([AM], pp.
298 ff.). We assume that we are given
a symplectic manifold (M, w),
a symplectic operation 0: G x M M of a Lie group G on M,
4.3. Reduction of phase spaces by the consideration of symmetry 105

and for g = Lie G


an associated Ad'-equivariant moment map 4' :

Then we denote by G,, for a u E g' the isotropy group


Gµ := {g E G; Ad9_,µ = Ea}.
It is a general fact that this is a closed subgroup of G and therefore also a
Lie group. Since 4i is Ad'-equivariant, the space
Mr, := `y-1(1t)1GN
of GN-orbits on the fibers 4-1(p) makes sense. It is called the reduced space
associated to the triple M, 4i and p, and it is this reduced phase space which
will find application in the important special case of M = TQ. We will
begin by giving a few technical conditions that insure that Afµ is at least a
smooth manifold.
i) Let p E g' be a regular value for 4?; that is, for all m E -D-1 (P), the
map T4i,,, (also written as (4',),,,) of the respective tangent spaces TmA1
in Tg' = g is surjective (which, because of Sard's theorem, must be true
for almost all p). Then, with the methods described in Appendix A (see
ABRAHAM-MARSDEN [AM], p. 49), one may show that the fibers of 45-1(p)
form a submanifold of M. and that this manifold has dim 44-1 (µ) = dim Al-
dim G.

ii) GF, operates without fixed points and properly on 4>-1(p). Here prop-
erly means the following: if (mj) and (¢,,,m3) are convergent series in Al,
then ¢(gg) has a convergent subsequence in C. This condition is, for exam-
ple, automatically satisfied when G is compact. It is then a fundamental
statement (see, for example, ABRAHAM-MARSDEN ([AM], p. 266)) that
the above introduced space M. = 4r1(µ)/G,, is a manifold and that the
canonical projection
7rµ . 4;-1(ld) - Mµ = 4'-1(1,)/Gµ
is a submersion. That Al. in this situation is symplectic is the content of
the following theorem.
THEOREM 4.24. Let (M, w) be 9ymplectic with a symplectic G-operation
and an Ad* -equivariant moment map satisfying the conditions given above.
Then Mµ = 4t-1 (µ)/Gµ has a uniquely defined syniplectic form wµ with
T,,w,,

where ir, : 4D (y) - Mµ is the canonical projection and i,, : (P -1(µ) M


is the inclusion.
106 4. The Moment Map

The proof requires the following statement.


LEMMA 4.25. Form E 4)-t(p) and Gm := {O9m; 9 E G} we have

o)

Tm(Gpm) = Tm(Gm) flTm($-I{ft)},


ii) T.($ (µ)) and Tm(Gm) are w-orthogonal complements of one
another.
EXERCISE 4.26. Prove the lemma. In case of emergency, one may consult
ABRAHAM-MARSDEN ([AM], p. 299).

Proof. Now the proof of the theorem follows in several steps.


a) For v E Tm (f-I (µ)), let [v] = (ir,,).m(v) be the associated equivalence
class in Tm(f-I(p))ITm(G, m). The equality irµw,, = i;w says that
w,,([v], [w]) = w (v, w) for all v, w E Tm(-6-1(µ))
Since n and (a,,), are surjective, w,, is clearly uniquely defined.
b) It follows immediately from part ii) of Lemma 4.25 that w,, is well-
defined.
c) w is closed, and so
d (ir,;w,,) = d(iµw) = i7,dw = 0.
and so also irµdw, = 0. Then, because 7r,, is surjective, we can conclude that
&A',,= 0 is closed.
d) w,, is non-degenerate. Thus from
w,,([v], [w))=0 for allwETm''1(i.c)
it follows that
w(v,w)=0 for all wETm4b-I(,u),
thus v E Tm(Gm) from part ii) of the Lemma 4.25, and in this situation
v E Tm(Gpm) from part i); that is, [v] = 0. 0
It should be remarked that when w = dd and t9 is G-invariant, w. need
not be exact.
Remark 4.27. Since it is symplectic, the manifold M. has even dimen-
sion. Because of general principles that we cannot go into here, it turns out
that
dim M. = dim 4i-1(u) - dim G. = dim M - dim G - dim G,,.
4.3. Reduction of phase spaces by the consideration of symmetry 107

Remark 4.28. If p is a regular value of fi, the operation of Gm is lo-


cally free. Then the reduction described in Theorem 4.24 can be taken, at
least locally, in known interesting cases; this remains true even if the global
conditions are not satisfied.

For the sake of completeness, we will now specialize the construction to


the important case for physical applications of M = T'Q. Here G oper-
ates on Q and then, as discussed in the previous section, also on M. The
associated moment map is, as in Theorem 4.17, given by
aq) = aq((tQ)q) for e E g, q E Q, aq E TqQ
Let the conditions for Theorem 4.24 be satisfied; then, moreover, G. oper-
ates without fixed points and properly on Q, so that
Q,.:= Q/GN
is again a manifold.
THEOREM 4.29. Let am be a Gµ -equivariant 1 form on Q with values in
that is, with for all. E g. Given the canonical
symplectic 2 -form wo on T'Q, the form
S2µ := WO - 7r'dam

is then also a symplectic form on T'Q. This then further induces a sym-
plectic form on T*QM, and there is a symplectic embedding
x1,:MM- 7"Q-
to a subbundle over Qm. Ya is then a diffeomorphism of T'QM precisely when
9 = 9N = Lie GM.

Sketch of proof. For


Fm:={(q,aq)ET'Q; aq((i;Q)q)=0 for all E 9M}
we have, using, for example, results about quotients of bundles,
T*QM = Fµ/GM.
From the definition of the fiber 4we get (see Theorem 4.17)
-1(µ) = {(q, aq) E T*Q; for all t; E g}.
Now define
zGµ : b (ie) - Fµ,
(q, aq) '-' (q, aq - (c )q),
a symplectic map with V)µ w0 4 _t():
108 4. The Moment Map

Q Q/Gµ=Q.
Vim is clearly an embedding, and is surjective precisely when g = gµ. We
also have that am is G.--equivariant. With this pr o ioN factors through the
quotients Mj, and so defines x,.. From the definition of symplectic structure
on M. Xv is symplectic. 0
Remark 4.30. ABRAHAM-MARSDEN ([AM], Section 4.5) discuss more
concrete systems with symmetry; these examples really bring Theorem 4.29
to life. They also construct differential forms aµ of degree 1 for this theorem.
It is worth remarking that for p = 0, one may assume that aµ = 0. When
Gisabelianwehave g=gM,andsoM,,-T'Q,,.
Examples.
EXAMPLE 4.31. The theorem of Jacobi and Liouville discussed at the
beginning of this section now takes the following form. Let (M, w) be a
symplectic manifold, and let
fl, ... , fk E F (M) with { fi, ff } = 0 for all i, j.
Since the associated flows Ki, respectively K,, to X f, and X f; commute (this
is a consequence of Corollary 3.16), we get (locally) a symplectic operation
of G = Rk defined on M. The associated moment map is then 4t = Kl x
... x Kk. It can then be taken that the dfi are independent at every point,
so that every u E Im 4; is a regular value of 0. Since G is abelian, we have
Gµ = G, and it gives rise to a symplectic manifold M. = 4i-1(p)/G of
dimension 2n - 2k.
It then remains to show (see ABRAHAM-MARSDEN ([AM), p. 304)) that
every invariant Hamiltonian system on M canonically induces a Hamiltonian
system on M,,. In the important special case n = k, the system is called
completely integrable.
In this sense, an example of a completely integrable system is the Hamil-
tonian system (M = T'R3, wo, Hv) belonging to a central force field, thus
with
Hv(q, p) = (1/2)p2 + V(q),
where V (q) is an SO(3) invariant potential. Then
fi =H, f2=111112, f3=I3 with I=qxp
are integrals with vanishing Poisson bracket.
4.3. Reduction of phase spaces by the consideration of symmetry 109

EXERCISE 4.32. Verify this last statement.


EXAMPLE 4.33. Let XH be a Hamniltonian vector field on M. Its flow
gives a symplectic operation of R on M. whose moment map 4) is then the
function H itself. Thus, for a regular value E E R of H, we have a symplectic
structure on H-'(E)/R =: ME. In these quotients every orbit and every
solution to the H-associated Hamilton's equations can be seen as a point:
ME is therefore called the manifold for the solution of constant energy E.
We have dim ME = dim M - 2.
EXAMPLE 4.34. For G = SO(3). the adjoint operation of G on g = R3
can be written as in Example 4.22. For µ E R3. p 34 0. G,, = SI then
corresponds to the angular velocity of the µ-spanned straight line. The
reduction of M to = Al,, for the associated angular momentum
4i goes back to JACOBI. We have dim Al,, = dim M - dim G - dim G,, _
dim Al - 4 = 2.
EXAMPLE 4.35. Example 4.33 can be further extended to assign sym-
plectic structures to specially given manifolds which can be considered as
quotients in this scheme. This is the case for complex projective space
IP" = P"(C) discussed in Section 2.6. and gives us a proof that this is a
symplectic manifold. Let Al = R2("+I) = 7"R"+' with the canonical sym-
plectic form wo = E=l dqi A dpi, and let
n+1
(1
H (q, P) = 1 2J (q? +P?)
-1
be the Hamiltonian function of the harmonic oscillator. Then we have (see
Remark 3.3)

Xy(9 p) =
n+I
Pi
a a l.
- 9i api
i=I aqi
and the associated flow is
Ft : (q, p) s -. (q cos t + p sin t, p cos t - q sin t).

EXERCISE 4.36. Prove this last statement.


Since Ft is periodic with period 2a, Ft defines a symplectic operation of
SI on M. Because of the compactness of SI. the operation is proper and
clearly fixed-point-free. The value 1/2 is a regular value for H, and we have
H-I(1/2) = Stn+I. Theorem 4.29, considered in the case of Example 4.33,
then says that
H-I(1/2)/R = H-I(1/2)/S' = S2n+,/SI = Pn(C)
is a symplectic manifold of real dimension 2n.
110 4. The Moment Map

EXAMPLE 4.37. We may also construct examples of symplectic mani-


folds with the help of the coadjoint orbit using the procedures discussed in
Section 2.5, which can be, at least partly, adapted here. As in Example 4.23,
let G be a Lie group, and A : G x G - G the action of G on itself by left
translations; thus
(9, h) -As(h).
A then denotes the usual extension of this operation on M = G. Then,
as shown in every example so far, the associated moment map 4i is given by
t (9, ag) = 4%.
Every µ E g' is a regular value for ', and we have
4-1(µ) = {(9, a9) E T* G, a9 ((p9.)9 () = µ (e) for all E g};
that is, here a is the right invariant 1-form aµ whose value at e is p. Thus
a = a,, with (ar.), =,a.
Moreover, it can be seen that
G,,={gEG, A9ap =aµ}
and yet further that
-1(µ)lGv = G/G,u = G*µ E g',
where the last isomorphism is given by the coadjoint representation. In
this way the coadjoint orbit G#µ is seen to be a symplectic manifold. This
is cited as the Kirillov-Kostant-Souriau theorem in ABRAHAM-MARSDEN
((AM], p. 302). They go on to explicitly construct the associated symplectic
form w,,. The result which we gave as Theorem 2.27 then Comes out.
EXERCISE 4.38. Let M = C" be considered as a real symplectic manifold
with w(x,y) := Im(t4) for x,y E C. G = S' = {( E C, I(I = 11 operates
on M by multiplication:
((,x)'lb((x)=(x for (ES', xEM=C".
Make explicit the above concepts in this simple situation and prove that 4C
is a symplectic operation, to which, for x E C" and Y E g = Lie S' -_1fI,
(x)(Y) := (1/2) 11 x 11' Y
gives a moment map. Are these Ad'-equivariant? What does the associated
reduced symplectic space look like?
Here we stop so that we can save room for the last chapter, which will
treat quantization. The reader interested in further concrete examples is
unconditionally recommended to look over the remainder of Section 4.3, as
well as Sections 4.4 and 4.5, in ABRAHAM-MARSDEN [AM]. This material
should now be fairly easy to understand.
Chapter 5

Quantization

As we have already encountered in Section 0.6, the process of quantization


converts the description of the time progression of physical systems on a
syinplectic manifold (in particular on that of the phase space) as described
in classical mechanics to a similar progression on Hilbert space in quan-
tum mechanics. The book by WALLACH [Wa] is dedicated to a critical
account of these problems and includes many historical comments. Also
well worth reading is the book by WOODHOUSE [Wo]. Yet another fully de-
tailed account of these topics can be found in ABRAHAM-MARSDEN ([AM],
Section 5.4, pp. 425 ff.). Shorter introductions occur in KIRILLOV ([Ki],
Section 15.4) and in GUILLEMIN-STERNBERG ([GS], Section 34). However,
these treatments require more supplementary materials from functional anal-
ysis and/or representation theory then we have covered in this treatise; a
proper treatment of this preliminary material would carry us too far afield
of the material of this text. The coverage of the general case in Section 5.5
is therefore, maybe, not based as soundly on the fundamentals as it could
be. Fortunately the easiest conceivable situation, namely M = R2n = T'1gn
can be covered without a great deal of theoretical apparatus, and yet in it
we can already see much of the problems and concepts coming into play. To
make possible the introduction of the Heisenberg and Jacobi groups, which
are equally important for physics and mathematics, we treat this simplest
case in Sections 5.1 to 5.4 in quite some detail.

5.1. Homogeneous quadratic polynomials and s(2


W e begin by letting M = R2,, with the coordinates (q,, ... , qn, pI, .... pn).
(However, when we really want to do some calculations, we will restrict to
the case n = 1.) The starting point will be the fundamental exact sequence

111
112 5. Quantization

for Lie algebras (see Sections 3.2 and 3.3)


0- R 0.
Here i is the embedding which identifies c E R with the constant function
f (m) = c for all m E M, and j assigns to H E F (M) the Hamiltonian
vector field
"
8H 8
XH = E 8H 8
-1
epjCk-av.9pj
F(M) generates an (infinite-dimensional) Lie algebra with the Poisson
bracket
Of 89 Of 89}
If,
j=1 aqi epj 8Pj 89j
as does Ham(M) with the Lie bracket [, ], for vector fields. Thus we have
[X1, X9] = -Xi1,9} for f, 9 E F (M).
Now in order to pass to the quantization, we search for an R-linear map
which assigns to elements of F (M) or, at a minimum, elements f from the
largest possible subset of F (M), self-adjoint operators f in a Hilbert space
?{ such that the Lie structure is preserved in the sense that
{*) U142}= 41,f21= c(flf2 - hfl),
where c = - (c is, up to the factor i, a factor ensuring the symmetry of
the operators, a constant from physics, and h is called Planck's constant).
Moreover, this map should extend
(s*) T=1:=id 7j.
In view of the fact that M can be understood as M = TR", we may take
?{ = L2(R"), which is here the smallest non-trivial Hilbert space. This
search can now proceed with the following general procedure.
We look for a Poisson subalgebra F° of .F(M) which is isomorphic to
the Lie algebra of a Lie group G.
a:Yo =+g=Lie G.
Then (see Appendix D) an irreducible unitary representation of G
7r : C -' Aut?{,
g - A(9),
has an associated infinitesimal representation of g, which for X E g is given
by

dir(X )v = dt ir(exp(tX)v) Ito.


5.1. Homogeneous quadratic polynomials and S12 113

This representation lives on the subspace 7{,,_, of smooth vectors v in 7{, for
which the differentiation process is practicable. There are many deep theo-
rems from representation theory (in particular, from NELSON and HARISH-
CHANDRA) which deal with this space and give answers to the question of
when 71 is dense in W. In the given case, 7t = S(R") is the Schwartz
space, and all works well.
By differentiation of the unitarity condition
(7r(exp tX)v, ir(exp tX)w) = (v. w)
for v, w E 7{x and ) the scalar product in f, we get
(da(X)v, w) + (v,dzr(X)w) = 0.
Thus the operators dzr(X) are skew hermitian. These become hermitian
after multiplication by ±i, and so. because
dir(a({f, g})) = [d r(a(f)), dir(a(g))),
the transformation
F° f a(f) = X da(X) fid7r(X) f
satisfies the condition
if, g) =±idir(a{f,g}) = T-i[f, g].
Choosing here the factor -i, we see that the condition (*) is also satisfied;
that is, once the constant is normalized to 1 by an appropriate choice of
units (as is often done in the case of physical problems). It is naturally easy
here instead of multiplication by -i to use multiplication by c-1. so that
the relation (*) is realized with the constant c.
As the most obvious candidate for 2 to which the general procedure
applies, we consider the Poisson algebra .F2 of the quadratic polynomials in
F. thus (on account of the simplifying assumption n = 1)
.7° = F2 = (gp,P2,g2)

with the relations


{g2,p2} = 4qp,

{gp,P2} = 2p2,
{gp,g2} = -2q2.
.F2 is, via j. isomorphic to the subalgebra Ham2(M) of Ham(M), which for
Ht = qp, H2 = (1/2)p2, H3 = (1/2)q2
is generated by the vector fields

XH,=qa--P op , XH2=Paq, Xjf,= -q .

o'! P
114 5. Quantization

Thus F2 is isomorphic to the Lie algebra g = S12 = Lie SL2(R), which can
be written as
s12=(H,F,G)
with
H=( -1), F=( 0 0 ), G=(1 0
and so
[F,G]=H, [H,F]=2F, [H, G]=-2G.

EXERCISE 5.1. It is recommended that the reader fill in the simple com-
putational details for this example.

The reader may find discussions of the representations of SL2(R) and


$12 in many sources (see in particular LANG: SL2(R) [L3], KNAPP [Kn],
Chapter II, or BERNDT-SLODOWY [BSI). We will also give some details in
Section 5.3.
EXERCISE 5.2. Let V be the vector space with basis (vjU E No}. For
µ E R\{0} let 42 act by
Hvj = (j + 1/2)vj,

Fvj = -(1/(21+))vj+2,
Gvj = (µl2)j(.j - 1)vj-2
Verify that this describes a representation of s[2.

It is worth bringing to the reader's attention that this representation is


isomorphic to an infinitesimal representation of a unitary projective repre-
sentation 7rw of the group SL2(R), the so-called Well representation, which
after a little preparation will be covered in Section 5.3.

5.2. Polynomials of degree 1 and the Heisenberg group


The treatment above allows (at least, in principle) the quantization of sys-
tems consisting of homogeneous quadratic Hamiltonian functions in p and
q. This should lead to, in particular, a description of the harmonic oscil-
lators, but it is not quite enough. A little more can be accomplished by
considering, instead, the case 0 = .F<2(M) of polynomials of degree < 2
in p, q. .F'<2(M) is also a Lie subalgebra relative to the Poisson bracket.
As preparation for considering this case, we next study the case of F1 (M),
which is the homogeneous linear polynomials in p, q. From the fact that
(1) {q+, pi}=b+j, {qi,q,}=0, {pi,pj}=0, i,j=1,...,n,
5.2. Polynomials of degree I and the Heisenberg group 115

we see that .F1(M) is not, a Lie subalgebra of 1(AI) (but .F<1(M) is). The
associated Hamiltonian vector fields are
a a
XP = and X.
57q 57p-

and so 11(Af) is bijectively carried by j to the constant vector fields


Hamc(M)=lag+1R--R2n.

Ham,,(M) is, in any case., a Lie subalgebra of Ham(M). which consists of


the infinitesimal translations of 1R2n. \N,,e thus have before us the following
situation: the Lie algebra 91 = II22n can be identified. via the map t. with
Ham (M). There is. however, a difference from the situation covered in
Section 5.1, where 512 could be identified with .F2(Af) and also with a sub-
algebra Ham2(.F) of Harn(F). Here there is no lifting of A : 91 Ham(Af)
to a Lie algebra homomorphism of 91 to.F(M) as in the following diagram:

0 IR .F(M) - J Ham(M) 0

91

However, there is a Lie algebra, namely .F<1(Af ), which is the image


under j of Ham,(AI) = 91: this is, however, not an isomorphism. but has a
kernel Ker j = R. This situation gives us cause to look at the group 1R2n of
all translations of IR2n, and to search for a group H (1R2n) and an associated
projection j on 1R2n so that for the associated Lie algebras j is carried to j.
Thus we are looking for H (R2") and j with

H(fl22n) 1R 2n

.F<1(M)--91=1R2n

that is, for a group satisfying the Heisenberg commutation relations (1)
This task is solved by
116 5. Quantization

The construction of the Heisenberg group.


Let (V, w) be a symplectic R vector space of dimension 2n. This gives
rise to an associated Heisenberg group which is developed in various ways
in the literature. We give a description of these.
a) The set
H(V)=VxR={h=(v,K); vEV,KER}
is made into the Heisenberg group by the multiplication rule
(3) h1h2 = (vl + V2, x1 + K2 + W(ul, V2)).

EXERCISE 5.3. Some simple calculations will show the following:


i) H(V) is in fact a group. We have
h-1 = (-v, -K)
and

hhlh-1 = (vi, Kl +2w(v, v1)).


ii) H(V) can be realized as the group of (2n + 2)-rowed matrices. In
particular, for v = (A, µ) (A, p E Rn as rows), we have
1 0 0 µ
p
h- A 1 K
0 0 1 'Ca
0 0 0 1

iii) The one-parameter subgroups of H(V) are of the form


G,,,,, := {(tv, ts), t E R}, v E V, K E R.

iv) For the/ associated Lie algebra h = b (V) we have 4 = V x R with


(4) [(ul, t9l), (u2, 192)] = (0, 2w(ul, U2)) for (ttl,'91), (U2, $2) E h

v) For the adjoint representation of H(V) on l1 we have


Ad(,,,.) (u, 19) = (u, 0 + 2w(v, u)).

vi) Forn=1,l =(P,Q,R)iswith [Q,R]=[P,R]=0and[P,Q]=


2R,
0 0 0 0 0 0 0 1 0 0 0 0
0 0 0 0 1
P- 1

0 0 0 -1
0
' Q=
0 0 1
0o00
0
R-
0
0000
0 0 0 0 0 0 0 0 0 0 0 0
5.2. Polynomials of degree I and the Heisenberg group 117

And now returning to the various constructions of the Heisenberg group:


b) For the second treatment, consider, for S' = {( E C, I(I = 1}, the
set
H'(V)=V xS' ={h'=(v,(),vEV,(ES1},
made into a group by
(3') hi h'2= (v1 + v2, (I(2e(w(vi, tri))), e(u) := exp(21riu).
Clearly we have an isomorphism of groups
H(V)/Z ^- H'(V)
with
h=(v,K)'-'h'=(v,e(a))
c) GUILLEMIN-STERNBERG ([GS], p. 94) consider V x S' with the
product
(3*) hlh2 = (vl + v2, (1(2 exp((i/2)uj(v1, v2)))
This is written as H*(V).
d) Often (see for example the fundamental article of A. WEIL [Wel] or
the book of LION-VERGNE [LV]), given an arbitrary K vector space W with
dual space W* and associated canonical bilinear form (,) : W x W' - K,
one takes as the associated Heisenberg group
Heis(W) := W x W' xK={h= (q, p,tc),gEW,pEW ,KEK}
with
hlh2 = (ql + q2, P1 + P2, Kl + K2 + (ql, p2)),
or

Heis'(W):=WxW*xSl={h'=(q,p,(); qEW, pEW*,(ES1)


with
hlh2 = (q1 + q2, Pl + p2, (1(2e((gl,p2))).
Here Heis (W) can be realized as a matrix group, specifically as a subgroup
of GLr+2(K), via
1 q K
h -- 0 1 tp
0 0 1

In what follows we will most often take as our starting point the descrip-
tion in a), which is also the one most often seen in the modern literature.
The Lie algebra h of the Heisenberg group will naturally be called the
Heisenberg algebra. For the natural basis elements (n = 1)
P=(1,0,0), Q=(0,1,0), R=(0,0,1)
118 5. Quantization

of b we have, on account of (4),


(5) [P, Q] = 2R, [R, P] = [R, Q] = 0 ,
and thus
o(1) = 2R, o(q) = P, r(p) = Q
defines a Lie algebra isomorphism o of F<I with h.
For the process of quantization, as formulated at the the beginning of
Section 5.1, it is now interesting (although those who study representation
theory do not need this additional motivation) to search for representations
of 4 and H(V). This is also covered in the literature in a variety of ways. We
begin by stating an analog to Exercise 5.1, that for n = 1 and a E R\{0},
an irreducible representation of 4 = (P, Q, R) on the space W spanned by
vj, j E No, is given by
(6) Pvt = vj+I, Qvs = -pjvi-I , Rvv = (µ/2)v,.
For quantization, this is not yet particularly helpful. It is important, how-
ever, that with just a little legwork a non-trivial unitary representation of
the Heisenberg group can be found. And it can be fairly easily calculated
that, for h = (A, µ, rc.) E H(R2n) and each m E ]R, a unitary representation
of H(R2n) on L2(R') can be given by
(7), (mss (h)f)(x) = em(K + (2x + ))tp)f (x + A), f E L2(flt"),
which for m 36 0 is called the Schrodinger representation of index m.
EXERCISE 5.4. Verify the necessary computations.

In the general theory this can be made to arise as a simple example


of an induced representation in the sense of the theory of MAcKEY (see
Appendix D; for more information, see, for example, KIRILLOV ([Ki], pp.
218-219), the already mentioned book by LION-VERCNE [LV] or the article
Of CARTIER [Ca]). GUILLEM N-STERINBERG ([GS], p. 97) give the following
representation r of H'(R2), which there arises as a result of a discussion
on geometric optics(!). It is given with h = (q, p, () E H'(R') by the
expression
(-I e-iy°neu`p f (x - q) for f E L2(R").
(7') (r(h)f)(x) =
This corresponds to the representation 7rs for m = -1 given above. The
fundamental theorem of Stone and van Neumann says that such a repre-
sentation rrs (respectively r) of the Heisenberg group through its action on
the center (that is, here through ors (0, 0, rc) = em(tc)) is, up to equivalence,
uniquely defined. For a more precise formulation of this theorem and an
accessible proof, the reader is referred to LION-VERGNE ([LV], pp. 19 if.)
5.2. Polynomials of degree 1 and the Heisenberg group 119

It is yet another easy exercise to calculate that from (7) it follows that
(on account of simplicity we take n = 1)

iS
7((t, 0, 0)) f (x)It _0 fi(x),

((0, t, 0)) f (x) It=o = 4irimxf (x),


d 7r'

emirs ((0,0,t)) f (x)It-0 = 2aimf(x).


For the skew hermitian operators

das (P} = das (Q) = 4nimx, do (R) = 21rim


the commutation rules (5) are satisfied. After multiplication by -i (and
normalization by 4am = 1) we get self-adjoint operators

(7) 1 = 1, 4=(1/i), P = x,
which is the first step in the process of quantization. Physicists may be
uneasy that the position variable q represents a differential operator while
p corresponds to a multiplication operator. But this follows from the par-
ticular form of the Schrodinger representation used here, which is the usual
one from the standpoint of function theory. This can be easily converted to
the form more familiar to physicists by considering instead a 7rs -equivalent
representation. The form we have chosen will very soon lead to the usual
formula for the Well representation, and so we will keep this possibly unfa-
miliar usage for now and make the appropriate transformation only in the
larger context.
From the general theory (see Appendix D) it makes sense to carry this
over to the complex numbers. In particular, let
Yf := (1/2)(P±iQ), Zo := -iR.
Then tic =b®RC=Y±,Zo' with
(5') [Z4,Y±]=0 and [Y+, Y- Zo,
as well as
1 d
(r) s (Y) = 2 dx
21rmx and * (Zo) = 2irm.

EXERCISE 5.5. Verify that for p > 0 on V = (vj)jEpb an irreducible


representation of ll, is given by
Y+v, = vi+i, Y vj = -jpvi-I, Z0vj = uv,, j = 0,1, 2, ... .
120 5. Quantization

Show that this may be realized as functions on L2(R), beginning with

th = fo, with fo(x) = e-2Rmx2


and then letting Yf and Zo act as differential operators, as given by (7), to
specify vj (j E N).

5.3. Polynomials of degree 2 and the Jacobi group


We will again focus, on account of its simplicity, on the case n = 1. For the
most part, the following is not difficult to generalize to n > 1. Since F<I
and F2 are Lie algebras, the subspace .F<2 of all polynomials of degree < 2
is also a Lie algebra relative to the Poisson bracket. It is isomorphic to the
Lie algebra
0J=sl2+h
of the Jacobi group GJ(R), which can be brought to light in the following
way. The group SL2(R) operates on the right on H(R) via
(M, h) F--4 hm = (vM, #c) for M E SL2(R) and h = (v, ,c) E H(R).
This gives us justification to form the semidirect product
GJ(R) := SL2(R) x H(R) = {g = (M, h)IM E SL2(R), h E H(R)}
with the multiplication rule
gg' = (MM', vM'+v','

EXERCISE 5.6. Verify the following:


a) We have gJ = (H, F, G, P, Q, R) with
[F, G] = H, [H, F] = 2F, [H. G] = -2G,
[P, Q] = 2R, (H, Q] = Q, IH, P] = -P,
[F, P] = -Q, IG, Q] = -P
and all the other Lie brackets are 0.

b) The map a: -F<- 2 -+ gJ with


a(1) = 2R, a(q) = P, a(P) = Q.
a(qp) = H, a(q2) = -2G, a(p2) = 2F
is a Lie algebra isomorphism.
5.3. Polynomials of degree 2 and the Jacobi group 121

Hint for a): GJ(R) can be realized as the subgroup of Sp2(R) in which
H(1R2) is embedded as described in Section 5.2 while introducing the Heis-
enberg group, and SL2(R) is embedded as follows:
a 0
labl
b 0

M=
0100
c d JJ ~ c 0 d 0
0 0 0 1

Further details on this and for all that will follow in Section 5.3 can be
found in BERNDT-SCHMIDT ([BeS], Chapters 1 and 2).
The operation of SL2 on H fixes, in connection with the already men-
tioned theorem of Stone and von Neumann, a distinguished projective repre-
sentation of SL2, the so-called Weil representation. Using the Schrodinger
representation, r := 7rs on L2(lR), from Exercise 5.5, we can let every
M E SL2 give a rIVI, which acts by
rM(h) := T (hM) for h E H,
giving a representation of H. Since this representation has the same effect as
r on the center of H, it follows from the theorem of Stone and von Neumann
that they are equivalent; that is, for every Al E SL2, there exists a unitary
operator U(M) such that
(8) U(M) T(h) U(M-I) = r(hM ) for all h E H.
This relation fixes U up to a constant factor by a general statement from
representation theory (the so-called Schur's lemma, see for example, KIR-
ILLOV ([Ki], p. 119)). So we have
(9) U(MI)U(M2)=c(MI, M2)U(MIM12).
Therefore c(MI, M2) E CI = {z E C, I z I = 11, and for these c we have. on
the basis of the associativity law in SL2, the relation (see KIRILLOV ([Kl],
p. 218))
c(MI, M2)c(All M2, hi3) =c(MI, M2Af3)c(M2, M3).
This then says that
(10) SL23M- U(M)EAutf
defines a projective representation of SL2 on f.
Remark 5.7. A projective representation in the above sense g 7r (g) of
a group G into the space 7.1 induces a usual representation z in the projective
space P (71) associated to R. This is defined by
v,. - it (g) (v-) = (7r (g) v) _ for v. E 1P (7-l).
122 5. Quantization

This procedure accommodates well to quantum theory, since it is not the ele-
ment v of the Hilbert space that. has physical meaning, but the 1-dimensional
subspace spanned by any such v 0 0.

There is now a standard realization of (10), which is today most often


called the Weil representation, although it also goes under the names of the
Shale-Weil representation, the metaplectic representation or the oscillator
representation. Also the name of Segal appears in this connection. This
representation, here denoted by U instead of by 7rw, is fixed by giving the
actions of the generators of SL2. This leads to the following formulas,
which first appeared in WEIL's fundamental article [Well. which should be
consulted since it introduces many other important concepts as well (also
see, for example, LION-VERGNE ([LV], pp. 1-63), ([Mu2], pp. 133 f.), or
GUILLEMIN-STERNBERG ([GS], pp. 47 65)). For f E L2(R) we postulate

irw (d (a))f (x) = f (ax)IaI1/2

a
ford(a)= 0 a01 I, a#0,
7rw (n (v)) J (x) = J (x)e2amiVX2

forn(v)= I 0 1 I, vER,

7rw (w)f (x) = J(x) = f f (u)e2'riumrdu

for w =

The derivation of these formulas needs more general considerations, but


that they fix a representation can be verified by elementary means. Those
who do not want to try this themselves may find the calculations in Mum-
FORD ([Mu2l, pp. 134-135).
Together with the Schrodinger representation given in the previous sec-
tion, 7rs = r, we obtain via

7rsw(M, h) = 7rs(h)7rw(M)

a (projective) representation of the Jacobi group G-'(R), which we here call


the Schrodinger- Weil representation.
5.3. Polynomials of degree 2 and the Jacobi group 123

EXERCISE 5.8. It is not such an easy calculation to determine the infin-


itesimal representation corresponding to lrswy. For the already given opera-
tors (7) and (7') we have also
fril,(F) = 27rimx2.
(d
(11) 11'(G) = Barn r)2.
frit!(H) = 2+xd
for

F= I. G=I 1 0 I. H=(00 -1
C0 10

respectively, for the complexification


X :E :_ (1/2)(H ± i(F + G)), Z = -i(F - G).

frw(X±) = a + 2 x 1rmx2 I6anr (d)2,

frlv(Z) = 2amx2 - Sam ( T a;)2.

An extension of the last exercise allows us to obtain a realization of the


infinitesimal representation frsw on V = (vj) jEry,. where

9! _ (Zo, Yf, Xf. Z)


for u = 2irm. > 0 operates via
Zovj = µvj, Zvj (j + 1/2)v,,

Y+Vj = Vj+i, X+vj - 2p vj+2,


Y-vj = -pjvj-1. X-vj jj(j - 1)'vj-2
The reader may find the necessary calculations in the notation used here
in BERNDT-SCHMIDT ([BeS[, Chapters 1 and 2) as well as, among oth-
ers, LION-VERGNE ([LVJ, p. 198) and GUILLEMIN-STERNBERG ([GS]. p.
101), where one must ever pay attention to the fact that the group law of
the Heisenberg group has various formulations. In any case. we get here a
representation of the Lie algebra relative to { , }:
.F<2 = (1, p, q? 42, 9P, P2 )

through the differential operators (7) and (11). which operate on a dense
subspace of L2(R) and after multiplication by (-i) solve the quantizat.ion
task from the beginning of Section 5.1.
124 5. Quantization

5.4. The Groenewold-van Hove theorem


The investigations in Sections 5.1-5.3 dealt with the topic of quantization
in the case of polynomials of degree 2, leading to the specification of the
operators in (7) and (11). This then raises the question of whether this can
be somewhat extended. The theorem of Groenewold and van Hove says,
to a first approximation, that this is not possible. This will be made more
precise in a moment; but first, we prove a statement of general interest for
an arbitrary field K of characteristic 0.
THEOREM 5.9. The algebra of the quadratic polynomials p = K[q, PI <2
is, relative to the Poisson bracket, a maximal subalgebra of the algebra
K[q, p] of all polynomials.

Proof. i) The formulas


{(l/2)q2, p'qk} = jpy-lqk+l
-kpi+lqk-',
{(1/2)p2, p1gk} =

{pq, pigk} _ (j - k)pigk


show that by (possibly several-fold) left-Poisson bracket multiplication with
q2, respectively p2, every monomial p1gk, j + k = 1, in the set f1 of the
homogeneous polynomials of degree I can be carried to any other; even more,
that every homogeneous polynomial of degree 1 generates in this manner the
whole of the monomials pigk, j + k = 1.
ii) Let f E K [q, p] be a polynomial of degree > 2. Because
{q, f } _ and {p, f } _ aq -1

we can assume that, by taking left-bracket multiplication with p, respec-


tively q, often enough, an algebra a contained in the algebra p has a poly-
nomial with maximal homogeneous component of degree 3, and after sub-
traction of the quadratic part the polynomial is homogeneous of degree 3.
From i), then, the whole of F3 lies in the algebra a.
iii) Because
{q3, p3} = 9g2p2

is a polynomial of degree 4, we have from i) that the whole of F4 is in the


algebra. And now, since
(3n)g2pn-1,
{q3, pn} =

this can be extended inductively to show that all polynomials of degree


higher than 2 similarly lie in a. 0
5.4. The Groeneaold-van Hove theorem 125

From ('7) in Section 5.2 as well as (11) in Section 5.3. after the normal-
ization 47rm = 1, the map a :

p2 '---+ ix2.

9P'- \I xd +1/2
is a representation of p = .7 <2 which acts by skew hermitian operators on a
dense subspace of 9{ = L2(]R), and after multiplication by (-i) can be made
hermitian. This can be shown by direct calculation, without returning to
the formulas for the Schrodinger-Weil representation.
EXERCISE 5.10. Verify the above construction. and verify the statements
in the following construction.

To adapt our presentation to the one in G UILLEMIN -STERN BERG ([GS],


pp. 101-104), which goes back to CHERNOFF and will be here varied a bit,
and to simultaneously manufacture the properties most familiar to physi-
cists, we introduce a Lie algebra isomorphism q of p given by
77(1) = -1, TIM = -p, rl(p) = -q,
rl(g2) = -p2. n(p2) = -q2, i (qp) = -qp.
Then oil is a Lie algebra isomorphism of p, and A = iaq is given by
A(1) = 1,

A(q) = x,
A(p) _ -i aj,

(12)
A(q2) = x2,
A(p2) = -
A(qp) = -i(x/ +1/2).
Setting f = A f , we see that C77 satisfies the quantization conditions (* )
and (**) given at the beginning, and so
{f,g} _ -i[f,g] and 1=1.
126 5. Quantization

To simplify the notation, we now put


A(q) = x =: Q and A(p) = -i =: P
ai
(the old meanings of the letters P and Q will no longer be needed). Then
we clearly have
(13) A(q2) = Q2 and A(p2) = P2.
In this situation, Groenewold-van Hove's theorem is contained in the
following:
THEOREM 5.11. There are no linear maps A from K[q, p] into an as-
sociative K algebra that satisfy (*) and (**) as well as (13) with A(q) = Q
and A(p) = P.
This means in particular that the quantization map A cannot be ex-
tended to p, since every properly contained Poisson algebra of p is by The-
orem 5.9 equal to K (q, p].
Proof. i) The basic idea of the proof is that, because of the plethora of
relations satisfied by the Poisson brackets, the polynomials can be repre-
sented by a Poisson bracket in a variety of ways. In our case we particularly
consider
(14) q2p2 = (1/3){g2p,p2q}
and
(15) g2p2 = (i/9)(g3, p3}.

From these we will construct a contradiction which says that A assigns to the
same Poisson bracket differing values. This proceeds by a few calculations
and will be given in several steps, beginning with:
ii) We have
(16) A(qp) = (1/2)(QP+PQ).
In fact, from {q, p} = 1 it follows with (*) that
(17) [Q, P] = QP - PQ = i
and from {q2, p2} = 4qp
4A(qp) = -i[Q2, P2] = -i(Q2P2 - P2Q2);
with (17), this gives
4A(qp) = -i(QPQP + iQP - P2Q2)
and by the further repetitive application of (17) then
4A(qp) = -i(2i(QP + PQ)).
5.4. The Groeneuold-van Hove theorem 127

iii) We have
(18) A(q3) = Q3.
Setting A(q3) =: X. it immediately follows from {q3, q} = 0 that
[X. Q] = 0,
and, from {q3, p} = 3q2 with (*) and (13).
[X. P] = 3iQ2.
Trivially. in an associative algebra, we have
[Q3, Q] = 0.
and because of (17)

[Q3, P] = Q3P -PQ3 = Q2PQ - PQ3 + iQ2 = ... = 3iQ2.


From here, it immediately follows that y = X - Q3 commutes with P and
Q. and then, because
[Y, PQ] = YPQ - PQY = [Y. P]Q + P[Y, Q] = 0.
also with PQ, and, by entirely analogous reasoning. also with QP.
This can be used in the following manner. Because {q3, qp} = 3q3 we
get from (16) and (*) that
[X, (1/2)(QP + PQ)] = 3iA(q 3) = 3iX,
but since y = X - Q3 commutes with QP and PQ. we also get
[X, (1/2)(QP + PQ)] = [Q3 (1/2)(QP + PQ)]
= (1/2)(Q4P+Q3PQ - QPQ3 - PQ4).
and, after further repeated applications of (17).
[X, (1 /2) (QP + PQ)] = -3iQ3;
thus X = Q3.

iv) With an analogous calculation. it can be shown that


(19) A(p3)=P3.

v) We have
(20) A(g2p) = (1/2)(Q2P+PQ2).
From {q3, p2} = 6q2p it follows from (*), (18) and (17) that
6iA(q2p) _ [Q3 , P2] = Q3P2 - P2Q3 = Q2PQP + iQ2P - P2Q3
... = 3i(Q2P + PQ2).
128 5. Quantization

vi) And again by the same methods, we arrive at


(21) A(gp2) = (1/2)(QP2 + P2Q)

vii) On the basis of the formulas (18) and (19), we have


(1/9)A{q3, pt} = (_i/9)[Q3,P'J.
From this. we further get. after a few similar calculations (or, when we don't
care about the purity of our method, quickly by returning to (12)),
(1/9)A{g3.p3} = -2/3 - 2iQP+Q2P2.
And then by (20) and (21), from what. is in any case a tedious compu-
tation, we finally arrive at
(1/3)A{q2p, p2q} = (-i/12)[(Q2P+PQ2), (P2Q+QP2)1
= -1/3 - 2iQP + Q2P2.
From (14) and (15) we have the two equations giving the contradiction
promised in part i) of the proof.
EXERCISE 5.12. Supply the missing calculations in the above proof.
Remark 5.13. The operators used here operate on a dense subspace of
L2(R) and cache into being with the help of the irreducible representation
7r' . We can now attempt to extend the area of applicability of quanti-
zation, by enlarging the space on which the operators of the quantization
operate. For instance, we can try to avoid giving the operators as arising
from an irreducible representation. ABRAHAM--A'IARSDEN ([AM[, pp. 435
439) demonstrate the impossibility of extending the quantization to p for
the case of the space of functions with values in a finite dimensional vector
space., in which the steps presented here are just slightly generalized.

5.5. Towards the general case


Kirillov sketches ([Ki], pp. 241 f.) the following picture of the process for
converting from classical mechanics to quantum mechanics. Although this
has already been described at various occasions, we reproduce it yet, once
more to provide it unifying view for the general problems we will now handle.
In classical mechanics, the physical quantities are real functions f E
.F(A!). The symplectic manifold Al is, in the first view, the phase space,
thus the cotangent bundle T'Q to a configuration space Q; but in actuality
the conversion does not only concern itself with this situation. The time
course of a physical system is fixed by giving it. Hamiltonian vector field
X11 E V (AI), where H E .1(M) is the energy of the system. Then the
5.5. Towards the general case 129

traversal of an integral curve -y of X11 describes the time course of a quantity


f E F (141) and is given by the generalized Hamiltonian equation
J' _ (f. H}.
A group G is called a symmetry group of the given system, when it operates
via symplectic transformations on Al.
In quantum mechanics the phase space, or the symplectic manifold Al, is
replaced by the projective space P (il) to an appropriate Hilbert space ? { for
the system. The physical quantities f are then carried over to self -adjoint
operators f . which then operate on (or, more precisely, in) f{ (on the rather
subtle problem of the appropriate space of definition of these operators, we
have no room here to elucidate). The value of the quantity f or f that is
the state of the system is then described by a unit vector v1 E 11 which is
a random variable with the distribution function p (s) = (E9vl, v1), where
E4 is the spectral projection measure of the operator f. This means that in
order to describe the state of the system by vi, the quantities f must have
a fixed value a, for which vi is an eigenvector with eigenvalue a.
The time course is then given by a one-parameter group of unitary
operators on }{ of the form
U(t) = e 12".

where h is Planck's constant. H is here a self -adjoint operator, called the


energy operator. so that then the time course of a quantity f = F satisfies
ih [H, F].
G is a symmetry group of the system, when G operates as unitary operators
on W.
A quantization is now the process of constructing, for a given classi-
cal system, a corresponding quantum system. It is however very difficult
(when it can be done at all) to give a unique interpretation to the word
corresponding. Classical mechanics should, in all possible interpretations,
be understandable as the limit case of quantum mechanics for h - 0. It
should not be expected that the reversal of this process is per se unique.
Be that as it may, KIRILLOV showed that. this process is, in many known
and simple cases, independent of the choice of the quantization process. In
common with the sources we here cite KLRILLOV ([Ki], p. 243), ABRAHAM-
hIARSDEN ([AM], pp. 433 f.), and GUILLEMIN-STERNBERG ([GS], p. 265),
and, in view of the experience gained in Sections 5.1 through 5.4. we can
attempt the following scheme.
We will consider in F (M) the physical quantities of a given physical
system and select from them certain ones, the so-called primary quantities
130 5. Quantization

f, which relative to the Poisson bracket generate a Lie algebra p. The


transformation to a quantum system should now proceed so that
fHf
is an R-linear map of p into a Lie algebra of (self-adjoint) operators on a
Hilbert space 7t with
(*) {fi,f2} =cIfI,f21 for fl, f2 E p c .F(M).
In Kirillov [Ki], we have c = A., where, on the other hand, the Poisson
bracket has the opposite sign from the one we have here used. In ABRAHAM-
MARSDEN [AM] as well as in GUILLEMIN-STERNBERG [GS], c is set equal
to -i. Further, it is required that the constant functions are contained in p
and that we then have
(**) 1 = 1 = idx.
A rule which delivers exactly this is called a prequantization. Otherwise
(see, for example, ABRAHAM-MARSDEN ([AM], p. 434)) for a so-called full
quantization the operation of the i's must satisfy an irreducibility condi-
tion in the sense that there be no proper general invariant subspace, or at
least, only finitely many. For p = .F<2, the construction in Sections 5.1-5.3
gave a full quantization, since it arose as the infinitesimal representation
of an irreducible representation ni, of the Jacobi group. The statement
of the Groenewold-van Hove Theorem 5.11 says that the domain of this
quantization cannot be extended.
We can now follow a somewhat different line, which is here taken from
KIRILLOV ([Ki], p. 243). And this begins by taking, for M = 7"'R", the
subalgebra p of F(M) spanned by the linear functions in pi , ... , p,,, and the
commutative algebra po of functions in p which depend only on q,..., q,,.
This algebra is infinite-dimensional and has the property that
if, g} c po for all f E p, g E po.
In agreement with the terminology of Kirillov, this algebra p will be called
the algebra of primary quantities. Under known conditions, it gives rise
to a representation of the Poisson .algebra F(M), which then, on restric-
tion to p, satisfies a minimality condition for the representation space 7i,
which, it turns out, corresponds to the above-mentioned condition for full
quantization.
The construction of a representation of the Lie algebra of primary
quantities p.
i) The starting point for this construction is the map
.F(M) 3 f . -, Xf E Ham(M),
5.5. Towards the general case 131

which gives the vector field Xf, after Theorem B.4. as a derivation on F (M).
and therefore as a linear differential operator. This extends to
f'f= -iXf
and satisfies, because X{ f.y) _ -[X f, X9] (see Corollary 3.23), the condition
(*) with c = -i. Since the image of the constant functions in Ham(M) is
zero, this procedure must be altered if we wish (**) to be satisfied. Therefore
Kirillov takes, with the 1-form a on M.
(+) f:=-iXf+f +a(Xf).
In this way, (*) is satisfied so long as for a we have da = w.
EXERCISE 5.14. Using the relation 1) from Section 3.1, verify that
da (XI, X2) = Lx, (a (X2)) - Lxs(a (XI)) - a ([XI. X2]).
In the special case of M = T'Q one can take -a = t9 = E pidgi. and
then (+) has the form
n(Ofa_afa1 f_ n of
f = -i 99l apj J + pf gpj
j=1 ap, a9j j=1
and so in particular
a a
Pi = -igaj' qj = iapj +qj

ii) If w is not exact, one can use the above construction in a generalized
form. In this case, by using the fact that w is closed and thus locally exact,
a cocycle h (see Section A.2) is constructed which has an associated vector
bundle E of rank 1. thus a line bundle. Then the construction is generalized
by assigning sections f of this bundle to the primary quantities f as in (+).
This then goes through (see Theorem 1 of KIRILLOV ([Ki], p. 245)), in the
case that the appropriate normalization of the integral of w

Ic

over an arbitrary 2-cycle ry in M is a whole number (an integral multiple of


Planck's constant h in the normalization of Kirillov). It is further shown by
Kirillov that
THEOREM 5.15. There exists a one-to-one correspondence between the
equivalence classes of representations of p and the elements of the cohomol-
ogy group HI(M. C').
132 5. Quantization

In particular for connected M, H' (M, C*) is trivial, and therefore this quan-
tization is unique.
In the general picture constructed so far, nothing has been said about the
Hilbert space R. nor has it been required that the operators f should operate
as self-adjoint operators, or that the choice of f is required to be restricted
to p. The operators which appear in (+) are (differential) operators which
operate on F (M) or on the space r (E) of sections of the line bundle E.
As already mentioned in the discussion of thequantization concept, it is
desirable that the space on which the operators f corresponding to f operate
be as small as possible (so that the operation is irreducible or at least of finite
multiplicity). In the explicit discussion of the examples M = T`Q = R2"
in Sections 5.1-5.3, there appears, somewhat informally, the space N =
L2(R"), which then is the desired space; namely, for n = I we get the
quantization of g1 with the help of the Schrodinger-Weil representation agy
of the Jacobi group GJ(R). In the next most general case M = T'Q we see
that f = L2(Q, du), with an appropriate measure dµ, is the proper space.
Now for the most general situation we will greatly simplify the discussion,
but still hope to make it clear that the choice of p makes it possible to
manage with this small Hilbert space as the domain for the operators f .
When M is not the cotangent space to a manifold, one would expect a
Hilbert space of functions in n = (1/2) dim M variables. The local separa-
tion of variables in q and p via the form

w_ dq,Adpi
is perhaps not globally possible. Halving the dimension is however possible
with the help of the local concepts. This is given by defining on M a Lagrange
distribution L; that is, by defining at every m E M, a Lagrangian subspace
L,,, in TmM (on which u4n vanishes), so that these spaces vary from point
to point in a differentiable way. (This can, like the concept of differentiable
vector spaces, be made more precise). Analytically this leads to the choice
of a maximal commutative subalgebra Fo (U) C Y (U), relative to { , }, for
all neighborhoods U C M. Then we have
Fo(U)={fEF(U)IXf=0forallXEL},

and ['0(E, U) denotes the space of sections of r (E) for which the functions
f E Fo(U) are assigned operators f, in accordance with the rule in ii), that
operate as multiplication operators. This leads to a subspace [o(E) of the
space of sections of t (E), which can also be characterized as the space of
sections which are annihilated by covariant differentiation (see Section A.4)
along arbitrary X E L. The primary quantities f E p are excellent, because
5.5. Towards the general case 133

they satisfy
{ f,.F0(U)} C .F0(U) for all U.
This condition implies that the operators f associated to f leave r0(E)
stable. In the special case of M = T*R" this translates to the result that
the f are primary, that is, sums of arbitrary functions of qI,... , q" and
linear functions of Pi , ... , p". 'H is then the completion of I'c(E) relative to
a fitting scalar product. so that the operators are selfadjoint.
Remark 5.16. This last briefly mentioned theme is introduced in KIR-
ILLOV ([Ki], p. 241-248) with somewhat more, but still not full, detail, and
shows that at least the topics introduced in the previous chapters and in the
appendices can be collected into a general theory. The general theory goes
yet further. but requires yet more application of functional analysis.
I hope that the interested reader is now sufficiently prepared to study
the original sources: ABRAHAM-MARSDEN [AM], GUILLEMIN-STERNBERG
[GS], KIRILLOV [Ki], WALLACH [Wa] and WOODHOUSE [Wo]. In par-
ticular, we refer the reader one last time to WALLACH [Wa], which has
an appendix by R. HERMANN that is a readable introduction to quantum
mechanics with many historical and critical remarks.

je coji? Yo no se
si t.e coji, plums suavisima.
o si coji to sombra.
J.R. JIMENEL
Appendix A

Differentiable
Manifolds and Vector
Bundles

The following definitions are standard in function theory and differential


geometry, and also pass into symplectic geometry. The reader will find this
material covered in [AM], p. 31 if., or [A], pp. 77 if. and 163 if. The
following presentation is also influenced by the books of H. Cartan: Formes
Differentielles [C], Hollman and Rummler: Alternirmnde Differentiolfor-
men [HR], Sternberg: Lectures on Differential Geometry [St], and Chern:
Complex Manifolds without Potential Theory (Ch), as well as the article by
Kt filer: Der innere Di femntialkalkiil [K2].

A.1. Differentiable manifolds and their tangent spaces


The concept of differentiable manifold is fundamental to the study of sym-
plectic (as well as Riemannian) geometry, and so we begin with the definition
of a p-dimensional (p E N) differentiable manifold. Here, differentiable will
mean infinitely differentiable, although in many cases this will be stronger
than needed.
DEFINITION A.1. A connected Hausdorff topological space M with a
countable basis is called a p-manifold if every point m E M lies in an open
neighborhood t1 C M holneomorphic to an open neighborhood of R.

In particular, M is locally compact, and is also referred to as a locally


Euclidian space of dimension p.

135
136 A. Differentiable Manifolds and Vector Bundles

DEFINITION A.2. A chart for Al is a pair (<p, U) (also referred to simply


as V), with U a neighborhood in M and : U --+ Wv a homeo)norphisin onto
an open subset V(U) of RP.

U will be called a coordinate neighborhood, (p the coordinate neap, and cp,


(the. ith component of gyp) the ith coordinate function. The coordinates in w
of the image of U under V will be given by the coordinates x(XI,
=. . . , x ,).
).
So for have r,=cp,(m), i = 1,...,p.
DEFINITION A.3. Two charts cp : U 1R'' and gyp' : U' RP are called
compatible when the transformation functions
-I
`p o `p 1 I ;p(Urnu,)
and 'p o Io,(Iin11')

are differentiable.

In general, we will abbreviate this to Al is a differentiable manifold.


Analogous to this concept of real manifold is the concept of a complex
differentiable manifold of dimension p. In this case each chart V gives a
homeomorphism of an open neighborhood U C 141 to an open set of Cr.
One then requires that the transformation function between any two charts
(cp, U) and (,p', U') be a holomorphic function.
There is not, enough space here to cover the many examples that. can
and do arise in the literature. For this reason only those examples which
show up frequently in this text will be discussed.

i) Open submanifolds. Let Al be a differentiable manifold with atlas a,


and 1410 an open set in Al. Then the family (Al() fl U, cpjA with (cp, U)
in a. is an atlas ao on A-!0. Such A10, with the given differential structure,
is called an open submanifold.

ii) p-dimensional submanifolds of R". In FORSTER ([F]. p. 128) and


LANG ((L5], p. 363). one finds the following definition.
A. 1. Differentiable manifolds and their tangent spaces 137

DEFINITION A.4. A subset M C R9 is called a p-dimensional differenti-


able submanifold of IlPQ if for every, point m E M there exist an open neigh-
borhood V C IIt9 and differentiable functions f, : V IR (i = 1, ... , q - p)
with
(i) M n V = {x E V : fl(x) _ ... = fq_p(x) = 0},
(2) rank Df(in)=q-p,
where
fi
0XI
0(h....' f9-P)
Df = _
8(x1 ...,xq)
1 Of
8z
is the Jacobian off = (fl..... fq_p).
It can be shown that such a differentiable submanifold is a differentia-
ble manifold in the previous sense. Among the simplest examples are the
hypersurfaces, with q = p + 1. Important examples of this class are the
hyperplanes
HP = {x E RP+1 f(x) = 01,
where f is a linear function
p+l p+1
.f (x) = ao + F, aixi, ao, ... , aq,+l E lR, ai2 # 0,
i=1 i=1
as well as the p-sphere
P+1 1
r x72 _1 l
SP x E RP+1
I` 1

iii) Complex projective space PP(C). This is


PP(C) = {(x)_ : z E CP+1\{0} },
the space of all complex lines in CP+1 passing through the origin; that. is,
z - z' if there is a A E C` with z; = \zi for i = 1, .... p + 1. In this case. an
atlas a, for i = 1,...,p+ 1, is given by
Vi`i:U,={(z).: zi#0} - CP,
ZI zi zP+1
m = (z)- x =
Zi Zi Zi

(where the symbol - indicates that this entry should be omitted). Clearly
the transformation functions are holomorphic; for instance, V = p(l) and
138 A. Differentiable Manifolds and Vector Bundles

P' _ P(P+I) We have


Z2'. zp+I
<P(M) = x = zI zI
zI zP
p'(m) = x' _ zP+I zP+1

it follows that
i
'P O
1
I cp(UinUU+t) (x) = r 1 ,2I , ... , X P - 1 l
xP x P xP
is then holomorphic, since xP = zp+I/zI 0 0 in UI fl Up+I.

Morphisms of differentiable manifolds. The appropriate concept of


maps for differentiable manifolds are those maps which are compatible with
the differential structures.
DEFINITION A.5. Let M and N be differentiable manifolds of dimensions
p and q, respectively. A map F : M -+ N is called a differentiable map or a
morphism precisely when the following are satisfied:

i) F is continuous, and
ii) F is given locally by differentiable maps.
This means that there exist atlases o on M and a on N which fix the
underlying differential structures of the manifolds, so that, for charts (gyp, U)
from o and (u(', V) from f3, on w := U fl F- I (V) the composition
V(W) -, W Fi V -O(V)

is differentiable.
A.1. Differentiable manifolds and their tangent spaces 139

Remark A.6.
(1) The condition ii) is independent of the choice of atlases for M and
N.
(2) The phrase F is given by differentiable maps should have the mean-
ing that o F o ;p-I gives a q-tuple of differentiable functions in p
variables in the local coordinates.
(3) The composition of differentiable maps is again a differentiable
map.

DEFINITION A.7. A morphism of differentiable manifolds F : M -p N


is called an isomorphism or a dif feomorphism. when F is a homeomorphism
to N and F-1 : N - M is also a morphism.
As an indication of just how general the concept of a submanifold of R q
is, we have the following statement.
THEOREM A.8. (Whitney) A differentiable manifold of dimension p is
diffeomorphic to a submanifold of R2P+1.

Differentiable functions. For N = R, we get, as a special case of the


definition of a differentiable map, the concept of a differentiable function on
a differentiable manifold M. For an open subset V of M, denote by F(V)
the ring of differentiable functions defined on V, namely those functions
f : V - R8 such that, for every chart (ep, U) of an atlas a of M with
U r V 0, the function f o can be given by a differentiable
function in the p variables xl, ... , xp.
The function f o p-1 for a typical chart (gyp, U) will, in agreement with
the practice found in the literature, also be denoted simply by f. Thus
f (m) = f (x) for the value of the function f for which m E V in the chart
(gyp, U) corresponds to the point with coordinates x.
Analogously, we denote by O(V) the ring of holomorphic functions for
an open set V of a complex differentiable manifold.

Tangent spaces. There are many ways in which one may assign the tangent
space to a point in of a differentiable manifold M. The central point is the
demonstration that it is a vector space which has the same dimension as M.
The following definition formalizes the visual image that the tangent space
in m E M is the set of the tangent vectors to all curves on M which pass
through the given point.
DEFINITION A.9. Let m be a point in M. Then the tangent space of M
at m is
T.. M := CuRM/,,,,
140 A. Differentiable Manifolds and Vector Bundles

where Cu,,,M is the class of all pairs (J, y), where J is an open interval of
R containing 0 and y : J - e M is a differentiable map with y(0) = m, with
an equivalence relation defined on Cu,,,M by
(J, y) - (J', -/)
precisely when
d( , o'>')

for a chart (:p, U) with in E U.

With the help of the chain rule, it can be shown that this definition is
independent of the choice of the chart.
Remark A. 10. When M C R9 is a submanifold, this definition simplifies
as follows: v E R° is a tangent vector to M at m E M (thus E TmM)
precisely when there is, for a given e > 0, a map y e) - M with
y(0) = m and ''y(0) = v.

This remark prepares the general statement that TmM, in a natural way,
has the structure of a vector space of dimension p.
THEOREM A.11. For each chart (cp, U) with m E U there exists a bijec-
tion
h: TmM-+IPp
given by

Xm
(d(o)
(0))
i-l+ ..,p

Now, given a second chart (gyp', U'), with m E U' and v'(m) = y, and
transformation functions y = y(x) = cp' o cp-l I+p(UnU') (x), then we have for
the map h' associated to gyp', on account of the chain rule,

hi(Xm) _ E axt (x) hj(Xm) (i = 1,...,p)


j=1 j
It is thus clear that T,,, M has a canonical vector space structure which is
independent of the choice of coordinates. It is conventional, for a chart
(cp, U) containing m with the coordinates x, to take as a basis for T,,,M the
dual members of the canonical basis e of RP, so

Balm
13 = \ lm19
axl OXplm
A. I. Differentiable manifolds and their tangent spaces 131

(the elements here can only be read as symbols). Then we write


P
a (i=1....,p).
TrnM:) Xmaiaxilm with a,EIR
1=1
The transformation to another chart (p', U') with coordinates y then leads
to the following calculation:
8
Xm = E b3 I with bj = E ax, (x)a,
j=1 yi m r=1
This interpretation is consistent because of how partial derivatives behave
under transformation, and it makes clear the following alternative.
The interpretation of TmM as the space of derivatives. Denote by
.F(m) the collection of all the differentiable functions locally defined at m E
M, thus all pairs (U, f) with U an open neighborhood of m of Al and f a
differentiable function on U. Further, denote by .Fm the set of all equivalence
classes (U, f)- =: f with
(U, 1)-(U'.f')
precisely when there is a neighborhood U" of m contained in U fl U' with
f I p" = f'I u,,. Such an equivalence class (U, f ) - is called a germ of f . Fm
can be easily given the structure of a ring, and from the value f (m) of f at
m the germ (U, f)- also has a uniquely defined value 1(m) at m.
DEFINITION A.12. The tangent space of Al at m is
TmM = Der (.Fm, R).
Here Der (.Fm, lit) is the set of all R-linear maps
L :.Fm -+ R
with
L(f g) = f(m)Lg+g(m)LJ for all j, g E .Fm.
Such linear maps are called derivations and can be interpreted as direc-
tional derivatives on the basis of the following theorems, which show that
Definitions A.9 and A.12 are equivalent. Note that, as the space of linear
maps, the space of derivations has an R vector space structure in a natural
way.
THEOREM A.M. There is an isomorphism
v : Cum Ml- Der (.Fm, R)
given by
d(f
Xm = (J,'Y)- LX_ with
dt y) (0),
142 A. Differentiable Manifolds and Vector Bundles

where f is a representative for the germ f E .fm.


If yU) is a curve with cp; o-y(i) (t) = b;?t, we get the associated derivation,
here denoted by L3. Clearly
L3 f _ d(f o (i)) (0) _ 8(f -1)
(x),
oxi
which, by a slight abuse of notation, may be also written as
L,f jmf.
And so this is the derivative of f in the direction of 7(() (the xj-axis in the
chart (W, U)).

Cotangent spaces. As background for the later introduction of differential


forms, the following concept will be useful.
DEFINITION A.14. Let M E M. Then the cotangent space to the point
mis
TmM = Horn (T .. M, R),
i.e., the dual space of the tangent space, consisting of the R-linear maps an
from TmM to R. These maps will also be called cotangent vectors.
We now introduce some conventional notation. Let an E T;,,M be the
map
am : T,,,M -b R,
Xm -+ (Xm,am) := am(Xm)
Since TmM is a p-dimensional R vector space, this is also true for Tn, M.
The standard basis 1m of TmM associated to the chart cp with coordinates
x corresponds to a dual basis, which will be denoted as (dx)m. Thus (dx)m
denotes the p-tuple of linear forms on TmM with
j)m`f

\ 9 1., ( = bii-
An am E 7`,,,M with
am = > ai (dxi)m, ai E R (i = 1, ... , p)
would in the coordinates y of another chart cp` and the correspondingly
constructed basis (dy)m be written as
am = E bi(dyi)m.
With the notation
Vr-1(y)
x = x(y) = cp o
A.I. Differentiable manifolds and their tangent spaces 143

for the transformation functions between these two charts, we can evaluate
the coefficients:
P
c3xj
b, = E i(y)aj
j=1
Remark A.15. Every f E F(m), and thus every differentiable function
which can be defined on a neighborhood of m, gives rise to a cotangent
vector (l)m, namely the linear form defined by

(Xm, (df)m) = d( dt (0)


(here y again stands for a curve through m with the tangent vector Xm).
This is in agreement with the notation dxi in the case that f = xi is the ith
coordinate function of the chart cp, and with the usual concept of the total
differential. Since, relative to a basis (dx)m, (df ),,, can be represented by
(df)m = E ai(dxi)m,
we have that (dx)m is the dual basis to Im,
i)
a - \l axj
Im, (df )m> = (x) (more precisely, aj = a(fax (gy(m)).
j i
Thus
P
(tf )m = E 8x; (x)(dxj)m/ /

j=1
Maps of tangent and cotangent spaces. Let F : M N be a morphism
of differentiable manifolds, m E M and F(m) = n E N. Then in the
following way we get a map
(F.)m : TmM ,
of the tangent spaces. Given Xm E TmM represented by the curve (J, y) in
M and y(0) = m; then (J, F o y) is a curve in N with F(-y(O)) = n and its
tangent vector can be taken as the image of Xm. Thus
(F.)m(Xm) = (J, F o 1'),. =: Yn E TN,
and this is well defined. Dual to this stands the map
Fm : 7T N -. TTM
of the cotangent spaces given by the following procedure. For 8n E T;, N
we take for the image F;nj3 in T,,,M the form which for all Xm E FmM is
defined by
FmFn(Xm) =
Thus, in other words,
(Xm, FmF fl) _ ((F.)m)(m, A,).
144 A. Differentiable Manifolds and Vector Bundles

F is called a submersion if F.,,, is surjective for all m E M.

A.2. Vector bundles and their sections


The concept, which is general in physics and particularly frequently encoun-
tered in symplectic geometry, of a vector field X on a manifold M, is easy to
grasp. The concept consists of a map which assigns to each point a tangent
vector Xm E TmM. Somewhat more delicate to state precisely is when such
a vector field should be called continuous or differentiable. This succeeds
quite easily, however, with the help of the concept of vector bundle. We
will now study this concept and will follow the treatment by FORSTER in
Riemannsche Fldchen ([FRF], pp. 195 ff.), where one may find the proofs
which are here omitted. As a more complete coverage of this material,
ABRAHAM-MARSDEN ([AM], pp. 37 f.) and the early pages of CHERN
[Ch] can be recommended.
Grossly stated, a vector space bundle, or, more frequently as well as more
briefly, a vector bundle, is a manifold which assigns in a smooth way to each
point m of a basis manifold M a copy of a standard vector space V. Here
one may imagine, for instance, the collection of all m of M = S2, the sphere
in R3, with attached tangent spaces. This is then a 4-dimensional structure.
We now make this both more general and more precise, and examine some
of its properties.
We assume for now that K = JR or C.
DEFINITION A.16. Let E and it! be topological spaces and 7r: E -+ M
a continuous map. Assume further that each fiber E,,, = 7r-1(m), m E M,
has the structure of an n-dimensional K vector space. 7r : E - M or more
briefly just E is called a K vector(space) bundle of rank n over M precisely
when the following condition is satisfied. For every point m E M there are
an open neighborhood U and a homeomorphism h from E11 := 7r-1(U) to
U x K" with the following properties:

i) The projection map factors through h; that is, the following dia-
gram is commutative:

U
A.2. Vector bundles and their sections 145

ii) For every m E U, the map hIE,,, is a vector space isomorphism of

The map h : EU -' U x K" is called a linear chart of E over U. If


it = (U,)jEj is an open covering of M and the hi : EU, -+ Ui x K" are
linear charts, then the family 21 of the hi is called an atlas of E. A vector
bundle of rank n is called trivial when there exists a global linear chart.
h:E-MxK".
Remark A.17. A vector bundle is thus defined to be locally trivial, and
so, under local examination, the concept of a vector bundle delivers nothing
new. It is only with a global study that they become interesting.
THEOREM A.18. Let E -+ M be a vector bundle of rank n over M, and
i an element from an index set I; let
hi:EU,-4UixK", iEI,
be the linear charts of an atlas of E. Then there are uniquely defined con-
tinuous maps
tpij:UifU.-4GL"(K),
such that for the maps
hij:=hiohil:(UiflUj)xK"- (U1flUj)xK"
we have
hij(m, v) = (m, tli,j(m)v) for all (m, v) E (Ui fl Uj) x K".
On U; fl Uj fl Uk the cocycle relation holds:
thj Ojk = ''ik-
146 A. Differentiable Manifolds and Vector Bundles

Remark A.19. As a matter of notation, the maps ii j are called transfor-


mation functions, and the family (vG:I), i, j E I is called the atlas 21= (h;),El
associated to the cocycle.
DEFINITION A.20. Let M be a real or complex differentiable manifold,
E M an R or C vector space bundle of rank n over M, and
21=(h,:EU4-U,xK",iEI)
an atlas of E. The atlas is called differentiable if the associated transforma-
tion functions tb,, are differentiable. Two differentiable atlases 21 and 2l' are
called compatible if 21 U 21' is again a differentiable atlas.

It is easy to see that this compatibility is an equivalence relation. An


equivalence class of compatible differentiable atlases is called a differentiable
linear structure on E. A differentiable vector bundle is a vector bundle
E M equipped with a differentiable linear structure over a differentiable
manifold. This definition deals with the real case. When V is a C vector
space, everything that follows can be easily carried over.
Remark A.21. For a differentiable vector bundle on a differentiable man-
ifold with the projection map r : E - M, E is itself a differentiable manifold
and rr is a differentiable map.
A differentiable vector bundle E - M is called differentiably trivial if
the differentiable linear structure contains an atlas which consists of but a
single chart, E M x K".
DEFINITION A.22. Let E _ M be a differentiable vector bundle and
U C M. A differentiable section of E over U is a differentiable map
f:U-#E with rof=idr.
The collection of all sections will be denoted by I'(E, U) . This is again a
K vector space.
Analogously one can, for every r E No, define the sections of class Cr.
The resulting spaces are then denoted by Cr (U).

A.3. The tangent and the cotangent bundles


Differentiable vector fields can now be defined as differentiable sections of
the associated tangent bundle E = TM of a differentiable manifold M. In
this somewhat general framework, we will consider the construction of vector
bundles.
For an open subset U of the differentiable manifold M, denote by
GL"(-F(U))
A.3. The tangent and the cotangent bundles 147

the group of all invertible n x n-matrices with coefficients in the space .F(U)
of differentiable functions on U. If it = is an open covering of M.
then denote by
Z' (!t. GL,, (-F))
the set of all 1-cocycles relative to it, that is to say, the families (i'1j)i.jEl
with
li',j E GL,(.F(Ui n Uj))
and
1Vij7Jjk = L''ik over Ui n Uj n Uk for all i, j, k E I.
If 21 is a differentiable atlas of a vector bundle over Al, then the family of
the transformation functions of 21 generate such a 1-cocycle. In the reverse
direction, from every 1-cocycle from Zl (ft. GLn (.F)) a differentiable vector
bundle of rank n can be constructed.
THEOREM A.23. Let Al be a differentiable manifold, U = (UI),EI an
open covering of Al and (irij) a family from Z1(I1.GLn(.F)). Then there are
a differentiable vector bundle 7r : E - M of rank n and a differentiable atlas
K")iEl
of E whose transformation functions are the given $Yij.

Of the proof, we will only say enough to show that the bundle E - Al
is given by the space
E':=UU,xK' x{i}CAfxKnxI
iEI
by introducing an equivalence relation
(m. V, i) - (m'. v', i')
precisely when
m=m' and v=yii.(m)v'.
Then, with the help of a few facts from topology. one can show that E arises
as E'/
As an application we will now develop the concepts of the tangent bundle
TM and the cotangent bundle T' M. To this end, let M again be covered by
It = (U1)iE j and, as we did earlier, denote by Bp(i) : Ui -+ K" the coordinate
map on the ith chart. Then, for the tangent bundle TM, we can take for
the transformation functions t4'ij those matrices which are the Jacobians of
the transformations on the charts Bp(i) o thus
i'ij(m) = (cp(')(m))
148 A. Differentiable Manifolds and Vector Bundles

If ;pW has the coordinates y and 1p(i) the coordinates x, this says that from
h;;=hiohjl:(U,nU;)xKP -. (U;nUj)xKP.
(m, a) --+ (m,ap;i(m)a = b),
we may calculate that the vector a = (aµ) in the x-coordinates has, in the
y--coordinates, the components b = with
P "Y"

N=1
For the cotangent bundleT'M one correspondingly takes
O>> (m) =
which for the vectors a* and b' is the transformation given by

Remark A.24. As a set, these are


TM = U Tm M, respectively, 7-M = U T,,, M;
mEM mEM
that is to say, they are the union of all tangent, respectively cotangent
spaces, to all points m E M. A differentiable vector field, thus a global
differentiable section, X : M - TM can be written for a chart (gyp, U) with
the coordinates x = (x1,.. . , xP) (with a slight but already familiar abuse of
notation) by
Iu P
X = a.(x) with ap E F(U).
Cxµ
µ=1
For a chart (ip', U') with coordinates y = (yl,... , yP) we have, correspond-
ingly,
8
X I u, _ b,, (y) with b E F(U'),
v=1
and in the overlapping set U n U' we have (so long as it's not empty) the
transformation rule
bi(y) = '.(y)aµ(x(y)) (contra).

Correspondingly, a field of cotangent vectors can be considered, which we


call differential forms of degree 1 (or 1-forms), and can be introduced as
differentiable global sections
a:M-+T'M
A.3. The tangent and the cotangent bundles 149

of the cotangent bundle. For these, we have analogously


arJU = a,,(x)dxµ, a;E .F(U),
respectively

aJu, = Eb (y)dy,,, b;, E .F(U),


with the transformation rule
b,(y) -(x(y))aN(x(y)) (co).

DEFINITioN A.25. The differentiable global vector fields on M will be


denoted by V(M), and the 1-forms by 1'(M).
Remark A.26. In the older literature, particularly in many texts from
physics, one finds the following definitions: A contra, respectively covariant
vector (or one-fold contra, respectively covariant tensor) on M is a rule
which assigns to a chart (w, U) of M with coordinates x a system of p
differentiable functions A,, with the condition that for two charts (cp, U) and
(gyp', U') with U fl U' 36 0 a relation as in (contra), respectively (co), exists
between the function systems A,,, and A,,,,. In this sense the components
of a vector field describe a one-fold contravariant tensor, and those from a
degree 1 differential form a one-fold covariant tensor.
EXAMPLE A.27. For every differentiable function on M, its total differ-
ential df can be represented by the 1--form which in a chart (gyp, U) with the
coordinates x has the form
P Of

df = µ-1 - d x,.
OXI,

Maps of vector fields and 1-forms. The description of the maps of tan-
gent and cotangent spaces at the end of Section A.1 leads immediately to
the fact that vector fields and their dual objects are mapped by a differ-
entiable map F : M N in the following way. if F is injective, then to
X E V(M) is assigned a vector field F.X on F(M) C N such that at every
n = F(m) E F(M)
(F.X )n = (F.)mXm
is assigned. When F(M) is a submanifold, one can prove that this vector
field is again differentiable. And this is the case when F is an embedding,
that is, all F.m are injective and F is a homeomorphisin of F(M).
In order to pull back a 1-form 13 E 01(N) on M, the injectivity of F is
not needed. F*,3 E 01 (M) is defined at every m E M so that
(F"J3)m := FF[3n with n = F(m).
This assignment gives a differentiable section of T* M.
150 A. Differentiable Manifolds and Vector Bundles

With a little more formalism it can be seen (see, for example, ABRAHAM-
MARSDEN ([AM], p. 45)) that a differentiable map F : M N gives rise
to a differentiable map TF : TM - TN of the tangent bundles and
analogously to a differentiable map 7F: T'N -' T'M of the cotangent
bundles.

A.4. Tensors and differential forms


With the help of some multilinear algebra (see, for example GREUB (Gb],
MARCUS [Ma], or, for the complete background, BOURBAKI ([Bo], Chap-
ter III, §5)) one may carry this concept yet further. Given three (finite-
dimensional) vector spaces T, T' and W, and a number q E N, let
L (T x T', W) be the K vector space of bilinear maps
f : T xT'-+W.
The tensor product of T and T', T ®T', is then the (up to isomorphism,
uniquely determined) K vector space with
L(T x T', W) Hom (T 0 T', W) for all W.
Let Lq(T,W) be the K vector space of q-linear maps f : Tq W, and
Aq(T, W) the K vector space of all alternating q-linear maps f : Tq -+ W.
Then Aq T is the qth exterior power of T, thus the (up to isomorphism,
uniquely determined) K vector space with
q
Aq (T, W) = Hom (AT, w) for all W.
Now put Lq(T) := Lq(T, K) and Aq(T) := Aq(T, K); so that, in particular,
L1(T) = A, (T) = Hom (T, K) = T'
is the dual space of T and A.(T) = l\qP.
Let t = (t1, ... , tn) be a basis of T and t' t') a basis of T'.
Then
t 0 e_ (ti 0 t')i j=1,_,.,n is a basis of T ®T',
tq = Iti, 19 194-q)1<i,<n is a basis of ®q T,
Atq = (tit A ... A ti01<i,<...<i9<n is a basis of nqT,
and, as we've already used,
t' = (ti, ... , tn} with t; (t,) = (ti, tJ) = bij is a basis of T'.
With these algebraic constructs we may now obtain yet more vector bundles
E on M, whose fibers E,n are, for example, ®q T, M or A12 M. In the
generalization of the definitions of the previous section, an 1 -fold co and q-
fold contravariant tensor on M is then, for example, a differentiable section
A.4. Tensors and differential forms 151

X of this bundle E such that X is a map which assigns to each m E Al an


element
t 9 11

XmE®TmM®(®TmM)
every

and is differentiable in the following sense. For point m from a chart


o of an atlas a of M, the components X,,, relative to the standard bases
associated to w, given as described above by

t Im for T..M, respectively. t` = (dx),,, for T,*,, AT,


YX_

are differentiable. Moreover a change of charts carries this assignment via


an adequate assembly of transformation rules for both (co) and (contra).
Remark A.28. Without exploiting multilinear algebra, one can, as in
Remark A.26, also give a definition of these in the style of the old definition.
An I-fold co and q-fold contravariant tensor is a set of data which assigns
to every chart (gyp. U) of M a system
X ' 1<i <n; r=1 s=1,... 1
of differentiable functions in p(U) with the condition that for two charts cp
and cp' with U fl U' 36 0 between the systems X,, and X' a relation exists
which gives for every index i a (contra) transformation formula and for every
index j a (co) transformation formula.
EXAMPLE A.29. As an example of a 1-fold contra and 2-fold covariant
tensor, we have
i9x ax
(APy,j2(y) = E
(y)
rt,11.42
(x(y)) (x(?!))(A,)a;32(x)
ayj

Of greater significance, especially for physics, are those tensors which


moreover realize a known symmetry condition. Before introducing these, we
next consider an example of one of the fundamental objects of Riemannian
geometry.

The R.iemannian metric. For each m E M, let a scalar product ( , ) in

TmM be given-that is, a symmetric positive definite bilinear form. In the


canonical basis representation this means that for v, w E TmM
(u, w) = E vjwjg7j
ij
with
o;; (w(mll :_ _8 Ial for i. i= 1..... v.
152 A. Differentiable ,Manifolds and Vector Bundles

The assignment of an atlas a of M for each m E M and every chart (gyp, U)


allows one to assign to
m' -- i (gij(V (m))) for m E U
a system of functions, defining then a twofold covariant tensor. This tensor
is called a metric or a Riemannian fundamental tensor, when

a) the matrix (gi1(x)) for all x E cp(U) is symmetric and positive


definite,

b) the functions gil (x) are all differentiable, and

c) between the function systems of compatible charts, there is the


following relationship between the transformation functions:

9ij(y) _ a (x(y))j (x(y))9rs(x(1I)}.

Naturally, a metric tensor can be better understood as the differentiable


section of a bundle E over M which has as fibers Em the second symmetric
power S2T,M of the cotangent space 7 M. It is also frequently fixed with
the help of a degree 2 symmetric differential form in the following way (in
the standard--chart (gyp, U) with the coordinates x):
ds2
= F 9ijdxi dxl.
A Riemannian fundamental tensor in this form gives rise to a Riemannian
metric, which for m, m' E M and a curve y linking the two points defines
the distance between the points along -y by

f Vz.
With this the manifold receives yet another additional structure which is
often of great help. One may derive, from a metric fundamental tensor, the
Riemannian curvature tensor which is required for the formulation of the
theory of general relativity. The following theorem is therefore of particular
interest
THEOREM A.30. On a differentiable manifold there is but one Riemann-
ian fundamental tensor.

Proof. See HOLMANN-RUMMLER ([HR], p. 108) or LANG ([L2], p.231).


0
A.4. Tensors and differential forms 153

Of great significance is also the concept of pseudo -Rieman.niann mani-


fold. which comes from a slight weakening of the conditions on the metric
fundamental tensor. In particular one weakens condition a) in the above
definition to the requirement that the matrix (gi,(x)) be symmetric and
non-degenerate.
From total skew-symmetric covariant tensors one gets the (exterior) qth
degree differential forms: differentiable sections CY(q) of the bundle E over
M whose fibers E,,, are the qth exterior powers A9 T,,, Al' of the cotangent
space are called exterior qth degree differential forms or simply q-forms.
In the standard chart (V. U) with the coordinates x, they have the reduced
representation
a(q)Ic- = ai,...iq dxi, A ... A dxiq, ai1,...,iq E.F(U).
1<ij <...<iq<p

Using the notation I. :_ {(i1.... J J1 < i1 < ... < iq < p}, this is fre-
quently more briefly written as
a(q) a(i)dxi, A... A dxiq = a(i)dxq i)
(i)Elq (i)Elq

The set of exterior q-forms on Al forms an R vector space, in which the


value at the points m E M is declared to be the product. This space is
denoted by Q 9(M), and, putting Q°(M) we have the definition:
P
(DS29(1) _: QW).
q=o

The calculus of differential forms. The concept of differential forms


allows one to define yet other operations in addition to taking sums. Here
we will only give a short overview. Begin by letting U C RP be an open set.
It will later be made clear that all can be carried over to differential forms
on manifolds.

1) As already noted, two differential forms


a(q) = a(i)dx9) d(q) _ E 1lq(U)
can be added:
a(q) + Q(q) := t'(a(i) + a(i))dxl(i).
Moreover, the multiplication of a differential form by a function f E .F(U)
is well defined by
for(q) :_ E fa(i)dxq(i).
and therefore 119(U) can be considered as an .F(U)-module.
154 A. Differentiable Manifolds and Vector Bundles

Moreover, n(U) = ®St'(U) is itself a ring. The product is given by


setting
(dxi, A... A dxiq) A (dx A ... A dxx,) := X(ir)dxf, A ... A dxtq+r,
where X(a) = 0 whenever all the i and j are not different, and otherwise
X(a) is the sign of the permutation 7r which carries to
the ordered form (l1, ... , lq+r) with 1 < 11 < ... < lq+r. Linearity then
permits a product to be defined by
nq(U) x Str(U) .--y SZq+r(U),
(a(q), a(r)) a(q) A a(r)

This product is associative, but not commutative. It is called exterior mul-


tiplication and satisfies the commutation rule
a(q) A a(r) = (-1)gra(r) A a(q).

From the interpretation of the differential forms a(q) and a(r) as alternating
q-, respectively r-, linear forms on TmM, we get that a(q) A a(r), through
antisymmetrization of the operation of a(q) x a(r) on (TmM)q+r, defines a
(q + r)-linear map.
Remark A.31. Differential forms may be added coefficient by coefficient
and (by writing one after the other) may be associatively multiplied under
the commutation rules
dxi A dx, _ -dx, A dxi for i 36 j,
0 otherwise.
For the specification of differential forms, one can exploit the already used
ordered or reduced representation, so that an element of the skew-symmetric
representation is usually written as
P
a(q) = E 1 qb71 dx)1 A ... A dx79 q

31,...Jq=1

and so it then holds that


bil...jq = 0, when j1, ... , jq are not q different integers,
= X(7r)ail...iq, when jl,... , jq are pairwise different and
(i1, ... , iq) is carried by the permutation 7r
to (ii,...,jq).

2) The assignment discussed at the end of Section A.3, which assigns


to every f E .F(U) = Q°(U) a 1-form df E 11(U), can be extended to an
arbitrary q-form, and then delivers the exterior differentiation.
A.4. Tensors and differential forms 155

DEFINITION A.32. For a differential form a(Q) in the ordered represen-


tation
a(q) _ a(i)dxi, A ... A dxi,,,
the total differential is defined to be

do(q) da(i) A dxi, A ... A dxiq, da(i) ax(k) dxk.

For further calculations da(Q) is brought to the ordered or the skew-


symmetric forms. The operation d is then extended linearly to the whole of
11(U).
DEFINITION A.33. A differential form a E 11(U) is called closed when
dw = 0, and exact when there is a 0 E 11(U) with a = 0.
The total differential d satisfies the following easily verified calculation
rules for aES1Q(U) and0E1r(U):
d(a + 3) = da + d(3 (for q= r),
d(a A p) = da A /3 + (-1)Qa A d0,
d(da) = 0.
Poincare's lemma says that for convex U every closed differential is exact.
With the help of cohomology groups (see Section B.3) one may significantly
sharpen this important statement.
3) In the interpretation of the differential as a differentiable section of
a bundle, the appropriate behavior under transformation from one chart to
another is already built in. To make our notation compatible, we postulate
the following general behavior under substitution of variables. Let U C RP
be open with coordinates x = (x l, ... , xp), V C W open with coordinates
y = (yl,..., y,-), and
F:U -. V.
x -- F(x) = y,
a differentiable map with the components fi(xl,... xp), i = 1, ... , r. Every
differential
E E b(r)dyi, A ... A dyiq E 11(V)
q=0 (_)
on V then assigns a differential on U by

F ' ( 3 = /3F = > b(i) o Fdfi, A ... A dfsq.


q=0 (i)
It is not hard to prove that this. In the special case of the earlier coordinate
change with the given preconditions, the following holds.
156 A. Differentiable Manifolds and Vector Bundles

Remark A.34. The map F : U V induces a map


F' : O(V) St(U),
Q --+ F`p = OF,
which obeys the rules
(,3+Q')F = j3F+p'F,
(0 A 3')F = OF n YF,
and therefore is a ring homomorphism. Since substitution and differentiation
commute, we have
d(F'l3) = d(13F) = (d)3)F = F'(d13) for all Q E Q(V).
With this remark, it is clear that all the earlier operations defined only
on U C RP, thus locally defined, are also globally defined, and so hold for
differential forms on manifolds.

When a vector field X E V (M) is given, we may define yet another pair
of operations on the differential forms St(M) on M.
4) The inner multiplication i(X) decreases the degree of a differential
form a(q) of f&(M) by one. It is defined by assigning to i(X)a(q), for
X E V(M), the map
i(X)a(g)(X1,---,Xq-1) := a[9)(X,X1,...,Xq_1).
It is not hard to see (see ABRAHAM-MARSDEN ([AM], p. 115)) that
i(X)a(q) E Q9-1(M) really holds, and that i(X) satisfies the following rules:

i) i(X) is R-linear with


i(X)(a(g) A 13(r)) = i(X)a(g) A #(r) + (-1)ga(q) A i(X),3(r),

ii) i (f X)a(q) = f i(X)a(q),

iii) i(X)df = Lx f for f E.F(M), and


iv) F*(i(F.X)/3(q)) = i(X)F`f3(q) for F : M -. M' a diffeomorphism
and 1(q) E n (M).
5) The Lie derivative Lx doesn't change the degree of a differential form.
For Q(q) E f(M) the Lie derivative Lxa(q) is defined at every point m E M
by
(Lxa(g))m
:= d
((Fx(t)`a(g))mLo.

Here Fx means the flow associated to X E V(M). Grossly stated, this


is a one-parameter family Fx(t) of diffeomorphisms in the direction X.
A.4. Tensors and differential forms 157

Somewhat more precisely, a fundamental statement, which will be heavily


used in what comes next, and which can be proven with the help of the
theorem for the solution of systems of ordinary differential equations (see
ABRAHAM-MARSDEN ([AM], p. 67) or STERNBERG ([St], p. 90)), says that
the integral curves ryx to a differentiable vector field X with yx(0) = X,n'
for all m' = yx (0) from a neighborhood U of m form a flow F = Fx ; that is,
a family (Fx(t))t of diffeomorphisms from U to F(U), where the parameter
t E R is taken from an interval containing 0. Next for the complete X for
which U = M can be taken, later also made more general (see ABRAHAM-
MARSDEN ([AM], p. 90)) and thus made independent of U, we can make
the following definition:

Lxa(a) = (Fx(t)a(a))

dt t=o.
For these Lx we have the computation rules (see ABRAHAM-MARSDEN
([AM], pp. 113 ff.))

i) Lxda = dLxa for a E 1(M),


ii) Lxa = d(i(X)a) + i(X )da,
iii) Lxa is R-bilinear in X and a with Lx (aA/3) = LxaAO+a ALx,3,

iv) L fxa = f Lxa + df A i(X)a for f E .F(M),


v) Lxi(X)a = i(X)Lxa.
In complete analogy, one can define the Lie derivative for a vector field
Y E V(M) as
Lx Y:= dt (Fx(-t). Y)It_o'
This is, like the definition of Lxa(Q) above, a kind of symbolic notation
expressing the derivative in the direction of the vector field fx. In principle
one has to go back to the definition of maps of cotangent and tangent spaces
from the end of Section A.2, and arrive at

(Lx Y).. = dt ((Fx (t). 1)n Yn) Jt-o for n := Fx (t)m


and

(Lx a)m = d (Fx (t) m an)It-o


This will be used further in Section B.1, and more details can be found in
ABRAHAM-MARSDEN (]AM], p. 90/1).
158 A. Differentiable Manifolds and Vector Bundles

There is yet another important concept of derivative, the covariant


derivatives. However, these are only definable in the context of connections,
which will now be introduced in a general context.

A.S. Connections
In [Ch], p. 33, CHERN introduces the general concept of a connection on a
differentiable vector bundle E over a manifold M. Here we only consider
the case of the tangent bundle TM on M (see ARNOLD [A] and ABRAHAM-
MARSDEN [AM]). In what follows we give a formulation of under what
conditions a differentiable vector field X E V (M), viewed as a differentiable
global section of TM over M, can be understood as consisting of parallel
vectors. Or - possibly easier to visualize - how a vector Xm E TmM along
a curve y : J M passing through m can be displaced parallelly. This
can be accomplished by making this image more precise. That is, between
the tangent spaces T,,,M and Tm'M of two neighboring points m and m',
which are isomorphic only through an abstract isomorphism, we look for
a connection, i.e. a concrete isomorphism 'm,m' depending differentiably
on (m, m') and such that 0m,m = id. One way to make this more precise
consists of defining a map for a given vector field X E V(M) at every point
m E M, which to every Xm E TmM and every direction Ym E TmM assigns
a derivative (VyX)m of Xm in the direction Ym. Xm does not change in
the direction fixed by Ym if this derivative (VyX)m is 0. This then leads to
the following definition (see ABRAHAM-MARSDEN ([AM], p. 145)).
DEFINITION A.35. A connection V on a differentiable manifold M is a
map
V : V(M) X V(M) --f V(M),
(X, Y) - VyX,
satisfying
a) V is R-bilinear in X, Y E V (M),
b) VfyX = fVyX, and Vy(fX) = fVyX + (Lyf)X for all f E
F(M).
VyX is called a covariant derivative of X along Y (to the connection V).
Lyf means the Lie derivative of f, as covered in the last section. In a
chart (.p, U), with coordinates x and the usual basis ej = for TmM with
m E U, let the Christoffel symbol r?k of the connection V be introduced as
v
Vefek = rikei.
i=I
A.5. Connections 159

For X = aiei and Y = > bjej we then have


C78xai
(1) (DyX) _ ciei with ci = Y bj +E r.,kakbj.
i j j,k
Now let y : J -' M be a curve in M. Unfortunately the associated tangent
vectors (t) E TmM for m = -y(t) do not form a vector field Y on M, but it
still makes sense to see these as a part of such a vector field and to denote
by V (t)X the covariant derivative of X along y. We get
DEFINITION A.36. The vector field X E V(M) consists of parallel vec-
tors along y when
Ory(t)X = 0 for all t E J.
With the use of the frequently needed notation
X=EXiei
we have, from (1). for the covariant derivative of X along y

(2) (V7(t)X)i = dt(Xi(y(t))) +E r;k(y(t))Xk(y(t))7i(t) (i =1,...,p).


j,k
When now the tangent vectors y' (t) to the curve y are substituted for X,
we get, from (2), as a condition for the tangent vectors being parallel in the
sense of the direction fixed by the connection V, the system of differential
equations
(3) yi(t) + rj,k(I(t))y,(t)yk(t) = 0.
j,k
The most important special case is when the connection is fixed and M has
the structure of a Riemannian (or pseudo-Riemannian) manifold and then
one is given a metric fundamental tensor g = (gik), respectively, a symmetric
2-form
d32 = gikdXidxk
Then the solution curve y of the differential equation (3) is called a geodesic.
It is a fundamental fact that for such a curve the length of -y,

has an extreme value. For the given gik,respectively,


tg-I
(9'k) := = (9k,)-I,
we calculate the Christoffel symbol of the connections to g as
r = (1/2) E(ai9j1 + aj9ii - ai9ij)9tk
23

I
160 A. Differentiable Manifolds and Vector Bundles

Connections on vector bundles. Inspection of formula (1) shows that in


the definition of connections given above, the coordinates bj in the direction
Y' remain unchanged under the covariant derivative. As in CHERN ([Ch],
p. 33), one can give a definition for a connection in the general situation of
a differentiable vector bundle E of rank d over M with r (E) as the space of
global sections over M.
DEFLNiTION A.37. A connection V on E - M is a map
V:F(E)--.I,(T'M(9 E)
with
i) V( + 77) _ V£ + Vr) for all , p E r'(E),
ii) V(f) = df ®t; + fV for all t E I'(E), f E.F(M).
Locally on an open set U C M with a d -frame 9i, .. , sd (that is, sections
s{ E r(E, U) so that sl (m), ... , sd(m) are linearly independent for all m E
U) this is written as
d
Os;_F'w;s;, i=1,...,d,
j=1

or in matrix notation as Vs = ws, where the are 1-forms on U. For


E r(E, U), thus { = fist with j E F(U),
we then have because of i) and ii)
(4) vE _ w. + ti+d )s

Thus

(5) dC;+1: l1jktj)dxk, i=1,...,d.


J=1 k=1 j=1

For the case E = TM, the equations (4) and (5) can be equated with
equation (1), and then it can be seen that Definition A.36 is carried into
Definition A.37 when for X, Y E v (m) = r (TM) one substitutes
VyX = (VX,Y),
where here I'(T`M ® TM) x V(M) - V(M) is induced by the
standard pairing
a
dx; = 6{j.
axj
A.5. Connections 161

The generalization of Definition A.36 of connections makes it possible to


make sense of a vector field X being parallel to a curve. Thus, as a conse-
quence of Definition A.37, a section £ E r (E) will be called horizontal when
it satisfies

or, locally,
d
dpi +EwJtj = 0, i = 1,...,d.
i=1
This can be understood as a system of partial differential equations. In
general, they will have no solution, but carry over to a solvable system of
ordinary differential equations, when one searches for a horizontal section
over a given curve 7 on M.
Up to this point we have considered only one chart and a fixed d-frame
si. Suppose now that another d-frame is given by
s' = As,
where A is a nonsingular d x d matrix of differentiable functions on U. If w'
relative to s' is given by
Vs' ='O's"
we derive, as the transformation law of connections,
(6) w 'A = dA + Aw
We further introduce a d x d matrix fI of 2-forms, called the curvature matrix
relative to s, according to the formula
11:=dw+wnw.
Then we can deduce, after exterior derivation of (6) with the matrix Q'
corresponding to s', that the transformation law is
S2'A = AS2.

Here it can be seen that the vanishing of f2 is independent of the choice of


the d-frame. A connection w with
52=0
is called flat or integrable. When the connection is derived from a metric, the
curvature has the geometric meaning that is studied in differential geometry.
In any case, one can by the exterior derivation of the defining equation of S1
arrive at the Bianchi identity
dQ+11Aw-wAS2=0.
162 A. Differentiable Manifolds and Vector Bundles

Given a connection V on the tangent bundle TM, as in (4) and (5), with
the help of the Christoffel symbols I)k and the 1-forms written as

j=1
one may now give, in a natural way, definitions for covariant derivatives
of differential forms, tensors and, more generally, tensors constructed from
differential forms (see for example, ABRAHAM-MARSDEN [AM], p. 148)
and KAHLER [K2], p. 440)). For a differential form a E f2(Q)(M) and
h = 1,... , p, the covariant derivative in the direction his
as
dha r, A era
:= h - r
az
with
era = Q, when a = dz,. A$ + y,
where dxr does not arise in y; thus era = i(Xr)a with inner multiplication
discussed in Section A.4. It is but a short step to generate a (q + 1)-form
by the sum
v
dxh A dhO,
h=1
which then, on the basis of the symmetry of PA, agrees with the exterior
differential do, also discussed in Section A.4.
The covariant derivative of a p-fold contra and q-fold covariant tensor
t is a (p, q + 1)-tensor Vt given by
3... iy = a
(V t)i ...jqe - axh tjl...jq +
rtj..jq hi il..-l is
... + ,tjl...jgrhi
ii ..iy
E il...iy
I
tl.Y2...jgl'lhj1
- ... - L.. u...: hjy'
!

For a much more general treatment which avoids all these indices, the in-
terested reader is referred to Equations Diferentieiks a Points Singuliers
Reguliers by DELIGNE [D].
Appendix B

Lie Groups and Lie


Algebras

As in Appendix A, we shall explicate from this large and important area


of mathematics only the major results and concepts needed in symplectic
geometry. As a guideline, Chapter 6 of KIRILLOV's book Elements of the
Theory of Representations [Ki] and also the corresponding sections from
ABRAHAM-MARSDEN [AM] will serve well. The interested reader might
also want to keep at hand a book with Lie groups or Lie algebras in the
title. Particularly recommended are the classics of WIGNER [Wi], WEYL
[W] and CHEVALLEY [Ce].

B.1. Lie algebras and vector fields


Lie algebras made their first appearance as infinitesimal Lie groups, but
quickly acquired meaning as self-standing algebraic objects. In the follow-
ing, the ground field K will be assumed to be R or C. The theory of Lie
algebras can, however, be considered for a completely general field.
DEFINITION B.I. A K vector space g is called a Lie algebra if a Lie
product (or a Lie bracket) is defined on it; that is, a K-bilinear map
[ , ]:gxg-*g
satisfying

(Antisymmetry) [X, X] = 0 for all X E g,


(Jacobi identity) [[X, Y], Z] + [[Y, Z], X] + [[Z, X], Y] = 0
for allX,Y,Zeg.
163
164 B. Lie Groups and Lie Algebras

Remark B.2. If g is finite-dimensional with basis (X1..., Xn), it is


clearly enough to know the values of the Lie brackets [Xi, X j] for pairs of
basis elements. In other words, one must know the coefficients cikj (i, j, k =
I,-, n) from the expressions
= n
[Xi,Xj] EcjXk
k=1

These coefficients are called the structure constants of g (relative to the given
basis (XI,... ,
EXAMPLE B.3.
(1) The Lie algebra g is called abelian or commutative whenever
(X,Y]=0 for all X,YE9.
Every K vector space can be considered as a commutative Lie al-
gebra when equipped with this trivial Lie bracket.
(2) Every associative K-algebra A can be taken to be a Lie algebra,
by taking for the Lie bracket
[X,Y] := XY - YX for X,Y E A.
In this manner, for the space of all p x p matrices A = Mp(K) we
have the Lie algebra called gt,(K).
(3) Every (not necessarily associative) K-algebra A allows the con-
struction of a Lie algebra Der A, given by the K-linear maps called
derivations; that is, maps satisfying
D : A -* A, with D(XY) = (DX )Y + XDY for all X,Y E A
and [DI, D2] := DiD2 - D2DI as Lie bracket. DerA is called the
algebra of derivations, or also differentiation, on A.
By an easy refinement of Theorem A. 13, it can be shown (see ABRAHAM-
MARSDEN ([AIM, p. 83)) that:
THEOREM B.4. The space V (M) of differentiable vector fields on a dif-
ferentiable manifold M is isomorphic to Der (.1 (M)) as K vector spaces.
This isomorphism is given by assigning to a vector field X on M the
derivation Lx defined by
Lx f (m) = (cf )m) for f E .T(M) and M E M.
Lx f will be called the Lie derivative of f relative to X, and has already
been introduced at the end of Section A.3 as well as in the discussion of the
B.2. Lie groups and invariant vector fields 165

definition of tangent spaces in Section A.1, where it was written, because of


local considerations, as Lx f (m) = Lxmf.
The isomorphism of the theorem gives a Lie algebra structure on V (M),
where (X, Y] for X, Y E V (M) is the uniquely defined vector field given by
L(x,Yj = [Lx, LY]
For a diffeomorphism F : M M'; in accordance with the conclusions from
Section A.3. F.X E V(M') is taken as the image of X E V(M). The map
F. is compatible with the Lie bracket (see ABRAHAM-MARSDEN ([AM], p.
85)).
Remark B.5. F. is a Lie algebra homomorphism; that is, F. is K-linear
with
F.[X,Y] = [F.X,F.Y] for X,Y E V(M).
The Lie derivative of a vector field. It is not very easy to prove the
statement that the map defined in terms of flows, Fx to X E V(M), at the
conclusion of Section A.4.
Lx : V(M) V(M),
Y --i LXY := dt
(Fx(-t). Y)I
e.o,
is the same map given in terms of the Lie bracket as
LXY = [X, Y].

B.2. Lie groups and invariant vector fields


DEFINITION B.6. A Lie group G is a finite-dimensional manifold, on
which a group structure is defined, whereby the group operations are given
by differentiable maps; that is, the maps
G x G ,G, and G -- G,
(9, h) 9h, 9'--' .9-I
are differentiable.

Here, we will focus on the case of real Lie groups; analogous constructs
can be formed over the complex numbers.
EXAMPLE B.7.
(1) G =R" with addition as product.
(2) G = R>0 with multiplication as product; or analogously,
G=SI={z EC,jzj=1}.
166 B. Lie Groups and Lie Algebras

(3) G = GL(n, R) with matrix multiplication as product.


(4) G = G1 X G2, where both GI and G2 are Lie groups.
(5) G = H(R) = {h = (A, µ, tc) E R3), the Heisenberg group of degree
1 with the product
hh' = (A+A',p+µ',K+IC +Ap.'-A'p).

Every Lie group G gives rise to a Lie algebra g = Lie G, which can
be thought of as the linearization of G and whose structure is fixed by G,
here in a neighborhood of the unity. For this assignment there are several
equivalent constructions. Here we take as the starting point a concept which
also finds many other uses.
NOTATION B.B. For each go E G denote by Ay,p the function defined by
g . -. ay9 909,
the so-called left translation given by go, and by py,o the right translation
given by
9'-'Pa,9;=990

For every go, the maps A and ppu are diffeomorphisms of C to itself. They
can also be used to shift given vector fields X on G. This makes possible
the concepts of right-invariant and left-invariant vector fields on G.
DEFINITION B.9. X E V(G) is called left-invariant whenever
(a9o).X = X for all go E G.

The space VI(G) of left-invariant vector fields on G is clearly a vector


subspace of V(G), which can then be proved to be a Lie subalgebra of the
Lie algebra V(G) (and therefore is carried over to g = LieG ).
Remark B.10. The tangent space TeG to G at the unity element e E G
is isomorphic to V (G) as a vector space:
V ,(G) TG.
Here API is given by
pI (X) := Xe for X E V (G)
and 4P2 by
(Z;) := X with X9 :_ (A9).t for l E T,,G.
Then we clearly have
t =
VIV2W /// C
= Ae).
(`` = S tt
B.3. One-parameter subgroups and the exponent map 167

Moreover, from the left-invariance of X


(A9)sXe = Xg,
thus idTG and <p2v1 = idu,(G)
Remark B.11. If X, Y E V (G), then [X, Y] is also in V (G). Then,
because of Remark B.2, we have
[(A9.)X, [X, Y].

And this allows us to fix a definition:


DEFINITION B.12. The vector space TG with the Lie algebra structure
induced by the isomorphism with VI(G) is called the Lie algebra g = Lie G
of G.
EXAMPLE B.13.
LieGL,,(R) = M(R),
Lie {X E 0},
E
as as
o (3) := Lie 0(3) R3
with
X C
[X, YJ - C x n (the vector product in R3).

B.3. One-parameter subgroups and the


exponent map
As already indicated, the passage from a Lie group G to its Lie algebra g
can, to a certain degree, be reversed. To make this statement more precise
requires the following concepts, which can only be briefly mentioned here.
To each £ E TeG denote by X its associated left-invariant vector field
(with Xe = ). Then denote by
yf:R -b G,
t f--. exp t.,
the integral curve to X which for t = 0 goes through e and whose tangent
vector 7f(t) at every point yy(t) is equal to X.,,(1). That such a curve ex-
ists is, as already discussed at the end of Section B.1, a consequence of the
168 B. Lie Groups and Lie Algebras

existence and uniqueness theorem for solutions of systems of ordinary differ-


ential equations (and a simple corollary). For such a curve, it can be shown
(see, for example, ABRAHAM-MARSDEN ([AM], p. 255)) that
exp (t + s)£ = exp tH exp sl;;
that is,
ryf:R -+C
is a (differentiable) group homomorphism. 'yf is called a 1-parameter sub-
group of G. This now makes possible the next definition.
DEFINITION B.14. The map
exp:TeG G,
% (1) = exp
is called the exponential map of the Lie algebra 9 = TeG in G.
This map turns out to be differentiable and induces the identity map on
the tangent space To(TeG) ^- TeG. Therefore it is a local diffeomorphism,
but not a diffeomorphism on G.
Remark B.15. For a differentiable homomorphism F : H --+ G between
two Lie groups H and G and for the induced map
TeF :_ (F,)e : TeH -+ TeG
we have the commutation rule
F(expH g) = expc(TeF)rl for ail n E TeH = Lie H.
Since the map
ry : R -P F(expH ti?)
is a 1-parameter subgroup of G, it is of the form

-y(t) = expc t with d y(t)I t o = (TTF)(rl);


that is,
F(expH,) ='r(1) = 7f(1) = exp = expc(TeF)il.
An important special case is given by conjugation:
1C9 : C --i G,
h +-----+ ghg-1 = p9-,a9h for g E C.

This is a differentiable map, and actually an inner automorphism of G. Now


denote by
I a9) : TG - TeG
Ad9 := 1'e K9 = Te(P9
B.3. One parameter subgroups and the exponent map 169

the adjoint map associated to g E G. As a consequence of the last remark,


it follows that
exp(Adgl;) = rcg expe = g (expt;)g-1 for all f E TAG and g E G.

EXAMPLE B.16.
(1) G = R", with addition as product, has p = Lie G = R', and exp:
R" ---' R'& is the identity.
(2) For G = and its subgroups, the exponential map is the
generalization of the exponential function of matrices; thus
exp : f),
00

A i--. > Am/m! .

.=o
To every A E M,,(R) there belongs the 1-parameter subgroup -YA,
given by
00

1'A(t) = exp to = E tmAm/m!


M=0
From this, for C E GL,,(R), we derive the frequently used compu-
tation rule
exp (CAC-1) = C(exp A)C-1.
Appendix C

A Little Cohomology
Theory

Homology and cohomology groups are important tools for both describing
and characterizing objects in almost every area of mathematics. They are
introduced under the topics of homological algebra and/or algebraic topology.
Here we will only give a few definitions from various sources (for instance,
KIRILLOV [Ki], and GUILLEMIN-STERNBERG [GS]) and treat enough of
their elementary properties to suffice for our study of symplectic geometry.
For a systematic introduction, we recommend MACLANE'S book Homology
[ML] or GODEMENT's book Topologie algebrique et theorie des faisceaux
[Go], which has retained its value as a classical introduction to sheaves and
their cohomology.

C.1. Cohomology of groups


Let G and M be groups with M abelian, and let G operate on M from the
left, that is, there is a map

GxM -+ M,
(g, m) '-' gm,
with
(gg')m = g(g'm),
em = m,
g(m+m') =gm+grn'
171
172 C. A Little Cohomology Theory

for all g, g' E G and m, m' E M. Then (see KIRILLOV ([Ki , p. 21)) an
n-dimensional cochain c is an (n + 1)-linear map

n+1
with
for all 9,9o.. ,gn E G.

The collection of all n-dimensional (or, more simply, n-) cochains is a group,
and will be denoted by
Cn(G, M).
Then a coboundary operator d
d : C?(G, M) Cn+1(G,
M),
c '--i dc,
is given by
n+1
dc(go,...,9n+1) _ E(-1)'c(90,...,9i,...,gn+1)
i^o
If
c = db,
then c E Cn(G, M) is called a coboundary of the cochain b E Cn-1(G, M);
and c is called a cocycle, if
do=0.
Then we get in C' the subgroup of cocycles Z' and the subgroup of cobound-
aries B. The following statement is central.
Remark C.I. We have
dod=0.
EXERCISE C.2. Prove the remark.

We thus have
C"(G, M) D Z"(G, M) D Bn(G, M),
and we can form the factor group
H'(G, M) := Z'(G, M)/B'(G, M).
This is again a group, and is called the n-th cohomology group of the group
G with coefficients in M. Taking a direct sum, we obtain
H* (G, M) := ®Hn(G, M).
n
This is a graded ring or a graded algebra, if M is not only a group, but a
ring or an algebra as well.
C.2. Cohomology of Lie algebras 173

For practical computation. the cochain functions c can also be replaced


by c:
c(hl, ... , hn) := c(e, hl, hih2,... , hi ... hn).
For these c the coboundary operator is then
dc(h1,...,hn+1) = hlc(h2,...,hn+1)
n

+ D-1)'e(h1, .. . hi-1, hihi+1. hi+2, , hn+I )


i=1

EXERCISE C.3. Prove this last formula.

As a hint as to the usefulness of this concept, we inform the reader of the


following example. Given two groups Go and G1 with an operation of G1
on Z(Go) = {g E Go, 990 = 909, 9o E Go}. the center of Go, one can identify
H2(G1, Z(Go)) with the set of equivalence classes of central extensions
1

G1 by Go (see KIRILLOV ([Ki], p. 18)).

C.2. Cohomology of Lie algebras


Let g be a Lie algebra and V a g-module (see GUILLEMIN-STERNBERG
([GS]. p. 417)). Then denote by Ck(g, V) the collection of all n-cochains,
that is, the antisymmetric n-linear maps
j gx...xg-,V.
Then take as coboundary operator the operator
b : Ck(g, V) -, Ck+1(g, V)

defined by
k i
bf(F.O..... k) := E(-'Y 1f(So,...,.i.....l;k)
i=0

+ ,Sj,.. .l;k)
i<j
Thus, with this arrangement, we have for k = 0
df (4) = U,
for k = I
6f f1) = W (6) - W (to) - f ([to, 6 ]),
174 C. A Little Cohomology Theory

and for k = 2
1,f2) = fof(S1,S2) - S1f(So,S2) +

EXERCISE C.4. Verify that 62 = 0.

Thus, in analogy to Section C.1, we may define the cohomology groups


Hk(9, V) Zk(9, V)/Bk(g, V).
In the special case that V is a trivial g-module (that is, l;v = 0 for all
E g and v E V), the first terms of the coboundary operators do not appear,
and if V is simply K, we write
Hk(9) := Hk(9, K).
The vanishing of HI (g) and H2(g) allows for important reductions in sym-
plectic geometry (see, for example, the theorem of Kostant and Souriau in
GUILLEMIN-STERNBERG ([GS], p. 179), which is covered in Section 2.5).
Here let us but list the following general statements from GUILLEMIN-
STERNBERG ([GS], pp. 418 ff.)):

(1) H'(g, V) = {0} for all V precisely when every representation of g


is completely reducible.
(2) If the above is satisfied, it follows that H2(g) _ {0}.
(3) If g is semisimple (that is, it has no commutative ideals), we have,
as in 1), that H1(g, V) = 0 for all V.

C.3. Cohomology of manifolds


It is especially important to actually compute homology and cohomology
groups for topological and geometric objects, for example manifolds, vari-
eties, schemes, etc. This has led to a whole art of technical apparati making
such computations possible. Here we give a brief treatment of the following
related concepts (see KIRILLOV ((Ki]. p. 9)).
Let G be an abelian group, M a manifold, and U = {Ua}aEI a covering
of M by open sets. A k-cochain of U with coefficients in G is then a skew-
symmetric map c defined on every (k + 1)-tuple (no, ... , ak) E Ik+1 with
Ua,,n...nUQk00
to G. Thus
c(...,aiI...Iaj ....) = -c(.... a1,...,ai,...).
C.3. Cohomology of manifolds 175

Ck (it, G) denotes the group of these k-cochains. The coboundary map


d: Ck(it, G) -i Ck+i (it, G)
is given by
k+1
-
dc(ao,...,ak+1) = (-1)ic(ao,...,6i,...ak+1)
i=0
We then have that d2 = 0, and therefore, in complete analogy to what has
come before, we define
Hk(f1, G) := Zk(it, G)/Bk(it, G)
as the Cech cohomology groups of the open covering it. This shows a way to
give M itself cohomology groups. Namely given a finer open covering it' of
it, we get, in a natural way, homomorphisms
Tff : Hk(it, G) - Hk(it', G).
This construct allows us to define inductive limits over a directed system of
open coverings:
Hk(M,G) := limHc(it,G).
u
Fortunately these cohomology groups can already be computed from a single
covering of it (LERAY's Theorem):
Hk(M,G) = Hk(it,G),
when it = {UQ} and the sets UQ and their intersections all have trivial
cohomology in the dimensions n > 1.
A completely different viewpoint allows one to attach cohomology groups
to a real or complex CI-manifold M with the help of differential forms (see
Section A.4). For K = R or C, denote by Zk(M. K) the closed K-valued
k-forms on M; thus w E SZk(M) with dw = 0. Denote by Bk(M, K) the
boundaries of w, thus those for which there is a 19 E SZk-1(M) such that
dig = w. Then the de Rham cohomology groups HDR(M, K) are defined by
HDR(M, K) = Zk(M, K)/Bk(M, K).
De Rham's theorem says that these cohomology groups are the same as the
above-defined Cech cohomology groups:
HDR(M, K) = Hk(M, K).
Appendix D

Representations of
Groups

Many of the concepts from representation theory have already played a


role in the previous appendices. Here, we collect together in one place all
sorts of important concepts which are needed in connection with symplectic
geometry. As a guideline to this material we take §7 from KIFULLOV's book
[Ki]. An introduction to the parts of the theory most relevant to us is also
offered in LANG's book [L4].

D.1. Linear representations


Let G be a group and V a K vector space. Here again we are mostly thinking
about K = ][t or C, although much of what we will say holds also for a general
field. We call T a representation of G in V when T is a homomorphism of
G in Aut V, i.e. when
T (glg2) = T (gl) T (g2) for g1, g2 E G.
dim V is taken as the dimension of T. Two representations T and T' of
G in V and V', respectively, are called equivalent when there is a bijective
intertwining operator between the two; that is, a linear isomorphism U
V - V' with
UT (g) = T'(g)U for all g E G.
The set of intertwining operators U of V and V' will also be designated by
C (T, V). A representation T is called reducible if the representation space
V properly contains a non-trivial subspace V1 invariant under all T (g) with
g E G. The restriction of T (g) to Vl then defines a representation T1 of
G on V1, which will be called a subrepresentation of T. There is then also

177
178 D. Representations of Groups

a natural representation of G on the quotient space V/Vi, which is then


called the quotient representation (in KIRILLOV ([Ki], p. 110), this is called
a factor representation of T). If the invariant subspace V1 C V allows for an
invariant subspace complement T2, then the representation T on V is called
decomposable and is written as T = T1 i@ T2.
A representation is called (algebraically) irreducible if it contains no non-
trivial subrepresentation other than itself. A representation T in V is called
completely reducible if every invariant subspace of V possesses an invariant
complement. This is then equivalent to V being completely decomposable
into a sum of irreducible subrepresentations.
One of the major tasks of representation theory is to give a complete
list of all isomorphism classes of irreducible representations, or at least of
all irreducible unitary representations (see the next section). Another task
is to take a given completely reducible representation T and to reduce it;
that is, to write it as a direct sum of its irreducible parts T; along with their
multiplicities
mult(T T) := dim C (Ti, T).
A character X of G is a homomorphism
X:G-'Ci={CEC, 15I=1}.
For a given representation T of G in V, T' is called the contragredient
representation if T' in V' is given by
g'-' T'(9) T (9-1)'
Under these conditions, it is then a fact that for a map of representations
F : V -+ V the canonical map F' : (V')' --+ V' with
F'(v'')(v) = v'' (F (v))
again gives a map of representations.
It is also but a simple exercise to show that from representations Ti of
G in V (i = 1, 2) one can obtain new representations in the following two
ways.

i) The direct sum T1 ® T2, as a representation of G on V1 ® V2, is given


by

(T1 (DT2)(9)(vr + v2) = Ti(9)vr + T2(g)v2 for v1 E V1, V2 E V2.

ii) The tensor product T1 0 T2, as a representation of G of the tensor


product V1 0 V2, is given by
(T1 O T2)(9)(vl ®v2) = Tl(9)vl (&T2(9)v2 for v1 E V1, v2 E V2.
D.2. Continuous and unitary representations 179

It is a standard result of representation theory that the tensor product of


given irreducible representations can be decomposed into irreducible repre-
sentations.
Parallel to the concept of linear representations, we discuss the projective
representations. These are maps T : G -. Ant V satisfying
T (9192) = c(91,92)T (91)T(92) for all 91,92 E G,
where c : G x G - K is a function satisfying the functional equation
c(9i,92)c(9192,93) =c(g1,92g3)c(92,93) for 91,92,93 E G,
which then insures the equality
T (9192)T (93) = T (9i)T (9293)
Such a projective representation of G on V then induces a representation
on the projective space, P(V) associated to V, which is defined by
T (9)(v..) = (T (9)v), ,
for v- E P(V).

D.2. Continuous and unitary representations


In the applications in this text, we will most often encounter G which are Lie
groups, and so we have a differentiable and topological structure. Thus, we
will need all types of restrictions on the concept of representations of G. Here
we will assume the following data is fixed: T is a continuous representation,
thus a representation of a topological group G in a topological vector space
V (mostly a Hilbert space) with
GxV V,
(g, v) '--- T (g)v is continuous.
For such continuous representations, the definitions of the previous section
are extended in the appropriate sense. For example, in this sense, T1 would
be a subrepresentation of T on V1 C V when V1 is a closed invariant subspace
of V. The extension of the other definitions is entirely analogous.

Unitary representations. A representation 7r of a (not necessarily topo-


logical) group G in the space V is called unitary if V is a Hilbert space and
the operators T (g) for all g E G are unitary. Unitary representations are the
most important in applications to physics. They are particularly tractable,
since every unitary representation is completely reducible. The collection
of isomorphism classes of irreducible unitary representations of G is called
the unitary dual and is denoted by G. G can be supplied with a topological
structure (see KIRILLOV ((Ki], p. 113)).
180 D. Representations of Groups

EXAMPLE D.1. Elementary examples are the following:


i) For an abelian group G, every 7r E G is one-dimensional.

ii) For a compact group G, every 7r E G is finite-dimensional.

iii) For G = IIt" we have G = W", and for G = SI we have d = Z.

Schur's lemma says


LEMMA D.2.

a) If Tl and T2 are (algebraically) irreducible representations, then


every intertwining operator A E C (TI,T2) is either 0 or invertible.
b) A unitary representation 7r is irreducible precisely when

dimC(7r,7r) = 1.

D.3. On the construction of representations


There are several ways to approach the representations of a group, in par-
ticular for the unitary dual G. Here we give a few hints.

Reduction of the regular representation. For a locally compact topo-


logical group, denote by V = L2(G) the Hilbert space of quadratic integrable
functions on V with respect to a suitable measure, dA (a Haar measure, see,
for example, KIRILLOV ([Ki], p. 130)). Denote by g the representation
¢ E L2(G) given by

g (go) 0 (g) = 0 (ggo) for g, go E G


the so-called right regular representation. Correspondingly, the left regular
representation A is given by

,\ (go) 0 (g) _ O W, g) -

An intertwining operator U between g and \ is given by


(g_1).
044 ¢ with q (g) := O
In the general case, the question arises as to whether all 7r E G are subrep-
resentations of g respectively A. For compact G, this is the case (this is a
consequence of the Peter-Weyl theorem).
D.3. On the construction of representations 181

A variant. We now treat a variant of this. We work in the following situ-


ation. G operates contiuously on a differentiable manifold M; that is, there
is given a map
GxM M,
(g, m) '-' gm = co9(m),
where all the spy, for g E G, are continuous, and both

P9192 = `P93 V92 for all 91, 92 E G


as well as
V, = idM
hold. Then gyp, given by

0 (9) 0 (m) _ o (Vy-, (m)) for Q, E F (M),


is a representation of G on the space F (M) of all arbitrarily often differen-
tiable functions on M. This theme can now be varied in a myriad of ways.
To show the simplest case, let j be an automorphic factor, that is, a function
j:GxM liY',

(9, m) (9, m),


satisfying the functional equation
j (g1g2, m) = j (91, 92m) j (92, m) for all gl, 92 E G, m E M.
Then from i with
Pj(9) f (m) = 0 (9-' (m)) j (9-1, m)
we get another representation of G on F (M).
The standard example in this connection is the operation of G = SL2(R)
on 1R2 given by multiplication of the matrix g E G with the column vector
m E R2. Then 0 with
,p (9) f (m) = f (9-1 m) for f E V = R[q, p]
is a representation of G on the polynomial ring R[q, p] in two variables.
Clearly the subspaces of homogeneous polynomials of degree l > 0 remain
invariant and give irreducible representations.

Eigenspace representations. As in the previous example, G operates


contiuously on a differentiable manifold M. Let D be a C-invariant dif-
ferential operator on F (M), and let E (D. A) be the eigenspace of D with
eigenvalue A. Then gyp, defined as in the previous example, is a representation
on E (D, A), called an eigenspace representation. These representations have
been particularly propagated by HELGASON.
182 D. Representations of Groups

The infinitesimal method (for Lie groups). Let it be a continuous


representation of a real Lie group G in a topological vector space V. Then
denote by V,, the space of analytic vectors in V, that is, those v E V for
which
g - 7r (9) v
is a real analytic map. From results of HARISH-CHANDRA and NELSON, V,,
is dense in V. it assigns, as an infinitesimal representation d7r, a represen-
tation of the Lie algebra g of G on V, given by
d7r (X) v := Wt 7r (exp tX )vl t_e for X E g, v E
(A representation of a Lie algebra in a vector space W is a homomorphism
of Lie algebra-, g - . End V, where End V is a Lie algebra as explained
in Section B.1. That d7r gives such a representation can be shown as an
exercise.)
The infinitesimal method consists of these steps, but followed in the
reverse order:
a) Determination of the irreducible representations it of g, or of the
complexification g, = g ® C. (This is somewhat easy to do for the
Heisenberg algebra g = b and for g = s12. The elements for these
were provided in Sections 5.1 and 5.2; the rest is a recommended
exercise.)
b) Examination of which of these a can be integrated to unitary rep-
resentations 7r; that is, the determination of those representations
it of G with d7r = F.

In important cases it turns out that it is uniquely determined by dlr.


If G = H (iit) is the Heisenberg group, one may take for it the Schrodinger
representation, which was described in Section 5.2. For G = SL2(IR), it is
relatively easy to determine the unitary dual G (see, for example, LANG
[L4], where however an error occurs: known equivalent representations are
counted twice). The Weil representation, which appears in Section 5.3, plays
here a special role in that it is a projective representation of SL2(R).
Induced representations. With the use of induction, one may construct,
in great generality, representations of the group G from representations of
its subgroups. To this end. let H C G be a closed subgroup and K C G a
subgroup such that
HxK-
is a topological isomorphism. Further, let
a:H - AutV
D.3. On the construction of representations 183

be a finite-dimensional continuous representation. Then the associated in-


duced representation
r = Indgo
of G is given, in that G operates by right translation on a space fl (a), which
is spanned by the functions
0:G--+ V
with
0 (hg) = A ( h ) (h) 0 (g) for all h E H, g E G
and
11 0 II2:=
I0(k)I2dk < oo,
JK
where A = AH is the modular function of H, that. is, the function with
dr(xy) = IH(x) dry for all x, y E H,
whenever dry is a right-invariant Haar measure on H. The factor A1/2
works so that r turns out to be unitary whenever a is.
One of the most significant theorems from representation theory is the
subrepresentation theorem, which is a statement that all interesting repre-
sentations are captured as subrepresentations of known induced representa-
tions. As an example, the Schrodinger representation is relatively easy to
realize as an induced representation (see, for example, BERNDT [Be]). The
induced representations were particularly strongly propagated by the efforts
Of MACKEY, [M].

The adjoint and coadjoint representations. To every Lie group G there


belongs a Lie algebra g = Lie G, and also a natural representation of G on
g, called the adjoint representation Ad. It arises from the fact that through
conjugation
K9(go) = e9-Ag(go) = ggog-1
a diffeomorphism of G is defined, and so, for every g E G, K9 induces maps
of the tangent spaces T..G. In particular, for go = e, TeG is mapped to
itself. It is written as
Adg := (Kg)se = (Log-"Y.e.
The action
GxTG TeG,
(g, Y) Ad9Y,
is defined via the identification of TeG and g, because K9,gz = Kg, K92 is a
representation of G on g, in which the operators are also written as Ad(g).
One may also consider the contragredient representation, as in Sec-
tion D.1, on the dual space g' of g to this representation. This is called
184 D. Representations of Groups

the coadjoint representation of G, and is of particular significance for sym-


plectic geometry.
Since for all go E G, the rcg(go) are differentiably dependent on g, the
infinitesimal representation d(Ad) =: ad associated to Ad can be built as
was described above under the infinitesmal method. It is not hard to see
(see KIRILLOV ([Ki], p. 97)) that then
adX(Y) = [X,Y].
Bibliography

[AM] Abraham, R., and Marsden, J.E., Foundations of Mechanics, Benjamin-Cum-


mings, Reading, MA, 1978.
[Ae] Aebischer, B., Borer, M., Kahn, M., Leuenberger, Ch., and Reimann, H.M.,
Symplectic Geometry, Prog. Math., vol. 124, Birkhauser, Basel, 1994.
[A) Arnold, V.I., Mathematical Methods of Classical Mechanics, Springer, New York,
1978.
[Ar] Artin, E., Geometric Algebra, Interscience Publ., New York, 1957.
[Be] Berndt, R., Darstellungen der Heisenberggruppe and Thetafunktionen, Ham-
burger Beitri ge zur Mathematik, Heft 3, 1988.
[BeS] Berndt, R., and Schmidt, R., Elements of the Representation Theory of the Jacobi
Group, Prog. Math., vol. 163, Birkhhuser, Basel, 1998.
[BS) Berndt, R., and Slodowy, P., Seminar uber Darstellungen von SL2(F) and
GL2(F), F em lokaler Korper, Hamburger Beitrage zur Mathematik, Heft 20,
1992.
[BI] Blair, D., Contact Manifolds in Riemannian Geometry, Lecture Notes in Math.,
vol. 509, Springer, Berlin, 1976.
[Bo) Bourbaki, N., Elements de Mathdmatique, AlgCbre. Chapitre III: Algcbre multi-
lineaire. Hermann, Paris, 1948.
[C) Cartan, H., Differential forms, Houghton-MitHin, Boston, 1970.
[Cal Cartier, P., Quantum Mechanical Commutation Relations and Theta Functions,
in Algebraic Groups and Discontinuous Groups (A. Borel and G.D. Mostow, ed-
itors) Proc. Symp. Pure Math., vol. 9, Amer. Math. Soc., Providence, RI, 1966,
pp.361-383.
[Ch] Chern, S.S., Complex Manifolds without Potential Theory, Van Nostrand, Prince-
ton, NJ, 1967.
[Cel Chevalley, C., Thdorie des groupes de Lie, Hermann, Paris, 1951.
[CH] Courant, R. and Hilbert, D., Methods of Mathematical Physics, Volume 1, Inter-
science, New York, 1953.

185
186 Bibliography

[D] Deligne, P., Equations Difflrentielles a Points Singuliers Raquliers, Lecture Notes
in Math., vol. 163, Springer, Berlin, 1970.
[dCj do Carmo, M.. Riemannian Geometry, Birkhauser, Boston, 1992.
[E] Eichler, M.. Introduction to the theory of algebraic numbers and functions, Aca-
demic Press, New York, 1966.
[EG] El Gradechi, M.A., Theories classiques et quantiques sur l'espace-temps anti-de
Sitter et leurs limites a courbure null. These de Doctorat, University Paris 7,
Paris, 1991.
[FL] Fischer, W., Lieb, I., Funktionentheorie, 7. Auflage. Vieweg, Braunschweig-
W iesbaden. 1994.
[F] Forster. 0., Analysis 1, 3. Aufage. Vieweg, Braunschweig--Wiesbaden. 1984.
[FRF] Forster, 0., Lectures on Riemann surfaces, Springer, New York, 1981.
[Go] Godement, R., Topologie algebrique et theorie des Hermann, Paris,
1958.
[GW] Goodman, R., and Wallach, N.R., Representations and invariants of the classical
groups, Cambridge University Press, 1998.
[Gbj Greub, W., Multilinear Algebm, Springer, New York 1997.
[Gr] Gromov, M., Pseudoholomorphic Curves in Symplectic Manifolds, Invent. Math.
82 (1985), 307-347.
[GS] Guillemin, V., and Sternberg, S., Symplectic Techniques in Physics, Cambridge
University Press, 1984.
[HR] Holmann, H., and Rummler, H., Altcrnierende Difjerentialformen, Bibbliogr.
Inst., Mannheim, 1972.
[HZ] Hofer, H., and Zehnder, E., Symplectir. Invariants and Hamiltonian Dynamics,
Birkhauser, Basel, 1994.
[Jai Jacobson, N., Basic Algebra, Freeman, San Francisco, 1974.
[K] Kahler, E., Uber eine bemerkensurerte Ifermitesche Metrik, Abh. Math. Sem.
Univ. Hamburg 9 (1933) 173-186.
[K1] Kihler, E., Einfiihrung in die Theorie der Systeme van Differentialgleichungen,
Teubner, Leipzig. 1934.
[K2] Kiihler, E., Der inhere Difjerentialkalkul, Rendiconti die Matematica 21 (1962)
425 523,
[Ki] Kirillov, A.A., Elements of the Theory of Representations, Springer, Berlin, 1976.
[Kn] Knapp, A.W., Representation Theory of Semisimple Groups, Princeton Univer-
sity Press, 1986.
[Li] Lang, S., Algebra, Addison-Wesley, Menlo Park, NJ, 1984.
1L21 Lang, S. Fundamentals of Differential Geometry, Springer, New York, 1999,
[L3[ Lang, S., Linear Algebra, Springer, New York, 1985.
[L4] Lang, S., SL2(R), Springer, New York, 1989.
[L5] Lang, S., Undergraduate Analysis, Springer, New York, 1983.
[LL] Landau, L.D., and Lifschitz, E.M., The Classical Theory of Fields, Addison-
Wesley, Reading, MA., 1961.
[LV] Lion, G., and Vergne, M., The Weil Representation, Maslov Index and Theta
Series, Prog. Math., vol.6, Birkhauser, Basel, 1980.
Bibliography 187

[M] Mackey. G.W., Unitary Group Representations in Physics, Probability, and Num-
ber Theory, Benjamin/Cummings, Reading, MA, 1978.
[Mal Marcus, M., Finite Dimensional Multilinear Algebra (in two parts), Marcel
Dekker, New York, 1977.
[ML] Mac Lane, S., Homology, Springer, Berlin, 1963.
[MS] McDuff, D., and Salamon, D.A., Introduction to Symplectic Topology, Oxford
University Press, 1995.
[Mu] Mumford, D., Algebraic Geometry, I. Springer, New York 1976.
[Mull Mumford, D., Tata Lectures on Theta II, Prog. Math., vol. 28, Birkhiiuser, Bos-
ton, 1984.
[Mu2) Mumford, D., Tata Lectures on Theta III, Prog. Math., vol. 97, Birkhauser,
Boston, 1991.
[Sal Satake, I., Algebraic Structures of Symmetric Domains, Iwanami Shoten. Tokyo,
and Princeton University Press, Princeton, 1980.
(Schl Schottenloher, M., Geometrie and Symmetrie in der Physik, Vieweg, Braun-
schweig/Wiesbaden, 1995.
[Sill Siegel, C.L., Symplectic Geometry, Academic Press, New York, 1964.
(Si2l Siegel, C.L., Topics in Complex Fhnction Theory, Vol. III. Wiley-Interscience,
New York, 1973.
[SMI Siegel, C.L., and Moser, J.K., Lectures on Celestial Mechanics, Springer, Berlin,
1971.
[So] Souriau, J.-M., Structure of Dynamical Systems, A Symplectic View of Physics,
Prog. Math., vol. 149, Birkhauser, Boston, 1997.
[St] Sternberg, S., Lectures on Differential Geometry, Prentice-Hall, Englewood
Cliffs, NJ. 1964: 2nd ed., Chelsea. New York, 1983.
[V] Vaisman, I., Symplectic Geometry and Secondary Characteristic Classes, Prog.
Math., vol. 72, Birkhauser, Boston, 1987.
[Wa] Wallach, N.R., Symplectic Geometry and Fourier Analysis, Math. Sci. Press,
Brookline, MA, 1977.
[We] Weil, A., Varidtt;s Kohleriennes, Hermann Paris, 1957
[Well Weil, A., Sur certain groupes d'op6rateurs unitaires, Acta Math. 111 (1964)
143-211.
[W] Weyl, H. The Classical Groups. Princeton University Press, 1946.
[Wil Wigner, E., Group Theory and its Application to the Quantum Mechanics of
Atomic Spectra, Academic Press, New York, 1959.
[Wo] Woodhouse, N., Geometric Quantization. Clarendon Press, Oxford, 1980.
Index

angular momentum, 104 connection, 48, 158


atlas integrable, 161
of a vector bundle, 145 contact manifold, 86
automorphic factor, 181 weak, 85
contravariant or covariant vector, 149

basis coordinates, symplectic, 36


complex unitary, 28 cotangent space, 142
dual, 11 covariant derivative, 158
[-related unitary, 29 curvature matrix, 161
real unitary, 28
related symplectic, 22 derivation, 141, 164
symplectic, 12 difeomorphism, 139
Bianchi identity, 161 symplectic, 83, 84
bundle, differentiable function , 139
characteristic, 87 differentiable section, 146
cotangent, 45, 147 differential form
differentiable vector, 146 closed, 155
tangent, 147 covariant derivative of a, 162
vector, 144 degree 1, 148
exact, 155
qth degree, 153
canonical form, 12
characteristic line fields, 91 reduced representation, 153
Christoffel symbol, 159 skew symmetric representation, 154
coadjoint orbit, 110 differential system, 56
coadjoint cocycle, 96 involutive, 57
coadjoint orbit, 54 differentiation, exterior, 154
coboundary operator, 52, 172 dual space, 11
cocycle identity, 96
cocycle relation, 145 energy function, 74
cohomology groups equations of motion for a charged particle,
De Rbarn, 175 75
of a group, 172 Euler-Lagrange equation, 2
of a Lie algebra, 52, 174 exponent map, 168
of a manifold, 175
configuration space, 1, 45 fiber derivative, 102
conjugation, 53 Row, 41, 156

189
190 Index

form Lie derivative


closed of type (1,1), 49 of a differential form, 52, 72, 156
Kkhler, 50 of a function, 71
left invariant q, 52 of a vector field, 72, 165
Liouville, 45 Lie group, 165
of type (1,1), 49
presymplectic, 85 manifold
symplectic, 9, 35 hermitian, 49
fundamental Kiihlcr, 45
duality, 72 of the solution of constant energy, 109
exact sequence, 78, 83, 111 real, 1
Riemannian, 101
smooth, 35
geodesic, 159 symplectic, 35
group map,
general linear, 24 differentiable, 138
Heisenberg, 166 moment, 94
Jacobi, 120 symplectic, 36
orthogonal, 24 map of
symplectic, 15, 24 tangent and cotangent spaces, 143
unitary, 24, 28 vector fields and 1-forms, 149
matrix, symplectic, 16
Hamilton's equations, 3 metric
Hamilton-Jacobi equation, 4 Fubini- Study, 64
Hamiltonian equations, 74 hermitlan, 26, 28
Hamiltonian function, 3 Kiihler, 47
time dependent, 89 PoincarE, 33
Hamiltonian G-space, 94 Riemannian, 48, 151
Hamiltonian system, 74 w-compatible pseudohermitian, 26
Minkowski space, 75
completely integrable, 108
moment map
harmonic oscillator, 104
Ad' -equivariant, 96, 105
Heisenberg algebra, 117
momentum, 103
Heisenberg group, 116, 117
morphism, 55
horizontal section, 161
of differentiable manifolds, 138
hyperbolic pair, 19
stable, 17
hyperbolic plane, 19 strongly stable, 17
symplectic, 14
infinitesimal generator, 94, 97 multiplication
inner product, 6, 13, 72 exterior, 154
integral curve, 6 inner, 156
integral manifold, 57 Mumford's criterion, 48
intertwining operator, 177
isotropy group, 22 normal form
of bilinear forms, 10
of contact forms, 86
Jacobi identity, 163
observable, 7
Lagrange distribution, 132 one parameter subgroup, 168
Lagrange function, 2, 102 orientation, 12
left translation, 166 orthogonal, 13
Legendre transformation, 3
lemma phase space, 3, 45, 88, 99
Poincar6's, 155 reduced, 105
Schur's, 180 Poincard's lemma, 155
Lie algebra, 163 Poisson bracket, 79, 80
Lie bracket, 7, 163 position function, 101
Index 191

potential of a Kiihler form, 50, 66 subspace


prequantization, 130 coisotropic, 19, 21
primary quantities, 7, 130 isotropic, 19, 21
principle of least action, 2 Lagrangian, 19, 21
pseudoholomorphic curve, 68 real Lagrangian, 31
symplectic, 19
suspension of a vector field, 90
quantization, 7, 112, 129 symplectic
full, 130 capacity, 69
invariant, 17, 21, 70
operation, 54, 93
radical, 18 radius, 69
rank of a bilinear form, 10 reduction, 58, 104
regular energy plane, 88 space, standard, 14
representation transvection, 17
adjoint, 54, 95, 183 symplectomorphism, 15, 36
coadjoint, 54, 95, 184
continuous, 179 tangent field of ,b/' along Ft, 36
contragredient, 178 tangent space, 35, 139
induced, 183 tensor, 150
infinitesimal, 112, 182 Riemannian fundamental, 152
irreducible, 112, 178 theorem
left regular, 180 Darboux's, 42, 85
linear, 177 de Rham's, 175
of a Lie algebra, 182 Darboux's, 36
projective, 121, 179 Flrobenius', 57
Schrodinger, 118 Groenewold- van Hove, 124
Schrodinger Weil, 122 Gromov's, 68
unitary, 179 Jacobi's, 83
Weil, 121 Liouville's, 76
right translation, 166 of Kostant and Souriau, 52, 60
of Stone and von Neumann, 118
Witt's, 21
scalar product
Euclidean, 26 unit ball, 49
hermitian, 24
vector
Siegel upper half plane, 30, 32
analytic, 182
skew hermitian operator, 113
space smooth, 113
vector field
affine, 23
characteristic, 87
complex projective, 63, 109, 137 differentiable, 146
hermitian, 28
Hamiltonian, 6, 74
homogeneous, 22 left invariant, 166
hyperbolic, 19 local Hamiltonian, 77
reduced, 105 vector space
structure Kiililer-, 26
canonical Euclidian, 24 positive Lagrangian, 30
canonical symplectic, 24 symplectic, 10
compatible complex, 25 Hamiltonian, 6
complex, 2.5, 46 volume form, 12
contact, 85
differentiable linear, 146 Wirtinger calculus, 49
hermitian. 26
positive compatible complex, 30
structure constants, 164
submersion, 144
Symbols

Symplectic vector spaces


(V, w) symplectic vector space 14
W1 to WCV the orthogonal space relative to w 22
radW=wnwl radical of W 22
Wred = W/rad W the symplectic space associated to W 23
L=L1CV Lagrangian subspace 23
C(V) collection of Lagrangian spaces L C V 26
T(L) ={L'EC(V),L®L'=V} 27
J=J(V,w) space of w-compatible positive complex
structures J 36
s7n = Sp, (R)/U(n) Siegel upper half-space .7(R' , w0) 36
h(v, w) = g(v, w) + iw(v, w)
hermitian, Riemannian and outer form 32, 55
JEAutV complex structure (J2 = -id) 29

Symplectic manifolds

rn.EM point of a differentiable real


manifold of dimension 2n
cp:U-4R2o chart of a neighborhood U C M 144
V(m) _ (q, p) symplectic standard coordinates 40

193
194 Symbols

f E .F(M) differentiable function on M 147


X E V(M) differentiable vector field on M (= C(TM)) 155
a E 1l (M) differentiable exterior q-form on Al 161
W E i22(M) symplectic form on M, in particular 39
wo = E. dqi A dpi the standard form 40
19 pi dgi the Liouville form 49
W# fundamental duality fl'(M) = V(M) 78
Wb inverse mapping to w# 78
F:M-.M' diffeomorphism in i F(m) = m' 146
Fetrn mapping of the tangent vectors F.,,,(X,,,) = X;,,, 151
mapping of the 1-forms F;,,(a;,,,) = a,,, 151
i(X)a inner product of X E V(M) with a E 11 (M) 17, 78
Lxf Lie derivative off E.F(M) by X E V (M) 77
Lxa Lie derivative of a E f24(M)
by X E V(M) 56, 164
LxY = [X, Y] Lie derivative of Y E V(M) by X E V (M) 78, 173
V connection 166
Vxa, VxY covariant derivative 166
Ft flow of X E V(M) 45, 165
(f, h} Poisson bracket of f, h E.F(M) 85, 86
Xf E Ham (M) Hamiltonian vector field to f E ,F(M) 80
X el infinitesimal generator of X E 8 = Lie C
and the group operation ¢ 99, 103
/:Gx M-+M with O(g, m) = gin = 0s(m) _ m(9) 99
NO = 990 right translation of g with go 174

Ago = 909 left translation of g with go 174


r.9.) = 9o990 1 conjugation with go 57
4, moment map 100

Representations
Lie group 173
associated Lie algebra 174
continuous representation of a Lie group G 187
associated contragredient representation 186
Symbols 195

d7r associated infinitesimal representation


(of g = Lie G) 120, 190
fr representation of a Lie algebra g 190
Ad adjoint representation Ad (g) = (K9).
(thus =: Ad9) 57, 191
Ad' coadjoint representation Ad* (9) = (Ad(g-i))
(thus = Adg_1) 58, 191
7rS Schrodinger representation of the Heisenberg
group H(R) 126
7rw Weil representation (projective representation)
of SL2(R) 130
irsw Schrodinger-Weil representation (projective
representation) of the Jacobi group GJ(R) 130
v g part of the quantization mapping, A, assigning
to each f E .7° a self-adjoint operator f 120

You might also like