You are on page 1of 692

Birkhäuser Advanced Texts

Basler Lehrbücher

Edited by
Herbert Amann, U niversity of Zürich
Jose M. Gracia-Bondia
Joseph C. Varilly
Hector Figueroa

Elements of
Noncommutative Geometry

Springer Science+Business Media, LLC


Jose M. Gracia-Bondia Joseph C. Värilly
Departamento de Fisica Departamento de Matemäticas
Universidad de Costa Rica Universidad de Costa Rica
2060 San Jose 2060 San Jose
Costa Rica Costa Rica

Hector Figueroa
Departamento de Matemäticas
Universidad de Costa Rica
2060 San Jose
Costa Rica

Library of Congress Cataloging-in-Publiestion Data

Gracia Bondfa, Jose.


Elements of noncommutative geometry I Jose M. Gracia-Bondfa, Joseph C. Varilly
Hector Figueroa.
p. cm.- (Birkhäuser advanced texts)
lncludes bibliographical references and index.
ISBN 978-1-4612-6569-6
1. Geometry, Algebraic. 2. Noncommutative rings. I. Varilly, Joseph C., 1952- II.
Figueroa, Hector, 1957- 111. Title. IV. Series.

QA564.G625 2000
516.3'5-dc21 00-057955
CIP

AMS Subject Classifications: 46L87, 58B34, 8IT10

Printed on acid-free paper. üS)®


© 2001 Springer Science+Business Media New York Birkhäuser aw>
Originally published by Birkhäuser Boston in 2001
Softcoverreprint of the hardcover 1st edition 2001

All rights reserved. This work may not be translated or copied in whole or in part without the written per-
mission of the publisher, Springer Science+Business Media, LLC, except for brief excerpts in connec-
tion with reviews or scholarly analysis. Use in connection with any form of information storage andre-
trieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or
hereafter developed is forbidden.
The use of general descriptive names, trade names, trademarks, etc., in this publication, even if the
former are not especially identified, is not to be Iaken as a sign that such names, as understood by the
Trade Marks and Merchandise Marks Act, may accordingly be used freely by anyone.

SPIN 10716483

ISBN 978-1-4612-6569-6 ISBN 978-1-4612-0005-5 (eBook)


DOI 10.1007/978-1-4612-0005-5

Typeset by the authors in 11\TEX.

9 8 76 5 4 3 2 1
To Hellen Maria, ]esusita and Orietta
Contents

Preface xi

I TOPOLOGY 1

1 Noncommutative Topology: Spaces 3


1.1 Continuous functions on a locally compact space 4
1.2 Characters and the Gelfand transformation 5
1.3 Trading spaces for algebras 9
1.4 Homotopy in noncommutative language 16
1.5 Exponentials and cohomology 0 17
1.6 ldentifications and attachments 21
l.A C* -algebra basics 26
l.B Hopf algebras and Tannaka-Krein duality 34

2 Noncommutative Topology: Vector Bundles 49


201 Vector bundles 49
202 The functor f 56
2.3 The Serre-Swan theorem 0 59
2.4 Trading bundles for modules 60
205 C*-modules 64
206 Line bundles and the Bott projector 74
20A Projective modules over unital rings 79
viii Contents

3 Some Aspects of K-theory 83


3.1 Endomorphisms of C*-modules 83
3.2 The Ko group 92
3.3 The importance of being halfexact 103
3.4 Asymptotic morphisms llO
3.5 The Moyal asymptotic morphism 113
3.6 Bott periodicity and the hexagon 120
3.7 The K1 functor 127
3.8 K -theory of pre-C* -algebras 133

4 Fredholm Operators on C* -modules 141


4.1 Fredholm operators and the Atiyah-Jänich theorem 141
4.2 Fredholm operators on C*-modules . 145
4.3 The generalized Fredholm index 148
4.4 The noncommutative Atiyah-Jänich theorem 156
4.5 Morita equivalence of C* -algebras 159

II CALCULUS AND UNEAR ALGEBRA 169

5 Finite-dimensional Clifford Algebras and Spinors 171


5.1 The eightfold way 171
5.2 Spin groups 180
5.3 Fock-space representations . 184
5.4 The exterior algebra viewpoint 195
5.5 Pfaffians and Gaussians 204
5.A Superalgebras 210

6 The Spin Representation 213


6.1 Infinite-dimensional Clifford algebras 213
6.2 The infinitesimal spin representation revisited 218
6.3 The Shale-Stinespring theorem 224
6.4 Charged fields 238

7 The Noncommutative Integral 251


7.1 A rapid course in Riemannian geometry 251
7.2 Laplacians . 258
7.3 The Wodzicki residue 264
7.4 Spectral functions 272
7.5 The Dixmier trace 284
7.6 Connes' trace theorem 293
7.A Pseudodifferential operators 298
7.B Homogeneaus distributions 306
7.C Ideals of compact operators 310
Contents ix

8 Noncommutative Differential Calculi 319


8.1 Universal forms . . . . . . . . . . 320
8.2 Cycles and Fredholm modules . . . . . 326
8.3 Connections and the Chern homomorphism 335
8.4 Hochschild homology and cohomology 343
8.5 The Hochschild-Kostant-Rosenberg-Connes theorem 355

111 GEOMETRY 367


9 Commutative Geometries 369
9.1 Clifford modules 370
9.2 Spine structures: the algebraic way 372
9.3 Spin connections and Dirac operators 382
9.4 Analytical aspects of Dirac operators 391
9.5 KR-cycles and the eightfold way . . 399
9.A Spin geometry of the Riemann sphere 407
9.B The Hodge-Dirac operator . . . . 423

10 Spectral Tripies 429


10.1 Cyclic cohomology . . . . . . . . . 429
10.2 Chern characters and entire cyclic cocycles 446
10.3 Tameness and regularity of spectral triples 460
10.4 Connes' character formula . . . . . . 470
10.5 Termsand conditions for spin geometries 481

11 Connes' Spin Manifold Theorem 487


11.1 Commutative spin geometries revisited 487
11.2 The construction of the volume form 491
11.3 The spin structure and the metric 500
11.4 The Dirac operator and the action functional 506
11.A The Riemann sphere as a spectral manifold 513

IV TRENDS 517
12 Tori 519
12.1 Crossed products . . . . . . . . . 522
12.2 Structure of NC tori and the Moyal approach 526
12.3 Spin geometries on noncommutative tori 539
12.4 Morita equivalence and crossed products 549

13 Quantum Theory 557


13.1 The Dirac equation and the neutrino paradigm 558
13.2 Propagatars . . . . . . . . . 564
13.3 The classical Dyson expansion in QED . . . 570
x Contents

13.4 The Rules 575


13.5 The quantum Dyson expansion . . . . . . . . . 584
13.A On quantum field theory on noncommutative manifolds 592

14 Kreimer-Connes-Moscovici Algebras 597


14.1 The Connes-Kreimer algebra of rooted trees 598
14.2 The Grossman-Larson algebra of rooted trees 613
14.3 The Milnor-Moore theorem . 618
14.4 Duality in Hopf algebras . . . . . . . 622
14.5 Hopf algebras of Feynman diagrams 627
14.6 Hopf algebras and diffeomorphism groups 630
14.7 Cyclic cohomology of Hopf algebras 635

References 641

Symbol Index 667

Subject Index 673


Preface

Our purpose and main concern in writing this book is to illuminate classical
concepts from the noncommutative viewpoint, to make the language and
techniques of noncommutative geometry accessible and familiar to practi-
tioners of classical mathematics, and to benefit physicists interested in the
uses of noncommutative spaces. Same may say that ours is a very "com-
mutative" way to deal with noncommutative matters; this charge we readily
admit.
Noncommutative geometry amounts to a program of unification of math-
ematics under the aegis of the quantum apparatus, i.e., the theory of ope-
rators and of C*-algebras. Largely the creation of a single person, Alain
Connes, noncommutative geometry is just coming of age as the new century
opens. The bible of the subject is, and will remain, Connes' Noncommuta-
tive Geometry (1994), itself the "3.8-fold expansion" of the French Geome-
trie non commutative (1990). Theseare extraordinary books, a "tapestry" of
physics and mathematics, in the words of Vaughan jones, and the work of
a "poet of modern science," according to Daniel Kastler, replete with subtle
knowledge and insights apt to inspire several generations.
Despite an explosion of research by some of the world's leading math-
ematicians, and a bouquet of applications-to the reinterpretation of the
phenomenological Standard Model of particle physics as a new spacetime
geometry, the quantum Hall effect, strings, renormalization and more in
quantum field theory-the six years that have elapsed since the publica-
tion of Noncommutative Geometry have seen no sizeable book returning
to the subject. This volume aspires to fit snugly in that gap, but does not
xü Preface

pretend to fill it. It is meant rather to be an introduction to some of the


core topics of Noncornrnutative Geornetry.
Our guiding principle has been to show how the tree of noncommutative
geometry sinks its roots in the soil of classical analysis, and we devised
the term "sources of noncommutative geometry" for great theorems that
effect the correspondence between classical constructs and noncommu-
tative ones (Gelfand-Naimark, Serre-Swan, Tannaka-Krein, Connes' trace
theorem, ... ), as well as theorems that are better proved by noncommu-
tative methods, or better understood from the vantage point of the new
paradigm.
This tropical tree unexpectedly sprouted roots from the branches and
many sprawling vines clinging to neighbouring mathematical plants. Sev-
eral strands or themes run through the book. It is ushered in by non-
commutative topology: the lore of algebraic counterparts for spaces, vec-
tor bundles and other topological constructs. Central among those are
C* -modules, and a second recognizable strand is the Morita equivalence
of C*-algebras, afforded by these modules. No work on noncommutative
matters can exclude K -theory, but since many excellent books are now avail-
able, our treatment is relatively brief. We take the leisure to introduce there
the Moyal product.
Noncommutative topology predates Connes' discoveries. Noncommuta-
tive geometry proper was born when he taught us how to do integrodiffer-
ential calculus by operatorial methods-and thus to extend it to noncom-
mutative spaces. How this comes about constitutes a fourth strand. The
heart of the book is the identification of a template for "noncommutative
geometries", necessarily preceded by cyclic cohomology and Connes' char-
acter theorem. Here again we are anchored in the classical realm, by proving
that this template, when applied to the algebra of smooth functions on a
spin manifold, reproduces all of (differential, Riemannian, spin) geometry
ofthat manifold. Last, but not least, a quantum field theory strand appears,
first as a "source", then as a playground and finally as a full-blooded scien-
tific partner.
The book is divided into four parts, comprising 14 chapters. The final
sections of some chapters are labelled by letters rather than numbers; they
contain prerequisite or, less often, supplementary matter.
Part I deals mainly with the task of building a corridor between ordinary
topology and algebra.
Chapter 1 develops the basic translation lexicon for the commutative
and noncommutative languages; we draw from the standard constructs of
algebraic topology. For the reader's convenience, we supply a summary of
the basics of C*-algebras, with a proof of the Gelfand-Naimark theorem.
Then we prove the Tannaka-Krein theorem and argue why Hopf algebras
are expected to play an important role in noncommutative geometry.
Preface xiii

Chapter 2 introduces vector bundles from scratch. A proof is given of the


Serre-Swan categorical equivalence between vector bundles and projective
modules over unital commutative algebras. Then we consider C* -modules
and the Bott projector, an important player throughout.
Chapter 3 starts with the C*-algebra of endomorphisms of a C*-module.
The K0 -group of a C* -algebra is defined in this framework, and its equiva-
lence with the purely algebraic K0 -group is established. We define the K1 -
group in the context of the long exact sequence. Bott periodicity is proved
by means of Moyal quantum mechanics. We dealalso with the K-theory of
pre-C* -algebras.
In Chapter 4, we mop up several remaining foundational topics in non-
commutative topology: Fredholm operators on C*-modules, K-theory as a
generalized Fredholm index, and Morita equivalence of C*-algebras. The
theorem of Atiyah and Jänich portraying the set of Fredholm operators as
a classifying space is recast in a noncommutative mould.
Some will find those first chapters, with their relentless pursuit of ab-
straction, less than easy going. Part II may provide some relief, by focusing
first on the linear algebra underpinnings and then on the real analysis foun-
dations of noncommutative geometry.
Chapter 5 contains the linear algebra needed for the theory of Dirac ope-
rators, i.e., Clifford algebra and spin representations.
Chapter 6 extends Clifford algebra to the infinite dimensional context,
with an eye towards quantum field theory. Charged fields, a source of
Connes' Fredholm module theory, are formulated in that language.
Chapter 7 shows how noncommutative geometry is grounded in spectral
analysis. After a quick review of Riemannian geometry, we examine the
Wodzicki residue and explain its appearance as a logarithmic divergence of
a kernel. The Kastler-Kalau-Walze theorem, relating Riemannian curvature
with the residue of apower of the Laplacian, is given. We then introduce
the noncommutative integral, showing that in the commutative context it
reduces to the ordinary integral by means of Connes' trace theorem.
Chapter 8 features several fundamental constructions in noncommuta-
tive geometry: universal forms and connections, cycles and their Chern
characters, Hochschild cohomology. The Chern isomorphism theorem is
proved. The chapter is crowned by a theorem of Connes, whose source is
the Hochschild-Kostant-Rosenberg theorem of algebraic geometry, estab-
lishing an isomorphism between the Hochschild cohomology of the algebra
of smooth functions on a manifold and the space of its de Rham currents.
Part III deals with the construction of noncommutative geometries.
Chapter 9 describes spin geometry from the noncommutative point of
view. Spine structures are defined as Morita equivalences between the Gel-
fand and Clifford algebras associated to a manifold. Dirac operators are
then introduced and exemplified, and their relation to charge conjugation
is analyzed in the context of Kasparov's K-homology of Banach algebras.
xiv Preface

In Chapter 10 we assemble the ingredients of noncommutative spin geo-


metries. First there is an account of cydic cohomology, which is the re-
ceptade of the noncommutative Chern character. The entire cydic coho-
mology formulation of the Chern character is also developed. We indude
here Connes' proof, heretofore unpublished, of the formula which gives the
Hochschild dass of the character as a "local" expression involving the non-
commutative integral. The promised template for spin geometries, namely,
spectral triples satisfying several natural conditions, is then laid out.
A fundamental theorem of Connes holds that this set of conditions, ap-
plied to spectral triples based on commutative algebras, yields the dassical
geometry of spin manifolds: Chapter 11 unfolds, for the first time in book
form, Connes' spin theorem.
Part IV scans some directions for further development. We hope that this
partwill appeal to people interested in the applications of noncommutative
geometry.
Chapter 12 begins with abrief overview of some of these applications,
induding its treatment of the Standard Model of partide physics. Then it
encounters the Moyal product again, and deals at length with the differen-
tial and geometrical structures ofnoncommutative tori, which constitute an
important dass of noncommutative manifolds. This is useful in string the-
ory with a background magnetic field, as there is a limit in which the string
dynamics is described by a gauge theory on a noncommutative torus.
In Chapter 13 we recover another physics strand. Connes' spin theorem
makes it possible, indeed imperative, to develop quantum field theory on
noncommutative spaces. In preparation, and based on the spin represen-
tation of Chapter 6, some aspects of perturbative quantum field theory are
translated into algebraic language. Linear quantized fields on noncommu-
tative tori are considered.
In the final Chapter 14, Hopf algebras resurface, in the context both of
renormalization theory and the pursuit of a local index formula in noncom-
mutative geometry.
Same topics arenot treated fully. For more detail on the reconstruction of
the Standard Modelas a noncommutative geometry, we refer to our [329).
Excellent versions of the noncommutative analysis of the quantum Hall
effect are available in [29,334,497) andin Connes' book itself. Also, index
theory is sidestepped: although it pervades our whole endeavour, it is never
squarely taken up-that requires another book.
All books are essays, and mathematics books are no exception. How-
ever, we have tried hard to reengineer this essay as a textbook, and we
believe it to be suitable for a one-year graduate course, after, say, a first
coursein functional analysis. We would also recommend our book to re-
searchers trying to enter the field. The text is leavened with discussions
and examples-although we rarely allow them to get in the way of the for-
mal development. The exercises are peppered throughout the text as they
Preface xv

crop up, rather than being tucked away at the ends of chapters; most of
them are easy, their main purpose being to allow the reader to keep check-
ing his understanding of the topics. The reader who goes "cover to cover"
will find many rewards, not least the pleasure of key concepts reappearing,
with enhanced pertinence and understanding.
We refer to a great many texts, original articles and survey papers. Some
of them simply sit in the junctions we perceive between commutative and
noncommutative mathematics; others deal with noncommutative geometry
proper. The latter Iiterature is already enormous, and wehavenot tried to
do it justice; we cite what we have been able to use, neither more nor less.
The most pleasant aspect of writing a preface is having the opportunity
to thank the many people who helped to improve this work from very rough
beginnings. Foremost, our thanks go to Mare Rieffel, who read through sev-
eral early drafts, and provided a lot of constructive criticism; and to Alain
Connes, who suggested several improvements, and liberally communicated
to us the proof of his character theorem given in Chapter 10. Jose Carifiena
and Alberto Elduque reminded us in time, respectively, of sections along
a map and of extensions of scalars, combining in the "noncommutative
pullback bundle" construction of Chapter 2. Mario Paschke explained to
us his ideas about the reconstruction of spheres and other homogeneaus
spaces in noncommutative terms. Ruy Exel supplied advice that simplified
our treatment of Fredholm operators on C*-modules. Luis Joaquin Boya
called our attention to his work on Bott periodicity, furnishing a link in the
theory of the charge conjugation operator.
Daniel Kastler helped with his own account of Clifford geometry, which
proved extremely valuable, and with discussions about the Milnor-Moore
theorem. The influence of the work by Ricardo Estrada is patent in all
of Chapter 7. Also, Stephen Fulling read through Chapter 7. Tomäs Kopf
helped to improve the formulation of the spin connection in Chapter 9.
Adam Rennie made available to us, prior to publication, his paper on the
proof of Connes' spin theorem. Stephan de Bievre and Askold Perelomov
taught us to recognize the advantages of the Moyal product approach to
noncommutative tori. Earl Taft made a crucial observation concerning the
antipode of the Kreimer Hopf algebra and pointed us to the work by Grass-
man and Larson. We thank Dirk Kreimer for sunny conversations over
branch and twig. Piotr Hajac, Gianni Landi, Fedele Lizzi and Franciscus
Vanhecke prodded us with many thought-provoking questions. We wish
to thank our reviewers, who suggested many useful amendments. c;a va
sans dire, we lay claim to such errors as remain.
Philippe Blanchard opened to us the vast resources and quiet atmosphere
of the Universität Bielefeld library-a great assistance for third-world re-
searchers. Florian Scheck enabled a first field-testing of the book. Peter
Bongaarts, Thierry Fack, Rainer Häußling, Antonio da Silva, Nicolae Tele-
man, Daniel Testard and Mark Villarino supplied important bibliographical
material.
:xvi Preface

This book was started in September 1997, thanks to an unexpected vaca-


tion of the first-named author, who thanks CONICIT and MICYT for financial
support.
Our former students William Ugalde and jose Rasales sat enthusiastically
through the lectures on which this book came to be based.
Ross Moore and Louis Vosloo provided timely TfXnical help.
Ann Kostant has been a most understanding and helpful editor.
It would be very difficult to properly mention all the "noncommutative
brethren" from whom we have been slowly learning in the course of the
years. We shall single out Jacek Brodzki, whose unforgettable course at the
University of Texas at Austin in the fall of 1991 started our affair of the
heart with noncommutative geometry.
The ultimate purpose of this book is to share the soul's joy derived
from this new approach to physics and mathematics; if it helps some in
the coming generation to understand why already mature scientists have
been fascinated by the noncommutative enterprise, our labours will prove
worthwhile.

San Pedro de Montes de Oca and Mainz


]une 2000
Elements of
Noncommutative Geometry
Part I

TOPOLOGY

The ordinary man wonders


at marvellous things;
the wise man wonders
at the commonplace

- Confucius
1
Noncommutative Topology: Spaces

The geometrical study of quadratic curves or surfaces, i.e., zero sets of


second-degree polynomials, proceeds by examining points of intersection
or tangent lines directly; but already for cubic curves it pays to examine first
the ideal of all polynomials that vanish on the curve: in this way the study
of an algebraic variety (the zero set of a given finite collection of polynomi-
als) is replaced by the study of the corresponding polynomial ideal. Such a
fundamental geometrical object as an elliptic curve is best studied not as a
set of points (a torus) but rather by examining functions on this set, specif-
ically the doubly periodic meromorphic functions: Weierstrass opened up
a new approach to geometry by studying directly the collection of complex
functions that satisfy an algebraic addition theorem, and derived the point
set as a consequence [51].
In noncommutative geometry, under the influence of quantum physics,
that idea of replacing sets of points by classes of functions is taken fur-
ther. In regular cases the set is completely determined by an algebra of
functions, so one can choose to forget about the set and obtain all infor-
mation from the algebra alone. When the associated set is too singular or
pathological, a direct examination frequently yields no useful information:
the set of orbits of group actions is generally of this type. In such Situa-
tions, when the matter is examined from the algebraic point of view -as is
donein Chapter 12 for the rotation of a circle by multiples of an irrational
angle- we often find an operator algebra holding the information we seek;
however, this algebra is not commutative.
The 1943 paper [194] containing the results nowadays known as Gel-
fand-Naimark theorems has become a cornerstone ofnoncommutative geo-
4 1. Noncommutative Topology: Spaces

metry. Gelfand and Naimark characterized the involutive algebras of ope-


rators, nowadays called C*-algebras, from the natural axiomatization for
the algebra of continuous functions, by just dropping commutativity. To
go from there to presuming that C*-algebras held the right generalization
of classical concepts of space still required a leap of faith (in the words of
Effros [152]), but one that has over time paid handsome dividends.
Thus, we proceed by first discovering how function algebras determine
the structure of point sets, and then learning which relevant properties of
those algebras do not depend on commutativity.

1.1 Continuous functions on a locally compact


space
Definition 1.1. A compact Hausdorff topological space X gives rise to a
natural commutative algebra C(X), consisting of all continuous functions
f: X - C This is a Banach algebra und er the sup norm

11!11 := sup lf(x)l. (1.1)


xEX

Moreover, C(X) has an isometric involution f - f* by defining f* (x) :=


j(x); and the norm satisfies the C*-property

11!11 2 = II!* !II. (1.2)

In other words, C(X) is a unital C*-algebra -the unit being the constant
function 1.
We shall have to deal with nonunital algebras as weil. To any Banach
algebra A we can adjoin a unit, denoted 1A (or simply 1 when no confusion
is feared), by taking A + := A x ([, with the obvious sum and adjoint and the
multiplication rule (a, ,\) (b, f.l) := (ab+ ,\b + f.Ja, ,\f.J); then 1A+ = (0, 1). For
the proofthat A + is indeed a C*-algebra when Ais a nonunital C* -algebra
to start with, and other necessary background on Banach and C* -algebras,
we refer to Section LA.
Every topological space considered in this book will be Hausdorff, unless
explicitly indicated otherwise. If Y is locally compact but not compact, then
obviously the algebra C(Y) is too big tobe of anyuse. A strategy to deal with
this case is to add a "point at infinity" to get a compact space y+ := Y~t~ {oo }.
Then the subalgebra of C(Y+) whose members satisfy j(oo) = 0 may be
identified, by restriction to Y, with the algebra C0 (Y} of continuous func-
tions "vanishing at infinity". It is clear that Co(Y) is a C*-algebra without
a unit, but then Co(Y)+ ""C(Y+). Conversely, if one deletes a non-isolated
point xo from a compact space X, the space Y = X\ {xo} is locally compact
but not compact, y+ :::::X, and Co(Y) "" { h E C(X): h(x0 ) = 0 }.
1.2 Characters and the Gelfand transformation 5

1.2 Characters and the Gelfandtransformation


Definition 1.2. A character of a Banach algebra A is a nonzero homomor-
phism J1 : A ~ (, which is necessarily surjective. The set of all characters
(that may well be empty) will be denoted M(A).
For example, if A = C0 (Y), the evaluation map Ey: f - j(y) at y E Y
clearly defines a character.
Any character J1 E M(A) extends to A+ by setting J1((0, 1)) := 1 (neces-
sarily). The zero functional on A also extends to the character (a, .\) - .\
on A+. Thus we identify M(A) u {0} with M(A+). We recall -see Sec-
tion LA- that the spectrum sp(a) of an element a in Ais the set of com-
plex numbers .\ such that a - .\1 is not invertible in A, or in A + if A is
not unital -in the nonunital case 0 always belongs to the spectrum. Then
Jl(a) E sp(a) for J1 E M(A) and a E A: otherwise 0 = Jl(a- Jl(a)l) would
beinvertiblein(!Therefore IJl(a)l 5 llall. so IIJ.lll 5 l.Infact, IIJlll = 1: since
Jl(a) = J1(1 · a) = J1(1)J1(a) for all a E A+, it follows that Jl(l) = J1(1) 2 and
J1(1) * 0.
Let A * be the Banach space of continuous linear functionals cf>: A ~ <C.
On A * we can consider the so-called weak* topology, that of pointwise
convergence on elements of A. For instance, if A = C(X), with X compact,
then A * is the space of complex measures on X with its standard topology.
The Banach-Alaoglu theorem -a corollary of Tikhonov's theorem [383]-
says that the unit ball Ai of A * is compact in the weak* topology.
Definition 1.3. If A is commutative, call M (A) the Gelfand spectrum of A.
Then M (A) .... Ai: the Gelfand topology is the relative topology determined
on M(A) by this inclusion.
Lemma 1.1. The Gelfand spectrum of a commutative Banach algebra, en-
dowed with the Gelfand topology, is a locally compact space.

Proof. We show that M(A) u {0} is weak*-closed in A* and hence is com-


pact. For fixed a, b E A, the map J1 - Jl(a)Jl(b)- Jl(ab) is weak*-continu-
ous; since it vanishes on M(A) u {0}, it vanishes on its closure.
The extra point {0}, corresponding to the "evaluation at the point at in-
finity", is isolated if Ais alreadyunital, since the weak'-continuous function
J1- J1(1) separates it from M(A). D

Definition 1.4. Let A be a commutative Banach algebra. The Gelfand trans-


form of a E Ais the function a: M(A) ~ ( given by

a(Jl) := Jl(a). (1.3)

In other words, ais the evaluation at a E A; this is continuous, by definition


of the weak* topology. The Gelfand transformation is the map (j: a - a
from A into Co (M (A)).
6 1. Noncommutative Topology: Spaces

For general cornmutative Banach algebras, the Gelfand transformation


is an extremely useful instrument, in spite of being in general neither in-
jective nor surjective. For instance, the convolution algebra A = L 1(IR{) of
Lebesgue-integrable functions on IR{ is a nonunital algebra whose charac-
ters are the integrals f - JIR e-itx j(x) dx for any t E IR{ [266, Thm. 3.2.26);
then j is the Fourier transform of j, and the Gelfandtransformation is the
Fouriertransformation that takes L1(1R{) into C0 (1R{): this is the Riemann-
Lebesgue lernma [316, §5.1). This map is injective, but not isometric nor
surjective (the Fourier transform of an integrable function has the extra
property of being absolutely continuous). There are also cornmutative Ba-
nach algebras that contain nonzero quasinilpotent elements, whose spec-
tral radius is zero (an example is the convolution algebra of continuous
functions on [0, 1), whose elements are all quasinilpotent [499, §11.2)); in
such cases the Gelfandtransformation §: A - C(M(A)) is certainly not
injective .
.,. The situation improves greatly when we Iimit ourselves to C* -algebras,
as we now hasten to do. The C*-algebras are distinguished among the ge-
neral clutter of Banach algebras with involution by the notable properties
of their selfadjoint elements.
Lemma 1.2. Leta beaselfadjointelementofaC*-algebraA. ThenJl(a) E IR{
for all J1 E M(A).

Proof. Suppose that a = a* and consider the exponential series u :=


exp(ia) := Lk=o(ia)k jk! that converges in A (or in A+ if A is nonuni-
tal) with the a prioribound llull ::::; ellall. Then u* = exp( -ia) and so
uu* = 1 = u*u; in particular, u is invertible with u- 1 = u*. The C*-
norm condition implies that llull 2 = llu*ull = 11111 = 1, and likewise
llu- 111 = 1. Since 111111::::; 1, we get IJ1(u)l::::; 1 and IJl(u)l- 1 = IJl(u- 1)1::::; 1,
so that IJl(U) I = 1. But by continuity and the homomorphism property we
obtain Jl(U) = Lk=O Jl( ia)k /k! = eij.l(a), and thus Jl(a) E IR{. D
Exercise 1.1. Prove, along the lines of the previous proof, that in a unital
C*-algebra the spectrum of a unitary element u (satisfying uu * = u *u = 1)
is part of the unit circle, and that the spectrum of a selfadjoint element
a = a * is real. <>

When a E Ais not selfadjoint, we write a = a1 + iaz with a1 := ~(a* +a),


az := 4(a*- a) selfadjoint. If Ais a C*-algebra and J1 E M(A), we then get

Jl(a*) = J1(a1- iaz) = Jl(ad- iJl(az) = Jl(a), (1.4)

or equivalently, Ci*(Jl) = a(Jl) for J1 E M(A); or better yet, Ci* = (a)*. In


other words, the Gelfand transformation a - a intertwines the involutions
of A and of C(M(A)).
We need one more lernma.
1.2 Characters and the Gelfand transformation 7

Lemma 1.3. Let A be a commutative C* -algebra, and Iet t\ belong to the


spectrum of a E A. If A is unital, or if A is nonunital and t\ * 0, there is a
character J1 suchthat Jl(a) = t\.

Proof. The kernel of any character J1 of a unital commutative C* -algebra


is a (proper) maximal ideal. Indeed, Al ker J1 = <C, a field; if] is an ideal
properly containing ker Jl, and if a E J \ ker Jl, then Jl(a) has an inverse;
consider then an element b E A such that J1(b)J1(a) = 1: both ba and
ba- lAbelang to ], so lA E ], and thus ker J1 is maximal.
Reciprocally, if I is a maximal ideal, it must be closed. Indeed, its closure
is also an ideal and, as argued in Section l.A, a proper ideal in a unital
Banach algebra cannot be dense. Therefore Al I is a commutative Banach
algebra without proper ideals, i.e., a field. Moreover, as the spectrum of any
element in this unital Banach algebra is nonempty, this field is Al I "" <C.
Thus there is a unique character Jl, namely the quotient map, suchthat
W 1 (0) =I.
The unital case follows, because the nontrivial ideal A(a- t\1) is con-
tained in a maximal ideal, which is an immediate consequence of Zorn's
Iemma [383]; for the nonunital case, use the same argument in A +. 0

Denote by r(a) the spectral radius of an element a of a Banach algebra.


In Section LA we remark that r(a) = !lall. when a is a selfadjoint element
of a C*-algebra.
The last ingredient we need is the Stone-Weierstrass theorem [131, 383].
This states that, if X is a locally compact space, and if B is a closed sub-
algebra of Co(X) that (a) separates points, i.e., whenever p * q in X, there
is some y E B with y(p) * y(q), (b) vanishes identically at no point of X,
and (c) is closed under complex conjugation; then Bis the whole C0 (X).
Theorem 1.4 (Gelfand-Naimark). If A is a commutative C* -algebra, the
*
Gelfand transformation is an isometric -isomorphism of A onto Co (M (A)).

Proof. The relation (1.4) shows that §: A - Co(M(A)) is a *-homomor-


phism. The isometric property follows from

lliill 2 = llii*iill = 11Ci*"a11 = r(a*a) = lla*all = llall 2 , (1.5)

for a E A, where we use Lemma 1.3 to justify the third equality. In partic-
ular, § is injective. Now §(A) is a subalgebra of Co(M(A)) that is complete
since A is complete and § is isometric, and therefore is closed. Clearly the
evaluation maps in §(A) separate the characters, do not all vanish at any
point, and (1.4) again shows that §(A) is closed under complex conjuga-
tion; the Stone-Weierstrass theorem then teils us that § is surjective. 0
A comment on the proof is in order. Wehave chosen to underline the
description of the Gelfand spectrum of a commutative C*-algebra in terms
of characters, rather than of maximal ideals, which is perhaps the more
8 1. Noncommutative Topology: Spaces

standard procedure. However, we did use the latter to establish the ex-
istence of a character producing any number in the spectrum of a given
element of the algebra. We could have appealed to the Hahn-Banach theo-
rem instead, but, leaving aside that this does not absolve us of Zornication,
there would remain the work of proving that the linear functional which
it yields is multiplicative. Such an approach is consistently taken in [175)
and [266), where it is shown that characters are just pure states (consult
Section l.A) for commutative C* -algebras; the subject is then inextricably
tied to representation theory.
To summarize the consequences of the theorem: being a commutative
C*·algebra is at least a necessary condition for an algebra tobe isomet-
rically isomorphic to some C(X) or C0 (Y). The beauty of the Gelfand-
Naimark theoremisthat this apparently entirely algebraic property is also a
sufficient condition for isomorphism with a space of continuous functions.
Actually, to be a C* -algebra is not wholly a matter of algebraic relations,
since completeness (in the metric determined by the C*-norm) involves lim-
its. At any rate -as we shall exemplify very soon- all topological informa-
tion about X is algebraically stored within C(X). We do gain the insight
to go beyond conventional topology by regarding a noncommutative C*-
algebra as a kind of "function algebra" for a virtual or "noncommutative
topological space". This viewpoint also allows one to study the topology of
non-Hausdorff spaces, such as arise in prohing a continuum where points
are unresolved [20, 304).
The second theorem of Gelfand and Naimark [194] says that any C*-
algebra can be embedded as a norm-closed subalgebra of a full algebra of
operators L(J-f) for a large enough Hilbert space Jf; thus all C*-algebras
are fairly "concrete". Fora discussion of this, see Section l.A; a full discus-
sion of both theorems and their offshoots is contained in the book [144).
lndeed, C(X) may be embedded in L(J-f), the algebra of bounded opera-
tors on the Hilbert space J{ with a countable orthonormal basis, in many
ways: if X is infinite but separable, take v to be any finite regular Borel
measure on X, and identify J{ with L2 (X, dv); then f E C(X) c L"" (X, dv)
can be identified with the multiplication operator h - fh on L2 (X,dv).
Developing the last remarks, one sees that the first Gelfand-Naimark theo-
rem is the basis of the spectral theorem for normal operators on a Hilbert
space [132, 383).
lt is in the spirit of noncommutative geometry to relax closure condi-
tions as much as possible when defining our algebras; for instance, if M is
a compact differential manifold, we would like to use the algebra of smooth
functions A = C"" (M), which is a dense subalgebra of C(M). lt is complete
in its natural topology, that of uniform convergence of functions, tagether
with their derivatives of all orders; but this yields only a Frechet algebra
(provided M is u-compact). Even so, it turnsout that the characters of this
algebra are simply the evaluations at points of M; that is, every character on
C""(M) extends to a character on C(M). Indeed, characters on C 00 (M) are
1.3 Trading spaces for algebras 9

automatically continuous, and therefore are distributions on M; moreover,


a positive distribution, as such a character is, is actually a measure [316,
§6.22]. Therefore, the involutive Frechet algebras of smooth functions on
0'-compact differentiable manifolds do characterize such manifolds. Unfor-
tunately, no one seems to know how, conversely, to characterize algebras
isomorphic to C"" (M) among involutive commutative Frechet algebras. We
shall return to consideration of C"" (M) shortly.
What happens if X is a (topological) group G? Then, at least in the com-
pact case, a dense subalgebra of C(G) is endowed with a much richer alge-
braic structure, allowing to recapture G as a group. This is the subject of
Section l.B.

1.3 Trading spaces for algebras


Definition 1.5. If f: X - Y is a continuous mapping between two compact
spaces, denote by Cf the mapping h ...... h o f from C(Y) to C(X). Then
Cf is a unital *-homomorphism, since (h + k) o f = (h o j) + (k o j),
(hk) o f = (h o f) (k o f), and h* o f = (h o j)* for h, k E C(Y), and clearly
lof=l.
If f: X - Y, g: Y - Z are continuous mappings between compact
spaces, then C(g o f) =Cf o Cg as *-homomorphisms from C(Z) to C(X).
Also, Cidx is the identity map idc(X) on C(X). We summarize by saying
that X ...... C(X), f ...... Cf is a contravariant functor from the category of
compact spaces and continuous maps to the category of commutative uni-
tal C*-algebras and unital *-homomorphisms. This functor is called con-
travariant since it "reverses the arrows"; we shall use the term cofunctor,
for short.
There is also a cofunctor going the other way. Recall that the Gelfand
topology of M(A) is, by definition, the weakest topology for which all the
functions a: M (A) - ((, for a E A, are continuous. Therefore, it has a
universal property, namely, that any function f: X- M(A) is continuous
if and only if each a o f: X - C is continuous.
Definition 1.6. If cf>: A - Bis a unital *-homomorphism between two com-
mutative unital C* -algebras, denote by M cf> the mapping 11 ...... 11 o cf> from
M(B) to M(A). Then Met> is a continuous map, since a o M(c/>) = ~ is
continuous for each a E A. If !Jl: B - Cis another unital *-homomorphism,
then M (1/1 o cf>) = M cf> o M !Jl.
We can compose these cofunctors in the obvious manner; if 11 E M(C(X))
then ker 11 isamaximal ideal of C(X), so there is at least one point x E X
where all the elements of ker 11 vanish. (Were this not the case, by using
the compactness of X, we could construct an invertible element of ker 11·)
Recall that Ex denotes the evaluation map at the point x; then ker 11 = ker Ex
10 1. Noncommutative Topology: Spaces

since both are maximal ideals. If f E C(X), then f- 11<fH E ker /1. so
0 = Ex<f- 11<f)1) = Ex<f)- /l(j). Thus the evaluations Ex are the only
characters of C(X).
Exercise 1.2. Prove that Ex: x ...... Ex is a homeomorphism between the orig-
inal compact space X and the compact space M(C(X)) with its Gelfand
topology. o
In fine, we have assembled homeomorphic maps Ex: X- M(C(X)) for
each compact space X suchthat Ey o f = MCf o Ex whenever f: X - Y
is continuous. In category-theory jargon, E is a natural transformation be-
tween the identity functor and the functor MC on the category of compact
(Hausdorff) spaces. In particular, all morphisms in the latter come from
*-homomorphisms of the algebras (i.e., the cofunctor M is full).
We take this opportunity toshorten the clumsy term "*-homomorphism"
to just "morphism". From now on, a morphism 1>: A - B will designate an
involutive homomorphism between the C*-algebras A and B. Also, in this
section, all morphisms between Unital algebras are supposed unital.
.,. The Gelfand-Naimark theorem provides another natural transformation
(j between the identity functor and the functor CM on the category of
unital, commutative C* -algebras. Indeed, the theorem states that a ...... a
is an isomorphism of A onto C(M(A)). Moreover, if 4>: A-Bis a unital
morphism, then for a E C(M(A) ), v E M(B), we get

(C(M4>)ii)v = a((M4>)v) = a(v o 4>) = v(4>(a)) = $(a}(v), (1.6)

so (CM4>)ii = GB(4>(a)) for each a, or equivalently, GB o 4> = CM4> o GA· In


particular, all unital morphisms from C(Y) to C(X) come from continuous
maps from X to Y.
Therefore, the categories of compact spaces and unital, commutative C*-
algebras are equivalent -or to be more pedantic, one is equivalent to the
opposite category of the other .
.,. When X and Y are only locally compact spaces, the correspondence h ......
hofwill not always map C0 (Y) into C0 (X).
Exercise 1.3. Show that h ...... h o f takes functions vanishing at infinity on
Y to functions vanishing at infinity on X iff f is a continuous proper map
(i.e., the preimage under f of any compact set in Y is compact in X). 0
It does not follow that there is an equivalence of categories between
locally compact spaces with continuous proper maps and commutative
C* -algebras with morphisms. For instance, the injective morphism embed-
ding C0 ([0, 1)) into C([O,a]), for any a;:: 1, does not come from any map
(proper or otherwise) from [0, a] into [0,1). To obtain such an equivalence,
one restricts to C*-morphisms which are proper, that is, send approximate
units into approximate units.
1.3 Trading spaces for algebras 11

At any rate, instead of the former category, we shall use an equivalent one
whose morphisms are easier to deal with, namely, the category of pointed
topological spaces. For that, we systematize our previous remarks about
compactifying locally compact spaces by adding a point at infinity. This
will be in tune with later homotopy-theoretical considerations.
Definition 1.7. A pointed compact space isapair (X,*), where Xis com-
pact and * E X is a distinguished element, the basepoint. A morphism
from (X,*) to (Y, *) is a continuous map f: X- Y suchthat f(*) = *·
We write f E Map+ (X, Y).
Any locally compact space Y determines a pointed space ( y+, oo); any
continuous proper map f: Y - Z is extended to a morphism f+: y+ - z+
by setting f+ ( oo) := oo. Conversely, if (X,*) is a pointed topological space,
then X\ { *} is locally compact, and the restriction of a morphism to X\ { *}
is proper.
We identify Y to y+ \ { oo} and (X, *) to (X\ {*}) +. This allows us to omit
mentioning the basepoint when it is unambiguous.
Let X, Y be compact spaces with Y ,;;; X; we call (X, Y) a cornpact pair.
From them we can always construct a pointed space XIY :=(X\ Y)+; one
can think that Y has been srnashed to become a base point for the new
space. Let c: X - X I Y be the collapsing map; the restriction of c to X\ Y is
a homeomorphism. Note that X I 0 = x+ (the basepoint of x+ is an isolated
point since X is already compact).
~ We next explore a few consequences of the equivalence of categories.
Proposition 1.5. Two cornrnutative C* -algebras are isornorphic if and only
if their character spaces are horneornorphic.

Proof. Suppose the C*-algebras A and Bare both unital. Then morphisms
cf>: A - B, tjJ: B - A such that tjJ o cf> = idA, cf> o tjJ = idB yield continuous
maps Mcf>: M(B) - M(A) and MtjJ: M(A) - M(B) suchthat MtjJ o Mcf> =
idM(B) and Mcf> o MtjJ = idM(Al; thus Mcf> is a homeomorphism.
Conversely, homeomorphisms of compact spaces f: X- Y, g: Y- X
suchthat g o f = idx and f o g = idy yield unital morphisms Cf: C(Y) -
C(X) and Cg: C(X) - C(Y) suchthat Cg o Cf = idc(Y) and Cf o Cg =
idc(xl; thus Cf is a *-isomorphism.
If A and B are both nonunital, they are isomorphic if and only if A + == ß+
if and only if M (A +) "" M (B+) by a homeomorphism that takes E"" E M (A +)
to Eoo E M(B+). D

Corollary 1.6. The group of autornorphisrns Aut A of a cornrnutative C*-


algebra A is isornorphic to the group of horneornorphisrns of its character
space. B

Note that there are no nontrivial inner automorphisms in Aut A.


12 1. Noncommutative Topology: Spaces

One can make the parallel argument for the Frechet algebra C"" (M). Im-
plicit in the previous proof is the property that morphisms between C*-
algebras are automatically continuous (consult our remarks in Section l.A).
Butthis is also true of C""(M). Indeed, Klee's theorem [412, Thm V.S.S] as-
serts that every positive linear form on an ordered Frechet space F such
that F = p+ - p+ is continuous. This is easily seen to be the case for
C"" (M, ~). and then it follows easily [454] that involutive homomorphisms
C""(M) ~ C""(N), for a-compact manifolds M and N, are continuous.
Corollary 1.7. I(M is a compact manifold, then the group of automorphisms
of the algebra Aut C"" (M) is isomorphic to the group Diff(M) of diffeomor-
phisms of its character space.

Proof. Any real-valued smooth function a E C"" (M; ~) can be written as


c - (c - a) if c is any positive constant such that -c :s; a :s; c; thus
C"" (M; ~) is an ordered Frechet space generated by its positive cone. Thus
each positive linear functional on C"" (M; ~). or indeed on C"" (M), is con-
tinuous. If cp: C"" (M) - C"" (M) is an algebra isomorphism and x E M,
then Ex o cp: a ,_. cp (a) (x) is a character of C"" (M) and its continuity makes
it a positive distribution on M; it therefore extends to a character Ef(x) of
C (M). Now f is a homeomorphism of M onto itself suchthat Cf extends cp,
and so f preserves the smooth structure of M. D

If f: Y - Z is continuous and injective, then two continuous maps


g: X - Y, h: X - Y are equal if and only if f o g = f o h. Thus two
unital morphisms cp: C(Y)- C(X), (/1: C(Y)- C(X) are equal if and only
if cp o Cf= (/1 o Cf, so the range of Cf must be all of C(Y). Conversely, if
Cf is surjective, then cp o Cf = (/1 o Cf implies cp = (/1, so f o g = f o h
implies g = h and thus f is injective. In particular, if Y is a closed (hence
compact) subset of a compact space X, then the inclusion j: Y - Z is injec-
tive, so the restriction morphism Cj: C(Z) ~ C(Y) is surjective. In other
words, any continuous function on a closed subset of a compact space can
be extended to a continuous function on the full space.
Exercise 1.4. Show that f: X- Y is continuous and surjective if and only
if Cf: C(Y) - C(X) is an injective unital morphism. o
To a large extent, this chapter constitutes a kind of training course in
Gelfand gymnastics, i.e., the art of rendering topological properties of spa-
ces in algebraic terms, which is the first step in mastering the language of
noncommutative geometry. As in any language course, we need a dictio-
nary. The succeeding paragraphs make new entries in our dictionary.
If Z is an open subset of a compact space X, then C0 (Z) is an ideal of
C(X). To see that, consider X\Z. There is a surjective morphism rr: C(X) -
C(X \ Z) given by restriction. Then ker rr is an ideal of C(X) that can be
identified, in an obvious manner, with C0 (Z).
1.3 Trading spaces for algebras 13

Exercise 1.5. Conversely, if 1 is an ideal of C(X), then 1 "' Co(Z) for some
opensubsetZ c X. 0
An essential ideal I in a C*-algebra A has, by definition, nontrivial inter-
section with any other nonzero ideal.
Exercise 1.6. If Z c X, prove that Z is open and dense in X if and only if
Co(Z) is an essential ideal of C(X). 0
Proposition 1.8. 1 is an essential ideal of A if and only if a1 * 0 for any
nonzero a E A.
Proof. Let 1 be an essential ideal of A and let ]1- := { a E A : a1 = 0 };
clearly, 1 j_ is an ideal. Now, if a E 1 n 1 j_, then aa * = 0, so 0 = II aa * II =
llall 2 , hence 1 n ]1- = {0}; therefore ]1- = {0} since 1 is essential.
Conversely, assume that 1J. = { 0}, and let I be another ideal such that
In 1 = {0}. If a E I and b E 1. then ab = 0 since ab E In 1. Hence I s;;;
1J. = {0}. Consequently, a nonzeroideal must have nontrivial intersection
with1. D
At this point, it is instructive to have another look at compactifications.
In general, if Y is a locally compact space, a compactification of Y is a pair
(X,j), where Xis a compact space and j: Y ...... Xis a homeomorphism
from Y into a dense, necessarily open, subset of X. By Exercise 1.6, C(X)
contains Co(Y) as an essential ideal.
Now, if Y is locally compact but noncompact, it is at any rate a "com-
pletely regular" space, that is, any closed set Z c Y and point x E Y \ Z
can be separated by a continuous function. An alternative to considering
the algebra of continuous functions vanishing at infinity is to take the alge-
bra Cb(Y) of bounded continuous functions. This is a unital commutative
C*-algebra; let us write ßY := M(Cb(Y)).
Exercise 1.7. Show that the canonical map j: Y ...... ßY is indeed a homeo-
morphism into its image. 0
Moreover, it is clear that if f E C(ßY) and fiJ<Yl = 0, then f= g for
g E Cb(Y) with g(y) = g(Ey) = 0 for each y E Y, and so f = 0; there-
fore Y is densein ßY. Consequently, the space of characters of Cb (Y) is a
compactification of Y: it is the so-called Stone-Cech compactification [30].
The fundamental property of ßY isthat each continuous map from Y to a
compact space X extends uniquely to a continuous map from ßY to X.
Exercise 1.8. Prove this extension property using Cb(Y) "' C(ßY) and an
algebraic argument. 0
The Stone-Cech compactification is maximal, in the sense that all pos-
sible compactifications are in one-to-one correspondence with closed self-
conjugate subalgebras of C(ßY) that separate points of Y. lts construc-
tion, as just presented, is already a fine example of the noncommutative
14 1. Noncommutative Topology: Spaces

viewpoint at work. The noncommutative counterpart of compactification


is unitization of a C*-algebra A, by which we mean any embedding of Aas
an essential ideal in a unital C* -algebra. (A has a nontrivial unitization in
this sense if and only if it is not already unital.) We require that A, as an
ideal in a larger C* -algebra, be essential, since otherwise we could not find
a "largest" unitization. The counterpart of the Stone-Cech compactification
is the concept of a multiplier algebra.
Definition 1.8. The maximal unitization of a nonunital C* -algebra A is
called :M(A), the multiplier algebra of A. In general, if Ais a C*-subalgebra
of another C* -algebra B, a multiplier of A in Bis an element b E B suchthat
bA ~ A and Ab ~ A. The set of all such multipliers is called the idealizer
of A in B; it is evidently a C* -subalgebra of B containing A as an ideal.
Let us leave aside for now the question of existence of :M(A), which is
beautifully solved by means of C*-module theory: see Section 3.1.
Lemma 1.9. Let A be a C* -algebra and Iet j: A - L (J{) be an injective
representation of A suchthat j(A):H is densein the Hilbert space :H. Then
j extends to an isomorphism between :M(A) and the idealizer of j(A) in
L(:H).

Proof. lf TE :M(A), then j(a)(/1 ~ j(Ta)(/1 isalinear operator on j(A):H


bounded by II Tll, so it extends by continuity to a bounded linear operator
j(T) E L(:H); by definition, j(T)j(a) = j(Ta). lts adjoint j(T)t equals
j(T*), since

(j (b ); 1J (Ta) TJ) = (; IJ (b* (Ta)) 11)


=(;I j((T*b)*a)J7) = (j(T*b)~ I j(a)J7) (1.7)

for a, b E A, ~. 17 E :H. Thus j: T ~ j(T) is a morphism from :M(A) to


L(:H) that obviously extends j, is injective, and whose image is contained
in the idealizer C of j(A). If SE C satisfies Sj(A) = 0, then Sj(A):H = {0}
and by continuity S:H = {0}, so S = 0: thus j(A) is an essential ideal in C.
The injective morphism ß: C- .J.t(j(A)) "":M(A) given by ßb(a) := ba,
for b E C, a E A, inverts j, so j: :M(A) - Cis an isomorphism. D

Proposition 1.10. lfY is a locally compact space, then :M(Co(Y)) ""Cb(Y).

Proof. Let -1! 2 ( Y) be the Hilbert space of square-summable functions ~: Y -


( that vanish outside a countable subset of Y. This space carries a rep-
resentation j of Co(Y) by multiplication operators: if f E Co(Y), then
j(j)~ := J; lies in -1! 2(Y) also, with llf~ll2 5 11!1111~112· lt is easily checked
that j(C0 (Y)).f! 2 (Y) is densein .f! 2 (Y). By Lemma 1.16, :M(Co(Y)) is iso-
morphic to the idealizer of j(Co(Y)) in L(.f! 2 (Y)). Multiplication by any
g E Cb ( Y) gives a bounded operator on -1! 2 ( Y) that is clearly a multiplier
of j(Co(Y)).
1.3 Trading spaces for algebras 15

Conversely, any such multiplier is realized by a bounded function h: Y -


C, as can be seen by considering its effect on any ~ E -t 2 ( Y) supported
by a single point. If K c Y is a compact set, there is an f E Co(Y) that is
identically 1 onK, so hf E C0 (Y) coincides with h onK; thus the restriction
of h to any compact subset of Y is continuous. Since Y is locally compact,
this implies continuity of h on all of Y, therefore h E Cb(Y). D

Therefore, the construction of the multiplier algebra is a noncommuta-


tive generalization of the Stone-Cech ;:ompactification.
In this context, it is convenient to recall the extended definition of "mor-
phism" in the category of C*-algebras given by Woronowicz [495]: given
C*-algebras A and B, a morphism in this sense is an involutive homomor-
phism 17: A - .11-'l(B) such that the subspace ry(A)B is dense in B. Those
are clearly the noncommutative analogues of continuous maps between
noncompact spaces that arenot proper.
Exercise 1.9. Prove that, in the commutative case, Woronowicz morphisms
are the Gelfand duals of continuous functions that are not necessarily
proper. 0
There are other instances in which modifying the Straightforward notion
of morphism of C*-algebras proves useful: see Section 3.4 .
.".. So far, our dictionary has the following entries [481]:
TOPOLOGY ALGEBRA
continuous proper map morphism
homeomorphism automorphism
compact unital
a-compact a-unital
compactification unitization
Stone-Cech compactification multiplier algebra
open (dense) subset (essential) ideal
closed subset quotient algebra
metrizable separable
Baire measure positive functional
The second-last entry is justified as follows.
Proposition 1.11. A compact topological space X is metrizable if and only
if C (X) is separable.
Proof. If Xis metrizable, there is a countable family of open balls {Un}
generating its topology. Let fn(X) := d(x,X \ Un); clearly fn E C(X), and
this sequence of functions separates the points of the Hausdorff space X.
The selfadjoint algebra of functions generated by the constant functions
16 1. Noncommutative Topology: Spaces

and these fn is therefore densein C(X) by the Stone-Weierstrass theorem,


and obviously contains a countable dense subset.
Conversely, if C(X) is separable, it contains a dense sequence of con-
tinuous functions Un}neN; we can assume that llfnll 5 1 (replace fn by
fn I ( 1 + lfn I) if necessary). This sequence must separate points, since oth-
erwise it could not approximate a continuous function that separates two
given points of X (such functions exist by Urysohn's Iemma). Then the stan-
dard recipe d(x,y) := 2:;;': 0 2-nlfn(X)- fn(Y)i defines a metric an X
whose balls are open subsets of X. The identity map idx is thus a contin-
uaus bijection from the compact space X (with its original topology) to a
Hausdorff space (X with the metric topology of d) and so is a homeomor-
phism. D

The last entry corresponds to the Riesz-Markov theorem [406, Thm.


2.14]. Wehave already mentioned that characters arepure states. The mod-
ern presentations [132] of measure theory use the functional approach;
thus that entry becomes a truism.

1.4 Homotopy in noncommutative language


The translation of concepts of homotopy theory into algebraic language
is a quite trivial endeavour. Butan important one -and excellent Gelfand
gymnastics, too.
Notation. Given a C*-algebra A and a locally compact space Y, we shall
sometimes denote by AY the C*-algebra Co(Y -A) of continuous functions
of Y into A. For instance, if I is the unit interval [0, 1], then AI= C(l-A).
If a E AY and y E Y, the notations a(y) or ay E A for evaluation of a
at y are clear. Foreacht EI, let Pt: AI - A be the restriction morphism
a ,_. a(t).

Definition 1.9. If A, Bare two C*-algebras, two morphisms 17: A-B and
cjJ: A - B are homotopic, written 17 ~ cp, if there is a morphism 'I': A - BI
such that Po o 'I' = 17 and p 1 o 'I' = cp. A morphism 17: A - B is called a
homotopy equivalence when there is a morphism ~: B - A suchthat ~ o 17
and 17 o ~ are homotopic to the respective identity maps of A and B. When
~ o 17 = idA and 17 o ~ ~ idB, 17 is called a retraction of A into B. The algebra
A is contractible when idA is homotopic to the zero map.
If f, g: X - Z are maps between compact spaces, and if F: X x I - Z is
a homotopy between them, then CF: C(Z) - C(X x I) "' C(X)I, given by
Pt(CF(h)): x ,_. h(F(x, t)) for all h E C(Z), provides a homotopybetween
the morphisms Cf and Cg from C(Z) to C(X). If R .... Xisadeformation
retract and f: X - R is a retraction, then Cf is a retraction in the C*-
algebra sense.
1.5 Exponentials and cohomology 17

Warning. If a compact space X is contractible (to a point), however, the


commutative C*-algebra C(X) is not contractible. This is because idc and 0
are not homotopic via morphisms of (, since the only available morphisms
are idc and 0 and they cannot be linked continuously. In fact, no compact
space possesses a contractible algebra; but some locally compact spaces
do. For instance, C0 (0, 1] is contractible: a homotopy from 0 to id is given
by 'Y: Co(O, 1]- Co(O, 1]/, where, for 0 5 s 51,
('Yj)r(s) := j(st), for 0 < t 51; ('Yf)o(s) := 0.
Here Pt o 'Y maps C0 (0, 1] into C0 (0, t] by rescaling; 'Y is continuous since
elements of C0 (0, 1] vanish at 0. We yield to the noncommutative world
and say that (0, 1] is contractible, whereas ~ is not. On the other hand, if
X is contractible to a point, then AX and A are homotopically equivalent.
Exercise 1.10. Prove the last assertion. 0

The following lemma will eventually play an important role.


Lemma 1.12. If (X, Y) is a compact pair and Y is contractible to a point,
then the collapsing map c : X - X I Y is a homotopy equivalence.
Proof. Let f: Y - Y be a constant map and f ~ idy. We can construct
g: X- X proionging f with g ~ idx, and any such g factorizes as h o c,
with h: X/Y- X. Clearly, c o h: X/Y- X/Y is homotopic to idx;v. D

.,. In homotopy group theory one deals with morphisms f between pointed
spaces (X,*) and (Y, * ); denote by [X, Y]+ the corresponding space of
homotopy classes.
Given two C*-algebras A and B, denote by [A,B] the set of homotopy
classes of morphisms from A to B (that need not be unital); if A and B are
unital C* -algebras, denote by [A, B] + the set of homotopy dass es (via unital
homotopies) of unital morphisms. Note that [A,B]"' [A+,ß+]+. With A =
Co(Y), B = Co(Z), this becomes [Co(Y),Co(Z)] "' [Co(Y)+,Co(Z)+]+ "'
[Z+,Y+]+.
Recall now that the higher homotopy groups of (X, *) are defined as
1Tn(X, *) := [§n,X]+, for n <::. 0; therefore

Connes [91, II.A] suggests a generalization, by replacing Co(~n) by the al-


gebra Mk(Co(~n)).

1. 5 Exponentials and cohomology


Definition 1.10. Denote by Ax the group of invertible elements of any uni-
tal algebra A. Recall, from Section LA, that if A is a unital Banach algebra,
18 1. Noncommutative Topology: Spaces

then A x is open in A. An element a of A is called logarithmic if it can


be expressed in the form a = exp b, for b E A. Logarithmic elements are
invertible, with (exp b)- 1 = exp( -b), but the converse does not hold in
general.
For any Banach algebra A, the subgroup H of A x algebraically generated
by logarithmic elements is the component of the identity in A x (in partic-
ular, if Ais commutative, then exp(A) is the neutral component of Ax). In
order to prove that, note first that H is path-connected, since each element
exp b is connected to 1 by the arc t ...... exp tb, 0 :5 t :5 1. Now, if a E A
satisfies lla- 111 < 1, the power series

b := - f (1 - a)n
n=1 n
converges to a logarithm for a. Therefore there is a neighbourhood U of 1
in Ax suchthat U c H. Foreach c E Ax n H, the neighbourhood Uc of c
lies in H also. Hence H is open in A x. An open subgroup of a topological
group is always closed (since all its cosets are open too); therefore H must
be the whole neutral component of A x.
Exercise l.ll. Prove that in CCU") the logarithmic functions are precisely
the nowhere-vanishing functions of degree zero. Show also that any invert-
ible function in C(lfY can be expressed in the form
j(z) = zng(z),

where n is the degree of f and g is logarithmic. 0

Exercise 1.12. Prove that in C([0,1]) all functions that never vanish are
logarithmic. 0
Exercise 1.13. Prove that the neutral component of the unitary group of
a C* -algebra is generated by logarithmic elements of the form exp b, with
b = -b*. 0

~ Important topological information about a space M is encoded in its


integral Cech cohomology groups. It is natural -and good training, too-
to try to relate them to the structure and/or the invariants of the algebra
C (M). Familiarity with the basics of Cech cohomology is presupposed from
the reader; a lucid expositionis found in [478]. Also, [41,82] and [447], from
which the "long proof" of Theorem 1.13 below is taken, are quite useful.
First, let us recall that an idempotent in C(M), where M is compact,
is a continuous function e satisfying e 2 = e, and so e (X) ~ {0, 1}; thus
M = e- 1 (0) w e- 1 (1) disconnects M unless e is constant. So we can add
a new entry to our dictionary: "connected" translates into "without non-
trivial idempotents". More generally, if e is an idempotent (e 2 = e) in any
C*-algebra, then exp(iAe) = 1 + e(exp(iA) - 1) and thus exp(2rrie) =
1 + e(exp( -2rri) - 1) = 1.
1.5 Exponentials and cohomology 19

Theorem 1.13. Let A be a unital commutative C*-algebra. There are natu-


ral isomorphisms:

H 0 (M(A), ll) "'ker(exp: A- Ax), (1.8a)


H 1 (M(A),7l) "'coker(exp: A- Ax) = Ax I expA. (1.8b)

For instance, from Exercises 1.11 and 1.12, H 1 (§ 1 , ?l) = 7l and H 1 (I, ?l) = 0.

Short Proof. Denote by !:_ and _g:x the sheaves over the compact character
space M = M(A) such that, for any open set U c M(A), !:_(U) is the additive
group of continuous maps from U to C and _g:x (U) is the multiplicative
group of continuous maps from U to cx := C \ {0}. There is a short exact
sequence:
0-ll~C~C-0 - - I

where 2rri is the scaled inclusion m ,_. 2rrim of constant functions. We


thereby get a long exact sequence in cohomology:

Here odenotes the Bockstein homomorphism [140]. An easily verified prop-


erty of Cech cohomology isthat Hr (M, !:_) = 0 for r > 0 (see the discussion
below for the case r = 1). The initial portion of the long exact sequence is
then

which gives the conclusion of the theorem. From the succeeding terms we
extract, for future use,

(and so on, for whatever it is worth). D

The long proof consists of spelling this out in words, in particular con-
structing the Bockstein homomorphism explicitly, which is an instructive
exercise; we fashion the proof of (1.8b) in a way that can be adapted to
check (1.9) by repeating arguments almost verbatim.

Long Proof. A function f E C(M) satisfies exp(2rrij) = 1 if and only if it


is integer-valued. On the other hand, any continuous (i.e., locally constant)
integer-valued function g determines an integral Cech 0-cocycle, say g. lt is
clear that g ..... g is one-to-one and onto. This proves (1.8a). Note in passing
that H 0 (M, ll) = 0 if and only if A has no nontrivial idempotents.
Let g E C(M)X, defining an element g E Z 0 (M,~x). We choose a finite
open covering 11 = {UJ, ... ,Um} suchthat g(Ud is contained, for all i,
20 1. Noncommutative Topology: Spaces

in a disk in C not containing the origin. We can find smooth functions


fJ: Uj- C suchthat exp(2rrijj) = glur Define

(1.10)

whenever Ui n Ui * 0. Then exp(2rriaij) = 1, so aii is l-valued. Thus


a = {aii} is an element of C1 (11, l). Now (1.10) says that a = 8f in C 1 (11, {),
so (8a) ijk := aij- a ik + aki = 0 on Ui n Ui n Uk; these are algebraic relations
among ;z'-valued functions, and so 8a = 0 in C2 (11, l); hence a E Z 1 (11, l).
lf we take a different set of logarithms J; := ki + fi (we can work with the
same covering 11, by passing to a common refinement of two coverings if
necessary), then ki is integer-valued and a;i := aii + ki- kj. In other words,
we modify a to a + 8k in Z 1 ('U, l), and the dass [a] is unchanged. There-
fore we obtain a well-defined homomorphism a: [g] ..... [a] from Z 0 (M, r;;_x)
to H 1 (M, l). Now, if g has a (global) logarithm, clearly the correspond-
ing cocycle is zero. Therefore we get a well-defined homomorphism from
C(M)x / exp(C(M)) to H 1 (M, 7!.).
To see that this is an isomorphism, let {1./Ji} be a smooth partition of unity
subordinate to 11. That is to say, {1./J i} is a family of smooth functions, the
support of each 1./J i is contained in some open set Uj belanging to 11, the
family is "locally finite" in the sense that each x E M has a neighbourhood
Vx on which all but finitely many of the 1./J i vanish, and lastly, Lj 1./Ji = 1;
the local finiteness implies that this is actually a finite sum at each x. Now
suppose a E Z 1 (M,l) is any Cech 1-cocycle; define f E C0 ('U,{) by fJ :=
Lr ajri./Jr· Then

fi- fJ = :L<air- ajr )1./Jr =


y
L aijf./Jr = aij
y

on using 8a = 0; hence 8f = a. By considering the element g E C 0 ( 11, r;;_x)


defined by Bi := exp(2rrifi ), we see that a is onto. On the other hand,
suppose that a[g] = 0. This means that fi- fJ = ki- ki on Ui n Ui with
ki integer-valued. Clearly we can then define a globallogarithm f for g by
taking j(x) := jj(X) + ki on each Uj. 0

We have in effect established that an element of H 1 (M(A), l) repre-


sents a homotopy class of maps of M (A) into cx. We shall soon deal with
H 2 (M, l), in the context of line bundles. The group H 3 (M, ;z') will emerge
when we come to Connes' approach to Riemannian spin geometry in Chap-
ter 9.
This discussion of Cech cohomology and commutative C* -algebras is
almost trivial. Far deeper is the fact that the Gelfand transformation ac-
tually induces the analogaus isomorphisms of Theorem 1.13 for arbitrary
commutative Banach algebras; these constitute, respectively, the Shilov and
Arens-Royden theorems [447].
1.6 Identifications and attachments 21

1.6 Identifications and attachments


Let (M,N) be a compact pair, Pa third compact space, and f: N - Pa
continuous map. In the disjoint union P wM we identify the points x E N
and f(x) E P; the space so obtained is called the attachment of M toP
along N by means of f and is written P u f M. lt is clear that the quotient
map p: P w M - P u f M restricts to a homeomorphism of P onto p (P) and
of M \ N onto (P UJ M) \ p(P).
Exercise 1.14. Show that Pu f M is Hausdorffand compact. Show moreover
that it is connected if M and P are connected. 0

The attachment construction is of foremost importance in homotopy the-


ory, Morse theory, and related fields in algebraic and differential topology.
A primary task is to find a correlated concept in C*-algebra theory: this is
provided by the pullback or fibred product construction.
Definition 1.11. Given a C*-algebra B and morphisms cp: A1 - B and
*
11: A2 - B, the pullback of B along ( cp, 11) is the -subalgebra of A1 EB A2
defined by

Note that A(B; cp,17) comes into the world with natural maps to A1 and A2:

A(B; cp,17) _,.. A1

!
A2
ry !~
B.

In most cases one of the maps cp, 17 (or both) will be onto.
lt is clear that C(P UJ M) ""A(C(N); Cf, Ci), where i: N ..... M: just con-
sider the morphism '1': C(P UJ M) - A(C(N); Cf, Ci) given by 'I'(F) :=
(F o plp,F o PIM), which is weil defined since F o PIP o f = F o PIM o i. It is
an isomorphism: if (g, h) E A(C(N); Cf, Ci) so that g o f = h o i, '1'- 1(g, h)
is the function given by p(y) ..... g(y) if y E P, or p(z) ..... h(z) if z E M,
which is weil defined and continuous for the quotient topology determined
by p.
Definition 1.12. A closed ideal] in A gives rise to the standard short exact
sequence of C*-algebras:
j ry
0 - ] - A - A/]- 0.

When], A, B are arbitrary C* -algebras for which there is an exact sequence


j ry
0-]-A-B-0, (l.ll)
22 1. Noncommutative Topology: Spaces

then j(J) is an ideal of A (since it equals ker ry) and B is isomorphic to


the quotient of A modulo this ideal. For convenience, we shall sometimes
abbreviate (l.ll) as] J_ A-1 B.
The exact sequence of C*-algebras (l.ll) is split:
j r)
0-]-A-B-0,
(T

if and only if there is a morphism u: B - A such that 17 o u = idB. (In


the notation, the two maps 17 and u are combined on a single arrow.) This
happens if and only if A = j(J) + u(B), where j(J) and u(B) sit inside A
as subalgebras with zero overlap. In this case, there is also a (-linear map
rr : A - ] with rr o j = id1 , but rr is not multiplicative in general. Indeed,
rr is a morphism if and only if u(B) is also an ideal in A, whereupon we
write A = j(J) e u(B); the ideals j(A) and u(B) are said tobe mutually
"orthogonal".
As an instructive example, consider the unitization-augmentation se-
quence

(1.12)

and the maps u: ( - A+: .\ ..... (0,.\) and rr: A+- A: (a,.\) ..... a. The
sequence is split exact, and A + is the vector-space direct sum of A and C;
however, ( sits as an ideal in A + only when A is already unital, in which
case A+ ""A e ( -by means of (a, .\) ..... (a + .\1A) e .\- is an orthogonal
sum of C* -algebras.
Definition 1.13. Given two C* -algebras ], B, an extension of B by] is a C*-
algebra A together with morphisms such that there is an exact sequence
0 - ] - A - B - 0. An extension is called "trivial" when the sequence is
split exact.
Sets of (appropriately defined) equivalence classes of extensions have an
interesting algebraic structure; in particular, when B is a commutative al-
gebra, they yield a functor from locally compact spaces into abelian groups
and the resulting interplay between operator theory and algebraic topology,
when first introduced, solved some outstanding problems. This approach
was pioneered by Brown, Douglas and Fillmore in the seventies: see [146],
for instance.
The commutative example is, in view of Exercise 1.5,
j r)
0 - Co(Z)- C(X)- C(X \ Z)- 0,

for Z an open subset of X, and, in particular,


j I)
0 - Co(M \ N)- C(M)- C(N) -0,
1.6 ldentifications and attachments 23

for a compact pair (M,N).


~ A very interesting situation arises with a locally compact but noncom-
pact space Y; we takeforM some compactification of Y suchthat N := M\ Y
is its boundary. Then the various solutions of the extension problern of
C(P) by C0 (Y) correspond to the different ways to attach M toP along N.
This happens all the time in practice: for instance, in the construction of a
CW-complex, if Pisa suitable subcomplex, M is a closed unit ball and N is
its boundary, one must make an explicit attachment of the cell M \ N to the
subcomplex. A particular case in point is given by the short exact sequence
o- Co(~ 2 )- C(I!Jl)- ccn- o, (1.13)

where iD denotes the closed unit disk. Note the difference with the unitiza-
tion (i.e., one-point compactification) extension:
0 - Co(~ 2 )- C(§ 2 )..!.... <C- 0,
where E denotes evaluation at the point at infinity.
Weshall now apply the noncommutative pullback construction with the
morphism 17 being onto. Important particular cases are the associated map-
ping cylinder and mapping cone algebras of a morphism of C*-algebras.
Definition 1.14. The mapping cylinder of a morphism cf>: A - B is the
C*-algebra
ZcJ> := A(B; cf>, p!) := { (a,j) E A e BI: j(l) = cp(a) }.
The cone over a C*-algebra Bis the contractible C*-algebra CB := B(0,1].
If p;: CB - B, for 0 < t :s; 1, denotes the evaluation morphism f - j(t),
the mapping cone of the morphism cf> is the C*-algebra
CcJ> := A(B; cf>,pi) := { (a,j) E A e CB: j(l) = cp(a) }.
Clearly, CcJ> is an ideal in ZcJ> and ZcJ>/CcJ> "" B. Also, if cf> = idA is the
identity morphism on A, then Cct_4 ""CA.
Exercise 1.15. Find a homotopy equivalence between ZcJ> and A. 0

~ lt is timetobring in the suspension functor. We make the commutative


definition in the context of pointed spaces. The functor Y - y+ carries
cartesian products into smash products. This is seen as follows. The bou-
quet X v Y and the smash product X A Y of two pointed spaces (X,*),
( Y, * ) are defined as

X V Y := (X X { *}) U ( { *} X Y) = (X X Y) \ ((X \ *) X ( Y \ *));


X A Y :=(X x Y)/(X v Y).
In this way, X v Y becomes the base point of X A Y. Now, if X, Y are un-
pointed, (X x Y) + is the one-point compactification of the complement of
(X+ X { *}) U ( { *} X y+) in X+ X y+, hence equal to x+ 1\ y+.
24 1. Noncommutative Topology: Spaces

Exercise 1.16. Check commutativity and associativity for the smash pro-
duct. o
Exercise 1.17. Checkthat there is a bijection

for any triple of pointed spaces. 0

In particular, we define the suspension LX of a pointed space X as § 1 A X;


alternatively, one can form the one-point compactification of IR x (X\ *).
Note that (!Rn)+ = [R+ A • • • A [R+ (n times), that is, §n = Ln§o = § 1 A • · · A § 1
(n times). Moreover, Ln(X+) =(!Rn x X)+.

Definition 1.15. The suspension of a C*-algebra Ais the C*-algebra

LA:= {jEAI:j(O) =f(1) =0}


== {jECA:j(1) =0} ==AIR==Co(IR)®A. (1.14)

Thesuspensionofamorphism<j>: A- BisthemorphismL</> = idco(lll) ®</>:


LA- LB given by L</>(J) := </> o f.
Remark. The tensor product C*-algebra Co( IR)® Ais a completion of the
algebraic tensor product Co (IR) 0 A generated by simple tensors f ® a with
f E C0 (!R), a E A. Taking tensor products of C*·algebras is a surprisingly
delicate matter: the idea is to complete the algebraic tensor product A 0 B
in a normthat is both a cross-norm, i.e., II a ®b II = II a II II b 11. and a C* -norm;
there is always at least one such norm, but there may be several. For details,
see [352, Chap. 6]; a fine pedagogical walk-through is given in [481, App. T].
The matter is also dealt with briefly in Section LA. Happily, if A or B is
either abelian or is the C* -algebra X of compact operators on an infinite-
dimensional separable Hilbert space, the C*-cross-norm is unique and we
can avoid this discussion in those cases.
We probe the C*-algebraic notion of suspension with several exercises.
Exercise 1.18. Show that the suspension LAis contractible whenever the
algebra A is contractible. 0

Exercise 1.19. Show that L(Al) == (LA)/. Conclude that if the C*-algebras
A and B are homotopy equivalent, then LA and LB are also homotopy equi-
v~em. o
Exercise 1.20. If 1 is an ideal in A, then L1 is an ide~ in LA and L(A/]) ==
LA/L1. Every exact sequence 0 - 1 _j_ A ..!!_All - 0 induces an exact
sequence of Suspensions

'f.j 'Ir)
0 ~ L1 ~LA~ LA/L1 ~ 0. 0
1.6 Identifications and attachments 25

Exercise 1.21. Every split exact sequence 0 - ] _j_ A....!!.... A/]- 0 induces
u
a split exact sequence of suspensions:

I.j I.ry
0 - - - 'I.]--- "l.A ~ "l.A/"l.]--- 0.

Proposition 1.14. For every morphism of C* -algebras cf>: A - B, there is


an exact sequence
}' ß
0-"l.B-CcJ>-A-0. (1.15)

In particular, the following sequence is exact:

0-"l.A-CA-A-0.

Proof. The maps in (1.15) are y(j) := (O,f) and ß(a,f) := a. 0

We need one more exact sequence related to 0 - ] _j_ A...!!... B- 0.


Proposition 1.15. Given (l.ll), there is an exact sequence

O-]-C11 -CB-O.

Proof. Since 1J(j(c)) = 0 for any c E ], there is a map a:] ...... C11 : c -
(j(c), 0); its image is the kernel of the map (a,f) - f: C11 - CB. 0

11> The notions of mapping cylinder, cone, mapping cone and suspension
are algebraic counterparts of well-known topological notions [150). We il-
lustrate them by their roles in the definition of the Puppe sequence, which
we now recall. Let f: M - P be a continuous map between compact spaces;
it is known to be the first arrow of an infinite exact sequence

f jf f q I.f I.]f f I.q 2 ( )


M - P - c -"l.M-"l.P-"l.C - " l . M - · · · . 1.16
A A A A

In order to make sense of (1.16), we consider first the mapping cylinder


Mf fashioned by attaching Mx I toP along M x {1} by means of the map
(x, 1) - f (x) E P. Now, both M and P are identified to closed subspaces of
the mapping cylinder. Also, Mf retracts on P by the homotopy r: Mf x I -
Mf given by

r(p(x, t), u) := p(x, u + (1- u)t),


r(p(y), u) := p(y), for XE M, y E P.

It is immediately seen that Mf is the Gelfand dual of the C*-algebraic map-


ping cylinder: ZcJ = C(Mf).
The unreduced cone CM over M is obtained fromM x I by identifying
Mx {0} to apoint. Note that CM is a space contractible to a point: '1'5 (x, t) :=
26 1. Noncommutative Topology: Spaces

(x, st), for s EI, provides a homotopy between idcM and the constant map
of CM into its basepoint. The unreduced mapping cone cf of f is defined
as P u 1 CM; it is often convenient to see it as the quotient Mf /M. Also,
when M = P and f = idM, then cf is just the cone CM of M. Theseare
not quite the Gelfandduals of the cones considered in Definition 1.14. For
instance, CC(M), the cone algebra over C(M), which is not unital, cannot
coincide with C(CM), the continuous function algebra over the (compact)
cone CM. The latter is the unitization of the former: regard the elements
of CC(M) as functions on Mx I that vanish at the points (x, 0).
Exercise 1.22. Show that, if (M, N) is a compact pair, then MI N is homeo-
morphic to Ci I C N, where i: N .... M is the inclusion. 0
Finally, the unreduced suspension iM is obtained fromM x I by identify-
ingbothMx {0} andMx {1} to points: clearly iM"' CM /Mx {1} "'CM /M.
lt is natural to take the umeduced suspension of the empty set to be the
(two-point) sphere § 0 . Neither of the two commutative notions of sus-
pension is quite the Gelfand dual of the suspension considered in Defini-
tion 1.15. For instance, it is not true that the suspension :LX for a compact
pointed space (X,*) is the character space of :LC(X), since this algebra
is never unital! lnstead, :LC(X) = Co(~ x X). (In fact, what we did was
equivalent to defining :LX, for X locally compact, simply as ~ x X, hence
:LCo (X) = Co (:LX) in general.) Care needs tobe exercised, then, in switching
between these related concepts in spaces and algebras.
Coming back to the Puppe sequence: the map ] f: P - Cf is obtained by
composing the inclusion P .... Mf and the quotient map Mf- Mf IM; now
P is identified to a subspace of Cf and the next arrow q is the canonical
projection onto the quotient Cf /P "'iM. Finally, :Lf is the SUSpension of
f, sending the image of (x, t) into the image of (j(x), t). The rest of (1.16)
is clear.
By definition, the "Puppe sequence" of algebras, associated to a mor-
phism cf>: A - B, is of the form

ry rß rcf> y ß cf>
· · · ___,.. :L2A ___,.. :LCcf> ___,..:LA___,.. :LB ___,.. Ccf> ___,.. A ___,.. B. (1.17)

The maps ß and y are those given in Proposition 1.14.


The two constructions are quite similar, although not strictly parallel.
Wehave presented them in the form most suitable for our purposes, tobe
revealed in Chapter 3.

LA C* -algebra basics
We collect here, for the reader's convenience, several facts and theorems
about C*-algebras as background for the main text. There are many good
LA C*-algebra basics 27

textbooks on this subject: we recommend [129, 137,183,266,352,366,481],


in no particular order.
Definition 1.16. A Banachalgebra is an associative algebra over the field
([ of complex numbers that is also a complete normed space, and in which

llabll 5 llallllbll (1.18)

for all elements a, b of the algebra; this condition guarantees continuity of


the product. If a Banachalgebra contains a unit 1, we mayasweil assume
that 11111 = 1; for if not, the operator norm of the map b ,__ ab yields an
equivalent norm for which the unit has norm 1. Any Banach algebra can be
unitized by defining A+ := A x ([ as in Section 1.1, and extending the norm
in a convenient way, by setting ll(a,.\)11 := sup{ llab +Ab II: llbll 51}, for
instance.
An involution in a Banach algebra is an isometric antilinear map a ,__ a *
satisfying a** = a and (ab)* = b*a*. When a particular involution is
given, we speak of a Banach *-algebra. A C*-algebra is then a Banach *·
algebrathat satisfies the crucial equality lla*all = llall 2 for each element
aEA.
If A is a C* -algebra, then so is A +, since

ll(a,A)*(a,A)II = sup{ lla*ab + Äab + Aa*b + Ä.\bll: llbll 51}


~ sup{ llb*a*ab + Äb*ab + Ab*a*b + ÄAb*bll: llbll 51}
= sup{ ll(ab + .\b)*(ab +Ab) II: llbll 51}
= sup{ llab + Abll 2 : llbll 51}= ll(a,.\)11 2 ,

and the opposite inequality II (a, .\)* (a, .\) II 5 II (a, .\) 11 2 follows from (1.18).
If 1 is a closed (two-sided) ideal in a Banach algebra A, then the quotient
algebra A/] is also a Banachalgebra under the obvious norm lla + 111 :=
inf{ lla + bll : b E 1 }. lf Ais unital, then 1 + 1 is a unit for A/]. If Ais a
C*-algebra, so is Al 1.
Already in any unital Banachalgebra A, the geometric series c := L:k'~o bk
converges absolutelyif llbll < 1, since its normis majorized by L:k'~o llbk II 5
L:k'~o llbllk = (1-llbll)- 1 . Clearly, bc = cb = c-1, so (1-b)c = c(l-b) = 1.
Setting a := 1 - b, we find that a is invertible provided 111 - a II < 1. More
generally, if x is invertible and llx- yll < l/llx- 1 11. then lll-x- 1yll < 1, so
y is also invertible: the set A x of invertible elements is an open subset of A.
Therefore, a proper ideal in a unital Banach algebra cannot be dense, as then
it would contain an invertible element. Cantrast this with the nonunital C*-
algebra X of compact operators on a separable infinite-dimensional Hilbert
space, which has many dense ideals (see Section 7.C).
Definition 1.17. Fora E A, the vector-valued function Ra: .\ ,__ (,\- a)- 1 is
defined and holomorphic on an open subset of the complex plane C, with
28 1. Noncommutative Topology: Spaces

a convergent Laurent series Ik'=oA-k-lak on the annulus lAI > llall. Since
the sum of this series tends to 0 as lAI - oo, the function Ra cannot be
extended to an entire function on (( (since, by Uouville's theorem, it would
then have the constant value 0). Therefore, the set of values

spa := { ,\ E ((: (,\- a) is not invertible}

is a nonvoid, closed (therefore, compact) subset of { z E (( : Iz I :5 II a II }, and


is called the spectrum of a. The smallest disk (centred at 0) that includes
the spectrum has radius r(a) := limn-oo 11an1111n. Notice that a nilpotent
element satisfies r (a) = 0 and hence sp a = {0}.
Any polynomial equation f1 - j(z) = (,\ - z)g(z) entails f1 - j(a) =
(,\- a)g(a) in A, so if f is a complex polynomial, then f1 E spj(a) if and
only if f1 = j(,\) for some ,\ E sp a; in other words, sp j(a) = j(sp a).
There are several "functional calculi" that seek to extend this relation to
more general functions, the main problern being to suitably define the el-
ement j(a); at any rate, f ..... j(a) must be an algebra homomorphism
into the (commutative) closed subalgebra generated by a. For general Ba-
nach algebras, the best one can do is to replace polynomials by functions
holomorphic near sp a: if j(a) is defined by the integral

j(a) := _21. ,[ j(1;,) ((1- a)-1 d?;, (1.19)


rrt Jr
on a rectifiable contour r that winds once araund sp(a), then the spectral
mapping and homomorphism properties hold; this is called the holomor-
phic functional calculus.
When Ais a nonunital Banach algebra, the spectrum of an element a E A
is defined tobe its spectrum in A +; it is then automatic that 0 E sp a, since
elements of A are not invertible in A +. The polynomial and holomorphic
functional calculi still make sense in A, provided they are constrained to
functions f satisfying j(O) = 0.
~ For the rest of this section, weshall suppose that Ais a C*-algebra. In a
C* -algebra, the spectrum of a selfadjoint elementaisreal (see Exercise 1.1).
Moreover, its spectral radius r(a) is equal to its norm: this follows from
lla 2 11 = llall 2 and r(a) = limn-oo llanlllln. From this last equality it also
follows that r(ab) :5 r(a)r(b) if a and b commute. A normal element
a E A is one that satisfies aa * = a *a. For instance, any selfadjoint or any
unitary (a*a = 1 = aa*) element is normal. If a is normal:

llall 2 = lla*all = r(a*a) :5 r(a*)r(a) :5lla*llllall = llall 2 ,


so r(a) = llall, too. The definition of normality means that the C*-sub-
algebra generated by a is commutative. By applying the Gelfand-Naimark
theorem 1.4 to this subalgebra, we get a continuous functional calculus,
whereby f ..... j(a) is extended to all functions in C(spa), by matehing
l.A C*-algebra basics 29

uniform convergence of polynomials to norm convergence of elements of A


(see the remark on the spectral theorem in the main text).
A morphism of C*-algebras is by definition a *-homomorphism. Any
morphism cp: A - B extends uniquely to a unital morphism cp +: A + - B+.
Lemma 1.16. Any morphism of C* -algebras is norm-decreasing, and so is
continuous.
Proof. lt is enough to check this for a unital morphism cp: A - B of unital
C*-algebras.If(.\1-a)c = 1inA,then(.\1-rp(a))rp(c) = 1inB;therefore,
sp rp(a) c;; sp a. In particular, r(cp(a)) ::5 r(a).lt follows that
llr:fJ(a)ll 2 = llr:fJ(a*a)ll = r(cp(a*a)) ::5 r(a*a) = llall 2 . D
Definition 1.18. A positive element of A is a selfadjoint element a for
which sp a c [0, oo). lt has a (unique) positive square root; one can define
a 1 12 := j(a) using j(x) := JX on [0, !lall]. An elementaispositive if and
only if a = b*b for some b E A [137, §1.6], if and only if (~I a~} ~ 0 for
any vector ~ in a Hilbert space on which A acts faithfully.
We write c ::5 d for selfadjoint elements c, d E A whenever d - c is
positive. If Ais unital, any positive element satisfies 0 ::5 a ::5 llall1, since
sp(llall1- a) c [0, oo); or equivalently, since the functionj(x) := x on the
interval [0, !lall] satisfies 0 ::5 f ::5 !lall. A seifadjoint element is positive if
and only if II a - t 111 ::5 t for t ~ II a II, again by functional calculus. This
property can be used to show that the sum of two positive elements is
positive, so that the set of positive elements of A is a convex cone, and the
relation c ::5 d is a partial ordering on the set of seifadjoint elements.
Exercise 1.23. Show that if 0 ::5 a ::5 1 in A, then 0 ::5 a2 ::5 a. 0

Definition 1.19. An approximate unit in a C*-algebra is an increasingly


ordered net {u 01 } of positive elements of A such that every II u 01 II ::5 1 and
llbu 01 -bll- Oforeachb E A(and therefore llu 01 b-bll = llb*u 01 -b*ll- 0
too). Such nets certainly exist in nonunital C*-algebras; for instance, one
can take as index set all the positive elements of norm less than one, putting
Ua := a. It turnsout that such a net may be chosen from any dense ideal
of A, and can be chosen to be an increasing sequence when A is separa-
ble [137, Prop. 1.7.2]. More generally, a C*-algebra with a countable ap-
proximate unit is called a -unital.
A linear functional cp: A - I( is called positive if cp (a) ~ 0 for all a ~ 0
in A, or equivalently if cp (b * b) ~ 0 for all b. If A is unital, this implies
that 0 ::5 rp(a) ::5 !lall rp(l), so that cp is automatically continuous with
llr:PII = rp(l). (In the opposite direction, any continuous linear functional
satisfying II cp II = cp (1) must be positive.) In the nonunital case, continuity
is also guaranteed, with II cp II = lim 01 cp (U 01 ) for any approximate unit [366,
Prop. 3.1.4]; moreover, cp extends to a positive linear functional r:p+ on the
unitization A + just by setting cp + (1) : = II cp 11.
30 l. Noncommutative Topology: Spaces

If cjJ and l/J are positive functionals on A such that II cjJ II = lll/J 11. and if
cjJ -l/J is also positive, then cjJ = l/); for we may suppose that Ais unital, and
then it is enough to notice that II cjJ - l/J II = (cjJ - l/J) ( 1) = II cjJ II - lll/J II = 0.
A linear functional T: A- Cis called tracial if T(ab) = T(ba) for all
a, b E A.
Definition 1.20. A positive linear functional of norm one is called a state
of the C* -algebra. If A is unital, any state satisfies cjJ (1) = 1. A state cjJ is
called faithful if a ~ 0 and cjJ(a) = 0 imply a = 0.
A (normalized) trace on A is a nontrivial tracial state.
The space of states is a convex set; in the unital case, this follows from
the equality 11(1- t)cjJ + tl/JII = ((1- t)cjJ + tl/))(1) = (1- t) + t = 1 if c/J,l/J
are states and 0 ::::; t ::::; 1. The extreme points of this convex set are called
the pure states.
Any state cjJ of a C*-algebra A gives rise to a representation TT<f> of A,
by what is called the Gelfand-Naimark-Segal construction, or "GNS con-
struction" for short [426]. The starting point of this construction is the
observation that
(a I b)<f> := cjJ(a*b)
defines a positive semidefinite sesquilinear form on the vector space A. As
such, it satisfies the Schwarz inequality, in the form

lc/J(a*b)l 2 ::::; cjJ(a*a) cjJ(b*b).

Therefore

N<f> := { b E A: cp(b* b) = 0} = { b E A: cp(a* b) = 0 for all a E A}

is a closed left ideal in A. We say that cjJ is a faithful state if N<f> = 0. The
quotient vector space AI N <I>, with elements f! : = a + N <I>, is then a prehilbert
space und er the positive definite scalar product (f! 112.> <I> := cjJ (a * b). Denote
by J{<l> its completion to a Hilbert space.
If b E A, the map A- A: c- b*cb preserves positivity, and so a*a::::;
II a * a II 1 (valid in A +, if A is not unital) entails the inequality

b*a*ab::::; lla*all b* b (1.20)

among positive elements of A. Applying cjJ to both sides gives the inequality

cjJ(b*a*ab)::::; lla*all c/J(b*b). (1.21)

In particular, the maps on Al N<f> defined by

TT<f>(a): 12.- ab (1.22)

extend to bounded operators on J{<l> satisfying llrr<t>(a)ll ::::; !lall. The map
TT<f>: A- L(J{<f>) is an algebra homomorphism, and since (Jzlrr<t>(a)f.)<t> =
LA C*-algebra basics 31

cf>(b*ac) = (TT<f>(a*)f!. 1 f.)<f>, the adjoint operator to TT<f>(a) is TT<f>(a)t =


rr<P (a * ). Therefore, rr<P is an (involutive) representation of A on Jf4>.
If A is unital, write ~<I> := 1. Otherwise, take an approximate unit {Uoc}
and define ~</> := limoc .:!!oc; indeed, the estimate
ll.:!!oc- .:!!ßll~ = </>((Uoc- Uß) 2 ) ~ </>(Uoc- Uß) = </>(Uoc)- </>(Uß)
for oc ;?: ß, and the convergence cf>(uoc) - 1 shows that the net {.:!!oc} has
the Cauchy property, so it converges in Jf4>. (The inequality follows from
Exercise 1.23.) Now TT<f>(b)~<f> = !!_ for b E A, so that TT<f>(A)~<f> = A/N<f>.
This says that ~</> is a cyclic vector for the representation TT<f>, which is to
say that the TT<f>(A)-invariant subspace generated by ~</> is dense in Jf4>.
Furthermore,
(~<f> I TT<f>(a)~<f>)<f> = cf>(a) for all a E A. (1.23)
When the state <P is faithful, J{<P is the completion of A itself in the new
norm llall<t> := -J<f>(a*a). In that case, ~</> is also a separating vector, that
is, !!_ = TT<f>(b)~<f> = 0 in J{<P implies b = 0 in A.
The representation TT<f> is irreducible if and onlyifthe state </> is pure [137,
Prop. 2.5.4].
With the GNS construction in hand, we can state the second Gelfand-
Naimark theorem.
Theorem 1.17 (Gelfand-Naimark). Any C* -algebra has an isometric repre-
sentation as a closed subalgebra ofthe algebra L(Jf) of bounded operators
on some Hilbert space.

Sketch proof. One uses the Hahn-Banach theorem to show that for any
nonzero positive element b* b E A, there exists a state «fJ = «/Jb such that
«JJ(b*b) = llb*bll = llbll 2 [129, Lemma 1.9.10]. Then (1.23) shows that
llrr~P(b)~~PII~P = llbll. Therefore, we can find a family <I> of states for which
<P ( b * b) = 0 for all <P E <I> implies b = 0 - if necessary, take <I> to be all
states on A. Now form the direct sum of the corresponding GNS represen-
tations rr := EB<t>E<I> TT<f>, acting on J{ := EB<t>E<I> Jf<P; then llrr(b) II = llbll for
every b E A, so rr is an isometric representation of A on L(Jf). 0

In particular, if Ais a separable C* -algebra, we can take <I> = {«!Jh} where


{bk} is a countable dense subset of A, so that J{ is then a separable Hilbert
space.
Corollary 1.18. A selfadjoint element a E A is positive iff cf>(a) ;?: 0 for any
state </> on A.

Proof. For any unit vector 17 E Jf, the linear functional a ...... (IJI rr(a)IJ)
is a state of A, since b*b ...... llrr(b)IJII 2 ;?: 0. Therefore, rr(a) isapositive
operator on Jf, so it has a positive square root rr(a) 112 . Since rr is an
isomorphism, this is ofthe form rr(b) for a unique b = b* inA, and rr(a) =
rr(b) 2 implies a = b 2 • o
32 1. Noncommutative Topology: Spaces

If Ais a C*-algebra, then so is Mn (A) =Mn(([) ®A for any n = 2, 3, ... ; its


elements are matrices [aii] with entries in A. Each representation rr: A -
L(J-{) gives rise to a representation, say rr<n>: Mn (A) - L(([n ® J-{), given
by (rr<n>(a)l7)i := LJ=l rr(aij)l7j for 17 = (171. ... , 17n), and if rr is injective
then so is rr<n>.
Lemma 1.19. An element of [aij] E Mn(A) is positive if and only if it is a
sum of matrices of the form [ at a i] with a 1, ... , an E A.

Proof. If a E Mn(A) is the matrix whose first row has entries a 1 , ••• ,an
and whose other entries are 0, then a*a = [ataj], so such matrices are
positive in Mn(A). On the other hand, if a = b*b isapositive element of
Mn(A), then a = c1 + · · · + Cn, where Ck = [bZibkj]. 0

Proposition 1.20. An element of [aij] E Mn(A) is positive if and only if


If.j=l c{ aijCi is positive in A for all CI. ... , Cn E A.

Proof. If a = b* bis positive in Mn (A), then


n n n n
L c{aijCj = L c{bZibkjCj = L dZdk. where dk := L bkjCj.
i,j=l i,j,k=l k=l j=l

Conversely, if If.i=l c{aijCj is positive for all c1, ... , Cn, let cp be any state
of A. The vectors of the form 17 = (rr.p(CI)~.p •... ,rr.p(cn)~.p) make up a
dense subspace of cn ® J-{.p, on which

(n I rr~n>(a)n) = .t (~.p rr.p(c{aijCj)~.p)


I,J=l
I = c/J(.f.
l,J=l
c{aiici) ~ 0,

so that each rr~n>(a) isapositive operator. If rr = ffi.pe~ rr.p is the injec-


tive representation of A given by Theorem 1.17, then rr<n> = ffi.pe~ rr~n) is
also injective, and rr<n> (a) isapositive operator; therefore, aispositive in
Mn(A). 0

~ Finally, we briefly address the matter of tensor products of C*-algebras.


The issue in defining such tensor products is to find a suitable norm. Con-
sider first the tensor product of two Hilbert spaces J-{ and J-{'. The alge-
braic tensor product J-{ 0 J-{', namely the vector space consisting of finite
sums of simple tensors LJ=l ~i ® 'Ii· is a prehilbert space under the scalar
product (~I® '11 I ~2 ® '12) := (~I I ~2) (171 I '12). The Hilbert space J-{ ® J-{'
is defined as the completion of :J{ 0 J-{' in the corresponding norm, which
is a cross-norm, that is, II~ ® nll = 11~11111711 in all cases. If S E L(J-{) and
T E L(J-{'), the linear map ~ ® 17 - S~ ® T17 on J-{ 0 J-{' is bounded and
extends to a bounded operator S ® T on J-{ ® J-{'.
LA C*-algebra basics 33

For more general Banach spaees E and F, there may be several cross-
norms on E 0 F, eaeh one yielding a different eompletion. The most impor-
tant of these is

where an element z E E 0 F may be written in many ways as a finite sum of


simple tensors. The eompletion of E 0F in the norm y is ealled the projective
tensor product, usually written E ® F. Taking ,\(z) tobe the supremum of
I(j ® g) (z) I, where f and g are eontinuous linear funetionals of norm one
in the dual spaees E* and F* respeetively, yields another eross-narm ,\, and
in general any eross-narm on E ® F satisfies ,\(z) ::s; llzll ::s; y(z).
For C*-algebras A and B, we eonsider only eross-norms having the C*-
property: llc*cll = llcll 2 for all c E A 0 B. There is a smallest norm a and
a largest norm f.l in this family. The former is given by a(c) := sup{ (c/> ®
lfJ){z*c*cz)/(c/>® lfJ){z*z)}, where the supremumis over all states 4> of A
and lfJ of Bandall z E A 0 B for whieh (c/> ® lfJ)(z*z) > 0. The latter
is f.l(Lj ai ® bj) := supii'Li rr(aj)p(bj)ll. where the supremum isover all
pairs of commuting representations rr: A - L(.Jf) and p: B- L(.Jf) on
the same Hilbert spaee. lt turnsout that any C* -norm II ·II on A 0B is in faet
a eross-narm [267, Cor. 11.3.10] and it must satisfy ,\ ::s; a ::s; II · II ::s; f.l ::s; y.
For the proofs of these inequalities, eonsult [267, §11.3], [352, Chap. 6]
or [481, Appendix T].
Definition 1.21. The eompletion of A 0 Bin the norm a is a C*-algebra,
ealled the (spatial) tensor product A ® B. A C*-algebra Ais ealled nuclear
if, for any C*-algebra B, the algebraie tensor produet A 0 B has only one
C* -eross-norm, namely the spatial one. In that ease, A ® B is referred to as
the tensor produet of A and B.

lt is known [352] that finite-dimensional C*-algebras are nuclear, with


Mn(C) ® B :::: Mn(B); and that eommutative C*-algebras are nuclear, with
Co(Y) ®B"" Co(Y -B). The C*-algebra Xis also nuclear. The larger algebra
L(.Jf) is an example of a nonnuclear C*-algebra.
Example 1.1. An element c E A 0 B ean be expressedas a finite sum c =
Li a i ® bi in several ways; let

(1.24)

Then II · llh is a eross-narm (although not a C*-norm), and the eompletion


of A 0 Bin this norm is ealled the "Haagerup tensor produet" A ®h B. One
of its useful features is that the multiplieation m: A 0 A - A : a ® b - ab
extends to a norm-deereasing map from A ®h A to A (see Exereise 2.14).
This norm plays an important role in the theory of operator spaees, i.e.,
closed subspaees of L(.Jf); for baekground, see the survey [80].
34 1. Noncommutative Topology: Spaces

l.B Hopf algebras and Tannaka-Krein duality


In the unusual Situation when the compact space X is not Hausdorff, the
algebra C(X) is still a C*-algebra. Nevertheless, we do not recover X from
the algebra C(X). At the other extreme, when the space X not only is a
compact, Hausdorff space, but has some extra structure, then the space X
can sometimes be recovered from a smaller algebra. The algebra of smooth
functions on a manifold is a case in point.
If G is a compact topological group, then G can be recovered from the
algebra ofreal representative functions R(G), which turnsouttobe a Hopf
algebra. In this appendix section, we outline the basic theory of (real or
complex) Hopf algebras, and give an account of the reconstruction theorem
of Tannaka and Krein. This is a close analogue of the Gelfand-Naimark
theorem, but requires us to work over the real field. What we learn about
Hopf algebras here will be of use in Chapter 14, which is concerned with a
particular, important Hopf algebra.
From a mathematical point of view, we seek the algebraic objects corres-
ponding to topological groups in the quantization program as described,
for instance, in [152]. A satisfactory correspondence for locally compact
groups, including an extension of Pontryagin duality, has only recently been
found [300, 301]; it requires a considerable amount of C*-technology. Here
we shall outline the equivalent problern for compact groups, where the al-
gebraic objects in the commutative case turn out to be commutative Hopf
algebras, and we develop an equivalence between the categories of compact
groups and of these Hopf algebras.
A Hopf algebra is a vector space over a field IF of characteristic 0, taken
here as C, IRl. or Q, on which there is both an algebra and a coalgebra struc-
ture related by some compatibility conditions. To emphasize the duality
between algebras and coalgebras, we shall start by describing the former in
terms of arrows and diagrams. Thus, a unital associative algebra is a triple
(A, m, u), where Ais a vector space over IF, m: A ® A- A and u: IF- A
are IF-linear maps such that the following diagrams commute:

(1.25)

In this Section ® always means the algebraic tensor product. Commutativity


of this diagram gives the usual associativity of the algebra product. Also:

(1.26)
l.B Hopf algebras and Tannaka-Krein duality 35

The unnamed arrows denote the natural identifications A ® IF = A = IF ® A


given by a ® .\ - .\a and .\ ® a- .\a. The diagrams (1.26) provide the unit
lA := u(l!F) for the multiplication m. The commutativity of an algebra can
be expressed by the commutativity of the diagram

(1.27)

where a is the flip operator a ® b b ® a. The commutativity of the


following diagrams:

A c/J A'

~~·
IF

describes a unital homomorphism cf.>.


Definition 1.22. The prefix "co" stands for reversing arrows in the dia-
grams. A coalgebra, then, is a triple (C, Ll, E), where C is a vector space
over IF, Ll: C- C ® C and E: C- IF are IF-linear maps satisfying the reverse
of (1.25) and (1.26):
~®idc
C®C®C--C®C

idc ®~ 1 ~ 1~ (1.28)

C®C C

and
E®idc
IF®C--C®C

C
t idc ~~
C,
(1.29)

The maps Ll and E are called the coproduct and the counit respectively, and
the property described by diagram (1.27) is coassociativity. Furthermore,
we say the coalgebra is cocommutative if the opposite diagram to (1.27)
commutes:

c.
36 1. Noncommutative Topology: Spaces

The tensor product of two unital algebras is again a unital algebra, where
1A®A' = 1A ® 1A' and the product is given on simple tensors by
mA®A'((a ® a') ® (b ® b')) :=ab® a'b'.

In terms of arrows, mA®A' and UA®A' are given by

A®A' ®A®A' idA®u®idA' A®A®A' ®A' ~ A®A', (1.30)

and IF- IF ® IF ~ A ® A'. Similarly, the tensor product of two coalgebras


is a coalgebra, where the coproduct ~C®C' is obtained by reversing (1.30):

C ® C' ~ C ® C ® C' ® C' idc ®u®idc' C ® C' ® C ® C'.

More concretely, if ~(c) = Li c; ® C:' and ~' (d) = LJ dj ® dj', then

~C®C (c ® d) = L. c; ® dj ® c;' ® dj', (1.31)


i,j

and the counit Ec®c is given by C ® C' ~ IF ® IF- IF; that is to say,
EC®C' (C ® d) := E(C)E' (d).

~ We now consider the situation where a vector space has both an algebra
and a coalgebra structure, with the obvious compatibility requirement.
Definition 1.23. Abialgebra is a quintet (B, m, u, ~. E), where (B, m, u) is
a unital algebra and (B, ~. E) is a counital coalgebra, suchthat the maps ~
and E are also unital algebra homomorphisms.
Exercise 1.24. Show that the compatibility condition is equivalent to m
and u being counital coalgebra morphisms, where such a morphism is an
IF-linear map .e : C - C' making the following diagrams commute:

c f C'

\:I IF.
0

A subbialgebra of B is a vector subspace D that is both a subalgebra


and a subcoalgebra; in other words, D, tagether with the restrictions of the
product, coproduct and so on, is also a bialgebra and the inclusion D ...... B
is a morphism of bialgebras.
Example 1.2. For q * 0, consider the algebra of polynomials in two vari-
ables x,y, with the condition xy = qyx. lt possesses a bialgebra struc-
ture, Setting ~X:= X® X, ~y := y ® 1 +X® y, E(X) := 1, E(y) := 0. This
is the "quantum plane" [326].
l.B Hopf algebras and Tannaka-Krein duality 37

Example 1.3. The tensor algebra 'I' (V) of a vector space V is a cocornrnu-
tative bialgebra, where the coproduct and counit are defined on v E V by

ß(V) :=V® 1 + 1 ®V, E(V) := 0. (1.32)

As given, ß: V- '!'(V)® '!'(V) is an IF-linear map, which, by the universal


property of tensor algebras, extends to an unital algebra homomorphism
ß: 'J'(V)- T'(V)®'J'(V);inparticular,ß(l) = 1®l.Now, (ß®id)oßand
(id ®ß) o ß are two unital algebra homomorphisms from 'I' (V) to 'I' (V) ®
'!'(V) ®'!'(V) that agree on V, giving v ,_ v ® 1 ® 1 + 1 ®v ® 1 + 1 ® 1 ®V.
By the uniqueness of extensions, (1.28) holds. The counit property (1.29)
likewise follows from
(E ® id)(ß(V)) = E(V) 1 +V= V= V+ E(V) 1 = (id ®E)(ß(V)).

Thecocornrnutativitycomesfroma(ß(v)) = v®1+1®v = ß(v)forv E V.


Since Eis an algebra homomorphism, it follows from (1.32) that E( v 1 ® · · · ®
Vn) = 0 for all n 2 1. Similarly, ß being an algebra homomorphism, one can
inductively write an explicit formula for ß(v 1 ® · · · ® Vn); see [278, 111.2.4],
for instance.
Example 1.4. The universal enveloping algebra 'U(g) of a üe algebra g is
the quotient of the tensor algebra 'I' (g) by the two sided ideal I generated
by the elements XY- YX- [X, Y], for all X, Y E g. Since ß is an algebra
homomorphism, (1.32) yields
ß(XY) = ß(X)ß(Y) = XY ® 1 +X® Y + Y ®X+ 1 ® XY,
and thus
ß(XY- YX- [X, Y]) = (XY- YX- [X, Y]) ® 1 + 1 ® (XY- YX- [X, Y]),

so ß(l) s;; I® H + H ®I. Clearly, E(J) = 0, too. These two conditions mean
that I is also a coideal in T' (g), and the quotient 'U(g) thus becomes a
bialgebra, which is clearly cocornrnutative. The cornerstone of the theory
of enveloping algebras is the Poincare-Birkhoff-Witt theorem [138, §2.1]
which yields a basis for the vector space 'U(g) in terms of a basis of g: if
X 1 , ..• , Xn is a basis of g, then the (ordered) products xr
1 X? ... x;k, with

ri E ~.form a basis for 'U(g). Naturally, 'U(g) becomes a üe algebraunder


the cornrnutator bracket, and there is an injective homomorphism of üe
algebras j: g- 'U(g) with the following universal property: if Ais an unital
associative algebra and if tjJ: g - A is a üe algebra homomorphism, then
there is a unique algebra homomorphism 'f: 'l1 (g) - A such that 'f ( 1) = 1
and tjJ = 'f oj. Moreover, if cf>: g - g' is a üe-algebra homomorphism there is
a unique unital algebra homomorphism 'l1 ( cf>): 'l1 (g) - 'U (g') lifting cf>. All
this follows trivially from the Poincare-Birkhoff-Witt theorem; in this way,
'U becomes a functor from the category of üe algebras into the category of
unital associative algebras.
38 1. Noncommutative Topology: Spaces

Any vector space V can be regarded as a Ue algebra with the trivial Lie
bracket, [u, v] := 0 for all u, v E V. In this case I is generated by the com-
mutators uv- vu, and the resulting enveloping algebra is commutative: it
is the symmetric algebra S(V) of the vector space V.
Definition 1.24. We pointout that there is a Ue algebra inside any bialge-
bra B. Indeed, bEB is called a primitive element if ~(b) = b ® 1 + 1 ® b.
If c is also primitive, then
~(bc) = bc ® 1 + b ® c + c ® b + 1 ® bc,
and therefore bc-cb is also primitive; the set P(B) of allprimitive elements
of Bis thus a Ue algebra. In addition, (1.29) shows that E (b) = 0 necessarily,
for b primitive.
Lemma 1.21. The set ofprimitive elements of'U(g) is g itself.
Proof. We claim that P('U(g)) isjust the embedded copy of g. Firstnote that
any element of 'U(g) can be written as a linear combination of powers xn,
withX e g, n E N. Forinstance, XY =~(X+ Y) 2 - ~X 2 - ~Y 2 +~[X, Y]; and
there is an analogaus expansion for any element of a PBW basis. If vn (g)
derrotes the subspace generated by { xn: XE g}, then 'U(g) = EB~=Ü vn(g)
is a graded bialgebra: since ~ (Xn) = Lk=O (~) xk ® xn-k. the coproduct is
compatible with this grading. If n ~ 2 and an element p = L,1 c1xj E vn (g)
is primitive, then the terms qk := Lj Cj (~)xJ ® x;-k of bidegree (k, n- k)
in ~(p) must vanish for 0 < k < n; but then (~) p = m(qk) = 0 in vn(g),
and (~) p * 0 since IF has characteristic zero. Thus the primitive elements
are only those in 'U 1 (g) ""g. D
Definition 1.25. Given an algebra A and a coalgebra Cover IF, we can define
the convolution of two elements f,g of the vector space of IF-linear maps
*
Hom(C,A), as the map f g E Hom(C,A) given by the composition

C~C®C~A®A~A.
Proposition 1.22. If A is an algebra and C a coalgebra, then the triple
(Hom(C,A), *• u o E) is a unital algebra.
Proof. By (1.25) and (1.28), the following diagram commutes:
<f*B)®h
C®C A®A

;/ ~ j®g®h ~ ~ A.
c C®C®C A®A®A

~C®C~ j®(B*h)
iÄ A®A
/m
l.B Hopf algebras and Tannaka-Krein duality 39

Since (j * g) * h corresponds to the rnap frorn C to A along the top edge of


the diagrarn, whereas f * (g * h) is the rnap frorn C to A along the bottarn
edge, the convolution is associative.
Moreover, if ß(c) = 2. 1 cj ® c'/ then c = (E o id) (2. 1 cj ® cj') = 2.1 E(cj )cj'
by (1.29). Then, since E is linear,

(u o E) * j(c) = m o [(u o E) ® f](2. 1 cj ® cj')


=m(2.1 (u o E)(cj) ® j(cj'))
=I E(cj)f(cj') = !(2.1 E(cj)cj') = j(c).
j

Similarly one checksthat f * (u o E) = f. 0

Lemma 1.23. If -1!: C - C' is a counital coalgebra morphism then -I!* : g -


g o -1! : Horn( C', A) - Horn( C, A) is a unital algebra morphism. Similarly, if
c/>: A - A' is a unital algebra morphism, then cjJ * : f - cjJ o f : Horn (C, A) -
Horn(C,A') isanother unital algebra morphism.
Proof. If g,h E Horn(C',A), then

-1!* (g * h) = m o (g ® h) o ß' o-1! = m o (g ® h) o (-/! ® -/!) o ß


=m o (-I!* g ® -1!* h) o ß = -1!* g * -1!* h.

Moreover, -/!* (u o E') = u o E' o -1! = u o E, so -1!* is unital. The proofthat cjJ*
is hornornorphic and unital is similar. 0

Definition 1.26. A Hopf algebra is a bialgebra H tagether with a (necessar-


ily unique) convolution inverse S for the identity rnap idH. The rnap S is
usually called the antipode of H. The property idH *S = S idH = u o E *
boils down to the commutativity of this diagrarn:

H®H~H~H®H
id®S! ~ UoE !S®id (1.33)

H®H~H~H®H.

In particular, if ß(a) = 2.1 aj ® a'j, then


E(a)lH =u o E(a) = m o (id ®S) o ß(a) = 2.1 ajS(a'j). (1.34)

andlikewise E(a)lH = 2.1 S(aj)a'j.


A morphism of Hopf algebras is a linear rnap -1!: H - H' that is both a
unital algebra hornornorphisrn and a counital coalgebra hornornorphisrn,
and satisfies the cornpatibility condition -1! o S = S' o -1!. Actually, the corn-
patibility condition is autornatic, as the following argurnent shows [236,
Prop. 37.1.10]; seealso [446].
40 1. Noncommutative Topology: Spaces

Proposition 1.24. LetHand H' be Hopf algebras and f: H- H' a bialge-


bra morphism; then -e o S = S' o -e.

Proof. By Lemma 1.23,

u o E = f* (idw *S') = f* (idw) * f* S' = -e * (S' o f),


u' o E' = f*(S * idH) = -e*s * f*OdH) = (f o S) * -e.
Associativity of the convolution then gives

Example 1.5. If G is a group, IFG denotes the group algebra of G over IF,
that is, the vector space over IF with G as basis; its product is defined by
extending linearly the group multiplication of G, so the unit in IFG is the
identity element of G.
Proposition 1.25. The group algebra IFG is a cocommutative Hopf algebra
where the coproduct, counit and antipode are the respective linear extensions
of the diagonal rnap ß(x) := x ® x, the constant rnap E(x) := lJF, and the
inverse map S(x) := x- 1 defined on G.

Proof. Let H = IFG; by linearity, it is enough to verify (1.28), (1.29) and (1.33)
on elements x E G. Coassociativity comes from the associativity of the
tensor product:

(ß ® id) o ß(x) = (x ® x) ® x = x ® (x ® x) = (id ®ß) o ß(x).

Cocommutativity is obvious. The counit property follows from

(E ® idH) o ß(X) = E(X)X =X= XE(X) = (idH ®E) o ß(X).

For the antipode,

m o (S ®id) o ß(x) = m(x- 1 ®x) = lH = m(x ® x- 1 ) = m o (id ®S) o ß(x),

and moreover, (u o E)(x) = u(liF) = lH, so (1.33) holds. 0

Conversely, an element a of any Hopf algebra H is called group-like if


Ll(a) = a ® a. The set G(H) of group-like elements is a unital semigroup,
since ß is an algebra homomorphism and ß(l) = 1 ® 1. lf a is group-like,
then a = (E ®idH) o Ll(a) = E ®idH(a ®a) = E(a)a, so E(a) = h; therefore,
by (1.34), aS(a) = S(a)a = u o E(a) = lA, so that a is invertible and
S(a) = a- 1 . Since ß(a- 1 ) must be a- 1 ® a- 1 , we conclude that G(H) is a
group, indeed a subgroup of Hx, within which S provides the inversion.
Moreover, it can be shown [446] that all the elements of G(H) are linearly
independent inH. Inparticular, whenH = IFG, G(H) is the original group G.
l.B Hopf algebras and Tannaka-Krein duality 41

Lemma 1.26. The antipode of a Hopf algebra is a unital algebra antihomo-


morphism.

Proof. Consider the maps p, T : H ® H - H given by

p(a ® b) := S(ab) and T(a ® b) := S(b)S(a).

To prove the lemma, it is enough to show that p is a left convolution inverse


and that T is a right convolution inverse for the multiplication m: H ® H -
H. Suppose that ß(a) = Lia; ® a;' and ß(b) = Ljbj ® bj'; since the
coproduct in H ®His given by (1.31), using (1.34) yields

m * T(a ® b) = m o (m ® T) (Li,j a; ® bj ® a;' ® bj')


= Li,j a;bjS(bj')S(a;') =Li a;E(b)S(a;')

= E(a)E(b) 1H = E(ab) 1H = U o EH®H(a ® b).

On the other hand,

p * m(a ® b) = m o (p ® m)(Li,j a; ® bj ® a;' ® bj') = Li,jS(a;bj)a;'bj'.


Now, since ß is an algebra homomorphism,

Using once more that Eis an algebra homomorphism and (1.34), we get

u o EH®H(a ® b) = u o E(ab) = m o (S ® idH) o ß(ab) = :Ls<a;bj)a;'bj'.


ij

Thus, p * m(a ® b) = u o EH®H(a ® b), as claimed. 0

As may be anticipated, S is also a counital coalgebra antihomomorphism,


by a similar argument [2 78]. We notealso that S 2 = idH when His commu-
tative or cocommutative. Many examples of Hopf algebras which are nei-
ther commutative nor cocommutative, with S 2 * idH in general, are found
in the Iiterature on quantum groups; the books [77], [278] and [324] give
comprehensive accounts of such algebras.
We now return to the tensor algebra example. Consider the IF-linear map
S: V- 'l(V)o given by S(v) = -v, where 'l(V)o denotes 'l(V) with the
reversed product: (w ®v) o : = v ®w. By the universal property of the tensor
algebras, S extends to a Unitalalgebra homomorphism S: 'l(V)- 'l(V) 0
,

Now, u o E, m o (S ® id) o ß and m o (id ®S) o ß are three unital algebra


homomorphisms from 'l(V) to itself that agree on V, since

mo (S®id) oß(v) = m(-v ® 1 + 1 ®V)= 0


= m(v ® 1 + 1 ® (-v)) = m o (id®S) o ß(v),
42 1. Noncommutative Topology: Spaces

and u o E(v) = u(O) = 0. By the uniqueness of the extensions, (1.33) holds,


and r (V) is a Hopf algebra. Since S is an algebra antihomomorphism, it
follows that S(vi ® · · · ® Vn) = (-l)nvn ®···®VI.
If g is a Lie algebra, the bialgebra ideal I of r (g) generated by the ele-
ments XY- YX- [X, Y] isS-invariant, since

S(XY- YX- [X, Y]) = YX- XY +[X, Y] = -(XY- YX- [X, Y]).

It follows that 11 (g) = T (g) I I is also a Hopf algebra .


.,. We shall now consider commutative Hopf algebras in more detail. We
start with a compact group G. One would like to reconstruct G from the
algebra of real-valued continuous functions C(G, ~). butthiswill not work,
since the algebra C(G x G, ~) is much larger than C(G, ~) ®C(G, ~). There-
fore, weshall use the smaller algebra X(G) of (real-valued) representative
functions, i.e., those functions f: G - ~ whose right translates Rxf: y ....
j(yx) generate a finite-dimensional subspace of C(G, ~). If x .... [aiJ(x)]
is a matrix representation of G, then each aiJ is a representative function.
There isanatural algebra homomorphism rr: X(G) ® X(G)- X(G x G):

rr(j ® g)(x,y) := j(x)g(y).

Lemma 1.27. rr is an algebra isomorphism.

Proof. Let F = LJ=I fJ ®Bi in X(G) ® X(G) besuchthat rr(F) = 0. Then


BI •... ,gn generates a finite-dimensional subspace of X(G); if ki, ... , kr
is a basis for it, we can find elements YI, ... , Yr of G such that kdYJ) =
8iJ. Then F = Li hi ® ki for some hi, ... , hr E X(G), so that hi(x) =
Li hi(x)kdYJ) = rr(F)(x,y1 ) = 0, and consequently F = 0. Thus, rr is
injective.
On the other hand, Iet FE X(G x G) and consider the function Fy: x ....
F(x,y). Since (RzFy)(x) = F(xz,y) = (R(z,uF)(x,y), the subspace ge-
nerated by all right translates of Fy is finite-dimensional, so Fy E X(G).
Similarly, the function px: y .... F(x, y) lies in R.(G). If ki, ... , kr is a basis
for the subspace of X(G) generated by the translates { RzFx : z E G}, we
can write px = LJ hi(x)ki for some hi(x) E ~. Choosing YI .... •Yr E G
as before, suchthat ki(y1 ) = 8iJ• we obtain hdx) = F(x,y1 ) = Fyi(x),
so that hi E R.(G). Therefore F = Li rr(hi ® kd; we conclude that rr is
surjective. 0

The Iemma allows us to translate the group structure of G into a Hopf


algebra structure on X(G). The counit is evaluation at the identity E(j) :=
j(l) and the antipode is given by Sj(x) := j(x-I ). The group product
(x,y) .... xy induces the map Cm: X(G) - R.(G x G), which yields the
coproduct ß := rr-I o Cm: X(G) - X(G) ® R.(G). 1t is easy to checkthat
(R.(G), ·, l,ß,E,S) is a commutative Hopf algebra, whose product · is the
l.B Hopf algebras and Tannaka-Krein duality 43

usual pointwise product. For a finite group, it is also easy to check that
R(G) is the dual bialgebra of the group algebra IFG.
Conversely, if (H, m, u, ß, E, S) isareal Hopf algebra, the set (j(H) of all
algebra homomorphisms cJ>: H - ~ is a group, under the convolution

(1.35)

where we use the identification ~ ® ~ "" ~. The counit E is the identity of


this group. The inverse of cJ> is cf> o S, since

cJ> • (cj> o S) = (cj> ® (cj> o S)) o ß = cJ> o m o (idH ®S) o ß = cJ> o u o E = E.

(j(H) becomes a topological group when equipped with the weakest topo-
logy that makes the evaluation maps Ef: cJ> ...... cJ> (j) continuous.
The compact group G carries a normalized Haar measure. Thus R (G) has
an extra structure, namely, the linear map ]: R(G) - ~: f . . . fc j(x) dx
that is invariant und er translations (i.e., ] o Rx = ] for each x), and satisfies
J (j2) > 0 whenever f * 0.
Definition 1.27. A commutative skewgroup is a commutative Hopf alge-
bra H tagether with a linear map ] : H - ~. called a Haar functional, such
that (] ® idH) o ß = u o ], and J(a 2) > 0 if a =1= 0.
When H = R(G), u(J(j)) is the constant function on G with value
](j). Furthermore, if Ryj(x) = j(xy) = LJ BJ(x)hJ(y), it follows that
](Ryf) = L. 1 ](BJ) hJ(Y) andß(j) = L. 1 BJ®hJ. therefore (J®id)(ß(j)) =
LJ ](BJ) h 1 , whose value at y is ](Ryj). The condition (] ® id) o ß = u o]
is thus a reformulation of (right) translation invariance. (The analogaus
left-invariance condition would be (id ®]) o ß = u o ].)
In this way, to every compact group G we associate the commutative
skewgroup R(G), and, conversely, to each commutative skewgroup H we
associate the topological group (j(H). Weshallshow that the functors R
and (j areinverse to each other. For that, we first establish that (j(R(G))
is a compact group.
Lemma 1.28. If (H, m, u, ß, E, S,]) is a real commutative skewgroup, then

(a) the formula (b I c} : = ] (bc) defines a scalar product on the real vector
space H;

(b) for each cJ> E (j(H),] =] o (idH ®cj>) o ß;

(c) any cf> E (j(H) can be written as cJ> = Eo (idH ®cj>) o ß.

Proof. The assertion (a) follows immediately from the properties of ], but
we should point out that the commutativity of H is used. Property (b)
follows from the translation invariance of ], since if a E H and ß(a) =
44 1. Noncommutative Topology: Spaces

] o (id ®c/1) o l1(a) = L.1 ](aj) cp(aj) = cp(L. 1 ](aj) aj)


= cp o (] ® id) o l1(a) = cp o u o ](a) = ](a).

Property (c) follows from

E o (id ®c/1) o l1(a) = L.1 E(aj) cp(aj) = cp o (E ® id) o l1(a) = cp(a),

using the counit relation (1.29). 0

Proposition 1.29. lf (H, m, u, l1, E, S,]) is a real commutative skewgroup,


then the group §(H) is compact.

Proof. Foreach a EH there are a~, ... ,a~,a~, ... ,a~ suchthat l1(a) =
,n ' " ' are ort h onormal Wlt
L.J=l a 1 ® a 1 ; we can assume t h at a '1 , ... ,an · h respect
to the scalar product ofLemma 1.28. Since (id ®c/1) (l1(a)) = L.j= 1 ajcfJ(aj ),
the subspace of H generated by { (id ®c/1) (l1(a)) : cp E §(H)} is finite
dimensional, with basis a~, ... , a~.
Consider the subsets Sa := { cp(a) : cp E §(H)} c ~ for a E H. By
Lemma 1.28(c), cp(a) = L. E(aj)cp(aj). Introduce the linear map F: ~n -
~ : z ,..... 2: 1 E(aj)zl, and observe that Sa is included in the compact set
F(§a), where §a is the sphere of radius JJT(i'2). Indeed, by Lemma 1.28(b),

}(a 2 ) =Jo (idH®c/J) ol1(a 2 ) =J((LjajcfJ(aj)) 2 )


= Li,J(a; cp(a;') I aj cp(aj)} = L.1 cp(aj') 2 .
The Cartesian product T := naEH F(§a) is compact, by Tikhonov's the-
orem [383), and the map j: §(H) - T: cp ...... {cp(a)}aEH is an injective
homeomorphism. To conclude that §(H) is compact, it remains only to
prove that j(§(H)) is closed in T. Take any net {t,\} c j(§(H)) with t" - t
in T. Then t"(a)- t(a) E F(§a) for each a, and t,\(a) = cp"(a) for some
cp" E §(H). lntroduce cp: H- ~: a,..... t(a); it is clear that cp is an algebra
homomorphism, that cp = lim" c/1.\, and that j(cp) = t. Therefore, j(§(H))
is closed. 0

Theorem 1.30. Let (H,m,u,l1,E,S,J) be a real commutative skewgroup.


Then the map E: H - 'R(§(H)) defined by Ea (cp) := cp(a), for a EH and
cp E § (H), is an isomorphism of Hopf algebras.

Proof. lt is easy to check that E is an injective algebra homomorphism. If


cp, f./J E §(H), then

l1(Ea)(c/J ® f./J) = Ea(c/1 · f./J) = (cp ® f./J) o l1(a) = L.1 cp(aj) f./J(aj)
= 'L. 1· Ea'J c/1 Ea"J f./J = 'L. 1· (Ea'J ® Ea'')
J
( c/1 ® f./J)
= [(E ® E) o l1(a) )(cp ® f./J),
l.B Hopf algebras and Tannaka-Krein duality 45

so ß o E = (E ® E) o ß. Moreover, since the unit in (j(H) is the counit E of H,


then E (Ea) = Ea ( E) = E (a), so that E o E = E; thus, E is also a coalgebra
homomorphism.
To see that the image E(H) is R((j(H)), we notice that the algebra E(H)
is unital and separates points of the compact group (j(H); by the Stone-
Weierstrass theorem, E(H) is densein C((j(H),IR{). More precisely, E(H) is
densein R((j(H)) in the norm topology of C((j(H), IR{). Furthermore, E(H)
is a (j(H)-submodule of R((j(H)) under the action

cJ> • Ea := E o (id®c/>) o ß(a) = I. 1 c/>(aj)Eaj·


Indeed, if ß(aj) = Ii b;1 ® b;j, we obtain, using (1.35),

c/>1 · ( c/>2 · Ea) = c/>1 · I.j c/>2 (aj) Eaj = Iij c/>2 (aj') c/>1 (b;j) Eb;i
= E o (id ®c/>1 ® c/>2) o (ß ® id) o ß(a)
= E o (id ®c/> 1 ® c/> 2 ) o (id ®ß) o ß(a)
= E o (id ®(c/>1 · c/>z)) o ß(a) = (c/>1 · c/>z) · Ea.
Since all (j(H)-submodules of R((j(H)) are closed in the norm topology-
see, for instance, [54, Prop. III.l.4]- we conclude that E(H) = R((j(H) ). D

Theorem 1.31. Let G be a compact Lie group. Then the evaluation map
e: G- (j(R(G)) defined byex(j) := j(x), for x E G and f E R(G), is an
isomorphism of compact groups.

Proof. Let x,y E G and f E R(G). If ß(j) = I. 1 fj ® fj' in R(G) ® R(G),


then

exy(J) = j(xy) = ß(j)(x ® y) = I. 1 Jj(x)jj'(y).


= I.1 ex(Jj)ey(Jj') = (ex ® ey) o ß(j) = (ex · ey)(j).

Thus, e is a group homomorphism. The Peter-Weyl theorem for compact


groups [54, Thm. III.3.1] implies that R(G) is densein C(G,IR{); therefore e
is injective. The continuity of e follows from that of each j: x ..... ex(j).
Let us abbreviate (j := (j(R(G)). The map E: R(G) - R((j) defined in
Theorem 1.30 is a right inverse for et: R((j) - R(G): F ..... F o e. Indeed, if
jE R(G) and XE G, then

et o Ej(X) = Ej(ex) = ex(j) = j(x).


Now E is an isomorphism, by Theorem 1.30, so et is also an isomorphism.
Since C((j, IR{) and C(G, IR{) are the respective norm-completions of R((j)
and R(G), we conclude that et: C((j, IR{) - C(G,IR{) is an isomorphism. In
particular, e: G- (j is surjective. D
46 1. Noncommutative Topology: Spaces

Theorems 1.30 and 1.31, taken together, are usually called Tannaka-
Krefn duality. In [241], one can find an exhaustive discussion that relates the
original works of Tannaka and Krein with more modern treatments. Here
we have followed Hochschild in [248]. Although the result runs in parallel
with (a real version of) the Gelfand-Naimark theorem, there are interesting
differences. The homomorphisms R.(G) - <C, for complex representative
functions, do not all extend to *-homomorphisms from C(G, <() to <C; in
the Lie group case, they form a complex group, whose Lie algebra is the
complexification of the Lie algebra of G.
The requirement that there exist a Haar functional, which is only used to
prove compactness of (i(H), seems suspect, for it might be redundant; after
all, existence of a Haar measure is automatic for compact groups. Certainly
in the Lie group case, it may be dispensed with (see the treatment in [54]).
In the category of "compact quantum groups", which are C*-completions
of certain Hopf algebras, Woronowicz has indeed established the existence
of a normalized Haar functional [495].
Another approach is the theory of Hopf C* -algebras developed by Vaes
and Van Daele [463]. In this category, the objects are (not necessarily unital)
C*-algebras, the commutative example being C0 (G) for any locally compact
group G. The coproduct is a morphism, in the sense of Woronowicz, from A
to 5\1(A ® A); the counit and antipode are only densely defined, in general.
The domain of the multiplication map is taken to be the Haagerup tensor
product A®hA, regarded as a dense subalgebra of A®A, on which continuity
of m is guaranteed.
The C* -algebraic approach can be pushed through to yield a full non-
commutative Pontryagin duality, recently worked out by Kustermans and
Vaes [300, 301]. Along the way, it provides a way to relate Haar function-
als l1 and lr which are respectively left and right invariant, whenever their
existence (and faithfulness) can be assured. These are densely defined pos-
itive functionals on a Hopf C* -algebra A, and one can construct a positive
group-like element a affiliated with A suchthat lr = 11 (a 112 ( · )a 1i 2 ); this
plays the role of the modular function of locally compact groups. Finally,
Connes and Moscovici [116] have extended this notion tothat of a modular
pair (8, a) for a Hopf algebra H, where 8 is a character of Hand a EH is
a grouplike elementsuchthat 8(a) = 1: see Section 14.7.
~ Hopf algebras play an important role in the description of symmetry by
algebraic means. To see why this is natural, consider a group G acting on a
space X. This automatically gives a homomorphism G - Aut(F(X) ), where
F(X) is a suitable algebra of (IF-valued) functions, such as the Gelfand-
Naimark algebra. In the presence of symmetry, as pointed out in [237], it is
a very good idea to pass to an algebraic description of X; for instance, if X =
G, the action is indecomposable and so contains essentially no information,
whereas the linear action of Gon F(G) is generally decomposable (think of
G = § 1 , for instance). This is the whole point of harmonic analysis. Now,
l.B Hopf algebras and Tannaka-Krein duality 47

more generally, the homomorphism G- Aut(F(X)) can be described by a


linear mapping IFG ®F (X) - F (X), with the added property that group-like
elements in IFG give rise to automorphisms.
Definition 1.28. Let H be a Hopf algebra. A (left) Hopf H-module algebra
Ais an algebra which is a (left) module for the algebra H such that h · lA =
E(h) lA and

h ·(ab)= 2: 1-(h'.J · a)(h':


J
· b) (1.36)

whenever a, b E A and h EH with ß(h) = LJ hj ® h'j.


In other words, if A ®Ais given the natural H-module structure, then m
and u are module maps. In the case H = IFG, where ß(g) = g ® g, the defi-
nition implies that g ·(ab) = (g · a)(g · b), i.e., H acts by endomorphisms.
So any action of Gon X gives a Hopf action of IFG on the algebra F(X), and
vice versa.
If h is a primitive element of H, the definition entails that h · lA = 0
and that h · (ab) = (h · a)b + a(h · b): primitive elements act by deriva-
tions. Therefore, (1.36) may be regarded as a generalized Leibniz rule. lt is
consistent with the algebra structure of H, since

h · (k ·(ab))= Li,J(h;kj · a) (h;'k'j· b) = (hk) ·(ab).

Definition 1.29. Let H be a Hopf algebra and A a (left) Hopf H-module


algebra. The smash product algebra A # H is the vector space A ® H with
the product
(a ® h)(b ® k) := LJ a(hj · b) ® hj'k.
Usually one writes a # h for a ® h in this context.
In the case A = F(X) and H = IFG, this can be written as

(1.37)

As indicated in [120], the smash product is dual to the principal bundle


construction.
This suggests that Hopf algebras provide a natural setting for study-
ing generalized symmetries, and an organizing tool in noncommutative
geometrical contexts. The last chapter of this book turns araund a very
important example of this. Moreover, Hopf algebras of generalized sym-
metries of finite noncommutative spaces have been constructed in [477].
There is no shortage of competition, however. For instance, groupoids and
hypergroups have also proved useful in understanding generalized sym-
metries [485, 486]. There are deep unexplored Connections among these
different approaches.
2
Noncommutative Topology:
Vector Bundles

We continue with our study of the duality between spaces and algebras
by considering modules over these algebras. As we shall shortly see, C*-
algebras have a supply of C*-modules on which they naturally act. In view
of the Gelfand-Naimark theorem, one should ask whether a compact space
M has topological partners corresponding to modules over C(M); under
suitable conditions, thesepartnersturn outtobe the vector bundles over M.
The heart of this chapter is thus the Serre-Swan theorem, establishing that
the categories of finitely generated projective modules over C(M) and the
category of vector bundles over M are equivalent.
We do not assume that the reader is thoroughly farniliar with the theory
of these bundles, and we begin by surveying it, in a form appropriate for
our purposes. The necessary purely algebraic notions relative to projective
modules over unital rings are given in Section 2.A, as an appendix to the
chapter.

2.1 Vector bundles


Definition 2.1. We recall that a (complex, unless indicated otherwise) vec-
tor bundle E .2:.. M is a fibration on a topological space M such that, for
each x E M, Ex := rr- 1 (x) is a (complex) vector space and there exist an
open neighborhood Wx of x, an integer r = rx and r mappings Si : Wx - E
suchthat rr o Si = id on Wx and suchthat the map
50 2. Noncommutative Topology:Vector Bundles

is a homeomorphism of Wx X er onto rr- 1 (Wx).


In other words, a vector bundle over M is a locally trivial continuous
family of vector spaces, indexed by M. The maps sJ are called local sections
and the package s := (s 1 , ... , Sr) is called a local frame over Wx. Clearly,
x .... rx is locally constant; if M is connected, rx is actually constant and
there is a "typical fibre" V suchthat Ex == V, whose dimension is called the
rank of the vector bundle. When the base space M is fixed, we often write
just E, to refer to a vector bundle E- M.
Partitions of unity are the essential tool in vector bundle theory; there-
fore we shall assume paracompactness of M at the outset. Any (Haus-
dorff) locally compact and second countable space (for instance, a finite-
dimensional manifold) is paracompact.
Morphisms in the category of vector bundles over a fixed M are continu-
ous bundle maps T: E - E' satisfying rr' o T = rr and so defining fibrewise
maps T x: Ex - E~ for each x E M, which are required tobe linear.
Definition 2.2. Let E, E' be two vector bundles over M. A vector bundle
equivalence between them is an invertible vector bundle map, which is
given by a homeomorphism T: E - E' suchthat each T x: Ex - E~ isalinear
isomorphism. We shall denote by [E] the equivalence dass of the vector
bundle E- M; for the set of equivalence classes with typical fibre V== er,
we write Vectr(M). A vector bundle E - M is trivial if it is equivalent to the
product bundle M x V~ M. We shall write 0 for the trivial vector bundle
Mx {0} of rank zero.

Let then E- M be a vector bundle with typical fibre V, and let {UJ}JEJ
be a trivializing open covering of M, so that each u1 carries a local frame
s J = (s J1, ... , sJr) of continuous sections over UJ that are pointwise linearly
independent. The family of continuous functions BiJ: Ui n u1 - GL(V),
defined whenever Ui and u1 overlap, suchthat Si = BiJ · s1 on Ui n u1 ,
satisfies the consistency conditions

(2.1)

That is to say, the transition functions g = {BiJ} for the vector bundle E - M
determine a Cech 1-cocycle with values in GL(V). The cocycle is trivial if
and only if BiJ = Jd1- 1 , for some family of functions fi: Ui - GL(V).
Therefore, isomorphism classes of vector bundles are determined by the
Cech cohomology classes of the transition functions; see, for instance, [2 71,
Chap. 1] for more detail. The upshat isthat Vectr(M) == H 1 (M,GL(r,e)).
In particular,

(2.2)

A more concrete handle on Vectr (M) will be obtained later.


2.1 Vector bundles 51

The set ofvector bundles over M admits operationssuch as duality, Whit-


ney sum and tensor product; the last two make it a semiring. A vector bun-
dle E' - M is a subbundle of a vector bundle E - M if E' ~ E and if
E~ = E' n Ex for each x E M; in particular, the rank function x .... dimEx
is locally constant, and the inclusion i: E' - E is an injective bundle map.
11> The theory rests on the following two chief propositions.
Proposition 2.1. IfM is paracompact, any exact sequence ofvector bundles
over M:

splits. In particular, the canonical sequence

0-E' ....!...E.!!...EjE' -0,

for a subbundle E' of E, splits.

Proof. At each point x E M, the short exact sequence E~ .!2. Ex!!.; E~ of


vector spaces splits (essentially because dimker ßx + dimimßx = dimEx);
moreover, if s is the rank of E and r = rankE", there is an open neigh-
bourhood Ux of x in which the bundle map ß is represented by a matrix
function b: Ux- Hom(CS,F) with rankb(x) = r. Since the set of maxi-
mal rank operators is open in Horn (( 5 , er), there is an open neighbourhood
v1 of x with v1 c Ux where rank b (y) = r for all y E v1, so that over v1
the sequence of bundles E' ..... E - E" splits. We may suppose that the sets
v1 form a locally finite covering of M. Denote by YJ a right inverse of ß on
v1 and choose a partition of unity {1./J J} subordinate to the covering {V1}.
Set y := LJ I.JJJYJ; we remark that rank y = rankßy = r, that is, y also has
constant rank r. This yields a right inverse for ß -and thus E "" E' e E".
(Note that, by a very similar argument, the bundle map oc also splits.) 0
Lest the reader be misled, we point out that if T: E - E' is a bundle map,
the set l.t!x ker Tx is not always a subbundle of E, the reason being that
x .... rank( Tx) is not locally constant in general. If we want to construct
a short exact sequence out of a bundle map, we need to make sure that
the rank is constant; this happens, for instance, if the bundle map is one-
to-one or onto. Weshall write kerT := l.t!xEMkerTx only if this disjoint
union forms a subbundle of E; then, on account of the splitting, im T :=
l.t!xEM im Tx "" EI ker T is also a subbundle of E, and vice versa.
Let us agree to call f(U,E) the set of sections over U ~ M, i.e., of con-
tinuous maps s: U - E such that rr o s is the inclusion of U in M (already
used in Definition 2.1). For the proof of the next proposition, we need to
assume that, given a vector bundle E over M, there isafinite open cover
{u1 } ~ 1 of M such that each EI u1 is trivial. This is obviously true for com-
pact spaces, and is also true for manifolds. A proof of the manifold case
is given in [82, Thm. 7.5.16], using results of [256]. (lf the manifold has
52 2. Noncommutative Topology:Vector Bundles

dimension nl the cardinality of the coverl which is not terribly important


herel can be taken to be equal to n + 1.)
Proposition 2.2. Let E" be a vector bundle over a paracompact space M
with the finite open cover property. Then we can find another vector bundle
E' - M such that the Whitney sum E" E!1 E' - M is a trivial vector bundle.
Proof. Let { U1, .. Um} be a finite open covering of M by chartssuchthat
'I

there are r = rankE" linearly independent local sections sJ1 sJr in each
1 ••• 1

f(UJ1 E" ). Let {f./111 ... f./lm} be a continuous partition


I of unity subordinate
to {UJ}j!, 1 • Define Gjk: M-E" as f./IJSJk on UJ and as 0 outside u1 ; notice
that the vectors G'Jk(X) span each fibre E::. Let n := rml and define a map
ß: Mx cn- E" by ß(xl t) := LJ,k tJkG'Jk(x); then ß is a surjective bundle
mapl giving the exact sequence
0-E' ~MX cn J!....E" -01
where E' := ker ß. By Proposition 2.1 this sequence splits yielding E' E~>E" ""
1

MxP. D

... In order to make the promised classification ofvector bundlesllet us now


consider vector bundles over different base spaces. A bundle morphism
from one vector bundle F .!!... N to another E ~ M is a pair of continuous
maps f: N-M and T: F-E suchthat rr o T = f o u from F to M; that is 1

the pair (T,f) makes the following diagram commute:

F~E

~!
N~M.
!rr (2.3)

A basic procedure in the handling of diagrams is "completing the square"


when two sides are given. In the vector bundle categoryl this gives rise to
pullback bundles.
Definition 2.3. Let E ~ M be a vector bundle and let f: N - M be a con-
tinuous map. Write f*E := { (u y) E Ex N: rr(u) = j(y)} and define
1

rr: f*E- N and j: f*E- E by rr(uly) := y and j(uly) := u. Since


f* E = (rr x j) - 1(LlM) I where LlM is the diagonal in Mx MI f* E is closed in
Ex N. Then rr- 1 (y) "" EJ<y) and so f* E..!!... N is a vector bundlel called the
I

pullback bundle of E ~ M by f. Moreoverl (j j) is a bundle morphism;


1

in other wordsl there is a commutative diagram of continuous maps:


2.1 Vector bundles 53

The pullback bundle has the following universal property: any bundle
morphism (2.3) with the same rr and f gives rise to a bundle morphism
(t,idN) with t: F- f*E, suchthat T = j o t, i.e., the following diagram
commutes:

F~
er f*E~E

1ft f !rr
N ____!____,... M.
In other words, any bundle morphism with base map f factors through the
pullback bundle. The recipe for the map t: F - f* E is v .... (O" ( v), T ( v)),
since f (O" ( v)) = rr ( T ( v)). If T is bijective on each fibre, then t is an iso-
morphism.
Exercise 2.1. If g: P - N isanother continuous map, show that (j o g) *E =
g*(j*E). 0

We refer to [257, Thm. 3.4.7] for the proof of the following standard
result.
Proposition 2.3. Let f, g: N - M be homotopic continuous maps. If E is a
vector bundle over M, the pullback bundlesf* E andg* E are isomorphic. 8
Corollary 2.4. IfM and N are homotopy equivalent, then there is a bijection
between Vectr (M) andVectr (N) for allr. In particular, a vector bundle over
a contractible space is trivial. 8

Among the good properties of the pullback construction, we hasten to


mention the following relations.
Exercise 2.2. If E, F are vector bundles over M and f: N - M is continuous,
prove that there are bundle equivalences:

f* (E e F) ~ f* E e f* F and f* (E ® F) ~ f* E ® f* F. 0

We see that there is a cofunctor from the category of (paracompact) to-


pological spaces and continuous maps to the category of semirings, given
by f ~ f*.
Notice that if E.!!... M is trivial, then f* Eis trivial. Also, f* Eis trivial when
f is constant. This indicates that pulling back tends to make a bundle less
twisted. What is more, there is a family of highly twisted complex vector
bundles from which other vector bundles can be obtained by pullbacks;
these are the Grassmanman bundles.
Definition 2.4. The complex Grassmannian Gr(CN) is the set of r-planes
in cN.
54 2. Noncommutative Topology:Vector Bundles

The unitary group U(N) acts transitively on Gr(CN). A unitary matrix


that sends an r-plane to itself will also leave invariant the complementary
orthogonal (N- r)-plane; thus the isotropy subgroup of U(N) at a particu-
lar r-plane is conjugate to U(N- r) x U(r). This gives a diffeomorphism

N U(N)
Gr(<C ) "" U(N- r) x U(r)

As the quotient space of a compact Lie group by a closed subgroup, the


Grassmanman is a smooth compact manifold. lts (real) dimension is then
given by N 2 - (N- r) 2 - r 2 = 2(N- r)r. Note that GI ((Y+I) = c_pr and
that Gr(CN) ""GN-r(CN).

Definition 2.5. Just as is done for projective spaces, we can define a tau-
tological bundle Tt' - Gr ( c_N) of rank r, the fibre at any element A of the
Grassmanman being the r-plane A itself.

A point A E Gr(CN) can be matched with a hermitian N x N matrix


PA suchthat P1 = PA and tr PA = r. To see that, choose an orthonormal
basis {VI, ... , Vr} of A, and let VA be the N x r matrix whose columns
are the vectors VI, ... ' Vr in c_N_ Then v;tvA = lr. while PA := VA v;t is
an N X N matrix for which { PAb : b E c_N} = {VAC : c E (Y} = A,
since VAc = LJ=I CJVJ· lt is clear that P1 = PA, and tr PA = trVAV1 =
trV;tVA = r. Thus, the range of PA isther-plane A, and PA is the unique
hermitian idempotent matrix (i.e., orthogonal projector) with this range.
Now, it is not imperative that PA be hermitian. If we just choose an ordinary
basis {VI, ... , Vr} of A, we can find an r x N matrix X so that X and XVA
have rank r (for instance, place columns from the standard basis of er in
positions corresponding to r linearly independent rows of VA, and fill the
remaining columns of X with zeroes). Thenreplace v;t by WA := (XVA)-I X,
so that WA VA = lr, and settle for QA := VAWA instead of PA. Clearly, QA
is idempotent, { QAb : b E c_N} = {VAc : c E C} = A and tr QA = r as
before.
Suppose that we are given a continuous choice of idempotents A .... QA
(the hermitian idempotents provide one such choice). Let oc denote any r-
element subset of {1, ... , N} and let Uoc be the set of r-planes A for which
the corresponding COlUmnS {vf, ... , V n Of QA are linearly independent.
From these columns we obtain matrices VX and WX, as before, satisfying
WXVX = lr and QA = VXWX. lt is clear that {Uoc} is an open covering
of Gr(CN) that trivializes the tautological bundle. When A E U01 nUß, de-
fine Bocß(A) := wxv! E GL(r, o. Then Bocß(A)Bßoc(A) = WXQAVX = lr
and Baß(A)gßy(A) = WXQAVX = 9ocy(A), so that the Bocß are a family of
transition functions for the bundle Tt'- Gr(CN).
2.1 Vector bundles 55

..,. Now, given a continuous map f: M- Gr(<CN), forM with the finite cover
property, we can form the pullback bundle of rank r:

By Proposition 2.3, the equivalence dass of this bundle depends only on


the homotopy dass of f. Therefore we get a map

It turns out that, for sufficiently large N, this map has an inverse, thus
solving the dassifying problern for vector bundles. We construct that in-
verse, following in part [173, §1.1]. Given a vector bundle E.!!.... M of rank r,
suppose that {U1 , ... , Um} isafinite open cover of M suchthat EI u1 is triv-
ial, let { Bii E GL(r, ([): i,j = 1, ... , m} be the transition functions for this
vector bundle, and let {«/J i} be a partition of unity subordinated to {Ui}.
For N = rm, let Q be the N x N matrix-valued function formed from r x r
blocks, where the ij-block is ..J«/Ji«/JiBii· Then the defining relations (2.1)
for the transition functions show that Q 2 = Q and trQ = r. Choose local
frames Si = (sn, ... , Sir) E f(Ui, E), and define fik: M- Eby fik := filii Sik
on Ui, extended by 0 outside Ui. By the proof of Proposition 2.2, E can be
regarded as a subbundle of M x ([N, where the fibre Ex is spanned by the
vectors f;k (x). Notice now that

fik = ~ .Jtfii «/J j Sik = ~ .Jtfii «/J j9ij,kl Sjl


j j,l

= ~ fiij qij,kl Sjl = ~ qij,kl fiz,


j,l j,l

where k, l = 1, ... , r are matrix indices, so each fibre Ex is identified to the


r-plane imQ(x) in cN. Wehave constructed a continuous map Q: M -
Gr(CN) as well as a bundle map T: E- Tf, which is an isomorphism on
each fibre. Now the universal property of pullbacks gives a bundle map
f: E - f* Tf, which is an equivalence of vector bundles.
In practice, N can often be taken much smaller than rm. For instance, if
M is a manifold of dimension n, the integer N need only be greater than or
r tn.
equal to + This stability result follows from the homotopy properties
of the Stiefel manifold, which is the total space for the principal bundle
counterpart of our universal vector bundle [257, Thm. 7.7.2]. However, in
order not to be bothered with that, and also to handle the dassification
of bundles in the general paracompact case, it is customary to take the
inductive Iimit (or "direct limit") on N of Gr (([N), denoted BU(r), to obtain
56 2. Noncommutative Topology:Vector Bundles

a one-to-one correspondence

Vectr(M)- [M,BU(r)]. (2.4)

The formal definition of BU(r) is as follows. Denote by (("" the space of


complex sequences in which all but finitely many entries are zero; in an
obvious manner, C c C2 c · · · (("", topologized by declaring a subset of ((""
tobe open if and only if its intersection with each cN is open. Now, BU(r)
is defined as the set of all r-planes of (("", also topologized so that a subset
of BU(r) is open if and only if its intersection with each Gr(CN), for N ~ r,
is open. An important particular case is BU ( 1) = CP"", the inductive Iimit
of CP 1 c CP 2 c · · ·. A canonical bundle over BU(r) is constructed just as
in the finite-dimensional case, and the correspondence (2.4) is established
by similar arguments.
Wehave been working in the "topological" category (where bundle maps
are continuous). The translation of our results to the smooth category is
straightforward. When the base M is a smooth manifold, nothing is won or
lost by working in the smooth category, as every continuous vector bundle
over a smooth manifold has an (essentially unique) compatible differen-
tiable structure (of any differentiability dass) [246]. This is a consequence
of the classification theorem itself: the classifying map f: M - Gr(CN)
can be approximated by (and thus is homotopic to) a differentiable map,
and, as the tautological bundle is certainly smooth, the original bundle is
isomorphic to a smooth one.
Other fundamental properties of vector bundles are similarly worked out
as consequences of the classification theorem; for instance, the construc-
tion of connections from universal connections [354] (see Section 8.3) and
the definition of characteristic classes as inverse images of universal classes
associated to the cohomology of the Grassmannians.
We now begin the lang turn to a more algebraic way of life.

2. 2 The functor f
The totality of continuous sections of a vector bundle E ~ M is denoted
by [(M,E), or by [(E) when we need not specify the base space M. Notice
that f(E) is a module for the commutative algebra of functions C(M); the
action of C(M) is just scalar multiplication in each fibre, which we write on
the right:
sa(x) := s(x) a(x), for s E f(E), a E C(M).

If T: E - E' is a bundle map, there is a C (M)-linear map fT: r (E) - r (E')


given by fT(s) := T o s. The C(M)-linearity of fT means that fT(sa) =
fT(s)a for s E f(E), a E C(M), which follows from linearity of each
Tx: Ex - E~, by fT(sa)(x) = T(s(x)a(x)) = T(s(x))a(x) = fT(s)a(x).
2.2 The functor f 57

The equalities f(idE) = lnE> and f(T o u) = fT o fu, with an obvious nota-
tion, are clear.
Lemma 2.5. The correspondence E .... [(E), T .... fT is a functor r from the
category ofvector bundles over M to the category o{C(M)-modules. E3

The f-functor carries the operations of duality, Whitney sum and ten-
sor product ofbundles to analogaus operations an C(M)-modules. (Tensor
products of modules over an algebra are defined in the appendix Section
2.A.)
Exercise 2.3. Verify that f(E) E9 f(E') "'[(E E9 E'). Show also that f(E*) =
[(E)~, where E* - M denotes the dual vector bundle to E- M, and f(E)~
is the dual C(M)-module Homc<M> (f(E), C(M) ). o
Proposition 2.6. If E, E' are vector bundles over M and A = C(M), there is
a canonical isomorphism of A -modules:
[(E) ®A [(E') "'[(E ® E').
Proof. If s E f(E), s' E f(E'), we provisionally denote by s 0 s' the section
x .... s(x) ® s' (x) of the tensor product bundle E ® E' - M; let s ® s' be the
element of f(E) ®Af(E') given by Definition B.3 (of the tensor product of A-
modules). Let (}: [(E) ® A [(E') - [(E ® E') be the A-linear map determined
by B(s ® s') := s 0 s'; the claimisthat (} is an isomorphism.
If U c M is a chart domain over which the bundles E and E' are trivial,
then E®E' - M is also trivial over U. Indeed, any local section s E f(U, E) is
of the form s = 2:k=l Skhk. where (sl, ... , Sr) is a local frame for E over U,
and h1, ... ,hr E C(U); in other words, the sections {s1, ... ,sr} generate
f(U,E) freely over C(U). If (t 1 , ... ,ti) is a local frame for E' over U, its
members generate [(U,E') freely as a C(U)-module, and it is clear that
{s1 0 tk: j = l, ... ,r; k = l, ... ,l} generate f(U,E ®E') as a free C(U)-
module, since { SJ (x)®sk (x)} is a basis for each fibre Ex®E~. In summary, (}
is anisomorphism whenever the bundlesE and E' (and consequently E®E')
are trivial.
In the general case, by Proposition 2.2 there are vector bundles F and F'
over M such that E E9 F - M and E' E9 F' - M are trivial. Let L: E - E E9 F :
u .... (u, 0) and u: E E9 F - E: (u, v) .... u be the extension and restriction
maps, and let t': E' - E' E9 F', u': E' E9 F' - E' be similarly defined. Then
u o L = idE, so ru oft= lnE> and also fu' oft' = lnE'l; thus, ft and ft' are
injective, whereas fu and fu' are surjective.
Ex® E~ is a direct summand of the vector space (Ex E9 Fx) ® (E~ E9 F~)
for each x; this yields bundle maps t": E ® E' - (E E9 F) ® (E' E9 F') and
u": (E E9 F) ® (E' E9 F') - E ® E' satisfying fu" oft" = lm®E'l· Finally,
define
[L ® ft' : [(E) ®A [(E') - [(E E9 F) ®A [(E' E9 F'),
fu ® fu' : [(E E9 F) ®A [(E' E9 F') - [(E) ®A [(E')
58 2. Noncommutative Topology:Vector Bundles

by s ® s' ..... ft(s) ® ft'(s'), and t ® t' ..... fO"(t) ® [O"'(t'), respectively. This
yields a commutative diagram:

!
f(E) ®A [(E') 8 [(E ® E')

[t®[t'! ft"

[(E EB F) ®A [(E' EB F') ~ [( (E EB F) ® (E' EB F'))

fu®fu'! !ru"
f(E) ®A [(E') 8 [(E ® E'),

where 8 is the isomorphism of free A-modules already obtained. The top


half of the diagram shows that () is injective, since ft ® ft' and ft" are
injective and 8 is bijective; the bottarn half shows analogously that () is
surjective. Thus () is an A-linear isomorphism in the general case. 0

Corollary 2.7. Each C(M)-linear map from [(E) to [(E') is ofthe form fT
for some bundle map T: E - E'.

Proof. The bundle maps T form the total space of Hom(E, E') - M, a vector
bundle whose fibres are Hom(Ex,E~) ""E: ® E~. Thus T can be identified
with a section of the bundle E* ® E' - M. On the other hand, a C(M)-linear
map from f(E) to [(E') belongs to HomA([(E),[(E')) ""[(E)~ ®A [(E') ""
[(E*) ® A [(E'). It is easily seen that fT E HomA ([(E), [(E')) corresponds
to e- 1 ( T) E f(E*) ® A [(E') und er those identifications. Since () is bijective,
these account for all C(M)-linear maps from [(E) to [(E'). 0

Corollary 2.8. The functor r is faithful, full and exact.

Proof. "Faithful" means that fT = r O" as module homomorphisms only if


T = O" as bundle maps; this is clear from the definition of r. "Full" means
precisely what Corollary 2. 7 affirms to be true: that T ..... fT is a surjective
mapping from Hom(E, F) to HomA ([(E), [(F) ).
"Exact" means that r preserves short exact sequences. Let then T be a
bundle map of constant rank. Note that r preserves kernels and images:
f(ken) = ker(fT) and f(im T) = im(fT). Furthermore, if E'...!.... E.!!:. E"
is an exact sequence of vector bundles, then im T = ker O", so im(fT) =
f(imT) = f(kerO") = ker(fO"), and thus f(E') .!!.f(E).!2"r(E") is an exact
sequence of C(M)-modules. Note that if y: E"- E splits O", then fy splits
[O", so this sequence of modules is also split exact. 0

When M is locally compact but not compact, we can introduce fo(E) =


fo(M, E), the subset of continuous sections that vanish at infinity, which
is clearly a Co(M)-module (and also a Cb(M)-module). For suitable M, the
previous propositions stand, when f(E) is replaced by fo(E) and C(M) by
2.3 The Serre-Swan theorem 59

C0 (M). However, in the next result, where we look ahead to modules over
unital rings, we need to assume that M is compact.
Proposition 2.9. Let M be a compact space. Then the C(M)-modules f(E)
are finitely generated and projective.

Proof. By Proposition 2.2, there is a vector bundle E' - M and a positive


integer n such that E Eil E' "" M x cn. Thus, there is a split exact sequence
0 - E' - Mx cn - E - 0, so by Corollary 2.8, 0 - f(E') - C(M)n
f(E) - 0 is also split exact, so that f(E) is a direct summand of C (M) n. D

By Proposition 2.22, there is an idempotent E E Endc(Ml (C(M)n) so that


f(E) "" EC(M)n, arising from the split exact sequence. Since the endomor-
phism algebra Endc(Ml (C (M) n) can be identified with the matrix algebra
Mn(C(M)), this means that there is a matrix e = e 2 E Mn(C(M)) suchthat
f(E) ""e C(M)n as right C(M)-modules.

2.3 The Serre-Swan theorem


This theorem appears first in the work by Serre [432], and more or less in
its present form in [445].
Theorem 2.10 (Serre-Swan). The r functor from vector bundles over a com-
pact space M to finitely generated projective modules over C(M) is an equi-
valence of categories.

Proof. In view of Proposition 2.9, it only remains to prove that every finitely
generated projective C(M)-module '.Eis of the form f(M, E) for some bun-
dle E- M. Now, '.E"" eC(M)n for some idempotent e E Mn(C(M)) and the
sequence
0- kere- C(M)n- '.E- 0
splits (see Proposition 2.21). As r is fully faithful and exact, the endomor-
phism e induces a bundle map T : M X cn - M X cn' such that the image
ofT is a subbundle E(e) of Mx cn. Then

f(M,E(e)) = {Tos: s E f(M x cn)} = ime = '.E. D

(On the matter of the imageofT being a subbundle, the doubtful reader
may reason directly that since T 2 = T, then id -T is also an idempotent
bundle map; moreover, rank ( 1n - Tx) = n- rank Tx since Tx is idempotent.
Thus x - rankT x is both lower and upper semicontinuous, hence locally
constant.)
One can understand in retrospect why Proposition 2.6 proved so sticky:
in fact, it incorporates the difficulty of the Serre-Swan theorem. This theo-
rem does for vector bundles what the Gelfand-Naimark theorem does for
60 20 Noncornrnutative Topology:Vector Bundles

compact spaces, inviting us to extend the fruitful notions in algebraic topo-


logy to the category of modules over (noncommutative) C*-algebraso And
vice versao We can summarize what we have learned so far by saying that,
if C*-algebras are the noncommutative generalization of locally compact
spaces, then one should regard the (upcoming) C* -modules as vector bun-
d/es over noncommutative spaceso
In the smooth category, one defines the f"' functor from manifolds to
smooth sections of a vector bundle, for which the analogue of Proposi-
tion 206 applies, and one can obtain analogaus Serre-Swan equivalence of
smooth vector bundles over a compact manifold M and finitely generated
projective modules over C"' (M)o

2.4 Trading bundles for modules


With the Serre-Swan theorem in hand, we may look for further algebraic
parallels with vector bundleso Also, in cantrast to the proof of the Gelfand-
Naimark theorem, which was at least partly constructive, here we pro-
ceeded by constructing a one-size-fits-all algebraic straitjacketo We would
like to recover the canonicity of the arguments of Chapter 1, where, given
the algebraic object, the corresponding geometric space was reconstructed
in a fully explicit mannero Weshall address these concerns after introducing
an indispensable algebraic tool.
Let .J\. and 'B be rings with unit, and let cf>: .J\. - 'B be a unital ring ho-
momorphismo Then any right 'B-module J" becomes a right 5\.-module by
defining
t a := t cf>(a) for t E J", a E .J\.0
0

If (j is another right 'B-module and 1./J E Hom 11 (J", (j), then 1./J(t a)
0

1./J(tc/>(a)) = 1./J(t) cf>(a) = 1./J(t) oa, so 1./J can also be regarded as a member
of HomJl (J", (j)o In this way, cf> defines a functor Rq, from the category
of right 'B-modules to the category of right 5\.-moduleso When cf> is the
inclusion map of a subring .J\. into a larger ring 'B, this functor is called
restriction of scalarso
The homomorphism cf> defines another functor Eq, -called extension of
scalars when .J\. isasubring of 'B- from right 5\.-modules to right 'B-mod-
ules, as followso
Definition 2.6. Suppose cP: .J\. - 'B is a unital ring homomorphism and 'E
is a right 5\.-moduleo Then 'B becomes a left 5\.-module by a · b := cf>(a)b;
thus we may construct the tensor product Eq, ('E) := 'E ®Jl 'B of'E and 'B over
.J\. by means of cf> -one of the key notions in this book. This is the abelian
group whose elementsarefinite sums L.1 s1 ® b 1 with s1 E 'E and b1 E 'B,
subject only to the relations
sa ® b = s ® cf>(a)b, for each a E .J\.0
2.4 Trading bundles for modules 61

There is a right action of 'Bon 'E ®5\ 'B defined by (s ® b)b' := s ® bb'. If :f
isanother right .Jl.-module and if T E Hom.Jt ('E, :f), then E<t>(T) := T ® id11:
'E ®.Jt 'B- :f ®.Jt 'Bis 'B-linear. The additive map cp~ : s ,_. s ® 1 from 'E to
'E ®.Jt 'B intertwines the module structures: cp~ (sa) = sa ® 1 = s ® cp(a) =
cp 1(s)cp(a), and it is "natural" insofar as (T ® id 11 ) o cpl = cp' o T.
The right 'B-module 'E ®.Jt 'B has a universal property. Let 'E, :f be right
modules overunital rings 5\, 'B respectively. By a module morphism (0, cp)
we mean a pair consisting of a unital morphism cp: 5\ - 'B and an additive
map e: 'E- :f that intertwines the module actions: if s E 'E, then (}(sa) =
(}(s)cp(a).

Lemma 2.11. Given a right 5\-module 'E and a right'B-module :f, ifO: 'E-
:f is an additive map satisfying (}(sa) = (}(s)cp(a) for s E 'E, a E 5\, then
there exists a unique 'B-linear map e:
'E ®5\ 'B- J suchthat (} = {J 0 cp;.

Proof. Just put e(s ® b) := B(s)b on simple tensors; this is well defined
since e(sa ® b) = (}(sa)b = (}(s)cp(a)b = e(s ® cp(a)b), and it is clearly
'B-linear. D

Examples will soon appear; meanwhile, we remark that the previous con-
struction has a nice application in linear algebra: when cp is the inclusion
~ ..... C and V is a real vector space, then E<t> (V) = V ®JR C is its complexifi-
cation vc. The "quaternionification" VIHI can be analogously defined.
Exercise 2.4. Prove that the quaternionification of C as a real algebra is
C ®~~t IHI ""Mz(C). 0
Exercise 2.5. Show that the functors R<t> and E<t> are mutually adjoint, i.e.,
Hom11 (E<t>('E),:f) ""Hom.Jt('E,R<t>(:f)) naturally, for any 5\-module 'E and
any 'B-module :f. 0
Exercise 2.6. Prove the uniqueness: if 'E# and cp# satisfy the respective
properlies of 'E ® 5\ 'B and cp', find an isomorphism a: 'E ® 5\ 'B - 'E# such
that cp# = a o cp' . 0
Exercise 2.7. Prove that R<t> sends projective modules into projective mod-
ules. 0
Exercise 2.8. If 'E is a finitely generated projective right 5\-module, show
that 'E ®.Jt 'Bis a finitely generated projective right 'B-module. 0
Exercise 2.9. We remark that an 5\-bimodule is a right (5\ o ® 5\)-module,
where .Jl. o is the opposite algebra of 5\, that is, 5\ o := { a o : a E 5\}
with aobo := (ba) Work out the bimodule case of extension of scalars
0 •

by cp: 5\ - 'B, to conclude that the E<t> ( 'E) is given by 'B ® 5\ 'E ® 5\ 'B. 0

.,. Now we apply all this to the case in which cp: 5\ - 'B is the morphism
Cf between the algebras C(M) and C(N) associated to a continuous map
f:N-M.
62 2. Noncommutative Topology:Vector Bundles

Proposition 2.12. There is an isomorphism ofC(N)-modules,

[(N,j* E) ""f(M, E) ®c(M) C(N).

Proof We proceed by interpreting geometrically the right hand side. We can


regard simple tensors like s ® g as functions N-E: y .... s(j(y))g(y) of
a particular type, to wit, as sections of E- M along the map f: N-M. A
section along a map f: N - M is defined as a continuous map G': N - E
suchthat rr o G' = f. For instance, a vector field X over N defines a vector
field along f, namely T f o X. Now, there is a one-to-one-correspondence,
in fact an isomorphism of C(N)-modules, between sections G' of E - M
along the map f: N-M and sections of the pullback bundle f* E- N:

given by y .... j- 1 G'(y); recall that j is an isomorphism on each fibre. D

A couple of comments are in order. The point of the identification made


in the previous proof is that we construct the module of sections of the
pullback bundle without invoking the commutativity of the algebra: we
simply apply the functor EcJ to the C(M)-module [(M,E). Thus we see
that the tensor product construction of Definition 2.6 effectively produces
"noncommutative pullback bundles".
Proposition 2.12 and its proof have obvious analogues in the smooth cat-
egory, on replacing continuous functions and sections by smooth functions
and sections; that will be usefullater on.
The concept of sections along a map, so natural from the noncommuta-
tive viewpoint, is unfortunately underused as yet in differential geometry.
Until now, it has been employed mainly in theoretical mechanics [70). For in-
stance, if we pull back the cotangent bundle T* M ..!:. M via the bundle pro-
jection rr itself (here E = N = T* M), we get semibasic differential forms,
i.e., elements of [(T*M,rr*T*M), locally of the form Ljfi(q,p)dqi on
phase space. The Uouville form oc on T* M is the semibasic form corres-
ponding to the identity section ofT* M along rr : T* M - M. Sections along
a map also play a crucial role in supermechanics, where their geometrical
surrogates prove tobe inadequate [69).

Exercise 2.10. Given a continuous map f: N - M and a vector bundle


E- M, prove that f(N,f* E*) ""[(N,j* E)~, where the latter is the C(N)-
module dual to f(N, f* E). 0
2.4 Trading bundles for modules 63

There is a canonical map cp from f(M,E) to f(M,E) ®c(M) C(N), given


by s ...... (y ...... s(j(y))), for y E N. Consider the diagram

f*E~E
f-st ls
N~M.
There is an obvious mapping of sections f~: f(M, E) - f(N, f* E), given by
Ps: y ,_ ] - 1 (s(j(y))) for s E f(M,E); on identifying f(M,E) ®c(M) C(N)
and f(N,j* E), it is clear that (Cf)~ - P.
We checkthat Pisamodule morphism:
f~ (sa)(y) = ] - 1 [sa(j(y))] = ] - 1 [s(j(y)) a(j(y))]
= j- 1 [s(j(y))] a(j(y)) = Ps(y) Cj(a)(y).

For instance, the inclusion of a closed subspace, j: N ..... M, will induce the
restriction map
j~: f(M,E)- f(N,EIN),

since j*E =EIN· In that case j~f(M,E) = f(N,j*E), but in general this
equality need not hold. As an example, in mechanics not every semibasic
differential form is a basic one, i.e., of the form rr~ w with w E f(M, T* M),
therefore locally of the form 2.1 fJ (q) dqi on phase space.
~ We can now have a second go at understanding Serre-Swan equivalence.
Given e = e 2 E Mn ( C (M)) associated to the projective module '.E, the sub-
bundle E(e) can be constructed in the following way. Form the tensor pro-
duct module '.E ®c(M) C "' eC(M)n ®c(M) C by means of the evaluation
map Ex: C(M) - C at x: this is a complex vector space Ex of dimen-
sion ~ n, namely, the fibre over x! We still have to assemble the vector
bundle from its pieces: consider the fibration E(e) - M as the disjoint
union E(e) := l±!xeM Ex. To checkthat this defines a vector bundle, we must
show first that it has (locally) constant rank. We extend Ex to a morphism
Ex: Mn(C(M))- Mn(C) in the obvious way. Then Ex is identified with the
subspace Ex(e)cn, so that dimEx is the rank rx of the matrix Ex(e). Now
r y ~ r x for all y in a neighbourhood of x, because linearly independent
columns of Ex(e) remain independent near x. Also, Ex(e) is anidempotent
in Mn(C), and Ex(l - e) = ln- Ex(e) is the complementary idempotent;
replacing e by l - e gives n- ry ~ n- rx for y near x. Therefore dimEx
is locally constant on M. We may topologize E(e) as a quotient space of
Mx cn; this makes E(e) a topological space for which the obvious projec-
tion E(e) - M is continuous, and indeed makes E(e) - M a (locally trivial)
vector bundle.
We finally checkthat '.Eis isomorphic to f(E(e)): if t E '.E, then El(t) E
'.E ®c(M) C =Ex. so st(x) := El(t) defines a section St E f(E(e)) and t ......
64 2. Noncommutative Topology:Vector Bundles

St is dearly an injective C(M)-module homomorphism [198]. To check its


surjectivity, suppose that {t1o . .. , tm} generates 'E; any section s E [(E (e))
is of the form s(x) = I}: 1 ti ®/j (x), for some fi E C(M). If t := I}: 1 ti!J•
then
m m
st(x) = E~(t) = L E~(tj)Ex(jj) = L (tj ® 1) · /j(x)
j=l
m
= 2:: tj ® IJ<x) = s(x).
j=l

.,. Yet another view comes from the dassification theorem. The last discus-
sion leads us to see the dassification theorem and the Serre-Swan theorem
as essentially equivalent. This comes ab out because an idempotent element
e of Mn ( C (M)) can be regarded as a continuous function into the set of
matrix idempotents. We remark, then, that E(e): x ..... Ex(e) gives a map
M- G := Granke((n); identify the idempotent Ex(e) with its range space
Ex(e)(n =Ex as elements of the Grassmanman tautological bundle T- G.
The vector bundle corresponding to 'Eis just the pullback E(e) = E(e)*T.
Thus [(M, E(e)) == [(G, T) ®c(G) C(M), and the latter is, by Proposition 2.12,
the C(M)-module of sections along the map E(e). Now, if a: M-T is such
a section, then a(x) E Ex c cn, so we can write a(x) = (ji(x), ... .fn (x))
with fJ E C(M), and a ..... e(fi, ... .fn) gives a C(M)-linear isomorphism
from [(G, T) ®c(G) C(M) to e(C(M))n == 'E.
Note that the transition functions of a pullback bundle are just the pull-
backs of the transition functions of the original bundle. This gives a uni-
versal formula for transition functions:
Bcxß(X) := WX(x) V{(x)•
where A(x) := imEx(e) is the range of a matrix idempotent. Reciprocally,
given the transition functions, one can construct the matrix idempotent
(and show that its homotopy dass depends only on the vector-bundle iso-
morphism dass), as already donein Section 2.1; seealso [173, §1.1].

2.5 C*-modules
The formal arguments concerning vector bundles expounded so far pre-
serve an "old algebra" flavour. Wehave refrained from introducing Hilbert
space-like structures over the fibres, which indeed would have simplified
some proofs, at the price of obscuring their geometrical roots. In order to
use the full power of the operator algebra methods, nevertheless, it is high
time that we bring inner products into our constructs.
Any complex vector bundle E - M can be endowed, in many ways, with
a Hermitian metric. For that, we can define a positive definite sesquilinear
2.5 C*-modules 65

form h on each trivial subbundle Elui' for UJ belanging to a trivializing


open covering of M; and then splice the metrics tagether using a partition
of unity. Or, simpler still, we can restriet the natural scalar product on
each fibre of MX cn to [(M,E) "" eC(M)n. Thereupon, we get a pairing
fo(E) x fo(E)- Co(M), given by (r I s) = h(r,s): x ..... hx(r(x),s(x)), for
x E M and r, s E fo (E), which is Co (M)-sesquilinear, conjugate-symmetric,
and positive definite. With this pairing, fo(E) is called a "pre-C*-module".
The pairing does not require or use the commutativity of the C* -algebra
Co(M). We now formulate this structure in the noncommutative context.
Definition 2.7. A (right) pre-C*-module over a C*-algebra Ais a complex
vector space 'E that is also a right A-module (not necessarily finitely gene-
rated or projective) with a sesquilinear pairing 'E x 'E - A satisfying, for
r, s, t E 'E and a E A, the following requirements:
(r I s + t) = (r I 5) + (r I t),
(r I 5a) = (r I s) a,
(rls)=(5lr)*,
(s I 5) > 0 for s * 0. (2.5)

In other words, the pairing is A-linear in the second variable, conjugate-


symmetric and positive definite. Notice the consequence that (rb I 5)
b* (r I 5) if b E A. One can complete 'Ein the norm

lllslll := ~ll(s I s)ll (2.6)

where II · II is the C*-norm of A; the resulting Banachspace is then a (right)


C*-module. When the underlying C*-algebra must be mentioned explicitly,
we call it a right C* A-module.
If Aisunital and s E 'E, then

(sl- 5 I 51- s) = l(s I s)l- 1(5 I 5)- (5 I s)l + (s I 5) = 0, (2.7)


and so the relation 51 = 5 follows from (2.5): there is no need to add this
condition as a postulate.
As mentioned in Section LA, any nonunital C*-algebra has an approx-
imate unit {uoc}, suchthat llauoc- all - 0 for each a E A. An obvious
adaptation of (2.7) shows that (5Uoc - s I suoc - 5) - 0 for each s E 'E;
therefore 'EA is densein 'E.
Call ('EI 'E) the linear span of the elements (r I 5) with r,s E 'E; it is
obviously an ideal of A. We say that 'E is a full C* A-module when the
closure of this ideal is all of A. When A is unital, 'E is full if and only if
('E I 'E) = A. In any case, 'E('E I 'E) is densein 'E: use an approximate unit
on the closure of ('E I 'E) in A.
The category of C*-modules -sometimes called "Hilbert modules"- en-
compasses both Hilbert spaces and the kind of Banach spaces that underlie
66 2. Noncommutative Topology:Vector Bundles

C*-algebras. One could say that a C*-module obeys the same set of axioms
as a Hilbert space, except that the inner product takes its values in a general
C*-algebra. This seemingly innocent generalization leads to a lot of tech-
nical trouble: cherished Hilbert space properties, like Pythagoras's formula
or selfduality, must be relinquished, and, to see our way out of difficulties
in proofs, familiarity with the properties of the cone of positive elements
in a C*-algebra (see Section LA) is essential. But, upon scratching the sur-
face, we shall find striking similarities with Hilbert spaces in the way the
corresponding algebras of operators behave. C*-modules were introduced
in the seventies by Rieffel, among others, to deal with the theory of induced
representations [387], and subsequently developed as a fundamental tool
by Kasparov [276]. They bring unity to many apparently unrelated notions
of current use in noncommutative geometry and physics.
The two most obvious examples of C*-modules already hintat their va-
riety.
Example 2.1. C* C-modules are ordinary Hilbert spaces, where the pairing
is just the Hilbert space scalar product ( · I · ).
Example 2.2. Any C* -algebra A is a C*-module over itself: if we define
(b I a) := b*a, then (2.5) is obvious; and the norm is the same because of
the C* equation, lllalll := lla*all 1 12 = llall.
Example 2.3. Another obvious example of a C* A-module is An: if a =
(a1, ... , an)t and b = (b1, ... , bn)t are typical "column vectors" in An, then
(b I a) := bia1 + · · · + b~an. [If A = Co(M), this is the same as the space of
sections vanishing at oo of the trivial vector bundle Mx cn.] We can regard
An as the (vector space) tensor product of cn with A.
Example 2.4. If we reorganize the direct sum of n copies of A as a set of
"row vectors", denoted by nA from now on, we get a right module over
the C*-algebra Mn(A). Define the pairing of two such row vectors a =
(a1, ... ,an) and b = (b}, ... ,bn) as the matrix (a I b) := [atbJ] whose
(i,j)-entry is at b 1. To see that this pairing is positive definite, it is enough,
in view of Proposition 1.20, to check that
n
2: ctataJCJ = (2:;a;c;)*C~: 1 aJCJ);:: 0
i,j=l

for any c 1 , ... , Cn E A (strict positivity is obtained by taking c1 = a j ). Thus


nA is made a right C* Mn (A)-module.
Example 2.5. The sum of countably many copies of A plays a central role
in the theory. lts definition needs a little care.
Notation. In this book, J{ will denote the Hilbert space with a countably
infinite orthonormal basis, unless explicitly indicated otherwise; that is, we
shall assume that J{ is infinite-dimensional and separable. For its scalar
product, we use the ( · I ·) notation. lt is modelled by the sequence space
-f 2 or by L 2 (~).
2.5 C*-modules 67

Definition 2.8. The algebraic tensor product J{ 0 A of J{ and a C* -algebra


A, consisting of finite sums of simple tensors L.j': 1 ~J ® aJ with ~J E J{
and a J E A, is clearly a right A-module and it is endowed with an A-valued
pairing given on simple tensors by

(1J ® b I~® a) := (17 I~) b*a. (2.8)

This pairing can be shown to be positive definite, and the remaining prop-
erties of (2.5) are obvious; thus J{ 0 A is a pre-C* A-module. Call J{ ® A
its completion to a C*-module.
Exercise 2.11. Checkthat the pairing (2.8) is indeed positive definite on
J{0A. 0
Example 2.6. Let A be a commutative C* -algebra with Gelfand spectrum M.
To obtain C* -modules over A, one is not restricted to spaces of sections
of vector bundles over M. One can consider infinite-dimensional vector
bundles, or bundles whose fibres are no Ionger locally trivial. For instance,
Moyal quantization may be described with the help of a continuous field
of Hilbert spaces, over a space related to the set of orbits of the coadjoint
action of the Reisenberg group on the dual of its Ue algebra; this field
forms a C* -module. Infinite-dimensional vector bundles are instrumental
in mathematical formulations of quantum field theory. These generalized
bundles are all instances of the following definition [137, Chap. 10).
Definition 2.9. Let Y be a locally compact space. A continuous field of
Hilbert spaces over Y is a fibration H.!!.... Y suchthat Hy := rr- 1 (y) is a
Hilbert space for each y E Y, tagether with a <C-linear subspace ß of the
Set of general sections nyEYHy, SUChthat
(a) for each y E Y, the subspace { ~(y): ~ E ß} = Hy;
(b) for any ~ E ß, the function y ...... ll~(y) II is continuous;

(c) ß is "locally uniformly closed": if ;y E nyEY Hy and if for each E > 0


and each Yo E Y, there is some ~ E ß for which ll;y(y)- ~(y)ll < E
on a neighborhood of y 0 , then ;y E ß.

The condition (b) may be replaced by the requirement that, for any ~. 17 E
ß, the function y ...... (~ (y) 117 (y)) be continuous. We shall use the notation
H := { Hy: y E Y} to refer to such a continuous field of Hilbert spaces.
Local uniform closure of the continuous section space ß is needed in
orderthat ß be stable under multiplication by continuous functions on Y.
Indeed, if ~ E ß and if f: Y - <C is continuous, then given Yo E Y and
E > 0, let V:= { y E Y: lf(y)- f(yo)l ::; E, ll~(y)- ~(yo) II ::; 1 }; then for
y E V,

llf(y)~(y)- j(yo)~(y)ll::; Ell~(y)ll::; E(ll~(yo)ll + 1),


68 2. Noncommutative Topology:Vector Bundles

so that ~~ is close to j(y0 )~ E ß over V; by (c), ~~ E ß too.


The condition (a) can be replaced by the apparently less demanding con-
dition that the subspace { ~(y) : ~ E ß} be merely densein Hy. However,
local uniform closure then guarantees that this dense subspace is all of H Y•
by the following approximation argument [137, Prop. 10.1.10].
Lemma 2.13. For each y E Y and 11 E H y, there is some ~ E ß such that
~(y) = 11·
Proof. Approximate 11 by the partial sums of an absolutely convergent se-
ries 11 = Lk 11k. whose members lie in the dense subspace { ~(y) : ~ E
ß }. For each k, choose a section ~k E ß such that ~k (y) = 11k. and de-
fine a continuous function fk: Y - (( by fk(x) := 1 if ll~k{x)ll 5 1117kll.
fk(x) := 1117kll/ll~k(x)ll otherwise. Then ~dk E ß, and the sum ~(x) :=
Lkfk(x)~k(x) converges in Hx for each x E Y; clearly ~(y) = 11· Now
~ E ß follows from local uniform continuity, since if E > 0, then

~~~(x)- ~ (~kfk)(x)jj = II L (~kfk)(x)ll


k~o k>m
5 L ll~k{x)fk(x)ll 5 L 1117kll < E,
k>m k>m

for large enough m and all x E Y. 0

Also, for any n E ~ the subset { y E Y : dim H y ~ n } is open: choose


~~ •... , ~n E ß suchthat the ~J(Y) are linearly independent in Hy; then
they are linearly independent in the neighbourhood of y where the Gram
determinant det[(~j(x) I ~J(x))] remains positive.
An isomorphism of two such continuous fields H and E is a family lJl =
{ l/ly : y E Y} of linear isometries l/ly: Hy - Fy that maps ß(H) onto
ß(E). Such a field is trivial if it is isomorphic to a constant field, i.e., one
in which Hy = J{ for every y and ß(H) = Co(Y -Jf). We say that H is
locally trivial if each point of Y has an open neighbourhood U such that
the restriction Hl u is trivial.
The point of Definition 2.9 is that every C*-module over C0 (Y) arises
from a continuous field of Hilbert spaces over Y. In the noncommutative
case, no such spatial picture is available in general. However, it should be
obvious that by making Definition 2.9 we open a Pandora's box of new
possibilities, as nothing forbids noncommutative C*-algebras to act on the
spaces of sections of a continuous field of Hilbert spaces. This "quantiza-
tion" is one of the themes of this book.
Definition 2.10. A continuous field of C*-algebras over a locally compact
space Y is a fibration A ~ Y where each fibre Ay is a C* -algebra, together
with an involutive SUbalgebra ß = ß(d) of 0yEY Ay that satisfies three
properties listed in Definition 2.9, namely, these sections fill out each fibre
2.5 C*-modules 69

and have continuous norms, and ß is locally uniformly closed. An isomor-


phism ljJ of two such continuous fields is required to have fibre maps ljJ y
that areC* -algebra isomorphisms.
In particular, if A and Bare two C*-algebras, a deformation from A toB
is a continuous field of C*-algebras { Ah : 0 :s h :s ho}. suchthat Ao = A
and Ah = B for h > 0, the restriction to (0, ho] being trivial.

.,.. With these examples in hand, we make some basic observations about
general C* -modules. First we need to checkthat (2.6) indeed defines a norm!
There is an analogue of the Schwarz inequality.
Lemma 2.14. If'E isa rightC* A-module, then ll(r Is)ll :s lllrllllllslll forall
r,s E 'E.

Proof. If a, c E A with c ~ 0 (i.e., c lies in the positive cone of A), then


a*ca :s llcll a*a by (1.20). In particular, a*(r I r)a :s lllrlll 2 a*a if r E 'E.
With this in mind, the usual proof of the Schwarz inequality goes through.
For r, s E 'E, a E A, consider the inequalities

0 :s (ra-s I ra-s)= a*(r I r)a + (s I s)- a*(r I s)- (s I r)a


:s lllrlll 2 a*a + lllslll 2 - a*(r I s)- (s I r)a.

Now take a := lllrlll- 2 (r I s) to establish the result. D

The triangle inequality follows immediately.


Corollary 2.15. II Ir+ slll :s lllrlll + lllslll for all r, s E 'E.

Proof. Since lllr + slll 2 = ll(r Ir)+ (s I s) + (r I s) + (s I r)ll. just apply the
ordinary triangle inequality for the C*-algebra norm II ·II and the preceding
Schwarz inequality. D

Remark. The inner product is separately continuous in each variable:

ll(r I s)- (r I Oll :s lllrlllllls- tlll.

In particular, the pairing of a pre-C* -module extends to an inner product


on its completion.
Exercise 2.12. Check the Banachmodule condition: lllralll :s lllrlllllall for
r E 'E, a E A. o
Exercise 2.13. Find counterexamples to Pythagoras's formula, using com-
mutative algebras. 0

Exercise 2.14. If a1, ... , an, h, ... , bn are elements of a C*-algebraA, show
that
70 2. Noncommutative Topology:Vector Bundles

Conclude that the multiplication map m: A 0 A - A : a ® b .... ab is norm-


decreasing if the algebraic tensor product A 0 A is given the Haagerup
norm (1.24). 0

.,. The morphisms of C*-modules will obviously be bounded linear maps


commuting with the module action of the C*-algebra. lt turns out that
these properties arenot enough; the existence of adjoints must be explicitly
demanded. On the other hand, this last requirement entails the others; we
make it a formal definition.
Definition 2.11. Let 'E and;: be C* A-modules. A map T: 'E - ;: is ad-
jointable if there is a map T*:;:- 'E, called the adjoint ofT, suchthat

(r I Ts) = (T*r I s) for all r E ;:, s E 'E. (2.9)

Proposition 2.16. If T is adjointable, then the adjoint is unique and is ad-


jointable: T* * = T. Moreover T, T* are bounded module maps and (ST) * =
T* S* when S is an adjointable map.

Proof. This is all routine. For instance,

(r I (Ts)a- T(sa)) = (r I Ts)a- (T*r I s)a = 0 for all r E;:


forces T(sa) = (Ts)a. Notice that T is automatically linear if (2.9) holds.
Also, the graph of T is closed, in view of the existence of T*; hence T is
bounded, by the closed graph theorem [383, Thm. III.12]. 0

The point ofDefinition 2.11 is that, in cantrast to what happens in Hilbert


spaces, not every bounded A-linear operator between C*-modules has an
adjoint. To see that, consider a C* A-module 'E with a closed submodule;:
and define its orthogonal submodule ;:1. in the usual way, as { r E 'E :
(r I s) = 0 if s E ;: }. Then it may happen that;: Eil ;:1. * 'E; and often
;: ~ ;:u. For instance, take 'E = A := CI and;: := C<C; then ;:1. = 0 and
;: 1.1. = 'E. A submodule ;: of 'E is said to be complementable when there

is an orthogonal submodule (j suchthat;: Eil (j = 'E. This does not happen


in general; but precisely because of that, a bounded A-linear operator does
not always have an adjoint. In fact, T: 'E - ;: is adjointable if and only if
its graph G ( T) satisfies G ( T) Eil G ( T) 1. = 'E Eil J. Then, as usual, G ( T*) =
{ (r, s) E ;: Eil 'E : (s, -r) E G (T)l.} = { (r', s') E ;: Eil 'E : (s', -r') E G(T)} 1..
Moreover, a closed submodule is complementable if and only if it is the
range of an adjointable operator [481, Cor. 15.3.9].
We write HomA ('E, :f) to denote the vector space of all adjointable maps
T: 'E- ;:. When;: = 'E, we use the notation EndA('E) for the algebra of
adjointable operators on 'E. It is clearly a Banach algebra, with norm given
as usual by IITII := sup{IIITslll : lllslll :::; 1}. We shall see in Chapter 3
that EndA ('E), just like the algebra of operators on a Hilbert space, is a
C*-algebra: this is a cornerstone of the theory.
2.5 C*-modules 71

~ We can now introduce a very useful concrete presentation of the C*-


module J-{ ® A. Denote by J-{A the C* A-module of sequences a = {ak}
in A suchthat Lk=O at ak converges in A, with the obvious pairing (a I b) :=
Lk=O at bk. The operators on J-{A given by
(2.10)

are clearly adjointable; indeed, p-:_ = Pn. Moreover, P~ = Pn, so they are
projectors on J-{A·
Exercise 2.15. Prove that I\ I~- Pn~l\1 I 0 as n - oo for any ~ E J-{A· 0
Exercise 2.16. Compute the norm of I~ 1 EJ ® a 1 in J-{ 0 A when the EJ
are orthonormal vectors in J-f. Then show that J-{ ® A may be identified to
J-{A· 0

~ Consider the "ketbra" operators of the form

lr){sl: t- r (s I t): 'E- :J,

for r E :J, s E 'E. Since r (s I ta) = r (s I t) a for a E A, these operators are


A-linear. Clearly, lr){sl is adjointable, with lr){sl* = ls){rl. Composing
two ketbras yields a ketbra:

lr){sl · lt){ul = lr(s I t)){ul = lr){u(t I s)l,

so all finite sums of ketbras from 'E to 'E form an algebra.


Exercise 2.17. Show that this algebra is a selfadjoint two-sided ideal of
EndA('E). 0
Anticipating a bit, weshall denote this ideal by End~0 ('E) and by End~ ('E)
its norm closure in EndA ('E). More generally, when 'E and :J are two C* A-
modules, the finite sums of ketbras Lk=ll Yk) {sk I may be called A-finite
rank operators from 'E to :J; they form a vector space Hom~0 ('E,:J). The
members of its closure Horn~ ('E, :J) may be called A-compact operators.
We warn the reader that such an operator is not necessarily compact in the
usual sense of the theory of Banach space operators.
Weillustrate these concepts by examining the commutative case: if A =
C(M) for a compact M and if 'E = f(M, E) is the space of sections of a finite-
dimensional hermitian vector bundle E- M, then End~ ('E) === [(M, EndE).
We first establish an easy lemma, of independent interest.
Lemma 2.17. If A is any C* -algebra, then End~ (A) === A.

Proof. If A is unital, T - T( 1) gives a bijection between bounded module


maps and elements of A. Obviously any such map a- b*a is adjointable
and of A-finite rank: it is just 11) {b 1.
72 2. Noncommutative Topology:Vector Bundles

Assurne then that A is nonunital. Consider the map

O:End~0 (A) -A:IilaJ}(bJI .... LjaJbj. (2.11)

This is well defined, because if Lj a i b ja = Lj ci d ja for all a E A, then


Lj ajbJ = Lj Cjdj. Moreover, e is a *-homomorphism:

O(la)(bllc)(dl) = ab*cd* = O(la)(bl) O(lc)(dl),


O(la)(bl*) = O(lb)(al) = ba* =(ab*)*= O(la)(bl)*. (2.12)

Now, e is anisometry, since IIIilaJ)(bjl(uC()II-IIIJaib}ll when {uC(} is


an approximate unit in A. The range of e is the ideal A 2 , which is densein A
(since auC( - a); so 0 extends to a *-isomorphism from End~ (A) to A. D
Clearly, End~ (An) = Mn (A) in the same way. Indeed, the following more
general result holds.
Lemma 2.18. If p E Mn(A) is a projector, pAn is a C* A-module and
End~(pAn) = pMn(A) p.

Proof. We show first that End~ (pA) = pAp when p is aprojectorinA itself.
The argumentrunsparallel to the proof of Lemma 2.17. If Ais unital, then
p is a unit for pAp, and T .... pT(p)p gives a bijection between End~ (pA)
and pAp, any such map pap .... pb*pap is of A-finite rank, since it equals
lp)(pbpl. Assurne then that Ais nonunital. Then

e: End~0 (pA)- pAp: Ljlpaj)(pbjl .... Lj pajbjp

is a well-defined *-homomorphism: replace a by pa, b by pb, etc., in (2.12).


lt is an isometry, since IIIJIPaj)(pbJI(uC()JJ- IIIJ paib}P11 when {uC(}
is an approximate unit in A. The range of e is pA 2 p, a dense ideal in pAp,
so 0 extends to a *-isomorphism from End~ (pA) to pAp.
The general case follows by an extension of the same argument, defining
e: End~0 (pAn)- pMn(A) p as the linear map that takes lpa)(pbl to the
matrix with (i,j)-entry Lk,l PikakbiPIJ· D

The reader who is already familiar with the K-theory of C*-algebras will
know that the K-theory of Ais intimately related to the properties of matrix
algebras over A; here we see that those are operator algebras on natural C*-
modules over A.
When A = C(M), we conclude that End~(f(M,E)) ""f(M,EndE) when-
ever f(M, E) is of the form p C(M)n for some orthogonal projector p in
Mn(C(M)). In order to justify the claim prior to Lemma 2.17, it remains
only to prove that in defining projective modules, idempotents can always
be replaced by projectors; in other words, to establish that any finitely gene-
rated projective module over a unital C* -algebra can be made a C* -module.
This will be proved in Section 3.1.
2.5 C*-modules 73

.,. This is perhaps a good place to say a few words about duality. For any
right module 'E over a C*-algebra A, consider the set 'E~ of bounded A-
module maps ~: 'E- A. This can be regarded as a right A-module, with the
(not very natural) operations

~a: s ..... a*~(s), (2.13)

for ~ E 'E~, i\ E C, a E A and s E 'E. Clearly there is an injective module


map 'E- 'E~ given by r ..... (r I ·) -in fact, the existence of this map could
be used to define C*-modules, as Daniel Kastler has suggested. We say
that 'E is selfdual if 'E "' 'E~ under this correspondence. By mimicking the
standard argument in Hilbert spaces, one can show that if 'Eis selfdual, then
any bounded module map from 'E to any C* A-module :J is adjointable.
lt follows that if :J is a selfdual submodule of a C* A-module 'E, then
:Je :Jl. = 'E. lt turnsout that Ais selfdual if and only if it is unital; and
then of course An is selfdual too .
.,. Now we arrive at a key junction. Suppose that a right C* A-module 'E
and a morphism of C*-algebras cp: A - Bare given. We want to import
to the category of C*-modules the functor Ecp of Section 2.4, so we need
to show that the algebraic Ecp('E) becomes naturally a pre-C* B-module. If
s 0 b denotes a simple tensor in the algebraic tensor product 'E 0 B (over C),
a B-valued pairing on 'E 0 B is naturally defined by

(s 0 b Ir 0 b') := b* cp( (s Ir)) b'. (2.14)

This pairing is not definite, and elimination of the nullspace is equivalent to


the conditionofgood definitionof Ecp('E), that is to say, sa0b = s0cp(a)b.
lt would be catastrophic if this failed! Let x := sa 0 b- s 0 cp(a)b. Then,
since cp((sa I sa)) = cp(a)*cp((s I s))cp(a), we check, with some relief, that
(x I x) = 0.
The converse is a bit trickier to prove. Suppose that x = LJ=l s1 0 b 1
satisfies (x I x) = 0. We can regard nx := 'E e · · · e 'E (n times) as a right
Mn (A)-module, where for "row vectors" s = (sl, ... , Sn) and t = (ti. ... , tn ),
s[aiJ] = t means Ir= I siaiJ = t 1 . Then the matrix [ (si I s1 )] satisfies, for all
a E An,
n
L a:(si I SJ)aJ = ('Lisiai I L. 1 sJaJ) ~ 0 inA,
i,j=l

so by Proposition 1.20, it is a positive element of Mn (A), with a positive


square root a E Mn(A). Now the morphism cp: A-B determines a mor-
phism cp(n) from Mn (A) to Mn (B), by cp(n) [aij] := [ cp(aij) ]. If ß = cp(n) (a),
then ß 2 = cp(nl(a 2 ) = [cp((si I Sj))]. Now
n n
L h7ßZißkJbJ = L b(cp((si I Sj))bJ = (x I x) = 0
i,j,k=l i,j=l
74 2. Noncommutative Topology:Vector Bundles

implies 2:1 ßkJbJ = 0 in B for each k. Finally, if r E ~ is determined by


roc = s, we get
n n n
2: (YkOCkj 0 bj- Yk 0 cP(OCkj)bj) = 2: Sj 0 bj- 2: Yk 0 ßkjbj =X.
j,k=l j=l j,k=l

Thus, the pairing (2.14) is degenerate only on the vector space spanned by
elements of the form sa 0 b -s 0 cf>(a)b. Therefore it descends to a positive
definite pairing on the quotient space, which we can rewrite as

(s ® b It ® b') := b* c/>((s I t)) b',

where s ® b, t ® b' aresimple tensors subject to the condition of good defini-


tion. In this way, the quotient space becomes a right pre-C*-module over B.
Weshall now denote by Eq,('E) the completion of this pre-C*-module.
This construction can be generalized to make an "inner tensor product"
of C*-modules and bimodules; weshall need to do that when we come to
the concept of Morita equivalence of C*-algebras in Chapter 4.
Note finally that, if Ais unital, then J{A = J{ ® A === E1 (A), where J{ is
regarded as a C* IC-module and L : I( ..._ Ais the natural embedding into the
multiples of the unit.
At this stage, the reader may have several unanswered questions. Per-
haps the most pressing one, as to whether any finitely generated projective
module over a unital C*-algebra can be made a C*-module, is implicitly
posed again by our last construction. Weshalldeal with C*-modules again
at the start of Chapter 3.

2.6 Line bundles and the Bott projector


For modules over a commutative unital algebra, we can consider the oper-
ation of (outer) tensor product. The tensor product distributes over direct
sums, so the tensor product of two finitely generated projective modules
is another of the same kind. Let P(A) denote the semiring of finitely ge-
nerated projective modules over a commutative unital C*-algebra A, up to
isomorphism: the identity is the A-module A. By Theorem 2.10, P(A) and
Vect(M(A)) are isomorphic semirings. The (isomorphism classes of) mod-
ules in P(A) for which there exists a tensorinverse form a group that we
call the Picard group Pic(A); that is, ['E] E Pic(A) if and only if there is an
A-module :J suchthat 'E ®A :J === A.
Now if 'E::: eAn and :J::: f Am for appropriate idempotents e, f, then

'E ®A :J = eAn ®A f Am= (e ® j)Anm,

where e®j is anidempotent inMn(A) ®AMm(A) = Mnm(A). Then 'E®AJ""


A can happen only if the matrix idempotent Ex (e ® j) = Ex (e) ®Ex (j) has
2.6 Line bundles and the Bott projector 75

rank 1 for all x E M(A). Thus, a necessary condition for [T] to lie in Pic(A)
isthat T have constant rank one; in other words, T = f(M(A), E) for some
(complex) line bundle E- M.
Conversely, if T has constant rank one, consider the dual A-module T~;
then T~ ®AT= [(E*) ®A [(E) is identified, as in the proof of Corollary 2.7,
with the A-module f(M(A),EndE). This rank-one module is freely genera-
ted by the nonvanishing global section x ..... idEx, and so f(M (A), EndE) ""
A. Thus T~ is a tensorinverse forT. In summary, we find that Pic(A) ""
Vecti(M(A)).
Moreover, in view of (2.2), on applying the Bockstein homomorphism of
Section 1.5 we get the following isomorphism.
Proposition 2.19. H 2(M(A),7L) ""Pic(A). E3

In the nontrivial context of commutative Banach algebras, this is Forster's


theorem [448]. Given a discrete abelian group G, an Eilenberg-MacLane
space for G is a path-connected space K ( G, n) suchthat TTk (K (G, n), *) = 0
for k ~ 1, k * n, and TTn(K(G,n), *) = G. We put K(G,O) := G. We know
that rr1 (§ 1 ) = ?L, as follows already from Theorem 1.13 and Hurewicz's
theorem [442, Thm. 7.5.2]: the homotopy classes of maps of the circle to
itself are given by the winding number. Higher homotopy groups of the
circle are all trivial. Therefore § 1 , or the homotopy-equivalent space cx, is
a K(?L, 1). Also, it is weil known that BU(l) = ([poo is a K(7L,2). In view of
Theorem 1.13 and the isomorphism (2.4) for r = 1, we can state:

Hk(M(A),7L)o:o[M(A),K(7L,k)] for k=0,1,2.

Indeed, these isomorphisms hold for all k E N; see [442, §8.1] .


.,. The example of line bundles over § 2 plays a central role in our story.
The complex projective space CP 1 is identified with the Riemann sphere
Coo = ([ l!:! { oo}, by identifying [z1 :zo] with z = Z1 I zo E ([ if zo * 0, and
[1:0] with oo. By writing z = x + iy, we can also identify Coo with the two-
sphere § 2, regarding the latter as a submanifold of IR{ 3 , via the (inverse)
Stereographie projection

h( ) ·= ( 2x -2y -1 + x 2 + y 2 ) (2.15)
z . 1 + x2 + y2 ' 1 + x2 + y2 ' 1 + x2 + y2 '

and h(oo) := (0,0,1), the north pole. (Actually, we make a reflection on


the second coordinate axis of IR{ 3 , in order to getan orientation-preserving
map.)
As such, the two-sphere is described by two charts on C, say UN and Us,
which omit respectively the north and south poles, with the respective local
complex coordinates

z = e-i</> cot ~. (2.16)


76 2. Noncommutative Topology:Vector Bundles

related by s = 1/z on the overlap UN n Us. The formulas


rief> eicf>
dz=-1 e(dfJ+isinfJdcf>), ds= 1 <dfJ+isinfJdcf>)
- cos + cos 8
and those that follow will be usefullater. We write
- 2
q == 1 + zz = 1 - cos e
(2.17)

for convenience. The Riemannian metric g and the area form n are given
by

g = dfJ 2 + sin 2 e dc/> 2 = 4(1 + zz)- 2 dz. di = 4(1 + S'()- 2 ds. d(,
(2.18a)
n = sin e dfJ 1\ dcf> = 2i(l + z.z)- 2 dz 1\ di = 2i(l + s()- 2 ds 1\ d(,
(2.18b)

or, in homogeneaus coordinates,


4(z1 dzo- zo dz1) (z1 dio- io di1)
9 = (lzol 2 + lz1l 2 )2 '
Q = 2i(Z1 dzo- Zo dz1) 1\ (Z1 dio- io di1)
(lzol 2 + lz1l 2) 2

.,. Herrnitian line bundles over § 2 correspond to projective C*-modules


over A := C(§ 2) of rank one; they are of the form 'E = pAn where p = p 2 =
p* E Mn (A) is a projector of constant rank 1. We follow the treatment of
Mignaco et al [337]: such a projector could be conveniently expanded in
terms of the identity and the Pauli-Gell-Mann herrnitian traceless matrices,
i.e., a suitable basis for the Lie algebra of the special unitary group SU(n),
say Pa: oc = 1, ... ,n 2 -1}, as

P= i<a + ba .\a),
where a and the ba are real-valued functions on the manifold, here § 2 •
When n = 2, these .\a are just the standard Pauli matrices u1, u2, 0"3,
namely

(7"2 := ( 0i -i)
0 I
(2.19)

For the sphere § 2 , in view of our remarks on the classification problern for
vector bundles, it is enough to consider the case n = 2. Then, apart from
the trivial solutions p = 1 and p = 0, the general solution is obtained as

n1-in2) =-1( 1 +n·u


- -)
1- n3 2 '
2.6 Line bundles and the Bott projector 77

where n is a continuous function from § 2 to § 2 . Wehave thus established,


by algebraic means, the existence of nontrivial line bundles over the two-
sphere, and, along the way, that the set of rank-one projectors in M2(0
forms a sphere.
Any homotopy between two such functions n yields a homotopy between
the corresponding projectors. This amounts to unitary equivalence, as is
further discussed in Chapter 3. Thus inequivalent finite projective modules
are classified by the homotopy group rr2 (§ 2 ), which by Hurewicz's theorem
is equal to 71. -indeed, TTn(§n) ""'71. for all n- the corresponding integer m
being the (Brouwer) degree of the map n. lf f (z) = (n 1 - in2) I (1 - n3) is
the corresponding map on Co after Stereographie projection (2.15), then
m is also the degree of j, while
1 (lf(z) 12 j(z))
p = 1+ lf(z)l 2 j(z) 1 ·

As a representative degree-m map, one could choose j(z) := zm or j(z) :=


1 I zm (notice that these maps are homotopic via a half-turn rotation of the
Riemann sphere araund the diameter through -1 and 1); the map j(z) :=
.zm has degree -m.
Parenthetically: to visualize better the elements of rr2 ( § 2 ), one can try
the following map [176]:
n(fJ,c/>) := (sinfJcosmcJ>,sinfJsinmcJ>,cosfJ).
This of course is not smooth at the poles. lf one insists on achieving diffe-
rentiability, one can get it at the price of an amount of stretching: substitute
for e a suitable function all of whose derivatives vanish at 0 and rr. For
instance,
n(fJ, cJ>) = (sinB(fJ) cos mc/>, sinB(fJ) sin mc/>, cos B(fJ) ),

J:
where
B(fJ) := rr e-1/t(rr-t) dt / forr e-1/t(rr-t) dt.

Returning to the main subject, Iet us exarnine the projector correspond-


ing to the identity map j(z) = z, of degree 1. We get

PB(Z) =q
1 (zi z)
i 1 , PB(oo) = (~ ~), (2.20)

the celebrated Bott projector. In general, if m > 0, suitable projectors for


the modules T(m), T<-ml of degrees ±m are given by
1 (zmz:m zm)
Pm (z) = 1 + zm.zm .zm 1 '
1 (zm.zm .zm)
P-m(Z) = 1 + zmzm zm 1 .
78 2. Noncommutative Topology:Vector Bundles

They all have trace 1, as befit rank-one projectors.


One can identify 1"(1) with the space of sections of the tautologicalline
bundle L - CP 1 , the fibre at [v] E CP 1 being the subspace Cv of C2 :

PB ( W1)
W2
= ("Z) .
\
n
' Wlt
h 1
1\
= 1 + W2
ZW1
+ZZ
- .

The projector corresponding to the dual of the tautological bundle, usually


called the hyperplane bundle H- CP 1 , is P-1· Now p_I(z) = PB(i); since
i and -1/i are homotopic (turn the sphere araund the diameter from -i
to i), and p( -1/ i) = 1- PB(Z), it follows that f(L) Ellf(H) is trivial, and that
L Eil H is homotopic to the trivial bundle of rank two. This can also be seen
by checking that the unitary matrix u = ( _01 ~) satisfies the intertwining
relation up_ 1 u- 1 = 1- PB, and noting that there is an obvious homotopy
from 1 to u through rotation matrices.
Exercise 2.18. Take for ii the antipodal map. What is the corresponding
module? o
Example 2.7. Let us show explicitly that Endc(§2) 'E(m) "" C(§ 2 ) for any pos-
itive integer m (the case m < 0 is similar). If u, v E C(§ 2 ), then

+ v (zm)
Pm (u) = 1zmu
V + zm zm 1 '

so that T(m) is generated, as a C(§ 2 )-module, by the single element (z;) =


Pm (z;'). If A E Endc(§2) 'E(m), there are functions a, b E C(§ 2 ) suchthat

= zmb + a (zm)·
A(zm)
1
=
Pm (b)
a 1 + zm zm 1 '

we may thus suppose that (!) oc (z;), namely b = zma, so that A(z;) =
a(z;). Therefore, A ..... a is an isomorphism from Endc(§2) T(m) onto C(§ 2 ).
Exercise 2.19. Line bundles over the sphere can be obtained by a general
"clutching" construction. Let L 1 , L 2 be trivial line bundles over the north
and south hemispheres. Over the equator § 1 , they are glued tagether by
the map

for some m E 71.. Show how the modules T(m) are identified to the spaces
of sections of the resulting bundles. 0
For future use, it helps to note that the unitary given by

u(z):= 1
~
v1 + zz
(z
1
2.A Projective modules over unital rings 79

intertwines the Bott projector with the constant projector PB ( oo), that is,

u(z)*pB(z)u(z) = (~ ~).
This is of course not a family of unitaries over the sphere (whereupon PB
would be trivial!), as u ( oo) is undefined; but is only an element of the unitary
group of the C*-algebra Cb(!Rl. 2 ).
Camplex vector bundles over the sphere always split into Whitney sums
of line bundles. This can be gathered, for instance, from another general
stability result [257]: from any vector bundle of rank r on a compact n-
dimensional manifold, trivial line bundles can be chipped off as long as
n + 1 :::; 2r. The reason is again homotopy-theoretic: cn \ {0} is (2n- 2)-
connected, so there is enough room to turn and construct a nowhere-van-
ishing section of the bundle. Therefore most relevant information about
vector bundles over the sphere is contained in the previous presentation.
Line bundles on super-spheres have been exploited by Landi [305, 306]
for giving unified algebraic descriptions of monopoles and instantons.
Our definition of the two-sphere is standard; it would be more in charac-
ter to define it instead as (the space associated to) a C*-algebra generated
by three commuting elements a, b, c of norm 1, with a and b positive and
c*c = 4ab. The readerwill have no difficulty in recognizing the corres-
ponding functions. Forthisalgebra tobe isomorphic to C(§ 2 ), it is suffi-
cient to assume, for instance, that it possesses a C* -norm invariant und er
the action of the group SU(2) in a natural sense [364].

2.A Projective modules over unitalrings


In this section, 5\ and 'B will denote arbitrary unital rings.
Definition 2.12. A right module 'E over 5\ is free if it has an .5\-basis, that
is, a set of generators T suchthat any relation t1a1 + · · · + trar = 0, with
t 1 E T and a 1 E 5\, implies that a 1 = · · · = ar = 0. We say that 'E is
finitely generated if it has a finite generating set. A free module is finitely
generated if and only if it has a finite basis.
A right .5\-module homomorphism cj>: 'E - :J is an additive map such
that cj>(sa) = cj>(s)a for s E 'E, a E .Jt; for brevity, we shall say that cJ>
is an 5\-linear map. The set of allsuch .5\-linear maps is commonly called
Hom.Jl ('E, :J), or End.Jl ('E) when 'E = :J.
Free .5\-modules behave essentially as vector spaces. For instance, if :J is
free, an 5\ -linear map cJ>: :J - 'E is determined by its values on a basis, via
c/>(L.1 t 1a 1 ) = L,1 cj>(t1 )a1. However, one should beware of some pitfalls.
For instance, consider the free l-module l. Then 3 ft 2l; nevertheless, 2
and 3 arenot "linearly independent" in the sense of Definition 2.12.
80 2. Noncommutative Topology:Vector Bundles

The Standard free .J\-module with n generators is .J\ n := .J\ Ea ••• Ea .J\
(n times); we may think of its elements as "column vectors" with entries
in .J\. The columns Uj := (0, ... , 0,1, 0, ... , O)t -with t for transpose, of
Course- with the identity in the jth p}ace, form the Standardbasis for .J\ n.
Any finitely generated free .J\-module is of the form :F "" .J\ n for some
n E ~. just by matehing bases.
As a notational device, we shall denote .J\ Ea • • • Ea .J\ organized as "row
vectors" by n.J\; this is a free left .J\-module with n generators. In this book,
we deal mainly with right .J\-modules.
Definition 2.13. We say that a right .J\-module 'Pis projective if, given any
surjective .J\-linear map of right .J\-modules 17: 'E - (j and any .J\-linear
map 4>: 'P - (j, there is an .J\-linear map (jJ: 'P - 'E suchthat 17«/1 = cJ>:

Any free right .J\ -module :F is projective. Indeed, given 17: 'E - (j and
4>: :F - (j, we may, for each element t of a basis of :F, choose St E 'E so
that 17 (St) = 4> (t); this determines (jJ: :F - 'E by (jJ ( t) : = St and .J\ -linearity.
Proposition 2.20. A direct sum EB i 'Pj of right .J\ -modules is projective if
and only if each summand is projective.

Proof. Assurne 'P = EB i 'Pj is projective and that .J\-linear maps 11: 'E - (j,
cPi: 'Pi - (j are given. We define 4> : 'P - (j on finite sums Lj Sj with
si E 'Pj by 4> (Li si) := Li 4> i (s i). As 'Pis projective, there exists a map (jJ =
{«/Jj}: 'P- 'E with 17«/1 = c/>. Clearly,I7«/Ji = cPi· The converse is obvious. D
Proposition 2.21. A right .J\-module 'P is projective if and only if any short
exact sequence of .J\ -modules 0 - 'E - (j ..!I. 'P - 0 splits; this happens if
and only if 'P is a direct summand of a free module.

Proof. Firstly, if 'P is projective, there is an .J\ -linear map (jJ: 'P - (j such
that 17«/1 = lp, thereby splitting the exact sequence:

Secondly, if such sequences do split, and if {Xj}jEJ is a generating subset


of 'P, Iet :F be the free right .J\ -module with a basis {ti} jEJ, and define
17: :F- 'P by E(tj) := Xj and .J\-linearity. Let X:= ker 17; then :F ""'P Ea X,
so that 'Pisa direct summand of a free module. Finally, any direct summand
2.A Projective modules over unitalrings 81

of a free module is projective, by Proposition 2.20. (Note that Xis therefore


also projective.) 0

Example 2.8. Notice that 71. 2 e 71. 3 == 71.6 as 71.6-modules. This shows that there
exist projective modules that are not free.
Example 2.9. Let 'B := Mn (.J't.) be the ring of n x n matrices over .Jt., acting
on the right on the row vectors n.Jt.. As a right 'B-module, 'B = n.Jt. e · · · e n.Jt..
Thus n.Jt. is projective, but is certainly not a free right 'B-module if n > 1.
Proposition 2.22. A right .Jt.-module Pis projective if and only ifit is ofthe
form P = E:f, where :F is a (ree right .Jt.-module and E is an idempotent in
End.J\ (J").
Proof. Consider any split short exact sequence

0 - ' E - : f -'I P - 0 ,
u

with :F free, P projective, and 17 split by O": P - J". The .Jt.-linear map
E := O" 17 E End.Jt (J") satisfies E2 = O" '70" 17 = O" 17 = E, so E is idempotent, and
P"" E:f. The map l.r- Eis also idempotent, and 'E ""ker17"" (l.r- E):f.
Reciprocally, if E = E2 E End.J\ (J"), the right .Jt.-module E(J") is projective
since :F = E:f e (l.r- E)J". 0

~ If 'Eis both a right .Jt.-module and a left 'B-module, with compatible ac-
tions, i.e., if b(sa) = (bs)a whenever s E 'E, a E .Jt. and b E 'B, so that
we can write bsa unambiguously, we say that 'Eisa 'B-.Jt.-bimodule. When
.Jt. = 'B, call it an .Jt.-bimodule.
Definition 2.14. The notation 'E181.JtJ" makes sense whenever 'Eisa right .Jt.-
module and J" is a left .Jt.-module; it is (at least) an abelian group, generated
by simple tensors s 181 t with s E 'E and t E J", subject only to the relations
s 1 181 t + sz 181 t =(SI+ sz) 181 t, s 181 t1 + s 181 tz = s 181 (t1 + tz), and

sa 181 t = s 181 at for each a E .Jt..

1t is clear that 'E 181 .J\ :F is a right 'B-module whenever :F is an .Jt. -'B-bimodule,
and that it is a left C-module whenever 'Eisa C-.Jt.-bimodule. In particular,
if both 'E and :F are .Jt.-bimodules, then 'E 181.Jt :Fis also an .Jt.-bimodule.
We leave the reader to verify that ('E 181.Jt J") 1811l (j and 'E I8I.Jt (J" 1811J (j)
are isomorphic whenever either of the two expressions makes sense, so we
may omit parentheses when concatenating tensor products over rings.
Example 2.10. If Mn(.Jt.) denotes the ring of n x n matrices with entries
in .Jt., then the right .Jt.-module .Jt. n is also a left Mn (.Jt.)-module, and the left
.Jt.-module n.Jt. is also a right Mn(.Jt.)-module. There are obvious bimodule
isomorphisms,
.Jt.n 181.Jt n.Jt. ""Mn(.Jt.) and n.Jt. 181Mn(.J\) .Jt.n "".Jt..
82 2. Noncommutative Topology:Vector Bundles

In this book, the unitalrings we use are generally (complex) algebras, that
is, complex vector spaces with a bilinear multiplication. The discussion of
their modules is the same as before, except that the ring homomorphisms
should be considered algebra homomorphisms, and the ward "additive"
must be replaced by "C-linear".
3
Same Aspects of K-theory

Noncommutative topology brings techniques of operator algebra to alge-


braic topology -and vice versa. lt is relatively difficult to extend the stan-
dard homotopy and (co)homology functors. On the other hand, Atiyah's
K-functor [12] generalizes very smoothly. Weshall continue with that.
The method of rephrasing concepts and results from topology using
Gelfand-Naimark and Serre-Swan equivalence, and extending them to some
category of noncommutative C*-algebras, will recur again and again. How-
ever, already in this chapter a difference of emphasis appears in our treat-
ment: on several occasions, the commutative case is mainly dealt with after
the fact. In other words, we argue that the deeper proofs of some properties
of objects in the commutative world are tobe found in their noncommuta-
tive counterparts. Bott periodicity will provide an outstanding example.

3.1 Endomorphisms of C*-modules


Webegin by further sharpening of our main tool, the C*-modules of Sec-
tion 2.5, looking at the structure of their endomorphism algebras.
Proposition 3.1. Let 'E be a right C* -module over a C* -algebra A. The al-
gebra of adjointable operators EndA ('E) is also a C* -algebra.

Proof. The norm on EndA ('E) is given by

IITII := sup{ IIITslll: lllslll !5: 1 }. (3.1)


84 3. Some Aspects of K-theory

lf TE EndA(T) and s E T, then

111Tslll 2 = ll(s I T*Ts)ll ::5 lllsiiiiiiT*Tslll ::5 IIT*TIIIIIslll 2 , (3.2)

in view of Lemma 2.14 (the Schwarz inequality); thus IITII 2 ::5 IIT*TII ::5
IIT* IIIITII. so IITII ::5 IIT* II. hence IITII = IIT* II as T** = T. Then it also
follows that IITII 2 ::5 IIT*TII ::5 IITII 2 , so (3.1) is a C*-norm.
To see that EndA (T) is complete, notice that if {Tn} is a Cauchy sequence
of adjointable operators, then {T;i} is likewise Cauchy, and the operator
T: T - T : s - limn Tns is adjointable, with adjoint T*: r - limn T;ir,
since
(r I Ts) = lim(r I Tns) = lim(T;ir I s) = (T*r I s).
n n
Then II Tn - Tll - 0, which establishes the completeness. 0

Proposition 3.2. The algebra of A-compact operators End~ (T) on a right


C* -module T is a C* -algebra, which is an essential ideal in EndA ( T).

Proof. The only thing to prove is essentiality. Let T be any nonzero ad-
jointable operator. Since T(TIT) is densein T, there are elements r, s, t E T
with T(r(s I t)) * 0. But T(r(s I t)) = (Tr)(s I t) = ITr)(slt, and so
Tlr)(sl = ITr}(sl * O.WehaveshownthatTEnd~(T) * OwheneverT * 0,
so by Proposition 1.8, End~ (T) is essential. 0

In particular, the algebra of compact operators X(J-f) is an essential


ideal in L(Jf) .
.". We digress for a moment to illustrate how C*-modules show their use-
fulness in C* -algebra theory itself: the multiplier algebra can be defined as
an endomorphism algebra.
Theorem 3.3. Let A be a nonunital C* -algebra. The algebra EndA (A) is an
essential extension of A, maximal in the sense that if B isanother C* -algebra
containing A as an essential ideal, there is a unique injective morphism (rom
B to EndA (A) restricting to the standard isomorphism A "" End~ (A) de-
termined by ac* - la}(cl. Moreover, EndA(A) is, up to isomorphism, the
unique maximal essential extension of A. In summary: .11-i(A) "" EndA (A).

Proof. The aforesaid isomorphism A"" End~ (A) is e- 1 , where 8 is defined


by (2.11). We define ß: B - EndA(A) by ßb(a) := ba, for b E B -as in
Lemma 1.9. Since c*ba = (b*c)*a for a,c E A, we see that ßb is ad-
jointable with adjoint (ßb)* = ßb*; and ß is clearly a morphism of C*-
algebras. If b E ker ß, then bA = 0; as A is essential in B, we conclude that
kerß = 0. It is clear that ß(ac*)(e) = ac*e = la}(cle for a,c,e E A; it
follows that ßiA = e- 1 . Finally, if Bis also a maximal extension, ß must be
an isomorphism of B with EndA(A). 0

Note the following generalization of Lemma 1.9.


3.1 Endomorphisms of C*-modules 85

Corollary 3.4. Let A, B be C*-algebras and 5 a right C* B-module. Let


j: A- EndB(:f) be an injective morphism suchthat j(A):J is densein J.
Then j extends to an isomorphism between .Jvt(A) and the idealizer of j(A)
in EndB(:f).
Proof. The argument of the proof of Lemma 1.9 goes through in full, on re-
placing J{ by 5 and the Hilbert-space scalar product ( · I ·) by the B-valued
pairing ( · I ·) on :J. The analogue of (1.7) shows that (j(b)r I j(Ta)s) =
(j(T*b)r I j(a)s) for r,s E :J, so that j(T) is an adjointable map in
End8 (:J), with adjoint j(T*). D
The following theorem, due to Green [214, Lemma 16), and rediscovered
by Kasparov [276), ties tagether even more neatly the multiplier and endo-
morphism algebras.
Theorem 3.5. If'E is a right C* A-module, then EndA ('E) "" .Jvf(End~ ('E) ).
Proof. The idealizer of End~ ('E) in EndA ('E) is EndA ('E) itself. To apply
Corollary 3.4 with j being the inclusion map End~ ('E) ..._ EndA ('E), it suf-
fices to checkthat End~ ('E)'E is a dense subspace of 'E; but we already know
that End~0 ('E)'E = 'E('E I 'E) is densein 'E. D

~ In particular: EndA(An) "" .Jvf(Mn(A)) "" Mn(:M(A)). Also, in the case


A = C and 'E = J{ (an infinite-dimensional, separable Hilbert space), we
conclude that L(J-f) "" .Jvf(X), where X := X(J-f) denotes the compact-
operator algebra (namely, the closure in L(J-f) of the subalgebra of opera-
tors of finite rank). To take the next step, we must introduce an appropriate
C*-algebra of "infinite matrices" over A, that is to say, the tensor product
of X with A; remernher that Mn (A) "" Mn ( C) ® A.
Definition 3.1. If Ais a C* -algebra, the tensor product X ®Ais the comple-
tion of the algebraic tensor product X 0 A in the (unique) C*-norm that it
carries (see Section LA). Weshall write As := X ®A and call this C* -algebra
the stabilization of A. Note that Mn (X) "" X and X ® X "" X (hence the
terminology). A C*-algebra Bis called stable when Bs ""B; the C*-algebras
B and C are said to be stably equivalent when Bs "" Cs.
Exercise 3.1. Show that End~ (J{ ® A) "" Endg (J-f) ® End~ (A) = As and
therefore EndA (J{ ® A) "" .Jvt(As ). 0
Definition 3.2. A unitary operator between two right C* A-modules 'E and
5 is an element U E HomA('E,:f) suchthat U*U = h and UU* = 1.r.
lf such a unitary exists, we say that 'E and 5 are unitarily equivalent C*-
modules and we write 'E::::; :J.

~ To complete the discussion of the questions left over from Section 2.5,
in particular the compact endomorphisms of C(M)-modules, we next look
at projectors on C*-modules. What makes the theory of projectors on a
86 3. Some Aspects of K-theory

Hilbert space comparatively easy is the close link with Hilbert subspaces.
Recall that a closed submodule :f of a C*-module 'Eis not (orthogonally)
complemented in general. But suppose that :f e :f.l. = 'E does hold. Then,
for any s E 'E, we can write s = t + u uniquely with t E :f, u E ;:.1.. Also,
s ...... t defines a projector (i.e., a selfadjoint idempotent) p E EndA ('E) with
range :f.
Exercise 3.2. Conversely, assume that p in EndA ('E) fulfils p 2 = p and
p = p*. Then prove that im(p).l. = im(1x- p) = kerp. 0
It follows that im(p) is complemented. Thus the complemented C*-
submodules of 'E are precisely the rang es of projectors in EndA ('E).
Definition 3.3. Let A be a unital C*-algebra. The A-compact projectors
on:JfA formasubset ofthe C*-algebraEndi (.1{A) = A 5 ; denoteitby P(As).
By Exercise 3.1, this is a norm-closed subset of the unit ball of As, since
p = p*p implies that IIPII = 1 or 0.
Among the elements of P(As) are found the projectors Pn given by (2.10).
As A is unital, they belang to Endi0 (.1{A) c Endi (.1{A) = As:
n
Pn = L iuJ) (UJ I with UJ := (0, ... , 0, 1, 0, ... ), 1 in the jth place,
j=l

and they forma countable approximateunit for As, in view ofExercise 2.15.
Indeed, if x = lr)(sl with r,s E :JfA, then llx- Pnxll = II Ir- Pnr)(slll s;
lllr- Pnrllllllslll - 0 as n- oo, and it follows that llx- Pnxll - 0 for any
X EAs.
Remernher the question: can an algebraically finitely generated and pro-
jective right module over a unital C*-algebra A be endowed with the struc-
ture of a C*-module over A? Conversely, do the elements of P(As) give rise
only to finitely generated projective modules over A, or do they include in-
trinsically more complicated objects? We answer the second question first:
it turns out that members of P(As) correspond to finitely generated pro-
jective modules only, and that all finitely generated projective A-modules
are described by elements of P(As)! We need a preliminary lemma, which
will be useful on other occasions, too.
Lemma 3.6. If p, q are two projectors in a unital C* -algebra B such that
II p - q II < 1, then there is a unitary u E B such that q = upu *.
Proof. Consider the symmetries 2p - 1, 2q - 1 in B: these are unitary
selfadjoint elements of B with spectra in {-1, 1}. Define r E B by 2r :=
(2q- 1)(2p- 1) + 1. Then qr = qp = rp, and
r*rp = r*qr = (qr)*r = (rp)*r = pr*r,
so p commutes with lrl 2 and also with Ir I. Now r is invertible in B since
llr-111 = ll2qp-q-pll = ll(q-p)(2p-1)11 s; llq-pll < 1.
3.1 Endomorphisms of C*-modules 87

Take u := rlrl- 1 ; this is unitary in Band satisfies upu* = rplrl- 2 r* =


qrr- 1 = q. D

Theorem 3.7. Let A be a unital C* -algebra. The right C* -modules pJ-fA, for
p E P(As ), are algebraically finitely generated and projective (equivalently,
they are isomorphic as right modules to direct summands of An for some
positive integer n).

Proof. First, we prove that for every p E P(As) there is a unitary Un in


EndA(J-fA) suchthat UnPU~ :5 Pn; moreover, Un - 1 in norm, as n- oo.
For that, we first note that for any given E > 0, we can find n such that
IIP- PnPPnll < E/3. Then the operator an := PnPPn is positive, indeed
0 :5 a~ :5 an in End~ (J-fA), and

If E < ~. then the spectrum of an lies in [0, 1] with a gap at ~. in fact


span c [0, 2E] w [1- 2E, 1].
Let Pn be the spectral projector of the upper interval [1 - 2E, 1]; then
Pn = j(an) for any positive continuous function f on [0, 1] such that
j(t) = 0 for 0 :5 t :5 2E and j(t) = 1 for 1 - 2E :5 t :5 1. Thus Pn E P(As ),
with IIPn-anll < 2E. Then IIPn-PII < 3E < 3/4.ByLemma 3.6, Pn = UnPU~
for some unitary Un in the Unital C* -algebra EndA (J-{A).
Finally, Pn = j(an) :5 Pn: indeed, by approximating f by polynomials
without constant terms, we see that Pnan = anPn = an entails PnPn =
PnPn = Pn· In consequence, Pn - Pn is a projector in P(As ). Now, as
PnJ-{A :::Anis free and finitely generated, PnJ-{A e (Pn- Pn)J-{A = PnJ-fA,
and s - Uns is a unitary equivalence of A-modules from pJ-fA to PnJ-fA;
thus pJ-{A is isomorphic to a direct summand of An, and so is finitely ge-
nerated and projective. D

Clearly, the finitely generated projective modules over A are closed sub-
modules of J-{A: they can be seen as images of idempotents in EndA (An)
and we identify An to Pn(J-fA). The converse of Theorem 3.7 hinges on
the question of whether the idempotent is adjointable. For more general
C* -modules, although it happens that the image of any idempotent q is a
complemented submodule, it is not true that q is always adjointable. These
questions are very well discussed in Chapter 3 of [303). Happily, in our case,
adjointability of the idempotent is guaranteed.
Theorem 3.8. Let 'E be an algebraically finitely generated and projective
right module over a unital C* -algebra A. Then it can be endowed with a
structure of C* -module over A, in such a way that it is isomorphic to pJ-{A
(or some p E P(As ).

Proof. The argument boils down to little more than linear algebra. First of
all, 'E = eAn, where e is an idempotent in Mn(A). As the range space of an
88 3. Some Aspects of K-theory

idempotent is closed, the standard C*-module structure on An restricts to


a C*-module structure on 'E. Then, as in linear algebra, e is adjointable: if
{u1, ... ,un} is the standard basis of An, so that 1A" = Ij= 1 1uj)(ujl. we
can write e as I}= 1 1euj)(ujl. and then e* = Ljluj)(euil· Now (ime).L =
kere* since e* s = 0 if and only if (er I s) = (r I e* s) = 0 for all r E An;
also, (kere*).L = ime since s E (kere*).L if and only if (r 1 s- es) =
(r- e*r I s) = 0 for all r, if and only if s =es. Likewise, (ime*).L = kere
and (kere).L = ime* since e* is idempotent.
Now kere = ker(e*e) since e*es = 0 implies (es I es) = (s I e*es) = 0 and
therefore es= 0. Taking complements, we get ime* = ime*e, so if s E An,
then e* s = e*et for some t E An, so we can write s = et + (s - et) E
ime + kere* = ime + (ime).L. Wehave thus shown that An= 'E e 'E.L, so 'E
is complemented. Therefore there exists a projector p E Mn (A) such that
'E = pAn. Since An"" Pn(J{A) and p ~ Pn under this identification, we can
also write 'E = pJ{A·
There is an explicit formula for p in terms of e, due to Kaplansky [269,
Thm. 26]. Consider r := ee* + (1- e*)(1- e) = 1 + (e- e*)(e*- e) =
1- e*- e + ee* + e*e. This is an invertible positive element of Mn (A), since
elements of the form 1 + a * a are always positive and invertible. Moreover,
it commutes both with e and e*: indeed, re =er= ee*e andre* = e*r =
e*ee*. Therefore r- 1 also commutes with e and e*. Now take p := ee*r- 1 ;
it is clear that p = p*, and p 2 = ee*ee*r- 2 = ere*r- 2 = p. Furthermore,
ep = p obviously, and pe = e since per = ee*e = er, so p is the range
projector of e. D

We conclude that End~([(M,E)) = EndA([(M,E)) "" f(M,EndE) for


finite-dimensional (made hermitian) vector bundlesE- M.
Notice that the Kaplansky formula for p holds in any C*-algebra, indeed
in any involutive Banachalgebra for which elements of the form 1 + x* x are
invertible. (The algebra A need not be unital, since if not, the construction
can be made in A + .) Also, the element 1 + e - p is invertible, with inverse
1-e+p; and ( 1 +e-p )e(1-e+p) = p. Hereisamore pictorial version ofKa-
plansky's argument [403], which clarifies the relation between idempotents
and projectors. Assurne that the C* -algebra A is faithfully represented on
a Hilbert space Jf; let J{e be the (closed) range of e. With respect to the
Hilbert space decomposition J{ = Jfe e Jf/, we can write

e= (~ ~)' e * = ( T*
1
~)' ee * = (1 + TT*
0
~)' (3.4)

where T : Jf/ - Jfe is a bounded operator. Now, Jfe = eJf = ee* J{


since 1 + T*T is invertible on Jfe, and sp(ee*) c {0} u [1, oo). Thus the
range projector for e is given by p := j(ee*) where f is any continuous
function suchthat j(O) = 0 and j(t) = 1 fort ~ 1; by functional calculus,
it is clear that p belongs to the same C*-subalgebra as e. Moreover, the
3.1 Endomorphisms of C*-modules 89

presentation (3.4) allows us to write Kaplansky's r and p in the form

r = ( 1 + 0TT* 0 )
1 + T*T ' p = (~ ~)' (3.5)

from which the invertibility of r and the equality im p = im e are transpar-


ent.
Exercise 3.3. If p is obtained from the idempotent e by Kaplansky's for-
mula, show that (1 - t)e + tp is idempotent for 0 =:; t =:; 1. Conclude that
the set of projectors in a C*-algebra is a deformation retract of the set of
its idempotents. 0

Exercise 3.4. An obvious equivalence relation on idempotents is similarity:


e- e' if and only if e' = xex- 1 for some invertible x (in A+, if necessary).
Using a := 1 + (2e* -1)(2e -1), find a projector q similar both to e and e*.
Show also, using a(t) := 1 + t(2e* - 1)(2e- 1) for 0 =:; t =:; 1, that e, q
and e* are homotopic via idempotents. o
Exercise 3.5. If two projectors are homotopic via idempotents, show that
they are homotopic via projectors. 0

.,.. lt was noted in the proof of Theorem 3.8 that any finitely generated
projective module over a unital C*-algebra Ais of the form pAn for some
projector p E Mn (A). An even more beautiful characterization of finitely
generated projective modules can now be revealed.
Proposition 3.9. A right C* -module 'E over a unital C* -algebra A is finitely
generated and projective if and only if the identity 1'E is an A -compact ope-
rator.

Proof. In one direction, it is trivial: if 'E = pAn for some n and some pro-
jector p E Mn (A), then 1x = L:.f= 1 1 puj) (puj I where u1, ... , Unis the Stan-
dardbasis of An; indeed, if s E pAn, then
n n n
I lpuj)(pujls = I puj(puj I s) = I puj(Uj I ps) = p1AnPS = ps = s.
j=l j=l j=l

Assume, on the other hand, that 1x E End~ ('E). Then End~ ('E) is a uni-
tal C*-algebra -and therefore End~ ('E) = EndA ('E), in view of Proposi-
tion 3.2- and so the dense ideal End~0 ('E) cannot be proper. This means
that 1x E End~0 ('E), so the identity operator is of the form
n
1x = Ilrj)(sjl. with r1. ... ,rn,Sl, ... ,sn E 'E.
j=l
90 3. Some Aspects of K-theory

Indeed, the trivial equality 1x = Lflx allows us to improve this presenta-


tion to
n
1x= Lltk)(tkl, with t1, ... ,tnE'E. (3.6)
k=l

To see that, write 1x = 1Tlx = L~j= 1 l5i (ri I ri)) (5j I and recall that the ma-
trix [ (ri Iri)] isapositive element of Mn (A), so it is of the form [ (n I ri)] =
mm* for some m = [aij] E Mn(A). As a result, (ri I rj) = Lk=l aikajk for
each i,j, so we can take tk := 2:r=l 5iaik·
Now define right A-module morphisms cp: 'E - An and (/J: An - 'E,
by letting c/>(5) be the column b with entries bi := (ti I 5), and setting
(/J(a) := Lk=l tkak. Clearly, cp and !fJ are mutually adjoint and 1/J(c/>(5)) =
Lk=l tdtk I 5) = 5 by (3.6), whereas cp(!fJ(a)) = pa, where p E Mn(A) is
the matrix p = [(ti I tj)]. Also, c/>(5) = pb, so cp('E) = pAn. Then p* = p,
and (3.6) yields p 2 = p, so cp('E) = pAn is a finitely generated projective A-
module. Note also that !fJ(pAn) = 'E, so that cp and !fJ are mutually inverse
unitaries between pAn and 'E. D

The set of I pu i) that give rise to the "reproducing kerne!" Li Ipu i) ( pu i I


clearly do not constitute a basis, since they are not linearly independent in
general. They constitute a (tight, normalized) frame, in the terminology of
wavelet theory. While C* -modules do not in general possess bases, much
of what is known about frames on Hilbert spaces extends to frames on
C*-modules [187].
Corollary 3.10. Ifp i5 an A-compact projector in End~ ('E), then in fact p E
End~0 ('E). Moreover, any idempotent e E End~('E) is actually an A-finite
rank operator.

Proof. The algebra pEnd~('E)p is a unital C*-algebra (whose unit is p it-


self) that contains p End~0 ( 'E) p as a dense ideal. Since this ideal cannot be
proper,wefindthatpisoftheformp = LJ=l plrj)(5jiP = 2:}= 1 1prj)(psjl.
so it is of A-finite rank. More generally, if e is an A-compact idempotent,
Kaplansky's formula provides an A-compact projector p suchthat pe = e,
so that e E End~0 ('E) also. D

Notice that for this corollary, A is not required to be unital. Actually, if


Ais a nonunital C*-algebra, and if 'E is a right C* A-module, the proof
of Proposition 3.9 still shows that 'Eis of the form pAn for some n and
some projector p E Mn (A) if and only if 1x is A-compact; in particular,
'E is finitely generated in that case. However, 'E will not be complemented
in An, since there is no complementary projector ln- p, so we cannot
conclude that 'Eisa projective A-module, although it is certainly projective
over A +. Even so, the conclusion justifies the introduction of the following
terminology [171], which will come tobe usefullater on.
3.1 Endomorphisms of C*-modules 91

Definition 3.4. Let A be a C*-algebra that is not necessarily unital. We say


that a C* A-module 'Eis of A-finite rank if lT E End~ ('E).
In particularl if A = C0 (M) is a commutative C*-algebral the spaces of
sections of vector bundles on M that aretrivial near infinity are of Co(M)-
finite rank; this characterization generalizes the Serre-Swan theorem.
.,. There are several equivalence relations that one may define on the set
of projectors in a C*-algebra A. If A is unitall similarity a ..... zaz- 1 pre-
serves idempotents but not projectorsl unless the conjugating element z
is unitary. Since finitely generated projective A-modules are determined by
projectors in P(As ) we may consider unitary equivalence within the unital
1

C*-algebra EndA(JfA).
Definition 3.5. Call two projectors p q E P(As) equivalent p - ql if and
1 1

only if q = upu-I for some unitary u E EndA(JfA). Call two projectors


r~ q E P(As) homotopicl provisionally written p!!. ql if and only if they are
connected by a norm-continuous path of projections.
Exercise 3.6. Construct a deformation retraction from the set of invertibles
Ax of a unital C*-algebra A to its set of unitaries 'U(A). o
In view of the proof of Lemma 3.6 and this exercisel homotopy equiva-
lence implies unitary equivalence. The converse implication depends on
being able to "border with zeroes" in P(As ).

Lemma 3.11. I(p- ql then (6 ~)!!. (ci ~); there(orel there is no need
to distinguish between p !!. q and p - q in P(As).

Proof. Say q = upu* I with u unitary. Consider the norm-continuous paths


of unitaries

Vt.-
._ (cos :g: t
. rrt
sm
- sin :g:
I!. t
t) and Vt := Vt (~ ~) V-t (~ :*). (3.7)
2 COS
2
I

The latter goes from ( ~ :*) to the identity. Then the unitary path t .....

Vt ( 6 ~) V -t goes from ( ci ~) to ( 6 ~). D

Proposition 3.12. The set o( equivalence classes V10P(A) P(As)/-


[<Cl As] is a commutative unital semigroup.

Proof. Let p q E P(As ). For conveniencel we shall denote by p EB q their


1

matrix direct sum:

P EB q := ( 6 ~) ·
92 3. Some Aspects of K-theory

The addition in V10P(A) is defined by the rule [p] + [q] := [p ED q]. This is
well defined, since upu- 1 ED vqv- 1 = (u ED v) (p ED q) (u- 1 ED v- 1 ) if u and
v (and thus also u ED v) are unitary in EndA (J{A). Moreover,

(6 ~)-(~ ~)(6 ~)(~ ~)=(ci ~)·


The unit of V10P(A) is obviously the trivial dass [0]. 0

lt is time to introduce K-theory.

3.2 The Ko group


The K-, KK- and E-theories of C*-algebras marry the methods of functional
analysis and algebraic topology. They are essential in the formulation of
index theory in noncommutative geometry. Here we examine the first of
these. As pointed out in [481], the very definitions and the proofs of basic
facts in K-theory tend to drown in matrix technicalities. Wehave tried to
avoid that, insofar as possible, partly by full use of the lessons about C*-
modules learned so far, partly by referring the reader in a couple of places
to Chapter 6 of that reference -for outright borrowing of its arguments.
We recommend that book by Wegge-Olsen for the foundations of K-theory
of C*-algebras. Our purposes are different, however. As stated in the in-
troduction, we want to regard K-theory as yet another meeting ground of
the commutative and noncommutative worlds.
Any commutative unital semigroup S determines an abelian group K,
called its Grothendieck group, together with a unital semigroup homomor-
phism 9: S - K such that the pair (K, 9) is universal in the following
sense: whenever G is a group and y: S - G is a unital semigroup homo-
morphism, y factors uniquely through (K, 9), i.e., there is a unique group
homomorphism K: K - G with y = K o 9:

By this universal property, K is unique up to isomorphism. lt may be con-


structed as the set of equivalence dass es [x, y] E (S x S) I- where (x, y) -
(x', y') inS x S if and only if x + y' + z = x' + y + z for some z E S. Define
9(x) := [x,O]; then [x,y] = 9(x)- 9(y) in K.

Definition 3.6. The zeroth K-theory group of the unital C*-algebra A, de-
noted K~0 P(A), is defined as the Grothendieck group of V10P(A).
3.2 The Ko group 93

Let .J\. be a unital ring. The algebraic K 0 -group of the ring .J\. denoted 1

Kg1g(.J\.) is defined as the Grothendieck group of the (direct sum) semi-


1

group of isomorphism classes of finitely generated projective right mod-


ules over .J\.. (Working with left modules insteadl one ends up with the same
object.)
Exercise 3. 7. Prove that the Grothendieck group of the semigroup of iso-
morphism classes of countably generated projective right .J\.-modules van-
ishes. (This is known as the "Eilenberg swindle" [403].) 0
There are two definitionsl thenl for the zeroth K-theory group of a uni-
tal C*-algebra. The relatively painlessl purely C*-theoretic definition in 1

terms of A-compact projectors on J{A1 that isl projectors in As1 elegantly


underlines invariance under stabilizationl which is one of the trademarks
of K-theory of C*-algebraS or topological K-theoryl as is most often said
1

nowadays. The general algebraic definition (the one usually found in texts)
is also a most elegant onel but it is awkward in operational terms. What is
needed is an approach based on idempotent matricesl using some equiva-
lence relation between idempotent matrices of different sizes. The problern
is solved by the following workhorse lemma.
Lemma 3.13. Let e E Mn(.J\.), f E Mm(.J\.) be matrix idempotents over a
unital ring .J\.. The corresponding finitely generated projective modules e.J\. n
and f .J\. m are isomorphic if and only if, after eventually enlarging the matrix
sizes of e and f by bordering with zeroes at the right and below, one can
find an invertible matrix a E MN(.J\.) suchthat a(e Gl ON-n )a- 1 = f Gl ON-m·
Proof. The condition is necessary: for suppose </>: e.J\. n - f .J\. m is an iso-
morphism. Construct module maps 1/J : .J\. n - .J\. m and 17: .J\. m - .J\. n given
respectively by extending cf> by 0 on (1 - e) .J\. n and cf> - 1 by 0 on ( 1 - j) .J\. m.
Then !Jl(s) = gs and 17(t) = ht where g E Mm,n(.J\.) and h E Mn,m(.J\.) are
suitable matrices over .J\.. Note the relations gh = f~ hg = el g = ge = f g
and h =eh= hf. Take now N := n + m and compute:

1-
h
!) (1-f
h 1- e) = (1
g 0
0)
1 I

and

(1~e 1~!)(~ ~)(1~f 1~e)={~ ~)·


The condition is sufficient: for if aea- 1 = f~ then ae.J\.N = fa.J\.N. D
Denote by Qn (.J\.) the set of idempotents in the matrix algebra Mn (.J\.) 1

and write GLn (.J\.) for the group of invertible elements in Mn (.Jt). There are
canonical identifications

Mn(.J\.) .... Mn+d.J\.): m ,._ (7: ~) 1


94 3. Some Aspects of K-theory

and likewise Qn(.Jl) .... Qn+d.Jl): e ,_ e EB 0. However, for invertible matri-


ces these will not do; instead we identify

Define
00

U GLn(.Jl).
00

U Qn(.Jl);
00

Moo(.Jl) := UMn(.Jl); Qoo(.Jl) := GLoo(.Jl) :=


n=1 n=1 n=1
Definition 3.7. Call two idempotents e,f E Qm(.Jl) equivalent, e - f, if
and only if they are conjugate via some v E GLoo (.Jl); that is, for some
n E N there is a v E GLm+n (.Jl) suchthat

V
0) _ (! 0)
( 0e On V
1
= 0 On · (3.8)

The addition on the quotient Qoo (.Jl) I- is defined by the rule

[e]+[fJ:= [(~ ~)] = [(~ ~)].


which is weil defined since e EB f - f EB e. Therefore valg(A) := Qoo (.Jl) I-
is a commutative semigroup, and the algebraic K-theory group Kg1g(.Jl) is
its Grothendieck group.
Now we areready to relate both definitions.
Theorem 3.14. There is an isomorphism Kg1g(A) ""K~0P(A) for any unital
C* -algebra A.

Proof. lt only remains to check that both equivalence conditions coincide.


Suppose e - f in Qoo (A). Then, for a suitably large n, the right A-modules
eAn and f An are isomorphic. By Theorem 3.8, there are projectors p, q E
P(As) so that eAn = pJ{A and !An= qJfA, where indeed p,q E Mn(A)
by (3.5). Thus p - e - f - q in Qoo (A), and so q = zpz- 1 for some z E
GLoo (A). In order to show that pJ{A "" qJfA, it is enough that q = upu- 1
for some unitary endomorphism u of J{A; in fact, if z = u[zl is the polar
decomposition of the invertible z in the C* -algebra EndA (J{A), then u is
unitary and lzlplzi- 1 = u*qu = (u*qu)* = [z[- 1p[z[, so p commutes
with [z[ 2 and thus also with [z[, so that indeed q = zpz- 1 = upu- 1 .
Conversely, if p,q E P(As) satisfy pJ{A "" qJfA. so that q = upu*
for some unitary u, then by the proof of Theorem 3.7 we can find n E N
and projectors Pn• qn E Mn (A) such that Pn - p - q - qn by unitary
conjugations, so that qn = Vpnv* with v unitary in EndA(J{A).
If it happens that v E GLm (A) for some m ~ n, then qn and Pn are
similar as elements of Qoo(A), so the modules pJ{A and qJ{A lie in the
samedass in Kg1g(A).
3.2 The Ko group 95

If not, then in any case we can suppose that v E A~, since v = UnUVn
where Pn = UnPU~, qn = v;rqvn and Un, Vn may be constructed via
Lemma 3.6 to lie in GLn(A) c A~. Thus, given E with 0 < E < 1/3, we
can find m 2: n, some w = PmwPm suchthat w*w = ww* =Pm and
some ,\ E C with 1"-1 = 1 so that llv - (w + ,\(1 - Pm))ll < E. Then
Cln := wpnw* E Mm(A) is a projector suchthat

llqn- 4nll = li(v- w)pn(V- w)* + WPn(V- w)* + (v- w)pnw* II


< E2 + 2E < 1,

and so there is a unitary Wm E Mm (A) such that qn = WmClnW~, by


Lemma 3.6. Now Zm := WmW E GLm(A) satisfies qn = ZmPnz;,l, from
which we conclude that [qn] = [Pnl in valg(A). D

From now on, we shall write simply V (A), Ko (A) to denote the (algebraic
or topological) V -semigroup and Ko-group of a C* -algebra A. In general,
the unital semigroup V(A) does not admit cancellation; that is, [p] + [q] =
[p'] + [q] does not imply [p] = [p']. Of course, the canonical map from
V (A) to Ko (A) will be injective only if V (A) is a cancellation semigroup, and,
in principle, tobelang to the same element of Ko (A) is a weaker equivalence
than projective module isomorphism. But the eventuallass of information
is worth the gain in enhanced algebraic agility. The dassie example is given
by the tangent bundle over the sphere: as a vector bundle, T§ 2 - § 2 is not
trivial, but [T§ 2 ] is trivial in (real) K-theory. Among the complex bundles,
it is known that V(lr 5 ) := V(C(l!" 5 )) does not admit cancellation [36].
The correspondence A ..... V(A) is a functor from the category of unital
C* -algebras to the category of commutative unital semigroups. Let cp: A -
B be a unital morphism. Recall that, if 'E is a finitely generated projective
C* A-module, then EcJ>('E) is a finitely generated projective C* B-module.
If 'E = pAn, then EcJ>('E) = cp(p)Bn, where cp is extended entrywise to a
unital morphism cp: Mn(A) - Mn(B). Notice that if p is a projector, then
so is cp(p); and if p - q, then cp(p) - cp(q). Moreover, to the composi-
tion of C*-algebra morphisms corresponds the composition of semigroup
homomorphisms.
Any element of Ko(A) is of the form [p]- [q], with p,q E P(As). We
write Kocf> for the map [p]- [q] ..... [cp(p)]- [cp(q)]. In other words, Kocf>
is induced by id ® cp: X ® A - X ® B.
Proposition 3.15. A ..... Ko(A) is a (covariant) functor from the category of
unital C* -algebras to the category of abelian groups. E3

Exercise 3.8. If Ais unital, show that each element of Ko(A) can be written
as [p]- [Pn]; moreover, [p] = [q] in Ko(A) if and only if p EB Pn - q EB Pn
for some n. o
Exercise 3.9. Prove that Ko (A) is countable if A is separable. 0
96 3. Some Aspects of K-theory

Exercise 3.10. Let pr 1 , pr 2 be the projections of the C*-algebra A 1 eA 2 onto


its summands; show Kopr 1 eKopr 2 is an isomorphism from Ko(AI e A2)
onto Ko(AI) e Ko(A2). 0

As for the examples, we point out first that the only invariant of a pro-
jective module over a field IF is its dimension, i.e., the rank of the corres-
ponding idempotent, which is given by the trace if the characteristic is zero.
Therefore Kg1g(IF) = 71.. The only C*-field is ([ itself, and we seealso from
the "topological" definition that Xo(O = 71.; indeed, V(([) = f\::1, since all
projectors in X are of finite rank (the unit ball of the range space must be
compact, hence finite-dimensional). On the other hand, since two projec-
tors in L(Jf) are equivalent if and only if they have the same rank, then
V(L(Jf)) = ~ u {oo}; from that, Ko(L(Jf)) = 0 is clear.
Now Ko(Mn(O) = Ko(C) = 71., since Mn(([) ®X "" X, or simply be-
cause all rank-one projectors in MN(([) are equivalent, so Ko(Mn 1 (([) e · · · e
Mn, (([)) = 71. e · · · e 71. = 71.r. The group Ko (A) provides a complete isomor-
phism invariant for a certain dass of C* -algebras, called "almost finite-
dimensional" algebras, or AF -algebras for short. An AF -algebra is, by defi-
nition, a C*-algebra having a dense subalgebra that is an increasing union
of finite-dimensional algebras; for instance, Xis the closure of Moo (([). Now
a finite-dimensional C*-algebra is just a direct sum of full matrix algebras,
A ""Mn 1 (([) e · · · e Mn, (C); but such algebras may be nested inside larger
ones in very many ways, and many nonisomorphic AF-algebras are thereby
obtained. There is a simple way to describe such nestings, called a Brat-
teli diagram [48], by specifying which matrix blocksofafinite-dimensional
subalgebra fit into which blocks of larger finite-dimensional subalgebras,
and this diagram determines the isomorphism type of the full C*-algebra
-but not conversely. Several interesting examples are worked out in [163).
By Propositions 3.18 and 3.19 below, the Ko-group of any AF-algebra is the
direct limit of such 71.r groups, with the corresponding nesting. Such Ko-
groups are ordered groups, whose positive semigroup is generated by the
set of classes of projectors { [p] : p E P(A)} (this set is called a "scale"
for the positive semigroup). The isomorphism invariant for the AF -algebra
A, according to Elliott's theorem [156), is the scaled, ordered group Ko(A).
This is thoroughly discussed and proved in [129, Chap. 4).
A particular example of an AF-algebra that models an interesting non-
commutative space is the classification algebra of the quasiperiodic Pen-
rose tilings of the plane. For a beautiful discussion of the tilings them-
selves, we refer to [225]; the relevant point hereisthat every finite patch
of tiles of one tiling occurs also in any other tiling, so that the ordinary
classifying space has an indiscrete topology. Connes [91, Il.3) explains how
to describe the set of tilings, up to local isomorphism, by an AF-algebra
that is the C*-inductive limit of a nested sequence of two-block algebras
Mr(C) e M5 (([) ...... Mr+s(C) e Mr(C) where r ~ s, starting from the alge-
bra ([ e C; the nth stage is therefore MFn+J (([) e MFn (([), where Fn is the
3.2 The Ko group 97

nth Fibonacci number. The details of the calculation of Ko(A) are given
in [129, IV.3], [163] and [304, §5.2]; the result isthat Ko(A) = 71. 2 , where
the positive semigroup lies on one side of the line y = - cf>x in ~ 2 , with
cf> = i<1 + .JS) being the golden ratio. The map (r,s) .... r + cp- 1 s is an
isomorphism of scaled ordered groups from K 0 (A) to 7l. ffi cf> -I 71..
Have a lookalso at the catalogue on page 123 ofWegge-Olsen [481]. Note
that there are cases for which Ko (A) has torsion (and so cannot be ordered).
This may also happen with commutative algebras.
~ We want now to extend the definition of Ko to nonunitat C*-algebras.
Would the same definition as the Grothendieck group of classes of projec-
tors work? It certainlyworks for the nonunitat algebra X, yielding Ko(X) =
7l. again. Same reflection shows that the same definition would work for any
algebra whose stabilization enjoys a countable approximate unit consisting
of projectors [36, §5.5]. Sadly, however, we realize that V(Co(X)) = 0, for
any connected, locally compact, but noncompact X, since any continuous
function from X to matrix projectors that vanishes at infinity must vanish
everywhere.
Therefore, we must take a more "functorial" tack. Whether A is unital
or not, we augment it. Recall the homomorphisms E: A + - A +I A "' ([ and
a: ([ - A+ : ,\ .... (0,,\) and the split exact sequence (1.12). Let Ko(A+)
-read: "reduced Ko of A+"- denote the kernel of KoE. Note that KoE is
surjective, since KoE o Kot = Ko(E o L) is the identity. The corresponding
sequence of Ko-groups
0-Ko(A+) -Ko(A+) -Ko(O- 0

splits, and so K 0 (A +) "' Ko (A +) ffi 71.. Therefore, if A is already unital, then


by Exercise 3.10, Ko(A) "'Ko(A+ ).
Definition 3.8. The group K 0 (A) is defined as Ko (A +) in all cases.
Note that the element of [p]- [Pn] of Ko(A+) actually lies in Ko(A) if
and only if the rank of the matrix E ( p) is equal to n.
Proposition 3.16. A .... Ko(A) is a functor from the category o(C*-algebras
to the category of abelian groups.
Proof. Fora morphism cf>: A-B, consider the following diagrams:

(3.9)

where cp+ (a + f.l) := cp(a) + f.l. The left handdiagram commutes, hence so
does the triangle of the right hand diagram. Therefore, Kocf>+ o }A maps
98 3. Some Aspects of K-theory

Ko(A) into ker KoEB = im}B, so there is a unique group homomorphism


Ko</>: Ko(A) - Ko(B) suchthat Ko</>+ o }A = }B o Ko</>; in other words,
Ko</> makes the right hand diagram commute. lt is now Straightforward to
check that if f.JJ: B - C is another morphism, then Ko ( f.JJ o </>) = Ko f.JJ o Ko </> :
Ko(A) - Ko(C). D

.,. One of the most striking features of K-theory isthat it lends itself to an
axiomatic treatment: the functor Ko satisfies a short Iist of characteristic
properties, that also hold for the other functors of the theory (we shall meet
them a little later on).
Definition 3.9. A functor H from C* -algebras to abelian groups is called
halfexact if, given the short exact sequence of C* -algebras

j II
0-]-A-A/]-0,

the corresponding sequence of abelian groups is exact at H(A):

H(J) !!J.. H(A)!!.!!. H(AJ]). (3.10)

Definition 3.10. A functor H from C*-algebras to abelian groups is called


a K -theory functor if it has the following properties:
(a) It is normalized: either H ( (() = 7L or H ( (() = 0;

(b) It is homotopy-invariant: if A and B are homotopy equivalent, then


H(A) ""H(B);

(c) ltisstable:H(As) =H(A);

(d) lt is continuous: it commutes with inductive Iimits (defined below);

(e) lt is halfexact: if J .____.A-B, then H(J) - H(A) - H(B) is exact


at H(A) .

.,. A construction that yields new C* -algebras from old is the inductive Iimit
of a directed system ofC* -algebras {AJ, f.JJkJ }, indexed by some directed set.
Here the A 1 are C*-algebras and morphisms f.JJkJ: A 1 - Ako for k ~ j, are
given, that satisfy f./Jzk o f.JJkJ = f.JJZJ whenever l ~ k ~ j.
Definition 3.11. Given such a directed system, form the Cartesian product
0 1 AJ and consider the subset .Jt c 0 1 AJ consisting offamilies a = {aJ}
such that, for some index j, ak = f.JJkJ(a 1 ) for all k ~ j; consequently,
az = f./Jlj(aJ) = f.JJzk(ak) for all l ~ k ~ j. lt is easy to checkthat .Jt is
an involutive algebra. Since the morphisms f./Jzk are norm-decreasing, the
Iimit limz II f./Jzk (ak) II is finite, but could be zero. The subset N of families
for which this Iimit is zero is an involutive ideal in .Jt; the Iimit defines a
3.2 The Ko group 99

C*-norm on the quotient .Jt/ N, whose completion is a C*-algebra, denoted


by A = li!!}.AJ and called the C* -inductive Iimit of the system [267, §ll.4].
Define morphisms !J.IJ: AJ - A by !J.IJ(aj) := a+N, where ak := !J.IkJ(aJ) if
k ~ j and ak := 0 otherwise. The definition of .Jt guarantees that !J.Ik o !J.IkJ =
!J.IJ for k ~ j.
This construction yields the following universal property.
Lemma 3.17. Let B be a C* -algebra and suppose that for each j there are
morphisms cf>J: AJ - B, satisfying cf>k o !J.IkJ = cf>J whenever k ~ j. Then
there is a unique morphism cf>: A - B such that cf> o !J.1 J = cf> J for all j. a
An important example of a C* -inductive limit is the algebra X of compact
operators. In fact, X = li!!}.Mn(O, where the morphisms !J.IkJ: Mj(<C) -
Mk(<C) are the inclusions aJ .... aJ $ Ok-J; in this case, .J\. is the algebra of
finite-rank operators on J{ and N = 0. lt is easily seen that if A = limAJ is
a C*-inductive limit, thenA+ = li!!}.Aj canonically.
There is a purely algebraic, analogaus definition of inductive or direct
limit for a directed system of algebraic objects, such as abelian groups.
The continuity property of the K 0 -functor may now be stated as follows.
Proposition 3.18. If A = li!!}.AJ is an inductive Iimit of C* -algebras, then
Ko(li!!}.AJ) ""li!!}.Ko(AJ ).

Proof. This is Proposition 6.2.9 in [481], or Theorem 7.3.10 in [352]. a


Proposition 3.19. The functor Ko is halfexact.
Proof. We follow the arguments of Theorem 6.3.2 in [481]. l f ] is a closed
ideal in A with quotient map 11: A - Al] and if a E A +, then a E ] if and
only if 17(a) = 0 and a E j+ if and only if 17(a) is a scalar. Consider the
sequence (3.10):
Ko(]) ~ Ko(A) ~ Ko(Af]).
To prove that imKoj ~ kerKo11. consider an element [p]- [Pn] E Ko(J)
with p E 'P(Jt) and E(p- Pn) = 0, where E: A+- A+ /A = ([ is the natural
projection. Then p- Pn EX®], so (Ko11 o Koj)([p]- [Pn]) = [17(p)]-
[11(Pn)] = [Pn]- [Pn] = 0.
To prove that ker Ko11 ~ imKoj, consider [q] - [Pn] E ker Ko11. with
q in Mm(A+) and q- Pn in Mm(A) for some m ~ n. This means that
for some land k := m + l there is a unitary u E Mk((AJJ)+) suchthat
u(17(q) $Pz)u* = Pn $Pz. Suppose that we can find a unitary v in M2k(A+)
suchthat 17+ (v) = u $u*; then the projector p := v (q $Pz $ Ok)v* satisfies

( ) = (u
11 P 0
0)
u*
(17(q) $Pz
0 Ok
0) (u*
0
0) =
u
(Pn+l
0
0) .
ok
This is a scalar matrix, so p belongs to M2k(j+). Therefore, [q]- [Pn] =
[q $ Pz] - [Pn+zl = [ p] - [Pn+zl, which is in the range of Koj.
100 3. Some Aspects of K·theory

In order to find such a v, we must lift unitaries in Mk ( (A/]) +) to uni-


taries in Mzk (A +). For that purpose only, we may simplify by assuming
that k = 1 and that A is unital; thus if u E Al J is unitary, we must lift
u E9 u * to a unitary v E M2 (A). But we already know, from the proof of
Lemma 3.11, that there is a continuous path of unitaries (3.7) from the unit
1z to u E9 u *. In view of the discussion in Section 1. 5, since u E9 u * is in the
neutral component of the unitary group of Mz (A/ ]), it isafinite product of
exponentials: uEau* = 0J(expxj), for Xj E Mz(A/J). Let b1 E Mz(A) such
that 1J ( b j) = X j for each j; then b : = n j ( exp b j) iS an invertible element in
the neutral component of M 2 (A)x suchthat ry(b) = u E9 u*. A unitary that
does the same job is v := b(b* b)-112. 0

Theorem 3.20. The functor Ko is a K ·theory functor.

Proof. The normalization K 0 (C) = 7L has already been established.


To prove the homotopy invariance, we show that if cf>t: A - B, 0 :::; t :::; 1,
is a continuous family of morphisms, then Kocf>o = Kocf>1. The homomor-
phisms Kocf>t: [p] - [q] - [ cf>t p] - [ cf>t q] are constant since [p] - [ cf>t p]
is constant, because II cf>t p - cf>; p II < 1 for small values of It - s 1. Thus, a
homotopy equivalence between A and B induces an isomorphism between
Ko(A) and Ko(B). For instance, this is true of a retraction. In particular, the
Ko-group of a contractible algebra is 0.
Stability is obvious, since (As)s === As. In particular, K 0 (A) === K 0 (B) if
As === Bs.
Finally, Propositions 3.18 and 3.19 respectively establish continuity and
halfexactness of the Ko functor. 0

Remark. The conditions of Definition 3.10 are essentially the ones Cuntz
employed to single out Ko among homology functors, for a large dass of
C*-algebras [124] .

.,. We now interrupt our rush into the noncommutative theory to steal a
Iook at the "geometrical" K functor. This yields many examples and is it-
self a notable example of commutative geometry treated from the noncom-
mutative point of view. Let M be a compact space, for the time being. The
set of isomorphism classes of complex vector bundles of any rank over M
forms a commutative unital semiring Vect(M), whose addition is the Whit·
ney sum and whose multiplication is the tensor product of vector bundles.
We canonically make from this semiring a commutative unitalring by using
the Grothendieck construction again.
Definition 3.12. The K 0 -ring of a compact manifold M is the Grothendieck
ring K(Vect(M)). We write it as K 0 (M); its elements may be called virtual
bundles. The notation KU 0 (M) is sometimes used when dealing simultane-
ously with the Grothendieck ring K 0° (M) of real virtual bundles.
3.2 The Ko group 101

Exercise 3.11. Fill up the details of the ring aspect of the Grothendieck
construction. What is the unit of the ring? Does the multiplication distribute
over addition? 0
Let cf>: N - M be continuous and consider the pullback bundle map
cf>*: Vect(M) - Vect(N). This is a semiring homomorphism. Therefore, it
descends to a ring homomorphism K 0 cf>: K0 (M) - K0 (N). In this way K 0
becomes a cofunctor from the category of compact spaces to the category
of unital commutative rings. By Proposition 2.3, K0 cf> depends only on the
dass of cf> in [N, M]; if M is a one-point space, or even if M is contractible,
then K0 (M) "' 71... As a matter of fact, K0 (M) contains a canonical copy of
the integers -more generally, of the group H 0 (M, il).
There is also a ring structure of Ko (C (M)), since C (M) is commutative.
In fact, for any commutative unital ring .Jl we can form the (outer) ten-
sor product of two .Jl-modules, and it is not hard to see that the tensor
product of two finitely generated projective modules is finitely generated
and projective; this tensor product clearly drops to a multiplication mak-
ing K~1 g(.Jl) a commutative unital ring. The Serre-Swan theorem yields an
immediate consequence.
Corollary 3.21. K0 (M) "'Ko(C(M)) as rings. E3

Denote by Ok - M the trivial vector bundle of rank k. Exercise 3.8 has


the following consequence.
Proposition 3.22. Each element of K 0 (M) can be represented as [F] - [ Ok]
for some vector bundle F and some k E N. Moreover, two vector bund/es
are in the same class if and only if they become isomorphic when a suitable
trivial bundle is added to both of them. a

.,. We can define a homomorphism called virtual rank from K0 (M) to il by


[E]- [F] ._ rank[E]- rank[F]. The virtual bundle [E EB Ok]- [E] is not in
the kernel of this map, but we would like to regard it as essentially trivial
in K-theory, since E and E EB Ok are homotopy equivalent by bundle maps.
To deal with this, we return to consideration of pointed spaces.
Definition 3.13. Let M be a compact pointed space, let i be the inclusion
of * into M at its base point and let c be the collapsing map of M onto *.
The reduced K0 -ring K0 (M) of M is the kernel of the natural projection
K0 i: K0 (M)- K0 (*),orthecokernelofthenaturalinjectionK 0 c: K0 (*)-
K0 (M). The following exact sequence splits:

since K 0 i is a left inverse for K 0 c, and therefore K0 (M) = K0 (M) EB il. It


is clear that K0 becomes also a cofunctor from the category of compact
spaces to the category of commutative rings: for cf>: M - N a morphism of
102 3. Same Aspects of K-theory

compact spaces, K0 <P is just the restriction of K 0 <P to K0 (N) whose image
lies in K0 (M). The construction is independent of the chosen base point
and K0 is homotopy invariant.
Definition 3.14. Two vector bundlesE and F over Mare called stably equi-
valent, written E "" F, if there are trivial bundles 0 1 and Ok such that
E E9 0 1 - F E9 Ok. We do not assume that j = k.
Proposition 3.23. Stahle equivalence classes are elements of the reduced
K -theory of M.

Proof. The map ~: Vect(M)- K0 (M): E- [E]- WrankEl is surjective and


~ (E) = ~ (F) if and only if E and F are stably equivalent. D

As already mentioned in our discussion of vector bundles over § 2 in


Chapter 2, in order to find a Whitney summand that trivializes a given
vector bundle, one need not go to very high ranks. There is the following
important result [257, Thm. 8.1.2], proper to the K-theory ofvector bundles,
known as the stable range theorem.
Theorem 3.24. Let Ek - M be a complex vector bundle ofrank k over an n-
dimensional manifold M, with k > ii := f ~ (n - 1) l, i.e., the smallest integer
not less than ~ (n- 1); then Ek - Fn E9 Ok-n for some complex vector bundle
Fn- M ofrank ii. a
Such bundles Ek are said to be in the stable range. The stable range the-
orem has the consequence that, as far as K-theory is concerned, nothing is
gained by considering bundles of very high rank, because as soon as the
stable range is reached, no new elements of K0 (M) arise. Also, in the stable
range, two bundles are isomorphic if and only if they are equivalent in the
K-theoretic sense. Thus, one can view K-theory as a kind of simplification
that occurs in the algebra of high-dimensional bundles. lt is worthwhile to
note that in real K -theory there is an analogaus statement, where the rank
condition isthat k > n; and also that results of this kind have been imitated
in algebraic K-theory. For instance, Bass' stabilization theorem says that if
P and P' are projective modules of rank k on a commutative, noetherian
ring .A of Krull dimension n < k, and if [P] = [P'] in Kg1g(.A), then P
and P' are isomorphic. Corach and Larotonda [122] have determined the
stable ranges of several Banach algebras. The related concept of "topolo-
gical stable rank" in the K -theory of C* -algebras has been developed by
Rieffel [393] and applied to the classification of projective modules over
noncommutative tori.
From the stable range theorem, we conclude that, for k > n/2,

K0 (M) = [M,BU(k)], or K0 (M) = [M,lxBU(k)]. (3.11)

Notice that K0 (M) = K0 (M+). Just as in the noncommutative case, this


trick is used to define the K-theory of a locally compact space M (more
3.3 The importance of being halfexact 103

precisely called the K-theory with compact supports) as the reduced K-


theory of its one-point compactification. An element of K0 (M) can be rep-
resented as a pair of vector bundles over M together with an isomorphism
between them, defined on the complement of some compact subset of M.
In connection with the remark after Definition 3.4, note that Corollary 3.21
generalizes: K 0 (M) "" Ko (Co (M)) as rings.
To close this longish section, we make a few remarks on the meaning
of Theorem 3.20 in the context of K-theory of vector bundles. Homotopy
invariance of the K 0 cofunctor has already been discussed. Continuity im-
plies that the K 0 group of the intersection of a directed family of compact
subsets of a compact space is the direct limit of the corresponding K 0 -
groups. Stability (in the C*-algebraic sense) is clearly a door opening into
the noncommutative world. lt will be strengthened and dignified later with
the name of Morita invariance: we shall eventually understand that classes
of Morita equivalent algebras have the same K-theory and that, although
two different commutative algebras are never Morita equivalent, a commu-
tative algebra may be Morita equivalent to many noncommutative algebras.

3.3 The importance of being halfexact


Halfexactness gives a form of excision in K-theory of vector bundles. Let
(M, N) be a compact pair, and consider the open set M \ N. Recall the exact
sequence of C*-algebras considered in Section 1.6:
0 - C0 (M \ N)- C(M)- C(N) -0.
Proposition 3.19 and Corollary 3.21 then give exactness at K0 (M) of
K 0 (M \ N)- K 0 (M)- K 0 (N),

which can be rewritten as

By choosing a base point of N (hence also of M), we immediately see that


K0 (MIN) -K 0 (M) -K 0 (N)
is exact; therefore the above reasoning can be applied to a locally compact
space M with a closed subset N as well. lt motivates the following definition
of halfexactness adapted to the needs of the commutative context.
Definition 3.15. A cofunctor F from the category of compact spaces and
homotopy classes of continuous maps into the category of abelian groups
is called halfexact if for any exact sequence N...!..... M ..!'..MI N, the corres-
ponding sequence
F(M IN) !..!!... F(M) .!..!.. F(N)
104 3. Some Aspects of K-theory

is exact. Note that F ( *) = 0 necessarily, if * is a one-point space.


Lemma 3.25. Let F be a halfexact cofunctor. I{ X, Y are pointed spaces, then
the natural homomorphism F(X v Y)- F(X) e F(Y) is an isomorphism.

Proof. The inclusions i: X - X v Y and j: Y - X v Y define the natural


homomorphism. Note that (X v Y)/X = Y and (X v Y)JY =X. Therefore

F(Y)- F(X v Y)- F(X)

is exact. As prx oi = idx and pry oj = idy, the first arrow is injective, the
second is surjective, and the sequence splits. 0

By no means is K0 the only example. For instance, if H denotes Cech


cohomology, the reduced cohomology cofunctor fi• ( ·, &:) := coker(&: -
H" ( · , &:) ) is halfexact [154]. Also, if HctR denotes de Rham cohomology on a
manifold, its reduced cohomology cofunctor Hd.R ( ·) := coker(~ - Hd.R ( ·))
is also halfexact [199) .
.,.. We want to prove that the K-theory of C*-algebras possesses the long
exact sequence, a powerful computational instrument. This is done by us-
ing (a particular case of) the Puppe sequence of algebras (1.17), i.e., by
imitating the classical methods by Puppe and Dold [139) to get lang exact
sequences for any "generalized" cohomology theory on topological spaces
-in which the previous definition of half-exactness plays the starring role.
Traditionally, a (co)homology theory for spaces was specified by seven ax-
ioms, beautifully laid out by Eilenberg and Steenrod in their book [154],
but the K-theory of spaces is then a generalized cohomology theory since
it does not satisfy the last axiom (the so-called "dimension axiom" that
H»(*) = 0 for n * 0). Likewise, the K-theory of C*-algebras is a general-
ized homology theory in the analogaus way. In our context, the Puppe-Dold
method was axiomatized by Schochet [418), whom the expositians in [36)
and [481) follow. By now, it should be folklore, but as we are not sure of
that, we outline it here.
If desired, the domain of the homology theory may be a subcategory of
C*-algebras, which should be closed under the various operations we dis-
cuss below, such as suspensions and quotients. lt should be noted that in
most treatments of K-theory, the definition of the functors K1 ,K2 , .•• is
given beforehand and the construction of the connecting or "coboundary"
homomorphisms 8 is hard work. Here we take a different path: namely, to
construct the homology theory using only the Ko and the stated proper-
ties, and then undertake the interpretation of the obtained functors and
connecting homomorphisms a posteriori.
Definition 3.16. We consider (covariant) functors from the category of C*-
algebras to the category of abelian groups that are halfexact and homotopy
invariant. A homology theory is a sequence of such functors H. = {Hn}
3.3 The importance of being halfexact 105

(indexed by ~ or 71., as the case may be) such that for every short exact
sequence 0 - 1 J..... A ...!!... B - 0 of C* -algebras there are group homo-
morphisms 8: Hn(B) - Hn-1 (]) connecting the short exact sequences
Hn (]) 1!!!1 Hn (A) l!!:.J Hn (B) into a long exact sequence

· · · ~ Hn (]) .!!.!:.!... Hn (A) .!!!!.!!.. Hn (B)


{j Hn-Ü Hn-111 {j
- Hn-d]) - Hn-dA) ~ Hn-dB)- · · ·

and the maps 8 are natural, i.e., to each commutative diagram of short
exact sequences

j 11
o-1-A-B-o

cPJ I . cPA! cPB! (3.12)


t /
o-1'-A'-B'-o
11'

there corresponds a commutative diagram of long exact sequences

Definition 3.17. The procedure will be to define Kn (A) := Ko(.~.:n A). This
may seem arbitrary, but there is no other choice, as the next result shows.
Proposition 3.26. Let H. be a homology theory. Then Hn+dA) "" Hn (I.A)
for all n.

Proof. In Proposition 1.14, we proved exactness of 0- I.A- CA- A- 0.


Since CA is contractible, we obtain Hn(CA) = 0 for all n (by homotopy
invariance). Therefore all the connecting maps corresponding to this short
exact sequence are isomorphisms. 0

This short proof shows how to proceed: we use the several exact se-
quences of C*-algebras involving cones, suspensions and mapping cones
already derived in Section 1.6 and apply suitable functors to them to get
improved behaviour of the arrows. A case in point is the next Iemma.

Lemma 3.27. Let 0 - 1 ..i..... A...!!... B - 0 be an exact sequence of C* -algebras


for which B is contractible. Jf H is a halfexact and homotopy-invariant func-
tor, then the map H j: H (]) - H (A) is an isomorphism.
106 3. Some Aspects of K-theory

Proof. lt is clear that Hj is surjective, since H(B) = 0. To show that it is


injective, we factor the map j: 1 - A in the form

1 1
c...,__.

}-
<I> pr 1 A
C17- t

in such a way that all the induced maps in homology are injective.
The map pr 1 occurs in the exact sequence ~B ...__ C11 ~~ A of (1.15) for
the mapping cone of 'IJ, and ~B is contractible since B is contractible; half-
exactness then shows that Hpr 1 : H(C11 ) - H(A) is injective. Now Iet

Dj :={jE AI: j(l) E j(]) },

and for c E 1. Iet L(c) E D 1 be the constant function with value j(c) E A.
To get </J: D1 - C11 suchthat pr 1 o</J o L = j, just take </J(j) := (j(O), 11 og)
with g(t) := j(1- t) for 0 < t ~ 1; notice that <Pis onto. Now D 1 retracts
on1; indeed, if y(j) := j- 1 (j(1)) for f E D 1 , then yo L = id1 and IJlsf(t) :=
j(s + t- st) defines a homotopy between L o y and idvr Therefore Ht is an
isomorphism.
Moreover, the map h ...... j o h embeds the cone C1 in D1 , and its image
is ker </J. This gives another short exact sequence C1 ...__ Dr:!:. C11 with a
contractible first term, so H <P is also injective. D

We arealmostready now to build the long exact sequence.


Theorem 3.28. Let H be a halfexact and homotopy-invariant functor and
Iet 1 ...:!_. A-2. B be a short exact sequence of C* -algebras. Then there is a
natural homomorphism 8: H (~B) - H (]) that makes the functors {H~n}
into a homology theory.
Proof. The trick is to splice the sequence

H(~j) ~ H(~A) .!!!:!!._ H(~B) -~~ H(])!!..!... H(A)!!.!!. H(B) (3.13)

with an arrow 8 formed from the composition

H(~B) .!!.P.. H(C11 ) ~ H(]), (3.14)

where ß and a are the injective maps in the short exact sequences

provided by Propositions 1.14 and 1.15, namely, ß(h) := (0, h) and a(c) :=
(j(c), 0). Since CB is contractible, Hais an isomorphism by Lemma 3.27.
Exactness of (3.13) at H(]) follows from noting that j = pr 1 oa, so that
Hj = Hpr 1 oHa and thus kerHj = (Ha)- 1 kerHpr 1 = (Ha)- 1 imHß =
im8.
3.3 The importance of being halfexact 107

Exactness at H(~B) is harder. Since ker8 = ker((Hoc)- 1 o Hß) = ker Hß,


we need to show that ker Hß = imH~'J. To do that, we decompose ß as
ß = J.loe where He is anisomorphismand thenprove that (He)- 1 ker HJ.I =
imH~'J. For that, in turn, we find another map K suchthat ker HJ.I = imHK
by half-exactness, and for which (He)- 1 o HK= H~'J.
Let C := { (j,h) E CA Eil CB: h(l) = t]{j(1))}, let K(j) := (j,O) and
e(h) := (0, h) be the obvious embeddings of ~A and ~B in C, and let
1\(j,h) :=fand J.l{j,h) := (j(1),h) E C17 • lt is clear that kerJ.I = imK
and that J.1 is onto, so we have manufactured two short exact sequences,
8 ,\ K jJ
~B......._C-CA and ~A......._C-C11 •

Now He is an isomorphism, since CA is contractible. Also, J.l(e(h))


(0, h) = ß(h) for h E ~B. as required. Finally, there isalinear homotopy
from K to e o ~'7 given by 1./Jsf := ((1- s)j,s '7 o j) for 0 ::5 S ::5 1 and
f E ~A. Thus, HK = He o H~t], and it follows that ker 8 = ker H ß.
Naturality of 8 is checked as follows. Suppose that a commutative dia-
gram (3.12), with exact rows, is given. From Propositions 1.14 and 1.15, we
get two more commutative diagrams with exact rows,

by defining tfJ(a,j) := (cf>A (a),j ocf>B). This gives the commutative diagram

H(~B) ~ H(C17 ) ~ H(J)

m:cf>B 1 H ' Hrp! ' Hcf>J 1


H(~B') ~ H(C11·) ~ H(J'),
from which it follows that

H(~B) ~H(J)

H~cf>B! Hcf>J!
H(~B') ~ H(J'),
since 8 = (Hoc)- 1 oHß. D

.,.. Therefore, our newly defined K-functors have the long exact sequence

· · · - K 2 (A/J) .i. KI(J)- KI(A)- KI(AJJ)


{j
- Ko(J)- Ko(A)- Ko(A/ ]). (3.15)
108 3. Some Aspects of K-theory

As a first application, we give the following result.


Proposition 3.29. Every split exact sequence
j ,.,
0-J - A - A l ] - 0
er

induces a split exact sequence of Ko groups

Ko11
o-Ko(J) -Ko(A) ~K
ocr
Ko(A/J) -o.
Proof. In the exact sequence

KI(J) .!!J.. KI(A) ~ KdA/J) ..!_ Ko(J) ..!0_ Ko(A) ~ Ko(A/J),

the last arrowis surjective (and splits), because Ko17oKo(J" = id onKo(A/J).


For the same reason, the second arrow K111 is onto; therefore 8 is the zero
map and the conclusion follows. 0

Let us now Iook at the commutative case:


~o ~o a ~o ~o ~o
(3.16)
· · · - K (IX) - K (IY) - K (X/Y) - K (X) - K (Y).

We notice that this Iooks very much the result of applying the K0 cofunc-
tor to the Puppe sequence (1.16), corresponding to the inclusion i: Y ~X.
That is indeed so, since it follows from Lemma 1.12 and Exercise 1.22
that K0 (X/Y) = K0 (Ci). The appearance of the unreduced suspensions
in (1.16) is unimportant here, in view of the homeomorphism IX "" IX 1I
and Lemma 1.12 again. The exact sequence (3.16) can also be rewritten as

Exercise 3.12. Prove that the inclusion and quotient maps

induce a split exact sequence of maps in K -theory of vector bundles:

.". We close with a result by Dold, concerning halfexact functors in the com-
mutative context, that will be used decisively in Chapter 8. We follow the
early treatment by Karoubi [270]. The action takes place in the category of
finite CW -complexes.
3.3 The importance of being halfexact 109

Definition 3.18. A finite CW-complex is a Hausdorff space K equipped


with a finite partition into a family of subsets Eo, E1 , ..• , EN, where each En
is a finite disjoint union of a family of n-cells { ej : j E ln }; an n-cell is
a homeomorphic copy of ~n (or of the open unit ball of ~n). Foreach cell
there is a continuous map Jp of the closed unit ball of ~n into K whose
image is the closure ofthat cell in K, suchthat the restriction of Jp to the
openball is a homeomorphism onto the cell and Jp(§n- ) is contained in
1

Kn- 1 := Um<nEm (the (n -1)-skeleton of K).


A CW -subcomplex L of a CW -complex K is the union of a set of cells of
K suchthat if ej c L, then the closure of ej is also a subset of L. The space
K/L has a natural structure of CW-complex. The n-skeleton Kn of K is a
CW -complex and En is open in Kn. We shall say that the CW -complex K is
of rank s; q if it can be constructed by means of at most q attachments of
cells.
We remark that when M = §n- 1 and f = Jp
is the map in the construc-
tion of a CW-complex, then Cf :=Luf CM is just the attaching of the cell
ej to the previous subcomplex [140, Prop. V.2.9].
Lemma 3.30. Let F and G be halfexact cofunctors on the category of finite
CW -complexes and Iet cJ>: F - G be a morphism of cofunctors such that
cf>(§n): F(§n) - G(§n) is an isomorphism. Then c/>(K): F(K) - G(K) is an
isomorphism for all objects of this category.

Proof We first prove, by induction on the rank of K, that cJ>CiK): F(iK) -


G (f.K) is an isomorphism. The assertion is clear for the empty set K = 0, as
then iK is § 0 and F(§ 0 ) :::: G(§ 0 ). If K is nonempty, we can choose a base
*
point E K and consider the reduced suspension ~K and the following
commutative diagram:

F(~K) ~ G(~K)

~
!
F(~K)
cJ>(tK)
__,..
!~
G(~K).

In view ofLemma 1.12, the vertical arrows are isomorphisms. To prove that
c/>(IK) is an isomorphism, it is enough to prove that c/>(~K) is an isomor-
phism. This follows from the inductive hypothesis, as a careful counting
of attachments shows that ~K has lower rank than iK. (Observe how the
existence of the two suspensions is essential for the argument!)
At each step in the construction of a CW-complex, the space obtained is
a mapping cone for an attaching map along spheres. We can thus prove the
Iemma by induction on the rank of K. For rank 1, the statement is trivially
true, as F ( *) = 0 and G ( *) = 0 for point spaces. For the induction, there
llO 3. Some Aspects of K-theory

is a commutative diagram with Puppe rows:

F(i.L) - F ( § n ) - F ( K ) - F ( L ) -F(§n-1)

cp(ll)! cp(§")! cp(K)! cp(L)! cp(§n-1)!


G(i.L) ~ G(§n) ~ G(K) ~ G(L) ~ G(§n-1).

Here K =Cf is obtained from L by attachment of a cell of dimension n by


means of f: §n- 1 - L. By the hypothesis in the statement of the theoreml
cp(§n) and cp(§n- 1) are isomorphisms. By the inductive hypothesis cp(L)1

is an isomorphism. As we have shown that cp(i.L) is then also an isomor-


phisml the result follows from the Five Lemma. D

3.4 Asymptotic morphisms


One of the tricks of the tradel both in topology and operator algebral is
to change categories by weakening the conditions on the arrows. This is
actually done in the definition of K -theory. It has been argued by Connes
and Higson that for applications of C* -algebra theory in topologyl the good
1

notion of morphism is much weaker than the ordinary one. The E-theory
of Connes and Higson [91 104] is abivariant homology and cohomology
1

theory for separable C*-algebras that satisfies the halfexactness or excision


property in both variables (giving rise to a pair of hexagon diagrams [128];
see the next section). The E-category has separable C*-algebras as objects
and the E( ·~ ·) groups as arrows; in order to define these groupsl one uses
the notion of asymptotic morphisml which in turn generalizes the older
notion of deformation of algebras. While E-theory is not on our agenda we 1

need asymptotic morphisms to construct important K-theory maps.


Definition 3.19. Let A, B be two C*-algebras. An asymptotic morphism
from A to B is a family of mappings T = { Th: A - B : 0 < h ::5 ho }, for
some ho > 01 suchthat h ,__. Th (a) is norm-continuous on (01 hol for each
a E A 1 and such that for any al b E A and ,\ E C:
1

Th(a +Ab)- Th(a)- ATh(b)- 0}


Th(a*)- Th(a)*- 0 as h I 0 1 (3.17)
Th(ab)- Th(a)Th(b)- 0

with convergence in the norm of B. An asymptotic morphism is called uni-


form if the convergence in (3.17) is uniform on compacts subsets of A.
We say that the asymptotic morphism is unital if A and B are unital and
Th(1A) = 1s for all hl linear if Th(a + ,\b) = Th(a) + ,\Th(b) and real if
T h ( a * ) = T h ( a) *I for all h. Two asymptotic morphisms TI S from A to B
3.4 Asymptotic morphisms 111

are equivalent if

lim IITh(a)- Sh(a)ll = 0 for all a E A;


hJO

they are homotopic if there is an asymptotic morphism 'I': A - BI suchthat


Po o 'I' = T and PI o 'I' = S. We denote by [A, B] the corresponding space of
homotopy dasses.
The choice of the parameter space for h is to a large extent arbitrary. For
simplicity, we shall take ho = 1 from now on. Connes and Higson choose
the interval .[ 1, oo), and so their t is 1 I h. Our convention accords with the
usual notation for an important dass of examples, the deformations of
Definition 2.10, in particular with Moyal Quantum Mechanics (where indeed
h can be interpreted as Planck's constant).
If A is not unital, one can extend T to a unital asymptotic morphism y+
from A + to (any unitization of) B by setting Th (a, ;\) := Th (a) +AlB.
Equivalent asymptotic morphisms T and S are homotopic, by s- sTh +
(1- s)Sh, for 0 5 s 5 1. Any point-norm-continuous path { IJh: 0 < h 5 1}
of morphisms from A to B gives rise to an asymptotic morphism, homo-
topic to the (constant asymptotic morphism defined by) 17 1 . Clearly, there
is a map from [A, B] to [A, B], which in general is neither surjective nor
injective.
Each deformation A = {Ah : 0 5 h 5 1 }, with Ao = A and Ah = B for
h > 0, gives rise to a dass of asymptotic morphisms from A toB, as follows.
Foreach a E A, Lemma 2.13 allows us to choose a section Sa E ~(,1) with
Sa(O) = a; then Iet Th(a) := Sa(h).

Lemma 3.31. I(T is an asymptotic morphism from A toB and if a E A, then


limsuphJO IITh(a)ll 5 llall.
Proof. It suffices to consider the case where A and B are unital. We show
that lim suph 1o r(Th (a)) 5 r(a), where r is the spectral radius. This then
implies that

limsup 11Th(a)ll 2 = limsupr(Th(a)*Th(a)) = limsupr(Th(a*a))


hJO hJO hJO
5 r(a*a) = llall 2 •

Now if llt.l > r(a), then (a- !t.)c = 1 for some c E Ax, so bh := (Th(a)-
!t.)Th(c) - 1 in B as h l 0. Thus, with a small enough h1, bh is invertible
for 0 < h 5 h 1 and so;\ rt sp(Th(a)).
For deformations, the boundedness condition is strengthened to equal-
ity: limhw IITh(a)ll = llall, since h- Th(A) is, by definition, a continuous
section. D

We can manufacture a genuine C* -algebra morphism from an asymptotic


morphism in the following way. Notice first that T and S are equivalent iff
112 3. Some Aspects of K-theory

h ..... Th(a) -Sh(a) belongs to the cone C*-algebra CB = Co((O, 1]-B). From
the definition and Lemma 3.31, we see that any asymptotic morphism maps
A into B-valued bounded continuous functions on (0, 1 ], that is, h ..... Th (a)
lies in Cb((O, 1]-B). Now define Boo as the quotient C*-algebra

(3.18)

Then the equivalence dass of h ..... Th determines a unique mapping


T: A - Boo; the properties (3.17) imply that T is a morphism of C* -algebras.
(Notice that Lemma 3.31 isjust a version ofthe proof ofLemma 1.16, show-
ing that T is contractive.)
lt turns out that this process can be reversed, in the sense that a mor-
phism T: A - Boo can be lifted to a (usually not linear) map T': A -
Cb((O, 1]-B) that gives an asymptotic morphism; and if T arises from an
asymptotic morphism T, then T' and T are equivalent. For that, we invoke
a theorem due to Bartle and Graves [21, Thm. 4] that provides a contin-
uaus lifting in the following situation. Suppose that E and F are two Ba-
nach spaces, and let Hom(E,F) be the space of continuous linear maps
from E to F (with the norm topology). If S is a continuous map from a
metric space A into Hom(E,F), there is an associated map a: C(A-E) -
C(A-F) given by a(j): a ..... S(a)[j(a)]. When S(a) is surjective, then
n(S(a)) := sup 11 yll=l inf{ llxll : S(a)x = y} is finite by the open mapping
theorem [383, Thm. III.10]. Nowifthe range of S consists of surjective maps,
and if SUPaEA IIS(a)ll and N := SUPaEA n(S(a)) are finite, then Bartle and
Graves showthat aisalso surjective byconstructing, foreachg E C(A-F),
a preimage f E C(A-E) suchthat llf(a) II ::;; Nllg(a) II for each a E A. For
the case at hand, let E := Cb((O, 1]-B), F := Boo and let S be the constant
function on A whose value is the quotient map q: Cb((O, 1]-B) - Boo.
Then supa IIS(a) II = llqll = 1 and N = n(q), so that TE C(A-Boo) lifts to
T' E C(A-Cb((O, 1]-B)). Any such lifting is a uniform asymptotic mor-
phism, and any two liftings are equivalent.
~ Next we show, following Higson [243], that (classes oO asymptotic mor-
phisms give rise to K-theory maps. Suppose that T isauniform asymptotic
morphism from A toB and that Ais unital. First, extend T to the asymp-
totic morphism T ® id from As to Bs, which we continue to call T. Let
p E P(As) be nonzero; then the elements { Th(P) : 0 < h ::;; 1} of Bs
satisfy limhw IITh(p)- Th(p) 2 11 = 0 and limh!O IITh(P)II = 1. Choose E
with 0 < E < ~; then we can find h 1 > 0 suchthat if 0 < h ::;; h 1 and
A E sp(Th(p)), then either lAI < E or lA- 11 < E. Let qh be the spectral
projector of Th(P) for the disk { A: lA- 11::;; ~ }; then IITh(P)- qhll < E.
Exercise 3.13. Show that if p and p' in P(As) are homotopic projectors,
then the corresponding projectors qh and qh are homotopic in P(Bs ). 0
Finally, set KoT( [p]) := [q!]. This is a well-defined homomorphism from
Ko(A) to Ko(B). (If Ais not unital, use the unital asymptotic morphism T+
3.5 The Moyal asymptotic morphism 113

and checkthat the corresponding K-theory map restricts to a map from


Ko(A) to Ko(B).)
The Iifting theorem has been used to produce uniform asymptotic mor-
phisms equivalent to given asymptotic morphisms. lt can, and will, also be
exploited in the following way. To define an asymptotic morphism T up
to equivalence, it often suffices to determine it on a dense subspace of A
in order to obtain the morphism T: A - Bco; the Iifting then produces an
extension of the original T (or an equivalent copy) to all of A.
Tensor products of asymptotic morphisms are important for us. Given
two asymptotic morphisms {Th: A- B} and {Sh: C - D}, let B', D' be
unitizations of B and D and let B' ® D' be the tensor product defined with
any C*-cross-norm (see Section l.A). Then { Th ® 1} and { 1 ® Sh} define
asymptotic morphisms, from A and C respectively, into B' ® D', yielding re-
spective morphisms from A and C into (B' ® D' )co. Their images commute,
and actually lie in the ideal (B ® D) co, and so define a morphism from the
tensor product A ®J.J C, with the maximal C*-norm, into (B ® D)co. Hence,
up to equivalence, there is defined an asymptotic morphism Th ® Sh from
A ®J.J C toB® D. A similar construction works to produce asymptotic mor-
phisms from A ®J.J C toB, say, provided that the images of the asymptotic
morphisms from A and C into B commute asymptotically.

3.5 The Moyal asymptotic morphism


An example of asymptotic morphism is given by Moyal quantization; this
is important on more than one account, so we give a detailed presentation.
Definition 3.20. Let X be a phase space (i.e., a finite-dimensional symplec-
tic manifold), f.1 a multiple of the Liouville measure on X, and J{ a Hilbert
space somehow associated to X. A Moyal quantizer for (X, f.l, Jf) is a map-
ping n of X into the space of bounded selfadjoint operators on Jf, such
that Q(X) is weakly densein L(Jf), and verifying

TrO(u) = 1, (3.19a)
Tr[O(u)Q(v)] = 8(u- v), (3.19b)

in the distributional sense. (Here, 8 (u - v) denotes the reproducing kernel


for the measure (.1.)
Ownership of a Moyal quantizer solves in principle all quantization prob-
lems: quantization of a (sufficiently regular) function or "symbol" a on X
is effected by

a ...... fx a(u)Q(u) df.l(U) =: Q(a), (3.20)


114 3. Some Aspects of K·theory

and dequantization of an operator A E L(J-f) is achieved by


A .... TrAO(·) =: WA(·). (3.21)
Indeed, from (3.21) it follows that lJ-f .... 1 by dequantization, and also

Tr Q(a) = fx a(u) dJJ(u). (3.22)

Moreover, using the weak density of Q(X), it is clear that W inverts Q:

WQ(aJ(U) = Tr[ (fx a(v)Q(v) dJ.I(V) )o(u)] = a(u),


so Q and W are inverses. In particular, WQol = 1 says that 1 .... 1.1{ by
quantization, and this amounts to the reproducing property:

fx Q(u) dJ.I(U) = l.J[.

Finally, (3.19b) and (3.20) combine to give

Tr[Q(a)Q(b)] = fx a(u)b(u) dJ.I(U) =: (a, b); (3.23)

note that (3.22) and (3.23) aretagether equivalent to (3.19).


The concept of Moyal quantizer was introduced in [211,467], where the
quantizer for spinwas worked out. In [467], it was baptized the "Stratono-
vich-Weyl quantizer". But that name did not stick, and so, since new appli-
cations for Moyal quantization crop up all the time, we rename it accord-
ingly. Most interesting cases occur in an equivariant context [205]; that is
to say, there is a (Ue) group G for which X is a symplectic homogeneaus G-
space, with J.1 then being aG-invariant measure on X, and G acts by a projec-
tive unitary irreducible representation U on the Hilbert space J-f. A Moyal
quantizer for the combo (X, J.1, J-{, G, U) is a map 0 taking X to bounded
selfadjoint operators on J-{ that satisfies both (3.19) and the equivariance
property
U(g)Q(u)U(g)- 1 = Q(g · u), for all g E G, u EX. (3.24)
One can think of X as an member of the dual space of G (i.e., the set
of coadjoint orbits). The quantizer then allows Fourier analysis to be per-
formed essentially as in the abelian case [181,467,471]: the "function"
E(g, u) := Tr[U(g)Q(u)]

works like the exponential kernel of the Fourier transform. Actually, E (g, u)
will in general be a distribution on the space of smooth sections of a non-
trivialline bundle over G x X, the nontriviality being related to the nonlin-
earity of the representation U. Using the properties of the quantizer, the
character x of U may be formally computed by

x(g) = fx E(g, u) dJ.1(u).


3.5 The Moyal asymptotic morphism 115

~ LetX = T*~nwithelementsu =: (q,p) andwithmeasuredJlh(u) =


(2rrh)-n dnq dnp. The (parametrized) Moyal quantizer (for nomelativistic
spinless particles, one would say) is given explicitly on J{ = L 2 ( ~ n), in the
Schrödinger representation, by the Grossmann-Royer reflection operators
[222,405], which are

(o.h (q, p)f) (x) := 2n exp ( ~i p(x- q)) j(2q - x). (3.2 5)

The properties (3.19) are easily checked. We shall work with symbols be-
longing to the Schwartz space S(T* ~n ). The integral (3.20) certainly makes
sense as a Bochner integral, and

(Qh(a)j)(x) := (2:n)n JT*Dln a(q,p)(O.h(q,p)f)(x)dnqdnp

= 1 f
(2rrh)n T*Dln
a(X+Y,p)eip(x-y)/hj(y)dnydnp,
2
(3.26)

for 0 < h :5 1, say. This is the basic formula of Moyal pseudodifferential


calculus [252] (a bitdifferent from the standard pseudodifferential calculus
reviewed later, in Section 7.A).It is immediate that the operators Qh (a) are
traceclass and

Moreover,
a(q, p) = Tr[Qh(a)O.h(q, p)],
from which (3.23) follows; for that, we compute

Tr[ Qh (a)O.h (q, p)]


= 1 f f a(x + 2q- Z, p')eip'(x-2q+z)fhe2ip(q-z)/h I dnp' dnx
(2rrn)n JDln JDln 2 x=z
= 1 f f a(q, p')e2i(p-p')(q-x)/h dnx dnp' = a(q, p),
(2rrn)n JDln JDln
by the Fourier integral theorem.
The group of translations of T* ~ n is represented by the Weyl operators,

(Wh(q, p)j)(x) := e-iqp/2heipx/h j(x _ q).

Theseare clearlyunitary. With u' = (q', p'), Iet s(u, u') := qp' -q'p denote
the standard symplectic form on T* ~ n. Routine calculations establish that
W is a projective representation of the group of translations,
116 3. Some Aspects of K-theory

and verify (3.24) in the present case:

W"(u)O"(v)W"(-u) = O"(v + u).

By a well-known result of von Neumann [359], this representation is irre-


ducible. lt can be proved that (3.24) is verified as well for the action of the
symplectic group Sp(2n, ~) on T*~n. when U is the metaplectic represen-
tation -see [184] for the latter.
Exercise 3.14. Show that

in keeping with the geometrical meaning of the operators. Verify also

O"(u)O"(v)O"(w)

= 22 n exp( ~i [s(u, v) + s(v, w) + s(w, u)] )o"(u- v + w). 0

Definition 3.21. The Moyal product a x,. b of two Schwartz functions a, b


is defined in such a way that Q,.(a x,. b) = Q,.(a)Q,.(b). In view of Exer-
cise 3.14, this is achieved by

(3.27a)

where the integral kernel is

L"(u,v,w) := Tr[O"(u)O"(v)O"(w)]

= 2 2 nexp(~i(s(u,v) +s(v,w) +s(w,u)}). (3.27b)

lf a,b E S(~ 2 n), then a x,. b E S(~ 2 n) and the product operation is con-
tinuous [210].
The Moyal product satisfies an important tracial identity:

(a,b) =Tr[Q,.(a)Q,.(b)] =TrQ,.(ax,.b) = (ax,.b,1) = (bx,.a,1).


(3.28)

The cyclicity inherent in this identity allows the extension of the Moyal
product to large classes of distributions via duality: if a, b, c E S(~ 2 n ),
then
(ax,.b,c) = (a,bx,.c) = (b,cx,.a).
ForT E S'(~ 2 n), we can then define T x,. a and a x,. Tin S'(~ 2 n) by
(T x,. a, b) := (T, a x,. b) and (a x,. T, b) := (T, b x,. a) respectively.
The Moyal product is very regularizing andin fact S'(~ 2 n) x,. S(~ 2 n)
and S(~ 2 n) x,. S'(~ 2 n) are made of smooth functions. Now :Mr := {TE
3.5 The Moyal asymptotic morphism 117

S' (~Zn) : T x b E S for b E S} is the left multiplier algebra; the right


multiplier algebra :MR is analogously defined. The Moyal algebra :M is
then defined as :M := :Mr n :MR. It takes no time to see that the functions
a(q, p) := q and b(q, p) := p belang to :M and of course they are quan-
tized as the selfadjoint operators of multiplication by x and as -ih a;ax,
respectively.
There is thus a fascinating interplay between the Moyal product and
distribution theory. We pause to recall the spaces of smooth functions
X(~n) C <9e(~n) C (')M(~n).

Definition 3.22. A smooth function cf>: ~n - ( has derivatives a01 cf> for
oc E Nn, with the usual multiindex notation. It belongs to (')M(~n) if each
derivative is polynomially bounded:

for some positive integer N 01 and some C01 ~ 0. Moreover, cf> lies in <'Je ( ~ n)
if we may take N 01 = m for some fixed m; andin the space X(~n) of
Grossmann-Loupias-Stein symbols, which was introduced [223] precisely
to meet the needs of Moyal quantum mechanics, if we may take N 01 = m-
Ioc 1. In this last case, we say that cf> is a GLS symbol of order m. Note, in
particular, that any polynomial of order m is a GLS symbol of the same
order.
Each of these spaces carries a suitable (locally convex) topology, for which
the inclusions X c <'Je c (')M c S' are continuous. For instance, if Xm
denotes the space of GLS symbols of order m, its topology is generated by
the seminorms

k E N,

and thus Xm ._. Xm' continuously if m ::5 m'. Notice that a01 cf> E Xm-lotl
if cf> E Xm. The space X(~n) is then the inductive Iimit of the spaces Xm
as m- oo.
If (') is either X, <'Je, (')MorS', it is true that (')(~n) n (')'(~n) = S(~n).
In even dimensions, it turnsout that <9e(~ 2 n) and X(~ 2 n) are involutive
subalgebras of :M; moreover, whenever a E <9M(~ 2 n) and b E <9e(~ 2 n),
their Moyal product a xh b makes sense and belongs to (')M(~ 2 n) [178].
~ We seek now an asymptotic development for the product (3.27); this is
of course well known, and can be obtained by several techniques; but we
want to pointout that perhaps the quiekest [167] uses the characterization
of the dual space X' (~n) as the space of distributions that satisfy the
moment asymptotic expansion [170], orthat have order -oo at infinity, in
the distributional Cesara sense. (These matters are treated in more detail in
Section 7.4, to which we refer.) In general, the moment asymptotic expansion
118 3. Some Aspects of K-theory

forT E X'(~n) can be written as

as T - oo,

where the f.lcx are the moments ofT, namely,

Since every polynomiallies in X(~n ), any distribution in X' (~n) has mo-
ments of all orders; this is already an indication that T decays rapidly at
infinity in some sense.
The Fourier kernel eixy, for x,y E ~n. belongs to X'(~ 2 n), and its mo-
ment expansion takes the form

as T - oo.

Let us now define, for oc E ~ 2 n,

Then, by taking T = 2/h, we immediately obtain that, as h l 0,

(a Xh 1 2n
b)(u) = ( 2rrh) J.J a(u + v)b(v + w) exp( 2"ht s(v, w)) d 2 nv d 2 nw

- L "h/2)1011
a 8(v) a 8(w), a(u + v)b(u + w) )
00
(
( l I 01 01

loti=O OC. v,w

= ~ (ih/ 2 )lcxl a a(u) §


01 01 b(u) + O(hN+l) (3.29)
lcxi=O oc!

for all u E ~ 2 n and all smooth functions a, b whose behaviour is not too
wild at infinity; for instance, a could belong to C9M(~ 2 n) and b to C9c(~ 2 n),
or vice versa.
When one of the factors is a polynomial, the development converges. In
particular,
ih aa ih aa
q Xh a = qa + 2 ap; p Xh a = pa - 2 aq . (3.30)

Weshall only need that

(a xh b)(u) = a(u)b(u) + O(h), (3.31)

that could have been obtained as the first term of a stationary phase approx-
imation to the integral defining a Xh b, around the critical point (v, w) =
3.5 The Moyal asymptotic morphism 119

(0, 0) -like in [4 76], for instance. At any rate, the task now is to convert the
pointwise estimates into norm estimates. With a still a Schwartz function,
the quantization rule (3.26) is given by a well-behaved integral kernel:

where
k~(x,y) := (2rr1'1)-n f~"a(x;y,p)eip(x-y)fhdnp.
Wegetanorm bound IIQI'!(a)ll ~ c whenever J~" lk~(x,y)l dny ~ c for
all x, since then

l(h I Ql'!(a)f}l 2 ~ r r lh(x)k~(x,y)f(y)l 2 dnydnx


J~n J~n
~ r clh(x) 2 dnx r clf(y) 2 dny
1 1
J~n J~n
by the Schwarz inequality in L 2 ( ~ n). Integration by parts N times gives
(IX- yi 2N + h 2 N)k~(x,y)
= n2N r eip(x-y)/h(!l2N + 1)a(x + y P) dnp
(2rrh)n J~" P 2 ' '
where flp :=- Ii:= 1 o2 /op~. Since a E S(~ 2 n), this yields the estimates

CNh2N-n
lk~(x,y)i ~ lx- yi2N + h2N'
for all N E ~ and some constants CN [243]. It follows that

l lk~(x,y)ldny~cNh 2 N-n0.n l oo yn-1 dr


2N h2N =CNO.n
i"" sn-1 ds
2N 1'
~" o r + o s +
where O.n is the volume of the unit sphere in ~n; since the right hand side
is finite for N > n/2 and is independent of h, it provides a uniform norm
bound for the operators Qh(a). In like manner,
c' n2N-n+1
lk~xhb(x,y)- k~b(x,y)l ~ lx ~yi2N + h2N'
in view of (3.31), and on taking N > n/2 again, we find that the norm of
the corresponding operator is O(h) as h l 0.
After noting that Qh is linear and real, we want to use the last result
to prove the existence of an asymptotic morphism from the C*-algebra
Co(T*~n) to the C*-algebra X(L 2 (~n)). No amount of distribution-theo-
retic wizardry will produce an extension of the Moyal quantization map to
a map from C0 (T*~n) to compact, or even bounded, operators; this is be-
cause boundedness on the operator side demands some degree of smooth-
ness on the function side. The best results in that sense are theorems of the
120 3. Some Aspects of K-theory

Calder6n-Vaillancourt type [178,184,255,395]: if a is a differentiable func-


tion with enough bounded derivatives, then Q11(a) is a bounded operator.
See also [466,476].
However, to establish the existence of a Moyal asymptotic morphism
from C0 (T*~n) to X, it is enough to note that the Moyal quantization
map (3.26) determines a *-homomorphism from S(T"IJ(I. 11 ) into Xoo. This
extends (see Corollary 3.42 in Section 3.8) to a morphism Q: C0 (T*~n)­
Xoo. Torecover an asymptotic morphism extending Q, one composes this
with a continuous section of the quotient map from Cb((0,1],X) to Xoo,
which is given by the Bartle-Graves theorem. In summary, the following
lemma holds.
Lemma 3.32. There is a uniform asymptotic morphism T (rom Co (T* ~ n)
toX(L 2 (~n)) suchthat

foralla E S(T*~n). B

We shall keep the notation Q11 instead of T11 for this asymptotic mor-
phism. It may be extended to an asymptotic morphism from Mn (Co ( T* ~ n))
to Mn (X), still denoted Q11, by working elementwise.
The Moyal asymptotic morphism is the first of a large family of exam-
ples, arising from strict quantizations of Poisson manifolds; several other
examples in this dass are constructed by Landsman in [309].
Before changing the subject, we remark that the very satisfying features
of Q11 give the means in principle to describe (nomelativistic) quantum
mechanics on the arena of classical phase space; it was introduced with
this aim by Jose Moyal [351]. The formula (3.28) allowed Moyal to compute
quantum mechanical expectation values as phase space averages. Indeed,
a natural step for the rigorous description of quantization processes is to
place the classical and quantum pictures of physical systems on the same
mathematical footing, with the aim of drawing an unbroken line between
them. Noncommutative geometry, on the other hand, accounts for classical
geometry in the operator language of quantum physics: we could regard it
as a gambit, in the opposite direction, for the same purposes of quantiza-
tion. However, from this standpoint, it has not yet lived up to its promise.

3.6 Bott periodicity and the hexagon


Two key results of algebraic topology, relevant for index theory, to wit,
Bott's periodicity theorem and the Atiyah-Jänich theorem, are proved in
this book by means of C*-algebra and C*-module techniques. In the pro-
cess, the power and reach of the noncommutative methods are revealed.
We count these as the sources of noncommutative geometry, alongside the
3.6 Bott periodicity and the hexagon 121

Gelfand-Naimark and Serre-Swan theorems, the Shale-Stinespring theorem


(to be expounded in Chapter 6) and the foundational results by Connes in
Chapters 7 and 8. The Atiyah-Jänich theorem belongs to the theory of Fred-
holm operators on C*-modules, which has a different flavour, and is dealt
with in Chapter 4. Bott periodicity is part and parcel of K-theory, as it is
equivalent to the most important property ofthat theory, namely, the exact
hexagon sequence.
Bott's original version of the periodicity theorem concerned the homo-
topy groups of the classical Ue groups. The fact that TTn(U(k)) is inde-
pendent of k for large enough k is referred as "stable homotopy". One can
define the direct limit group U(oo) := limU(k), via the obvious embed-
dings U(k) .... U(k + 1); the nth stable homotopy group of this group is
defined as TTn (U( oo)) := lim TTn (U(k) ), which equals TTn (U(k)) for large k.
The Bott periodicity result is then written as Trn(U(oo)) = Trn+2(U(oo))-
or as Trn(GL(oo)) = Trn+2(Gl(oo)), by the usual polar decomposition trick.
The argument rephrasing the classical version of Bott periodicity in terms
of the K-theory of the spheres is quite old, but it bears repetition here: let E
be a vector bundle over the sphere §n in the stable rangerankE = k > n/2.
We can cover the sphere with two hemispheres UN, U5 , both homeomor-
phic to ~n; therefore E is determined by a single transition function on
UN n Us, which is of the same homotopy type as §n-l; we can assume that
this transition function takes values in U(k). Therefore,
K(§n) = TTn-l (U(k)).
In view of (3.11), this (at least for k > n/2) implies that
TTn-dU(k)) = TTn(BU(k)),

a well-known result in homotopy group theory.


One ought to prove, then, that Ko(§n) = Ko(§n+ 2 ) for n ;e: 0. The original
proof of Bott periodicity used the calculus of variations to establish that
U ( oo) and the second-order loop space 0 2U ( oo) are of the same homotopy
type: we refer to [341] for a complete exposition, or to [135, pp. 498-508]
for a useful summary. In K-theory the scope of Bott periodicity is vastly
expanded, and we aim to prove that the abelian groups K2(A) := Ko(~ 2 A)
and Ko(A) coincide. Nevertheless, the moral of the story so far -especially
the discussion in Section 3.3- is that spheres are very important. They
keep their importance in the approach to Bott periodicity employed here
(which is essentially that of Higson [243] as perfected by [460]), one of
whose ingredients is the theory of asymptotic morphisms.
We need to establish that K0 (C) = K0 (§0 ) = K0 (§ 2). No big deal, the
reader may say. Indeed, K0 (§ 2) = 71. Eil 71., in view of our detailed study of
line bundles over the sphere in Section 2.6, and ker K 0 i (Definition 3.13) is
generated by [L)- [OI). The reduced K-group K0 (§ 2) "" K0 (C0 (~ 2 )) can
then be identified to the set of multiples of the virtual bundle [L]- [Oll,
where L - § 2 is the Hopf or tautologicalline bundle.
122 3. Some Aspects of K-theory

Definition 3.23. This virtual bundle deserves a name, as it plays a central


role in the theory: we call it the Bott element of Ko(C0 (1~ 2 )) and write
17 := [L]- [OI].

~ No big deal, then, except that the identification of Ko ( C) and K(§ 2 ) must
be made in a functorial way. For that, we concentrate on the case n = 1
of the Moyal asymptotic morphism Q, that yields a K-theory map K0 Q
between C0 (~ 2 ) and X. On the other hand, it is also true that K0 (X) ""
Ko(C) = 71, generated by the dass of (any) rank-one projector [p]. There-
fore, KoQ is determined by KoQ(1J) and it is an isomorphism if and only
if KoQ(1J) = ±[p], i.e., if and only if KoQ(1J) is represented by a rank-one
projector in X.
A representative for 1J in M 2 (Co ( ~ 2 )) is

PB- 0 (1 0)0 =
1 (-1 ~)'
1 + lzl 2 i

where PB is the Bott projector of equation (2.20).


Now, the Moyal quantization of z := q + ip is the "annihilation operator"
a12 := x + h(d/dx) of the theory of the harmonic oscillator in quantum
mechanics, acting on L 2 ( ~). The harmonic oscillator annihilation operator
has closed range since it is in fact surjective: ima 12 = L 2 (~); and a~a 12 is
the number operator, with spectrum { 2nh : n E ~}, so 0 is an isolated
eigenvalue. Indeed, ker a12 is the one-dimensional subspace spanned by the
ground state 10) of the harmonic oscillator, namely, the normalized Gauss-
ian function (rrh)- 114 e-x 2 1212 • Note also that (1 + a~a 12 )- 1 is compact.
We consider now the operator

a 12 (1 + a~a 12 )- 1 )
(1 + a~a 12 )- 1

a 12 (1 + a~a 12 )- 1 )
(1 + a~a 12 )- 1 '

that is clearly an orthogonal projector. There are several ways to think

idempotent e(b) := (~
J{} matches the graph of b.
n
of p(a/l). For any closed operator b on a Hilbert space J{, consider the
acting on J{ Eil Jf, whose range { (b~, ~): ~E

Exercise 3.15. Show that the orthogonal projector corresponding to e(all ),


computed by the Kaplansky formula of Section 3.1, is precisely p(a/l). 0
Alternatively, like in [243], one can consider the selfadjoint operator
3.6 Bott periodicity and the hexagon 123

compute the Cayley transform U := (Bh + i)(Bh- i)- 1 , checkthat S :=


(~ ~1 ) U is a symmetry, and form the corresponding projector ~ (1 + S).
The result is again p(ah).
In order to compute the K0 -theory dass of p(ah)- (~ ~). we use a
homotopy argument [159]. Replace a11 by a multiple ta11, with t ~ 1. The
projector p ( t a11) depends continuously on t (in the norm topology), and
one can compute the uniform limit limr-oo p( ta11) by functional calculus.
Notice that fr(A) := At I (1 + t 2 1AI 2 ) - 0 as t - oo, uniformly on sets of the
form { A: lAI ~ r} with r > 0; therefore ta110 + t 2 a~a 11 )- 1 - 0, since 0
is an isolated point of the spectrum of a~a 11 . Moreover, (1 + t 2 a~a 11 )- 1 -
10)(01, the projector with one-dimensional range ker(a~a 11 ) = kera11, and
t 2 a 11 (1 + t 2 a~a 11 )- 1 a~ = 1- (1 + t 2 a 11 a~)- 1 converges to the projector with
range (ker(a 11 a~)).L = (kera~).L = ima 11 = Jf. Thus p(tah) converges in
norm to the projector 1J{ e p with range im a11 e ker a11 in J{ e Jf. Finally,
[pah]- [1J-f e 0] = [0 e 10)(01], which is a generator of Ko(X) = Ko(C).
The readerwill have noticed the similarity between p(ah) and the Bott
projector PB· To prove that
KoQ(J7) = [0 e 10)(01],

and so to conclude that KoQ is indeed an isomorphism between Ko( Co ( ~ 2 ))


and Ko(C), it is enough to checkthat the dequantization of p(ah) tends to
PB as h 1 0. Because of (3.30), that can be rewritten as zx11a = za+h aa;az.
and i x11 a = ia- h aa;az, it is enough to show that the dequantization
of (1 + a~a 11 )- 1 tends uniformly to (1 + zz)- 1 as h 1 0.
Let us call this dequantization A11. Note that (1 + a~a 11 ) is the quanti-
zation of 1 + zi- h. We contend that A11 will be a function of zi. This is
obtained from Moyal quantum mechanics: on solving the harmonic oscilla-
tor problern on phase space, one finds a spectral decomposition in which
the eigenfunctions depend only on H = zi [22, 24, 210]. Denote by Al, and
A;; the first two derivatives with respect to zi; then

Using (3.29) to second order, this simplifies to


(1 + zi- h)Ah- h 2 (A/, + ztA;;) = 1, (3.32)

from which it is plain that there is a development for A11,

L nk ck(zt).
00

A11 = (1 + zz)- 1 +
k=1
The terms of the series can be obtained recursively; indeed, (3.32) can be
solved explicitly as a confluent hypergeometric function. Now, the Ck are
124 3. Some Aspects of K-theory

GLS symbols of order -2- 2k, hence are square summable; therefore the
corresponding operators are Hilbert-Schmidt with norms n- 1 i 211q 11 2 com-
puted by (3.23), using the measure J..lh that contains a factor n- 1 • A fortiori,
the operator norm of the series goes to zero uniformly as h l 0.
In conclusion, we have established the following result.
Proposition 3.33. The Moyal asymptotic morphism Q yields a K -theory iso-
morphism KoQ between Ko(Co(~ 2 )) and Ko(O that sends the Bottelement
to 1 E Ko(C). 8

.,. We develop some relations among K-theorymaps, along the lines of [13].
Suppose the C*-algebras A and Bare unital. There is an obvious map (gen-
erally not surjective)
J1: Ko(A) ® Ko(B)- Ko(A ® B): [p] ® [q] ...... [p ® q],
where p E P(As ), q E P(Bs ). [The notation A®B is ambiguous, since A0B
may be completed in several C*-norms, although if either factor is abelian
or X, the norm is unique: see Section LA. To ensure that the multiplication
map m: A ® A - A is continuous, we can use, say, the maximal C* -norm.]
lf A or B is nonunital, we define the analogaus map J1 by adjoining a unit
and restricting. Replacing both A and B by L2A = Co ( ~ 2) ® A, we denote
also by J1 the composed map

(3.33)

Ifwe write ](t) := j( -t), the automorphismf ...... j of C0 (~ 2 ) induces an


involution z ...... z of K2 (A) in the obvious way. The transformation (s, t) ......
(t, -s) of ~ 2 x ~ 2 is connected to the identity, since it just a rotation by rr /2,
and so the automorphism f ® g ...... g ® j of C0 (~ 4 ) induces the identity
homomorphism on K 4 (A). Combining this with (3.33) gives an equality
J..l(Y ® z) = J..l(Z ® j/), for all y,z E K2(A). (3.34a)
If we replace either copy of A by ([in (3.33), we may dispense with the map
K 4m; we thereby obtain homomorphisms 11: K2(C) ® K2(A) - K4(A) and
11: K2(A) ®K2(C)- K 4 (A). The argument that gave (3.34a) then also yields
J..l(Y ® z) = J1(Z ® j/), for y E K2(A), z E K2(C). (3.34b)
Finally, if x E K0 (A) and z E K2 (C), suspension gives the simpler relation
J..l(X ® Z) = J..l(Z ®X) E K2(A) .

.,. The group homomorphism Q* = K 0 Q: Ko(Co(~ 2 )) - Ko(X) induced


by Moyal quantization extends, via the asymptotic morphism Q~ (a ® b) :=
Qh(a) ® b, to a homomorphism Q~: K2(B) := Ko(C0 (~ 2 ) ® B)- K 0 (X ®
B) ""'Ko(B), for any C*-algebra B. We call this the Bott map. Keep in mind
that Q* is an isomorphism sending 1 E K2(C) to 1 E Ko(C). An argument
based on naturality now yields the full Bott periodicity.
3.6 Bott periodicity and the hexagon 125

Theorem 3.34 (Bott periodicity). For any C* -algebra A, the groups K2 (A)
and Ko (A) coincide.

Proof. Given any *-homomorphism (not necessarily unital) cp: B ..... A, we


write K 2 cp := K0 ~ 2 cp. Then the diagram

commutes. Indeed, (idx ®cp) o (Qh ®idB) = (Qh ®idA) o (idco(IJ!.ZJ ®cp) holds
on simple tensors; since the asymptotic morphism Q is linear, this is an
equality between commuting maps from C0 (1~ 2 ) 0 B to X 0 A. On comple-
tion, it yields maps from Co ( ~ 2 ) ®B to X® A that commute asymptotically;
and it then passes to K 0 cp o Q! = Q~ o K 2 cp at the K-theory level.
Because Q~ 08 = Q~ ® id8 , the following diagram commutes:
Jl
Kz(A) ®Ko(B) --Kz(A®B)

Q1®idB! !Q1"B (3.35)


Jl
Ko(A) ® Ko(B) ~ Ko(A ® B),

where the top arrow is J1: K0 (C0 (~ 2 ) ® A) ® K 0 (B) ..... Ko(Co(~ 2 ) ® A ® B),
and so on. The inverse map P~ from Ko (A) to K2 (A) will be given by

P~(x) := J1(TJ ®X),

where 11 = [p 8 ] - [1] is the Bott element, as in Definition 3.23. Indeed,


replacing A by I( and B by A in the diagram (3.35) gives

Q~ o P~ (X) = Q~(J1(TJ ®X)) = J1(Q* o idA){TJ ®X) = J1(1 ®X) = X E Ko(A).

Therefore, P~ o Q~ is an idempotent endomorphism of Kz(A); for it tobe


the identity, it is enough that it be an automorphism. If y E Kz(A), then

P~(Q~(y)) = J1(TJ ® Q~(y)) = J1(Q~(y) ® TJ)


= Q~zA(J1(Y ® TJ)) = Q~zA(J1(TJ ® jl))
= J1(Q*(TJ) ® y) = y,
where we have used two applications of (3.35):
126 3. Some Aspects of K-theory

as well as (3.34b) and Q* (TJ) = 1 again. Since y ...... y is an automorphism of


K2 (A), it must actually be the identity, and P: is a two-sided inverse of Q~.
In fine, Q~: K2 (A) - K0 (A) is a natural isomorphism. D

The proof of invertibility of Q* goes back to Atiyah (13], as adapted by


Trout [460]. There are other proofs of Bott periodicity for C* -algebras. The
original argument of Atiyah [12] for topological K-theory carries over nat-
urally to the C*-algebraic context [481, Chap. 9]. Another standard proof,
given by Cuntz [124] and cited in Connes' book [91, II.B.E] -see also Mur-
phy's exposition [352, §7.5]- is based on the Toeplitz operator algebra.
An asymptotic morphism construction has also been used by Higson, Kas-
parov and Trout to extend the topological periodicity theorem to the case
of an infinite-dimensional Euclidean space [244]. In all proofs, at least one
step needs considerable effort: in our case, it is the construction of the as-
ymptotic morphism Q; the Botttheoremexhibits the irreducible complexity
that marks the deepest results in mathematics.
Corollary 3.35 (The hexagon). Let ] be an ideal of A. The following dia-
gram is exact:

Ko(A)
/ ~
Ko(]) Ko(A/])

·1
Kr(A/])

Kr(])
(3.36)

~
Kr(A)
/
Proof. In the long exact sequence (3.15), the Bott map allows us to identify
K2 (A/]) and the arrow before it with Ko (A/]) and its preceding arrow, and
so the sequence is short-circuited. Exactness of (3.15) at K2 (A/]) yields
exactness of the hexagon at K0 (A/ ]). D

Actually, the hexagon is equivalent to Bott periodicity: in order to derive


periodicity from the six-term exact cyclic sequence (3.36), just apply it to
the suspension sequence of A. We remark that the Bott map Ko (A) - K2 (A)
is unique, up to a sign, since proof of Theorem 3.34 shows that a natural
transformation of K0 groups is entirely deterrnined by the specific value
takenon the generator [1] of Ko(C) ~ lL, or equivalently on the generator
1J of Kz (!(). For a natural isomorphism from Ko to K2, there are only two
options, given by [1] ...... ±TJ. For the same reasons, the connecting maps
K1 (A/]) ~ Ko (]) and Ko (A/]) ~ K1 (]) are unique up to a sign.
3.7 The K 1 functor 127

Exercise 3.16. Let I,] be ideals in a C*-algebra A, suchthat I+] = A.


Show that there is an exact "Mayer-Vietoris" sequence:

KoU) Eil Ko(J)


/ ~
KoU n]) Ko(A)

I J
0

... In the commutative case, define K 1 (X) := KI(C(X)) ""K 0 (X x ~). There
follows the exact cyclic six-term sequence:

(3.3 7)

Corollary 3.36. The relations TTn(U(k)) = TTn+2(U(k)) = 71. hold forn odd,
2k- 1 2:: n; and TTn(U(k)) = TTn+2(U(k)) = 0 forn even, 2k- 2 2:: n. B

Bott periodicity is indisputably the most important property in K -theory


and it holds for Banach algebras as well; in particular, for commutative
Banach algebras, this means that the K-groups are invariants of the char-
acter space, rather than of the algebra itself (see the comments in [447]).
Bott periodicity is perhaps the most illustrative example of a whole realm
of topological results established by algebraic means. To tell the whole
truth, however, its formulation is rounder in the K-theory of vector bun-
dles, where there is a ring structure on the K -functors, and Bott periodicity
is then beautifully implementedas multiplication by the Bottelement [12].

3. 7 The K 1 functor
lt is time to examine more explicitly the K1 group and the accompanying
maps. The last sections were quite abstract; to make up for that, we take
the liberty of giving a somewhat informal presentation here; the reader
128 3. Some Aspects of K-theory

will easily sharpen most of our remarks into proofs. We first have some
unfinished business to take care of.
Theorem 3.37. K1 is a K-theory functor.

Proof. If A and B are homotopy equivalent C* -algebras, then so are 2A and


2B, by Exercise 1.18, and then K1 (A) = Ko(2A) = Ko(2B) = K1 (B). Clearly,
(2A)s = Co(IR) ® A ®X = 2(As), so KdAs) = Ko((2A)s) = Ko(2A) =
K1(A). Halfexactness of K1 is clear from that of K 0 , since 2 is an exact
functor by Exercise 1.19. Continuity of K1 follows formthat of Ko and 2.
For the normalization, we shall soon establish that K1(I() = 0.
Note that, since 2 is exact, Proposition 3.29 also goes through for K 1. D

Now, in view of what we said about uniqueness of the hexagon in the


Section 3.6, it follows that if we find a functor behaving in every way as K1
should, it must be the K1, even if at first it does not resemble K 0 2. Assurne
first that Ais a unital C*-algebra. A promising candidate [402] is

(3.38)

It is a definition we favour because it exhibits stability. To make it more


intuitive, notice that [Co(m?.),As] ""[C(lf),Asl+· By Fourier analysis, C(lf)
is the C*-algebra freely generated by a single unitary element. Thus, a mor-
phism in K~0 P(A) just selects a unitary in As = (X® A)+. Hence we can
regard K~0P(A) as the group rro(U(As)) of path components of U(As ).
We would like to give this notion a more algebraic flavour. We recall that
X= li!!}.Mn(C). lt turnsout that

K top(A) "" 1. Un(A) =li GLn(A)


1 lm.un(A)o ____mGLn(A)o"

Here Un (.JI.) Stands for the group of unitary elements in Mn (..Jl), with the
subscript 0 meaning the neutral component in the corresponding group.
This is what the reader will find in the textbooks. Although the statement
seems quite plausible, the detailed manipulations are not so straightfor-
ward: after all, K~op (A) is abelian, whereas the intermediate groups are
not. A painstaking treatment along these lines is given in [481, Chap. 7]
and there is little point in repeating it here.
We see immediately that K~0 P(I() = 0. Indeed, U(X+) is connected be-
cause any unitary in x+ is of the form ,\(1 + k), with ,\ E lf and k com-
pact; it has discrete spectrum, and so is logarithmic if ,\ = 1. By multiply-
ing an exponential path from 1 to 1 + k with rotations from 1c to ,\, we
can reach ,\(1 + k) from 1 by a unitary path. As obviously K~0 P(A e B) =
K~op (A) e K~op (B), we conclude that K~op (A) = K~op (A +). This is a wonder-
ful equality, as it allows us to generalize to nonunital algebras at once.
Definition 3.24. Fora nonunital C*-algebra A, take K~0 P(A) := K~0P(A+).
3.7 The K1 functor 129

In the case of AF-algebras, the Krgroup adds no new information, be-


cause it is always trivial. Indeed, by continuity ofthe K1 -functor, it is enough
to notice that KI<Mn(()) = 0, either because KdMnUO) ""KI(C), or sim-
plybecause the unitary groups U(mn), m ~ 1, are connected. Another C*-
algebra with trivial K1 -group is the full operator algebra L(J-f). Any unitary
operator u is logarithmic in L(J-f) since one can define log u by Borel func-
tional calculus, using the Borel function eiiJ ..... (} for -rr < (} 5 rr; and then
t ..... exp(t log u) provides a unitary path from 1 to u. Thus KI(L(J-f)) = 0.
In fact, this argument shows that the K1 -group of any von Neumannalgebra
is zero.
~ It is amusing that (3.38) remains valid in the nonunital case, by the fol-
lowing argument.
Exercise 3.17. Let 0 - A - B - C - 0 be an exact sequence of C*-
algebras. Show that U(A+) isanormal subgroup of U(B+). 0

In the present case, we need to show that rr0 (U(A_5)) ""rr0 (U((A+)_5)).
As there is a split exact sequence

0-X®A-X®A+-X-0, (3.39)

we can consider the quotient U ( (A +) 5) 1U (As) that is identified with an


open subgroup of the connected group U +(X) of unitaries of the form
1 + k, with k compact, and so with all of U+(X). This yields a fibration

and the corresponding lang exact sequence of homotopy groups is

Since (3.39) splits, the second arrow is surjective, so the fourth arrow is an
isomorphism.
Exercise 3.18. Expand the previous remarks into a proper proofthat (3.38)
remains valid in the nonunital case. 0

To work in K~op (A) in practice is quite easy. The multiplication is defined


by

[u] [v] := [uv] = [ (~ ~)].


The second equality and commutativity of the product follow from the
homotopies

0) ~
u
(vu
0
0)
1 '
130 3. Same Aspects of K-theory

which are easily checked by using the tricks developed in proving Lemma
3.11. In particular, we may write the product of [u] and [v] as [uv] or as
[ u Eil v], according to convenience.
Let a E K~0P(A/J), where] is an ideal in a C*-algebra A. We can write
a = [u], with u E Un(A/J), since any unitary in Uco(A/J) is of the form
u E!l1 for some n. Then u Eil u * belongs to U2n (A/ J)o since u Eil u * !!, 12n· We
lift it to an element w of U2 n (A) 0 , as donein the proof of Proposition 3.18.
Then we define

8 (a) := [ w C; ~) w* Lo(J+)- [ C; ~) Lo(J+) (3.40)

which is a nontrivial element of K0 (]) in general. (We have implicitly as-


sumed A tobe unital, but it should be clear how to proceed in the nonunital
case.) As argued in [128], 8 is actually an index map.
Stealing a Iook at [481, §8.1], one checksthat the recipe (3.40) for 8 gen-
erates a well-defined group homomorphism. It is then clear that K~op enjoys
the lang exact sequence property. Hence, there is a twin for the connecting
homomorphism (3.13) from K1 (A/J) to Ko(J). In view of Proposition 3.26,
we are through: our K~op is indeed Ko~ (allowing us to drop our clumsy
notation in favour of simply KI). To verify directly the claim essentially
amounts to arguing that a unitary and a loop of projectors are the same
thing. Let u E Un(A). There are unitaries Vt, given by (3.7), that form a
homotopy between u Eil u* and 12n; define a map from K~op to K1 by

[u]- (t- [Vd1n Eil OnlVr*J- [1n Eil On]). (3.41)


Indeed t - IVt ( 1n Eil On) Vt*] is a loop, but it is not unitarily equivalent to the
constant loop, as Vo * V1 in general. It is clear that this construction closely
resembles the previous one of 8. The direct checkthat the map (3.41) is well
defined and is an isomorphism is given in [481, §7.2], although the author
recommends it "only for K-masochists".
We still have to develop a feeling for the other index map 8: Ko(A/I)-
K1 (I), but this is easier. Given a projector p E Mn (A/ 1), lift it to an element
x = x* in Mn(A). Then
8([p]) = [exp( -2rrix)]KJ(l)·
This is trivial if p lifts to a projector x, since then exp(-2rrix) = 1, as
already remarked in Section 1.5.
Asymptotic morphisms give rise to K1-theory maps as well. The argu-
ment is even easier that for K0 : if T is a unital asymptotic morphism from
A toB and a E Ax, then Th(a) E Ex for small enough h, and its path
component in Bx depends only on the path component of a in A x.
When A = C or Co(~). the natural map from [A,B] to [A,B] is bijective.
Therefore [C0 (~).As] = Kt(A) and, by Bott periodicity, [Co(~).~As]
Ko(A). This is the beginning of wisdom in E-theory [36].
3.7 The K1 functor 131

.,.. We may ask if a similar formulation could be given forageneralring .5\.


So far, the only topological ingredient in KI theory is precisely the role of
the connected component of the identity. We look then for an algebraic sub-
stitute for the notion of neutral connected component. We recall that the
multiplicative commutator of two elements a, b of a group G is aba -I b-I.
Consider the normal subgroup G' generated by such commutators; the quo-
tient group Gab:= G/G' is of course abelian. The algebraic Krgroup of .5\
is an abelianization in that sense:

Exercise 3.19. Show that a ring homomorphism cJ>: .5\ - ~ induces a group
homomorphism GLoo(..:.l) - GLoo(~) that restricts to a map K~1 g(..:.l) -
K~1 g(~). 0

Computing K~Ig is a difficult business, but when .5\ is a commutative


ring there is a trick that is sometimes helpful [53]. The usual concept of de-
terminant then makes sense in GLoo (.5-l): the inclusion maps commute with
taking determinants, as det 1n = 1 and det (u e> v) = det u det v. Thus, there
is a homomorphism det: GLoo(..:.l) - .5\x := GLI(..:.l). Denote by SLoo(..:.l)
the kernel of this map:

U SLn(..:.l),
00

SLoo(..:.l) := ker(det) =
n=I

with an obvious notation. Now GLoo(..:.l)'!;;;; SLoo(..:.l), and therefore det de-
scends to a group homomorphism K~Ig (.5-l) - .5\ x, also called det. Writing
SKI(..:.l) := SLoo(..:.l)/GLoo(..:.l)', the short exact sequence

1 - SKI (.5-l)- K~1 g(..:.l) ~ .5\ X - 1

splits, since det has a right inverse given by the map .5\ x .... GLoo (.5-l) -
K~ 1g(..:.l). Thus K~1g(..:.l) = .5\x e> SKI(..:.l), and so, in the commutative case,
the computation of K~1 g(..:.l) reduces tothat of SKI (.5-l). There are fortunate
instances in which SKI (.5-l) is trivial. This happens, for instance, when .5\
is a local ring or an Euclidean ring [403]. In particular, K~1 g(l) = lz and
K~1 g(F) = P for a field F.
Since any unital C* -algebra A is a ring, it makes sense to define its alge-
braic Krgroup. There is a comparison map K~1g(A) - KI(A). lt is induced
by the identity map GLoo (A)disc - GLoo (A)top on the group GLoo (A), en-
dowed respectively with the discrete and the usual topologies (one topol-
ogizes GLoo (A) by the usual method of taking the finest topology. that
makes the inclusion maps continuous; this topology is paracompact but not
metrizable). At any rate, since the commutator subgroup lies in GLoo (A)o,
the comparison map is surjective. Clearly it is not an isomorphism in ge-
neral, as K~1 g(([) = P. Thanks to the results of [242,444], summarized
132 3. Some Aspects of K-theory

in [404], it is now known that the comparison map is an isomorphism for


stable C* -algebras, as conjectured long ago by Karoubi. Therefore the al-
gebraic theory (in particular the K2 -group, related to central extensions of
groups), which could have tantalizing applications in quantum field theory
(see [105] in relation with our Chapter 13), keeps a normative character in
the K-theory of C*-algebras.
~ We close with a direct construction of an important -although quite
commutative- example of the index map, that is very instructive. We see
that KI(C("U")) = 71, since

(3.42)

By the way, we also get

Ko(C("U")) =Ko((IC)+) =71eKo(IC) =71eKI(C) =71.

However, we would rather have a more explicit presentation of KI(C("U")).


Comingback to det, we note that it maps GLn(A)o into (Ax)o = expA when
Ais commutative and unital, as discussed already in Section 1.5. Therefore,
there is the induced map det: K1 (A) - Ax I expA, with the obvious right
inverse given by the injection

AxlexpA = GLI(A)IGLI(A)o ..... KI(A).

By Theorem 1.13, H 1 (M(A), 71) is a direct summand of K1 (A). In our case,


C("U")xlexp(C("U")) "'H 1 ("U",71) "'71: as there is nothing else, the classes of
the unitaries eicf> -eine/> constitute K1 (C("U") ).
Now remernher the short exact sequence (1.13):

0 - Co(IR{ 2 ) - -
C([J))- C("U")- 0.

Here we identify IR{ 2 or (( with the open unit disk [J) by z = e-icf> cot ~ ......
( 1 - () lrr)e-icf>; that is, we regard IR{ 2 as the range of the UN chart for the
sphere § 2 = [J)I"U", as in Section 2.6, using the complex coordinate z given
by (2.16). For any integer m, we choose am =[um] in KI(C("U")) where Um
is the unitary e-icf> ...... e-imcf>, and compute 8(am) as in (3.40). For that, we
lift Um e uf:n to the following unitary element in U2 ( C (ii»)):

Wm(Z) :=
1 (zm -1)
-
.J 1 + zmzm 1 Z
-m ·

Since zm I .J1 + zm .zm - e-imcf> as Iz I - oo, this formula extends con-


tinuously to the boundary and coincides with Um e uf:n there. Note that
3.8 K -theory of pre-C*-algebras 133

O(am)(Z)=Wm(Z) (1 0) *
O O Wm(Z)- (1 0)
O O

1 (zmzm zm) _ (1 0)
= 1 + zmzm zm 1 0 0 .

That is to say, we get the K-theory elements of each and every line bundle
over the sphere! In particular, as already seen in Section 2.6, we get the Bott
element when m = 1. The point is that, since Bott periodicity is used only,
in (3.42), to guarantee that KI(C(lr)) contains no more than the unitaries
ein<J>, we could have obtained the Bottelement from the connecting homo-
morphism even if we had never heard of it before! We remark that, from the
long exact sequence too, 8 must be an isomorphism, and its image must
consist of nontrivial elements of K0 (§ 2 ).
Exercise 3.20. Prove that Ko(C(lr -A)) = K1 (C(lr -A)) = Ko(A) e K1 (A),
and then, as c (lrn) "" c (1r- c nn-l))' show by induction that
0

Exercise 3.21. The quotient C*-algebra Q_(J{) := L(J.f)/X(J{) is called


the Ca/kin algebra. Establish that K1 (Q. (J{)) "" 71.. 0

3.8 K-theory of pre-C*-algebras


It stands to reason that, as soon we get into the differential aspects of non-
commutative geometry, many of the algebras used will not be C*-algebras
at all, just as the algebra C"" (M) of smooth functions on a compact manifold
is only a dense subalgebra of the C*-algebra C(M). That is to say, they will
be represented faithfully as algebras of operators, but need not be complete
in the uniform norm. We would prefer that topological properties, such as
K-theory, remain unaffected by passage to such dense subalgebras. In this
prophylactic section we identify a dass of algebras, called pre-C* -algebras,
whose K -theory is the same as that of their C* -algebraic completions. They
include the algebras C"" (M).
Suppose that J\ is a dense *-subalgebra of a C*-algebra A, with inclu-
sion i: J\ - A. We wish to determine whether Koi: Ko(Jl) - K 0 (A) is an
isomorphism of abelian groups. First of all, Ko(Jl) has to be defined as
the algebraic Ko-group of Section 3.2, since the analogue of P(As) is not
available; but Theorem 3.14 shows that nothing is lost by doing so.
Recall from Section l.A that the holomorphic functional calculus is avail-
able for unital Banach algebras. If a E A has spectrum sp a, a compact sub-
set of (, and if f is holomorphic on an open neighbourhood of sp a with
134 3. Some Aspects of K-theory

smooth positively oriented boundary r, then the Dunford integral (1.19)


defines an element j(a) E A (which in fact depends only on the germ of f
on spa); also, f .... j(a) is an algebra homomorphism. (lf Ais a nonuni-
tal Banach algebra, we take the spectrum in A + and use only functions f
satisfying j(O) = 0.)
Definition 3.25. We say that a subalgebra B of a unital Banach algebra A
is stable under the holomorphic functional calculus of A if (i) B is com-
plete under some locally convex topology finer than the topology of A; and
(ii) whenever b E Band j(b) is manufactured by the integral (1.19) using
the spectrum of bin A, then j(b) actually lies in B. (When Ais nonunital,
this is required only when j(O) = 0.)
The completeness condition is only needed to guarantee that the vector-
valued integral in (1.19) makes sense in B whenever the integrand consists
of elements of B; see, for instance, [407, Thm. 3.27]. Given this complete-
ness requirement, it is clear that B is stable under the holomorphic func-
tional calculus if and only if whenever b E B is invertible in A, then b -I E B
(briefly: B n Ax = Bx).
Example 3.1. To see that some completeness condition is necessary, con-
sider the C*-algebra A = C(l) and the dense subalgebra B consisting of
(restrictions to I of) rational functions on <C with no pole on the interval I.
The set B n Ax = Bx consists of rational functions with neither poles nor
zeroes on I. However, the Dunford integral, applied to the identity func-
tion z .... z in B, reduces to the ordinary Cauchy integral formula, so it can
produce any function holomorphic in a neighbourhood of I; thus, Bis not
stable under holomorphic functional calculus.
Lemma 3.38. Let A be a unital Banach algebra and let B be a dense subal-
gebra with 1A E B. In orderthat B n Ax = Bx, it suffices that B n V s; Bx for
some open set V in A.
Proof. If b E B has an inverse a E A, then a V n V a is an open set in A
and so contains an element c E B. Then bc and cb lie in V, so there exist
y,z E B with bcy = zcb = 1; thus cy(= zc) is an inverse for b that lies
~B. D
Definition 3.26. A pre-C*-algebra is a subalgebra of a C*-algebra that is
stable under holomorphic functional calculus.
Proposition 3.39. If5\ is a pre-C* -algebra, then Mn (51.) is a pre-C* -algebra
for all n.
Proof. Let A be the C*-algebra completion of 5\. We show that Mn (Jl.) is
stable under holomorphic functional calculus in Mn (A). The original to-
pology on 5\ yields a topology on Mn (51.) for which it is complete, so it
remains to show that matrices over 5\ having inverses in Mn (A) are al-
ready invertible in Mn(Jl.). By the previous Iemma, it is enough to show
3.8 K-theory of pre-C*-algebras 135

that a E Mn (5\} is invertible whenever it lies in some fixed neighbourhood


of the unit 1n in Mn(A).
Now if a matrix a = [aiJ] is close enough to 1n, its inverse can be com-
puted by Gaussian elimination. The standard "pivoting" procedure, famil-
iar when A = C, provides a factorization a = ldu where d is an invertible
diagonal matrix and l and u are respectively lower and upper triangular
matrices with 1's on the diagonal; e.g., in the case n = 2,

More generally, l and u are products of elementary matrices (that differ


from 1n in only one off-diagonal entry) and the diagonal entries of d are
the "pivots" au, b22 := azz- az1a1la1z. C33 := h3- b3zb2:l bz3, and so on,
produced by the elimination. Take a E Mn (A) with each 111- akk II < 8, and
II aiJ II < 8 for i * j; then for small enough 8, the pivots a u, bzz, C33, ... are
close enough to 1 that they are indeed invertible in A; by hypothesis, their
inverses lie in B. Therefore, a- 1 = u - 1d-l[- 1 where u - 1, L- 1 are products
of elementary matrices and d- 1 = diag(a!l, bzi, c3l, ... ), so that a- 1 E
Mn(A). 0

Example 3.2. Consider S( ~ m), the algebra of smooth rapidly decreasing


functions on ~ m, whose topology is defined by the norms

This is a Frechet algebra whose C* -completion is Co ( ~ m ); notice that II · llo


is the usual sup norm. If f E S(~m) and 1 + f is invertible in Co(~m)+""'
C(§ 2 ), then (1 + j}- 1 = 1 + g where g E C0 (~m). Since g is smooth and all
its derivatives are bounded, because one can regard 1 + g as an element of
C""(§ 2 ), it follows that fg E S(~m), and the equationf + g + fg = 0 then
shows that g E S(~m). Therefore, S(~m) is a pre-C*-algebra .

.,.. To deal sensibly with locally convex algebras whose topology is not given
by a single norm, it is crucial to assume the following, which guarantees
compactness of the spectra of their elements.
Definition 3.27. We say that a unitallocally convex algebra Ais good if the
set of invertible elements A x is an open subset of A. In particular, every
Banachalgebra is good. (Kaplansky coined the name "Q-algebra" for these,
but the term seems to have fallen into disuse, and in any case nowadays
could be misunderstood by q-theorists.)
Outside the realm of Banach algebras, good behaviour is the exception
rather than the rule. For instance, the algebra of continuous functions C(Y)
on an open domain Y c C, with the topology of uniform convergence on
compacta, is a Frechet algebra, but if {zk} is a sequence of points in Y
136 3. Some Aspects of K-theory

converging to a boundary point Zoo, then the invertible function z -Zoo is


approximated by the noninvertible functions z - Zk.
Exercise 3.22. Show that if .Jl is a good locally convex algebra, then Mn (.Jl)
is a good locally convex algebra for all n. 0
Lemma 3.40. If a pre-C* -algebra .Jl has a locally convex topology stronger
than that of its C* -algebraic completion A, then .Jl is a good locally convex
algebra.

Proof. The hypothesis means that the dense inclusion i: .Jl - A is contin-
uous, so .Jl x = .Jl n Ax = i- 1 (Ax) is open in .Jl. 0

Definition 3.28. If .Jl is a unital pre-C* -algebra and A is its C* -algebraic


completion, the spectrum of an element a E .Jl is sp a : = { A E (( : ( A- a) is
not invertible in .Jl }. Since .Jl x = .Jl n Ax, this coincides with the spectrum
of a as an element of the C*-algebra A.
The spectrum of an element in a nonunital pre-C*-algebra .Jl is declared
to be its spectrum in .Jl +, which coincides with its spectrum in A +.
Lemma 3.41. Any * -homomorphism from a pre-C* -algebra .Jl into a C*-
algebra B extends to a morphism from the C* -algebraic completion A into B.

Proof. By adjoining units to A and B if necessary, we can assume that


we are given a unital *-homomorphism x: .Jl - B. just as in the proof
of Lemma 1.16, it follows that x shrinks spectra and is norm-decreasing:
llx(a) II .:5 llall for all a E .Jl. Therefore, x extends by continuity to its com-
pletion A in the C* -norm, and this extension is a morphism X.: A - B that
is likewise norm-decreasing. 0

We record an important special (commutative) case.


Corollary 3.42. Any *-homomorphism from S(IR{m) into a C* -algebra B is
automatically continuous in the sup norm, and so extends to a morphism
from Co(IR{m) toB. a

.,. In practice, the examples we shall meet are usually Frechet algebras; we
shall always assume that the Frechet topology on .Jl is stronger that the C*-
norm topology on A. For instance, if M is a compact manifold, fk - f in
C 00 (M) if and only if Dfk- Df unifonnly on M, for alldifferential opera-
tors D; by differential operators we mean sums of products of smooth vec-
tor fields on M. The smooth vector fields form a finitely generated (projec-
tive) module 1€(M) over coo (M); by taking finite products of generators, we
obtain a countable basis for the unital algebra of differential operators. The
topology of C 00 (M) is generated by the seminorms f ..... supxEM IDf(x)l,
as D runs over this basis; thus, coo (M) is a Frechet space.
3.8 K-theory of pre-C*-algebras 137

Lemma 3.43. lf .J\ is a Frechet pre-C* -algebra, there is an open neighbour-


hood U ofO suchthat the set ofidempotents Q(.J\) = {e E .J\: e 2 = e} is a
deformation retract of the open set Qu (.J\) = {v E .J\ : v 2 - v E U } .

Proof. Each element x E .J\ has compact spectrum, and if r(x) < 1/4,
the element v := j(x) provided by functional calculus from JUI.) := ~ -
~v'1 + 4,\ satisfies v 2 - v = x. (Here j(,\) is holomorphic for lAI < 1/4,
and we take the branch of the square root for which JI = + 1; note that
j(O) = 0.)
Since .J\ is good, there is a convex open neighbourhood U of 0 such that
1 +4x E .J\ x and r(x(1 +4x)- 1) < 1/4 whenever x E U. If v E Qu(.J\), then
(1-2v) 2 = 1+4xisinvertibleandy := -x(l+4x)- 1 = (v-v 2 )(1-2v)- 2
satisfies r(y) < 1/4. Thus we may define, foreachsuch v and 0 ~ s ~ 1,
1/Js(V) := v + (1- 2v) j(s(v- v 2 )(1- 2v)- 2 ).

Clearly, 1/Jo(v) = v, and


I/J 5 (v) 2 -!Jl 5 (V) = v2 - v + (1- 2v) 2 (j(sy) 2 - j(sy)) = (1- s)(v 2 - V),

so that !J1 5 (v) E Qu(.J\) for all 0 ~ s ~ 1 and !Jldv) E Q(.J\). D

Therefore, in a Frechet pre-C* -algebra, one can replace the set of idem-
potents Q (.J\) by a suitable open neighbourhood Qu (.J\) to which it is ho-
motopy equivalent; we may use the latter instead of the idempotents alone
to construct Ko(.J\). Now, a quite different argument shows that an open
set V in the C* -completion A is homotopy equivalent to the intersection
V n .J\. Namely, a theorem of Milnor [339] shows that an open subset of a
Frechet space has the homotopy type of a CW-complex. Therefore, to show
the inclusion j: V n .J\ - V is a homotopy equivalence, it is enough to show
that all of their higher homotopy groups are thereby isomorphic, i.e., for
anypointv 0 E Vn.J\,weneedtoknowthatrrk}: 7Tk(Vn.J\,vo)- 7Tk(V,vo)
is an isomorphism for each k.
The argument for k = 0 is easy; recall that the functor rro just counts
the path components of a space. Each connected component of V is open,
and so meets V n .J\ since .J\ is dense in A. On the other hand, if x and y
in V n .J\ lie in the same component of V in A, we can connect them via
a polygonal path, i.e., a chain of segments [x, vd, [v1, Vz], ... , [vm,yl,
with each Vi E V. Then we can find nearby points Ui E V n .J\ so that each
[x, ud, [u 1, uz], ... , [um.Yl isapolygonal path in V n.J\ linking x and y.
This shows that rroj: rro(V n .J\)- rro(V) is bijective.
The general case is then handled by a standard homotopy trick [40]. If *
is a basepoint of the sphere §k, we replace A by A<kl := {jE C(§k-A):
f(*) = vo}, and also A by .J\(kl and V by V(k), where these are the subsets
of A<kl satisfying j(§k) c .J\ andj(§k) c V, respectively. Take as basepoint
for these sets the constant map with value vo. Then 1Tk (V, vo) = [§k, V]+ =
rr0 (V(kl· vo); the bijectivity of rrk} is established by perturbing paths in V<kl·
138 3. Some Aspects of K-theory

With these preparations, we come to the main result of this section.


Theorem 3.44. If 51. is a Frechet pre-C* -algebra with C* -completion A, the
inclusion i: 51. ..... A induces an isomorphism Kai: Ka(51.)- Ka(A).

Proof. We may assume that 51. is unital, since the nonunital case follows by
the familiar argument. Choose a convex open neighbourhood U of 0 in A
so that 1 + 4x E Ax and r(x(1 + 4x)- 1 ) < 1/4 whenever x EU. If y E 51.,
the spectra of y for the algebras 51. and A coincide, since 51. n A x = 51. x, so
that U n 51. has the analogaus properties in the algebra 51..
The inclusions Q(A) ..... Qu(A) and Q(51.) ..... Qun5t (51.) are homotopy
equivalences and so is the inclusion Qun5t (51.) ..... Qu (A), since

Qu(A) n51. = {v E 51.: v 2 - v EU}= Qun51.(51.).

Thus the inclusion Q(51.) ..... Q(A) is yet another homotopy equivalence.
By Proposition 3.39, for each k = 2, 3, ... , Mk (51.) is again a Frechet pre-
C*-algebra that is densein the C*-algebra Mk(A), so the inclusion of their
idempotent sets Qk (51.) ..... Qk (A) is the final homotopy equivalence that we
need. It remains only to show that two homotopic idempotents in Qoo (51.)
are equivalent in the sense of Definition 3.7; this we leave for the next
exercise. Thus the induced map Kai: Ka(51.) - Ka(A) is bijective. D

Exercise 3.23. Show that if 51. is a unital locally convex algebra and two
idempotents e,f E Qm (51.) are homotopic, then they are conjugate via
some v E GLoo (51.). 0
It may happen that Kai: Ka(B) - Ka(A) is an isomorphism for a dense
subalgebra B of A that is not stable under holomorphic functional calculus.
Bast [40], whose treatment we have followed here, gives the following ex-
*
ample. Consider the Banach -algebra B consisting of functions continuous
on X:= { z E <C : ~ :::; [z[ :::; 2} and holomorphic on ~ < [z[ < 2, with the
involution f* (z) := j(l I z). Restrietion to the unit circle 1r yields a homo-
morphism j: B - C(lr) that is injective (by analytic continuation) and has
dense image by the Stone-Weierstrass theorem, since f ... f* yields com-
plex conjugation in C(lr). Thus j(B) is a dense subalgebra of C(lr) (with a
stronger topology: the sup norm in C(X) majorizes that of C(lr)). However,
if za EX\ lr, thenj(z- za) is invertible in C(lr) but not inj(B), so j(B) is
not a pre-C*-algebra. Even so, K0 j is an isomorphism, by [40, Thm. 1.1.1].
This example shows the importance of criteria to decide when a given
dense subalgebra of a C*-algebra is or is not a pre-C*-algebra. The most
useful criterion is the following.
Proposition 3.45. Let A be a C* -algebra, G a Lie group and ()(: G - Aut(A)
a strongly continuous action. The dense subalgebra A"" of smooth elements
of A under this action is a Frechet pre-C* -algebra.
3.8 K-theory of pre-C*-algebras 139

Proof The continuity condition is that, for each a E A, the map G - A : t .....
()(t(a)is continuous; A 00 consists of those a for which this map is smooth.
The Lie algebra g of G and its universal enveloping algebra 11(g) act on Aoo
in the obvious way, and the seminorms a -iiD(t . . . ()(t(a)) lt=ell. for D E
11(g), define a Frechet-space topology on A 00 • Since j(()(t(a)) = ()(t(j(a))
for any function holomorphic in the neighbourhood of sp a = sp ()(t (a), it is
clear that A oo is closed und er the holomorphic functional calculus in A. 0
Finally, we note that the Frechet algebra coo (M), forM a smooth compact
manifold, is a pre-C* -algebra. Indeed, any function f E coo (M) is invertible
in C(M) if and only if it does not vanish on M, and then its reciprocall I f
is also a smooth function. In particular, Ko (coo (M)) ""' K0 (M)!
4
Fredholm Operators on C*-modules

4.1 Fredholm operators and the Atiyah-Jänich


theorem
By definition, two operators S, TE L(J{) have the same image in the Calkin
algebra Q (J{) := L (J{) 1X (J{) if and only if S is a compact perturbation
ofT, i.e., S = T + K for some compact operator K. As we shall see later
on, the compact operators can be regarded as "infinitesimal elements" of
L(J.f), and it is of interest to know what properties of an operator are
unchanged by compact perturbations. For instance, an invertible operator
does not remaininvertible (think of 1-1~) (~1. where ~ E J{ is any nonzero
vector), but we may recall the following well-known result, called Atkinson's
theorem [367, Prop. 3.3.11].
Proposition 4.1. A bounded linear operator F E L(J{) has an invertible
image in Q(J{) if and only ifthere isanother operator G E L(J{) suchthat
1 - FG and 1- GF are compact, if and only if both the kemel and cokemel
of F are finite-dimensional and F has closed range.

Proof. If ry: L(J{) - Q(J{) is the quotient map, then ry(F) and ry(G) are
inverse if and only if '7 ( 1 - F G) = '7 ( 1 - GF) = 0, so the first two conditions
are equivalent.
We recall that coker F = J{ I imF, where the notation means the ordinary
vector-space quotient, and that ker pt = (imF).L is isomorphic to the quo-
tient of J{ by the closure of im F. Therefore, we can say coker F "" ker pt if
the range of F is closed. Assurne that the first two conditions hold. Then
142 4. Fredholm Operators on C*-modules

the closed unit ball of ker Fis invariant under the compact operator 1- CF,
so it is compact; therefore, ker F is a locally compact Hilbert space, which
entails that it is finite-dimensional.
To see that the subspace imF is closed, choose a finite-rank operator R

suchthat II (1- CF)- R II ::5 Then, for ~ E ker R,

ill~ll ::5 11~11 -II~- CF~II ::5 IICF~II ::o; IICIIIIF~II.


and the inequality IIF~II ~ (2IICII)- 1 11~11 implies that the restriction of F
to ker R has closed range; on the other hand, (ker R).L = imRt is finite-
dimensional since Rt also has finite rank, thus imF = F(ker R) + F(imRt)
is closed. Now kerFt is invariant under (1- FC)t = 1- ctFt, so it is
finite-dimensional; since coker F "" ker Ft, it has the same finite dimension.
Conversely, if ker Fand coker F arefinite-dimensional and imF is closed,
we can construct a suitable C by defining C(F(~)) := ~ for ~ E (ker F).L
and C(() := 0 for ( E (imF).L = ker Ft; this is well defined because
F: (ker F).L - imF is bijective. D
Definition 4.1. A bounded linear operator between Hilbert spaces, F: J{ -
J{', is a Fredholm operator if ker F and coker F are finite-dimensional and
F has closed range. lts index is then defined as the integer

indexF := dimker F- dimcoker F. (4.1)


With the obvious modifications, Proposition 4.1 shows that Fis Fredholm if
and only if there is another bounded operator C: J{' - J{ such that 1J{· -
FC and 1J{ -CF are compact. When J{' = J-f, we shall write Fred(J-f), or
simply Fred, for the space of Fredholm operators on J-f, with the topology
induced by the norm topology of L(J-f). Note that Fred is a multiplicative
semigroup.
Remark. lt is worth mentioning that the closedness of the range of F is
actually a redundant condition, in view of the following argument. If coker F
is finite-dimensional, then we represent each coset by elements of a finite-
dimensional supplement to imF; for example, ker Ft will do. If ( E J-f,
then ( + imF = p + imF for a unique p E ker Ft, so that ( = p + F~ for
some ~- Therefore the continuous linear map F: J{ $ ker Ft - J{ given
by F(~,p) := p + F~ is surjective. Now the open mapping theorem [383,
Thm. III.10) tells us that imF = F(J-f $ 0) is a closed subspace of J-f. Even
so, the condition is usually kept, as it is hard to tell beforehand whether
coker F is finite-dimensional. In fact, the most useful way to see that an
operator is Fredholm is to determine that both ker F and ker Ft are finite-
dimensional and (no Ionger redundantly) that F has closed range.
The formal difference (ker F - coker F) can be thought of as a finite-di-
mensional "virtual Hilbert space". lts "virtual dimension" is the index (4.1).
The theory of this index is developed in many good books (for instance,
[352) or [367)); here we shall only recall the main properties:
4.1 Fredholm operators and the Atiyah-Jänich theorem 143

index: Fred- 7L is a continuous map;


it is also a semigroup homomorphism, i.e., index(FIFz) = indexF1 +
indexFz;
the index is insensible to compact perturbations: index(F + K)
indexF if K EX;
index F = 0 if and only if F is a compact perturbation of an invertible
operator.
the standard (unilateral) shift operator [229] is Fredholm and it has
index -1.
Note that index(F) = dimker F- dimker pt = dimker pt F- dimker ppt.
1t is clear that index(Ft) = -index(F); to sum up, the index splits the
space of Fredholm operators into a countable number of connected compo-
nents, one foreachinteger value. This follows only because the zero-index
operators form a connected set, which happens because the group of all
invertible operators GL(J-f) is connected, which in turn is true because the
group of all unitary operators U (J{) is connected (in the norm topology).
This last statement holds because every unitary operator U is logarithmic,
i.e., U = exp ( i T), for some T = yt, by the spectral theorem, as was re-
marked in Section 3.7. Actually, a farnaus theorem of Kuiper [299] asserts
that U(J{) and GL(J{) are even contractible, in marked cantrast to the
finite-dimensional case, where rri(U(n)) = rri(GL(n,C)) = 7L.
Exercise 4.1. Develop these hints into a proof that the set of Fredholm
operators of index zero is connected. (Use the polar decomposition T =
UITI ofbounded operators to reduce from GL(J{) to U(J-f).) 0

~ A formula for the indexwill be handy later. The last part of the proof
of Proposition 4.1 shows that the operator G can be chosen so that 1- GF
and 1 - FG are not merely compact but of finite rank; indeed, with the
above choice of G they are the finite-rank projectors on ker F and ker pt
respectively. lt follows that

indexF = Tr(lJ{- GF)- Tr(1J-l•- FG). (4.2)

Now, there are other possibilities for G; for instance, when F is a Fredholm
operator of index zero, one may choose G invertible. But formula (4.2) still
holds, provided 1- GF and 1- FG remain traceclass. Let G' be one such
alternative. Then

Tr(1.1l- G' F)- Tr(1J-l•- FG') = Tr(1.1l- GF)- Tr(1J{·- FG)


+ Tr((G- G')F)- Tr(F(G- G')),
and the last two terms cancel. More generally, the next result holds [62].
144 4. Fredholm Operators on C*-modules

Proposition 4.2. Suppose that ( 1 - GF)N and ( 1 - FG)N are traceclass for
some positive integer N. Then the index ofthe Fredholm operator Fis given
by

indexF = Tr(1J{- GF)N- Tr(1J-f'- FG)N. (4.3)

Proof. Let us remark, before proceeding, that if the formula indeed holds
for a given N, it holds for all integers greater than N. We have already
dealt with the case N = 1, so assume that N > l. We replace G by G' :=
G + G(1- FG) + · · · + G(l- FG)N- 1 • Using the equality

(1 - X) ( 1 + X + · · · + XN - 1 ) = 1- XN,

we see that 1-G'F = (1-GF)N and 1-FG' = (1-FG)N, and the conclusion
follows from (4.2). D

Proposition 4.2 is weil known; we have more or less reproduced the argu-
ment by Hörmander [253]. See also [86, Appendix 1], for a spectral theory
proof of the last two results .
.,. This chapter is motivated by the following remarkable theorem of Atiyah
[12] and jänich [261]. Let X be a compact space; then

K 0 (X)"" [X,Fred].

This result is striking. One would naturally expect the classifying spaces for
K 0 -theory tobe objects of a geometrical rather than an operatorial nature;
after all, vector bundles are classified by Grassmannians, which are very
geometrical objects -recall (3.11). Instead, a hidden gate appears, lead-
ing to the world of operator algebras, and eventually to noncommutative
geometry.
Theorem 4.3. For any compact Hausdorff space X, there is an isomorphism
ofgroups
index: [X,Fred]- K 0 (X),
which is natural, i.e., for any continuous map 4>: Z - X,

index o 4>* = K 0 cf> o index.


lf X is a point, then rro (Fred) "" 71., the isomorphism being the ordinary index
map. 8

Here 4>* arises from the Gelfand mapping h ,_ h o 4> from C(X- Fred)
to C(Z- Fred), by descending to homotopy classes. We do not prove this
theorem yet, as our plan is precisely to supersede it by a noncommu-
tative version. However, it is instructive to look at what is involved. Let
F: X - Fred : x ,_ Fx be continuous. The field of virtual Hilbert spaces
x ,_ ker Fx - coker Fx ought tobe a virtual vector bundle. Unfortunately,
4.2 Fredholm operators on C*-modules 145

the dimensions of ker Fx and coker Fx arenot locally constant, so this map
does not define an element of K 0 (X) in general. Compactness of X allows
us to overcome this objection by the following device. Replace {ker Fx} xEX
by a trivial bundle with fibre J{ /V, where V is a dosed subspace of J{
of finite codimension suchthat V n ker Fx = {0} for all x; we can take
V:= n(kerFx;).L for suitably chosen points X1, •.. ,Xm. Write [JfJV] for
the dass of the trivial bundle X x (Jf /V) - X. It can also be arranged that
each Fxj (V) is dosed and that Ef := l!:!xEX J{ I Fx (V) is the total space of
a locally trivial vector bundle over X; this bundle replaces {coker Fx} xEX.
The index of the map F is now defined by
indexF := [J{ /V]- [Efl E K 0 (X). (4.4)
Exercise 4.2. lf V' is another subspace with the same properties, and if
V' ~ V (as we may suppose, because V n V' has again the same properties),
show that
[JfjV']- [Jf!V] =[V/V']= [Ef']- [Efl.
Condude that the right hand side of (4.4) is independent of V. 0
Exercise 4.3. If G: X x I - Fred is a continuous map defining a homo-
topy between Go and G1, the inclusions it: x - (x, t) E X x I induce
restriction isomorphisms ii: K 0 (X x 1) - K 0 (X). Show that ii (index G) =
index Gt. Condude that (4.4) yields a well-defined map (also called index)
from [X,Fred] to K 0 (X). o
The next steps in the proof, for details of which we refer to the Appendix
of [12], establish that (a) the map index: [X,Fred] - K 0 (X) is a group
homomorphism, (b) there is an exact sequence

[X,GL(Jf)] -[X,Fred] ~ K0 (X) -0,

and (c) an appeal to Kuiper's theorem shows that [X, GL(Jf)] = 0, since
GL(Jf) is contractible, so the index map is an isomorphism. In fact, the
full strength of Kuiper's theorem is not needed to prove (c); it is enough to
know that the unitary group of the C*-algebra C(X) ® L(Jf) is connected,
since any homotopy of continuous maps from X to U(Jf) can be regarded
as a continuous path in that unitary group.
Jänich [261] finishes the proof in a different manner: after embedding
Fred(Jf) in Fred(J{ e J{) via F - Fe 1, he shows that the index of F in
K 0 (X) depends bijectively on the dass of F e 1 in [X, Fred (Jf e Jf)] and
uses Kuiper's theorem to pullthisback to the dass of F in [X, Fred].

4.2 Fredholm operators on C*-modules


The previous characterization of K 0 (X) = Ko (C (X)) strongly suggests tak-
ing a Fredholm-operator approach to K-theory of C*-algebras by forming
146 4. Fredholm Operators on C*-modules

homotopy classes of Fredholm operators on C*-modules over C*-algebras


such as C(X), or indeed over noncommutative C*-algebras. We can hardly
expect that all properties of the usual Fredholm operators, such as closed-
ness of the range, hold more generally. However, "invertibility modulo com-
pact operators" clearly makes sense in a more general context, so we adopt
it as a definition.
Definition 4.2. Let 'E, :J be right C*-modules over a C*-algebra A, and Iet
FE HomA('E,:J). We say that Fis an A-Fredholm operator if 1.r- FG E
End~ (:J) and 1x- CF E End~ ('E) for some G E HomA (:J, 'E). An equivalent
definition, when 'E = :J, is to ask for invertibility in EndA ('E) I End~ ('E). We
write F E FredA ('E, :J), or FE FredA ('E) when 'E = :J. Note that we do not
suppose A to be unital.
Exercise 4.4. If F E HomA('E,:J) is A-Fredholm, it is also A+-fredholm
when regarded as an element of HomA + ( 'E, :J). 0

~ The range imF of an A-Fredholm operatorwill not be closed in general.


A nice example, still in the realm of commutative algebras, is the following.
Take A = C(I) and 'E = A. Define F: A - A by Fa(t) := t a(t) for 0 ::5
t ::5 1. Since A has a unit, End~ (A) = EndA (A) "" A, so any operator is A-
compact, and any operator, F in particular, is also A-Fredholm. But imF is
not closed: the function g(t) := .jf is clearly not in the range of F, although
it can be uniformly approximated by polynomials with constant term 0,
using the Weierstrass approximation theorem. (For instance, the Bernstein
polynomials L.:k= 1 (~)vk!n tk<1- nn-k will do.)
Exercise 4.5. Explain how FE C(X- Fred) determines an operator on the
C* -module Jfc(XJ and show by an example that this operator need not have
closed range. 0
lt turns out that this difficulty can be overcome by extracting the alge-
braic essence of the closed-range condition. Consider again the construc-
tion in Proposition 4.1, which for a given F E Fred determines G so that
1- FG and 1- GF are the respective projectors on ker Ft and ker F. Then
FGF = Fand GFG = G.
Definition 4.3. Given TE HomA('E,:J), a pseudoinverse forT is an Ope-
ratorS E HomA(:J, 'E) suchthat TST = T and STS = S. In that case, TS
and ST are continuous idempotents, and therefore have closed range. Also,
ker ST = ker T, as Tu = TSTu = 0 for u E ker ST. Moreover, the range of
1x -ST is ker T, and the range of TS is im T, as TSTv = Tv for Tv E im T.
By the same token, ker TS = kerS, imST = imS and im(1.r- TS) = kerS.
Operators endowed with a pseudoinverse are called regular.
The usual Fredholm operators in Fred(J{) are regular, since for each F
we have constructed a pseudoinverse G. The point is that, in the present
context, regularity ensures closed range.
4.2 Fredholm operators on C*-modules 147

Let F be a regular A-Fredholm operator, let G be as in Definition 4.2, and


let S be a pseudoinverse for F. Note that (l:r- GF)(1:r -SF) = (l:r -SF), so
that (1:r- SF) lies in End~ (T); and similarly for (l.r - FS). Therefore any
pseudoinverse S is itself a suitable G. Let e := 1:r- SF and let p E End~ (T)
be its range projector, given by Kaplansky's formula; then p E End~0 (T) by
Corollary 3.10, and e E End~0 (T) also, since pe = e. Similarly, (1.r- FS) E
End~0 (_J).
An important technical pointisthat the A-compact operators appearing
in Definition 4.2 can always be taken to have A-finite rank, whether or not
Fis regular.
Lemma 4.4. Let A be a unital Banach algebra and ] an ideal of A, whose
closure is J. Then, if an element a E A is invertible modulo ], it is also
invertible modulo ] .

Proof. Let b be such that 1 - ab E ]. Then there exists c E ] such that


111- ab- eil < 1, and so ab+ c is invertible. Let b' := b(ab + c)- 1 . Then
1- ab' = 1- ab(ab + c)- 1 = c(ab + c)- 1 ,

which lies in]. A similar argument shows that there exists b" such that
1 - b" a E ]. This implies that a + ] has both a left and a right inverse in
the quotient Al], so it is invertible. 0

~ To relate A-Fredholm operators to K -theory, we need, as an intermediate


step, to relate them to operators on the particular C* A-module J{A· The
key step is the next theorem, due to Kasparov [276], by which any finitely,
indeed countably, generated A-module may be "absorbed" in J{A· We follow
the treatment of Mingo and Phillips [343], simpler than the original proof.
Exercise 4.6. Let T E HomA(T,J") suchthat both T and T* have dense
range. Show that im(T*T) is densein T, and conclude that ITl:= (T*T) 1 i 2
has dense range. o
Lemma 4.5. Let T, _r be C* A -modules and suppose that there is some T E
HomA(T,J") for which T and T* have dense range. Then T and _r are
unitarily equivalent.

Proof. Consider ITI := .JT*T, which has dense range by the previous ex-
ercise. Define the A-linear map V: imT- imiTI by V(Ts) := ITis. This
operator is isometric, since (VT s I VT s) = (s I T* T s) = ( T s I Ts). Because
im T and im ITl are dense, V can then be extended to a unitary operator
in HomA(J, T). 0

Theorem 4.6 (Kasparov absorption theorem). If T is any countably gene-


rated C* A-module, then TEil J{A ""J{A as A-modules.
148 4. Fredholm Operators on C*-modules

Proof. The idea is to find an intertwining operator T between J{A and 'E Eil
J{A with T and T* having dense range. Assurne first that A is unital. Let
{ uk} be a countable set of unit vectors generating 'E, and let (sn) be a
sequence formed by repeating these generators in such a way that every
uk appears infinitely often; let ( ~n) be the canonical sequence of generators
of J{A· Define the A-compact operator T: J{A - 'E Eil J{A by
00

T := I 2-nlsn)(~nl e4-nl~n)(~nl·
n=l

Each timethat Sn = Uk, it happens that T(2n~n) = (uk,2-n~n). Because


this occurs for infinitely many n, we see that each (uk, 0), and then also
each (0, ~n), lies in the closure of the range ofT. Since T* (0, ~n) = 4 -n~n.
we get also that im T* is densein J{A· Lemma 4.5 now yields the result for
unital algebras.
When Ais nonunital, we augment it. We can regard 'E as an A+-module,
and we obtain a unitary isomorphism of A+-modules W: 'E e JfA+ - JfA+·
Now, the closure of JfA+A is JfA, that is, (Jf ® A+)A is densein Jf ® A
since (~ ® 1)a = ~ ® a. Also, the closure of 'EA is 'E, so W extends to an
A-module isomorphism from 'E e J{A to J{A· D
The absorption theorem shows that any A-finite rank C*-module 'E can
be regarded as a submodule of JfA. of the form pJfA with p E P(As) and
p 5 Pn for some n (compare with Theorem 3.8 in the Unital case). We may
regard p as a projector in P( (As )+) with E(p) = 0, where E comes from the
augmentation map A+ - ([; thus [p] isadass in Ko(A), which we denote
also by ['E].
There is a subtlety that should not be overlooked: if A is not unital, J{A
may not itself be countably generated as a C* A-module. In fact, even A
itself may not be singly generated: what is required is that A contain a
positive element h such that the right ideal hA be dense in A. lt turns
out [366, Prop. 3.10.5] that this is the case if and only if A is O"-unital,
that is, has a countable approximate unit. In particular, J{A e J{A "" J{A
whenever Ais O"-unital.

4.3 The generalized Fredholm index


In this section weshall show that Fredholm operators on C*-modules over
a C* -algebra A have an index that takes its values in Ko (A) as expected, and
has the homomorphism property. Our treatment follows the outstanding
article by Exel [171], for the most part. The goal is to obtain an isomorphism
result that can be regarded as the noncommutative Atiyah-]änich theorem.
Definition 4.4. If F E FredA ('E,J') is regular, then ker Fand ker F* are A-
finite rank C*-modules and determine elements of Ko(A). The index of F
4.3 The generalized Fredholm index 149

is defined as
indexF := [kerF]- [kerF*] E Ko(A). (4.5)
Exercise 4.7. If F in FredA(T,.T) is regular and has a pseudoinverse S,
show that F* and S arealso A-Fredholm and regular, and that indexF* =
- index F. Prove that ker S and ker F* are isomorphic as A-modules, and
conclude that index S = - index F. 0
Exercise 4.8. If f 1 E FredA(TI,J'I) and fz E FredA(Tz,.Tz) are regular,
show that F1 fDfz is regular in FredA (TI$ Tz, .T1 $.Tz) with index(fi fDfz) =
indexf1 + indexfz. 0
Lemma 4.7. If F in FredA (T, .T) is regular, and if U E EndA (T) and V E
EndA (J') are invertible, then both FU and VF are regular A-Fredholm Ope-
rators such that index FU = index V F = index F.
Proof. The regularity of FU is clear: if S is a pseudoinverse for f, then u- 1s
is a pseudoinverse for FU. Also, ker(fU) = u- 1(ker f) is isomorphic to
ker F as an A-module, while ker( U* F*) = ker F*, so index(FU) = index F.
The case of VF is similar. D

where TiJ E HomA(T1 ,.Td for j = 1, 2. We next need a criterion to decide


when two A-finite rank modules have the same K-theory dass; this is met
by the next definition.
Definition 4.5. Two right C* A-modules T and .T of A-finite rank are called
stably quasiisomorphic if there is an operator TE HomA (T $ J{A, .T $ J{A)
that is invertible -call S its inverse- such that 1J{A - Tzz and 1J{A - Szz
are A-compact.
Exercise 4.9. Prove that this is an equivalence relation. 0
Proposition 4.8. If T and .T are stably quasiisomorphic A -finite-rank C*-
modules, then [T] = [.T] in Ko(A).
Proof. Suppose that operators T and S = y-I are given satisfying the con-
ditions of Definition 4.5. Consider the following operators in EndA (T fDJ-fA ):

c := (oo S1z)
Szz ·
They satisfy FGF = f, GFG = G, and also

-S1zTzz )
1- SzzTzz '
(4.6)
150 4. Fredholm Operators on C* -modules

The relation ST = 1 irnplies that

1 _ GF = (SuTu SuT12) = (Sn S12) (1.r 0) (Tu T12)


S21 Tu S21 T12 S21 S22 o o T21 T22
= S(1.r E9 O)S- 1.

Since 1 - GF and 1.r E9 0 are projectors, taking adjoints yields 1 - GF =


(S* )- 1(1.r E9 O)S*, so 1.r E9 0 commutes with S* S. Thus, 1- GF and 1.r E9 0
are equivalent via the unitary U = S(S* s)- 112 in HornACf E9 J{A, 'E E9 J{A).
By fixing an isornorphisrn frorn 'E E9 J{A to J{A, we rnay regard Fand G
as operators on J{A· Since

1- s22T22 = (1- s22)(1- T22> + s22<1- T22) + o- s22>T22.


S12 = 1:rS12 E Horn~ (JfA, 'E) and T21 = T211:r E Horn~ ('E, J{A), we con-
dude that 1- FG and 1- GF lie in End~(J{A) = As. Thus FandGare
regular A-Fredholrn Operators in FredA(J{A).
K-theory now enters the picture through the lang exact sequence (3.15),
corresponding to the quotient rnap ry: EndA(J{A) - EndA(J{A)!As. The
connecting hornornorphisrn

given by (3.40), rnaps the dass of the invertible elernent TJ (F) to the dass
[V(1 E9 O)V- 1]- [1] in Ko(As ), where V E EndA (JfA E9 JfA) drops to ry(F) E9
T](G). Forthat purpose, we rnay take

F 1- FG)
V:= ( 1- GF G .

Then c5(ry(F)) is represented by the rnatrix

(1 !GF 1 -;G) (~ ~) (1-GFG 1 -FGF)- (~ ~)


=(FG0-1 1-0GF).

Clearly [1- FG] = [1:r E9 0] = ['E], and [1- GF] = [ 1.r E9 0] = Lfl in K 0 (A),
so that c5(ry(F)) = [}"]- ['E].
However, since T2 1 is A-cornpact, F itself differs frorn 0 E9 1.'1-{A by an
A-cornpact operator, so that ry(F) = 1 (the unit in A5), whose K1-dass is
trivial. Thus, finally, [}"]- ['E] = c5([1]) = 0 in Ko(A). D

The next thing to notice is that the index is unchanged by A-cornpact


perturbations.
Proposition 4.9. If F1, F2 E FredA ('E, }") are regular operators and if F1 -
F2 E Horn~ ('E,J"), then indexF1 = indexF2.
4.3 The generalized Fredholm index 151

Proof. Let S1 be a pseudoinverse for F1. Define A-linear operators U and R


.r
on 1' E9 by

Then R is regular and A-Fredholm while U is invertible; indeed, U2 = 1. It


is clear that indexR = indexF2 + indexS1 = indexF2 - indexF1. Therefore

This last matrix is a compact perturbation of the unit, since 1 - S1F2


1 - S1F1 + S1 (F1 - Fz ), and it is regular since R is regular.
lt remains to show that a regular operator F: 1'- 1' of the form 1:r + K
with K compact must have zero index. For that, let S denote a pseudoin-
verse for Fandlet (j := imS = imSF, a complemented submodule of 1'.
Note also that S is a compact perturbation of the unit, since F = FSF en-
tails S = 1 + K- KS- SK- KSK. Consider the operator from ker F E9 (j to
ker S E9 (j given by the matrix

1-FS)
s .
This is an invertible operator, with inverse

1-SF (1- SF)F)


(
SF SF 2 '

as is easily checked, taking into consideration the spaces where the maps
are defined and the idempotence of SF. Now, T E9 13(~ takes ker F E9 (j E9 J{A
onto kerS E9 (j E9 J{A· Since 1G- Sand 1G- SF 2 are A-compact, as soon
as we identify (j E9 J{A with J{A we get an invertible operator satisfying
the conditions of Definition 4.5, so that ker F and ker S are stably quasiiso-
morphic. Since ker S ~ ker F* as A-modules, Proposition 4.8 implies that
indexF = 0. D

~ lt is time to look at irregular Fredholm operators. The following result


serves as a preparation.
.r
Lemma 4.10. Let 1' and be C* -modules over a unital C* -algebra A, and
Iet FE FredA (1', ]"). Then there is an integer n E ~ and a regular operator
FE FredA(l' E9 An,_r E9 An) suchthat F 11 = F.

Proof. Choose an A-linear operator G: .r - 1' such that 1:r - GF =: R


and 1.r- FG are of A-finite rank. Write R = Ir= 1 lri)(sil with ri,Si E
1'. If {u1, ... ,un} denotes the standard basis for An, the notation Lr :=
152 4. Fredholm Operators on C*-modules

L~= 1 1ri)(ud: a .... L~=l riai determines amap Lr E Hom1(An, 'E) withad-
joint given by L~ := 2:~ 1 1ui)(rd. Now we can define F: 'E EB An- :J EB An,
as well as an operator S going the other way, by

~ (F ~)I
F:= Li s~ := (G0 Lr)
0 . (4.7)

Since LrLi = R = 1:r - GF, we find that FSF = F, SFS = S, and


~~ (1.r- FG -FLr ) (4.8)
1-FS = -L*G 1n- Li Lr ·
s

Since these are clearly A-compact operators, Fisregular and A-Fredholm,


with S as a pseudoinverse. D

If FE FredA('E,:J) is not regular, we may regard Fasan A+-fredholm


operator in the space HomA+('E,:J), and then amplify F to a regular ope-
rator F as in (4.7). It remains to show that indexF does not depend on the
particular choices made in its construction, and that it lies always in Ko(A);
once we have accomplished that, we may then extend the index (4.5) to all
A-Fredholm operators by setting indexF := indexF.
There is indeed considerable arbitrariness in the choice of G and thus of
the integer n and the elements ri, Si E 'Ein the construction of Lemma 4.1 0,
so the construction of a regular A-Fredholm operator is by no means canon-
ical. However, any other G' for which 1:r- G' Fand 1.r- FG' are of A-finite
rank will lead to an operator F' that differs from F by a finite-rank ope-
rator (as usual, we identify An with An EB On-m c Am if m > n). Thus
index F' = index F by Proposition 4.9.
Exercise 4.10. By suitably amplifying each A-Fredholm operator F to a reg-
ular operator F, establish the algebraic properties of the index in general:
namely, that indexF* = -indexF, index(Fl EB F2) = indexF1 + indexF2,
and that indexFU = index VF = indexF if U, V are invertible. o
Exercise 4.11. Prove that two A-Fredholm operators that differ by an A-
compact operator have the same index, without assuming regularity. 0

Exercise 4.12. Let F E FredA ('E, :J). If Gis suchthat l:r- GF and 1.r- FG
are A-compact, show that index G = - index F. 0

In the nonunital case, indexF, as defined,lies in Ko(A+). Happily, we can


improve this by showing that it actually lies in the subgroup Ko(A).
Proposition 4.11. Let 'E, :J be C* -modules over a nonunital C* -algebra A
and Iet F E FredA('E,:J). Consider a regular A+-Fredholm operator F in
HomA+ ('E EB A+n,:J EB A+n) with Fn = F. Then indexFlies in Ko(A).
4.3 The generalized Fredholm index 153

Proof. We reconsider the construction of F in Lemma 4.10, where A+ is


now used instead of A. Let E: A + - C be the augmentation; we must show
that KoE(indexF) = 0.
First of all, it follows from (4.7) that [ker F] = [1- SF] = [1nl in K 0 (A +),
which implies KoE[ker F] = n.
Next, let P := 1 - FS. This is an A-finite rank idempotent acting on :f EB
A+n, so P = I.f= 1 1xJ + Uj}(YJ + VJI with Xj,YJ E :fand Uj,Vj E A+n.
Since P 2 = P, we may assume that P(x1 + u 1 ) = x 1 + u 1 for each j; this
thenimplies thatthe matrix q E Mm(A+) given by qiJ := (Yi +Vi 1x1 +u1 ) is
idempotent. Moreover, imP and qA+m are isomorphic as right A+-modules
under the map cjJ: :f EBA+n- A+m given by cp1 (x + u) := (YJ + v 1 Ix + u).
These considerations allow us to compute

KoE[ker F*] = KoE[1- FS] = KoE[q] =rank of the matrix E(q)


m
= trE(q) = I E(Vj I Uj).
j=l

Now I.f= 1 1uJ}(VJI = P22 = 1n- L1Lr where r,s E 'En, and L1Lr is the
matrix in Mn (A) with entries (Si I YJ) in kerE. Therefore, KoE[ker F*] =
tr(E(P22 )) = tr(ln) = n. Finally, K0 E[indexF] = n- n = 0, as expected. D

~ The essential property of the ordinary Fredholm index is, of course, the
homomorphism property. We show that this holds also for our more gene-
ral index map. First, we establish the important continuity property of the
index.
Lemma 4.12. The set FredA('E,:f) is an open subset ofHomA('E,:f), and
the index map from FredA ('E, :f) to Ko (A) is locally constant.

Proof. First consider the case 'E = :f. If 17: EndA('E)- EndA('E)/End~('E)
is the quotient map, then FredA ('E) is the preimage of the invertible ele-
ments in the target C* -algebra, so it is open in EndA ( 'E). A similar result
holds when 'E * :J, since a small enough perturbation of an invertible ele-
ment in HomA ('E, :f) I Horn~ ('E, :f) is also invertible.
In view of the definition of the index, it is enough to show that the in-
dex map is locally constant in the neighbourhood of a regular operator
A-Fredholm operator F. LetS be a pseudoinverse for Fandlet F' be an-
other A-Fredholm operator with IIF- F'll < 1/IISII. Then IISF- SF'II < 1,
so that 1 - SF + SF' is invertible in EndA ('E). Since

FSF' = F(1- SF + SF'),

wefind thatindex(FSF') = indexFbyLernma4.7.ButFSF' = F' -(1-FS)F'


is a compact perturbation of F', so index(FSF') = indexF' also. D
154 4. Fredholm Operators on C* -modules

Proposition 4.13. If F E FredA (1.", J") and F' E FredA (]", ~), then F' F E
FredA (1.", ~) and index(F' F) = indexF' + indexF.

Proof. We use the absorption theorem to embed 1.", .r


and (j in J{A, so we
may consider F, F' and F' F as A-Fredholm Operators on J{A· We can then
replace F by F E9 1 in FredA (J{A E9 J{A) -the index remains the same, of
course. Then we may use the well-worn homotopy argument to show that
there is a continuous path in FredA (J{A E9 J{A) from F' F E9 1 to F' E9 F. For
instance, we can take Vt as in (3.7) and

._ (F'0 0)1 (F0 0)1


Ft .- Vt
*
Vt '

for 0 5 t 5 1. The only thing to notice is that if TJ is the quotient map


that kills End~ (J{A E9 J{A), then TJ(Ft) is invertible for every t, so this path
remains within the set of A-Fredholm operators. Now

index(F' F) = indexF0 = indexF1 = index(F' E9 F) = indexF' + indexF,

in view of the previous Lemma. D

We would like to establish that the index map from FredA(J{A) to Ko(A),
which we now know to be a semigroup homomorphism, in fact yields an
isomorphism between a suitable quotient group of these A-Fredholm ope-
rators and the group Ko(A). First, we check surjectivity.
Lemma 4.14. The mapindex: FredA(JfA)- Ko(A) is onto.

Proof. If 1." and]" are finitely generated C* A-modules, any F E FredA (1.", ]")
can be amplified to F E9 1 E FredA(1." E9 J{A,J E9 J{A) and then, by the
absorption theorem, F can be identified to an operator F' E FredA (J{A)
with the same index.
Now if Ais unital, any element of Ko(A) is of the form [1."]- []"] where
1." and.r are finitely generated and projective A-modules. Consider the
zero operator 0 E HomA(1.",J"); both 1x and 1.r are A-compact, so 0 is A-
Fredholm; and clearly, index(O) = [1."]- []"]. Therefore, the corresponding
operator in FredA (J{A) has the prescribed index.
Next, suppose that Ais not unital; then any element of Ko(A) is of the
form [p]- [q] where p,q E P(At) and KoE[p] = KoE[q]. This condition
means that E(p) and E(q) are unitarily equivalent projectors in .X; in fact,
for some n, E(p) and E(q) lie in Mn(O and E(p) = UE(q)u- 1 for some
unitary u E U(n). Regarding u as a unitary element of Mn(A+) allows us
to write UE(q)u- 1 = E(uqu- 1 ). We may replace q by uqu- 1 , so that now
E(p) = E(q); therefore, p- q lies in Mn(A).
With 1." = pAn,.r = qAn, define FE HomA(1.",J") and GE HomA(J, 1.")
by F(s) := qs, G(t) := pt. If Pi• for j = 1, ... , n are the columns of p (these
4.3 The generalized Fredholm index 155

lie in A+n), then PJUO< E An if {u()(} is an approximate unit for A. If s E 'E,


then

(1:r- GF)(s) = s- pqs = p(p- q)s


n n
=I p(p- q)pjpj s = li~ I p(p- q)pjUO<Pj S,
j=l j=l

and since p(p- q)pjuO<pj s = ip(p- q)pJ) (pjuO<is, 1:r- CF is A-compact.


Similarly, 1.r- FG is A-compact. This shows that FE FredA('E,J").
To compute index F, we amplify F to a suitable operator F from 'E E9 A +n
.r
to E9 A+n by bordering it with A+-finite rank operators, and we amplify
G to an operator S going the other way. For instance, we can take

F ·=
.
( qp ,
(1-q)p
q(l-p))
(1-q)(1-p)'
s·- ( pq
.- (1-p)q
p(l-q)
(1-p)(1-q).
)

Theseoperators satisfy SF = p E9 (1- p) = 1:r E9 (1- p), FS = 1.r E9 (1- q),


and so SFS = S, FSF = F, and 1 - SF, 1- FS are of A +-finite rank; thus Fis
a regular A+-fredholm operator, with Fn = F. Also, ker F = im(1- SF) ""
pA+n, kerS = im(1- FS) ""qA+n, and so indexF = indexF = [p]- [q], as
desired. D

Lemma 4.15. IfF inFredACE,J") withindexF = 0, then there isan integern


such that F E9 1A" is an A-compact perturbation of an invertible element.

Proof. First, suppose that A is unital. Then we can construct F and S as in


(4.7), using A-finite rank operators Lr: An - 'E and Li: 'E- An to border
Fand its partner G. By (4.7), 1- SF = 0 E9 1n, so that [im(1- SF)] =[An]
in K 0 (A). Since indexF = indexF = 0, we conclude that [im(l - FS)] =
[An], so that the A-modules im( 1 - FS) and An are stably equivalent. This
means that there is some m ~ n such that im(l - FS) "" Am as C* A-
modules. By taking Sn+ I = · · · =Sm = 0 if necessary, we can regard Li as
taking values in Am; thus we mayasweil suppose that m = n.
An explicit isomorphismfromim(l-FS) to Anis :Lr= 1 1ui) (ti +vil = LtEB
Lv, where ti E J", Vi E An and {tl +VI, ... , tn + Vn} is a set of n generators
for im(l- FS). The operator

u := ([; Lt)
Lv

.r
is an isomorphism from 'E E9 An to E9 An, and it is an A-compact pertur-
bation of F E9 1A".
When A is nonunital, the same argument works with A replaced by A +.
The only difference is that, in orderthat U- (F EB1A") be A-compact, Lv -1A"
must lie in Mn(A) or, more precisely, in the kerne! of E: Mn(A+)- Mn(<C).
156 4. Fredholm Operators on C* -modules

Now, it does no harm to suppose that the generators {ti + Vi} are orthonor-
mal with respect to the Hermitian structure of J" $ A +n (after applying the
Gram-Schmidt algorithm); so we can assume that (ti I tJ) + (Vi I VJ) = DiJ.
This implies that E(Vi I Vj) = DiJ• so the matrix v with columns E(VJ) is
unitary. Replacing each ti +vi by v*ti +v*vi allows to assume that v = 1n,
which in turn implies that Lv - 1A" E Mn (A). D

We finally have a criterion for two A-Fredholm operators to have the same
index.
Corollary 4.16. If FJ E FredA(TJ,JJ) for j = 1, 2, then indexF1 = indexF2
if and only ifthere is some n E N suchthat F1 $ F2* $ 1A" is an A-compact
perturbation of an invertible element. a

4.4 The noncommutative Atiyah-Jänich theorem


We could have stopped already, by declaring two A-Fredholm operators
equivalent if they satisfy the criterion of Corollary 4.16; then the index map
identifies the totality of such equivalence classes with Ko(A), and the rela-
tion between A-Fredholm operators and K-theory classes is fully specified.
However, this is not quite enough to recover the Atiyah-Jänich theorem,
which speaks of homotopy classes of operators. For that, we must invoke
an important generalization of Kuiper's theorem, proved by Troitskii [459]
in the unital case and by Mingo [342] in the a-unital case; the latter proof
was somewhat simplified by Cuntz and Higson [125]. Weshall not prove it
here; a good exposition can be found in [481, Chap. 16]. The precise state-
ment is as follows.
Theorem 4.17 (Kuiper-Mingo). If Ais a a-unital C*-algebra, the unitary
group o{EndA (J{A) is contractible. a
Actually, all the proofs show that the group of invertible elements in
EndA (J{A) is a contractible space. Using the polar decomposition T = U ITl
and the homotopy s - T5 := UITI 5 , this may be retracted to the unitary
subgroup. By using the theorem by Milnor mentioned in Section 3.8, that
this open set has the homotopy type of a CW -complex, this theorem is
reduced to the next proposition.
Proposition 4.18. If Ais a a-unital C* -algebra and if Xis a compact space,
the unitary group of C (X) ® EndA (J{A) is connected. a
Indeed, Mingo [342] proved a more general theorem: if B is any unital
C*-algebra and Ais a-unital, then the unitary group of B ® EndA (J{A) is
connected.
We come finally to the main result of this section. To lighten the notation,
we abbreviate FredA := FredA(J{A); and note that Fredc = Fred.
4.4 The noncornmutative Atiyah- Jänich theorem 15 7

Theorem 4.19 (Noncommutative Atiyah-Jänich theorem). If A is any


O"-unital C* -algebra, the index map induces a group isomorphism from
rro(FredA) onto Ko(A).
Proof. ByLemma 4.12, index: FredA - K0 (A) is locally constant. This gives
a well-defined semigroup homomorphism [F] ..... indexF from rro(FredA)
into Ko(A); its surjectivity is due to Lemma 4.14. If F E FredA and if G E
FredA is suchthat 1- GF and 1- FG are A-compact, then [F] [ G] := [FG] =
[1] since t ..... (1- t)l + tFG provides a continuous Fredholm path from 1
to FG; thus rro(FredA) is actually a group. (A tapalogist would say that
FredA is an H -space.)
lt remains to show that this homomorphism is injective, i.e., that F lies in
the neutral component of FredA whenever indexF = 0. From Lemma 4.15
with 'E = :J = J{A, we can suppose that F = U + K where U is invertible
and K is A-compact. Since the segment t ..... U + tK, 0::; t ::; 1, consists of
A-Fredholm operators, it is enough to show that any invertible U is path-
connected to 1 in EndA (J{A). Butthis follows at once from Theorem 4.17,
since we have assumed that A is O"-unital. 0

Exercise 4.13. Prove that this isomorphism index: rr0 (FredA) - Ko(A) is
natural, that is, if cJ>: A - B is a morphism, there is a group homomorphism
cf>* : rr0 (FredA) - rr0 (FredB) such that index o cJ> * = Koc/> o index. 0
When A = C(X) for a compact space X, Theorem 4.3 does not follow
immediately. But the situation is easily sorted out. We recall the notation
AX := C(X-A) if Ais a C*-algebra; likewise, we may abbreviate FredX :=
C(X- Fred).
Exercise 4.14. Given a compact space X, show that FredX is a subsemi-
group of Fredc(X)· 0
The semigroup Fredc(X) of Fredholm elements in the (huge) multiplier
algebra Endc<XJ(Jfc<xJ) is much bigger than FredX = Fredc<Xl nL(Jf)X.
However, their respective path components can be matched.
Theorem 4.20. If X is a compact space, there is an isomorphism of groups

[X, Fred] "" rro (Fredc(X)).

Proof. First of all, a homotopy dass of maps from X to Fred is just a


continuous path in FredX, so [X,Fred] = rro(FredX). Now the inclusion
j: FredX .... Fredc(X) induces a group homomorphism rr0 j: [X,Fred] -
rro (Fredc(X)).
The main point is that if F E Fred X is connected to the identity in
Fredc(X), then it can be connected to the identity in FredX. Assurne that
F can indeed be connected to 1 in Fredc<xJ; then Lemma 4.12 shows that
indexF = 0 in Ko(C(X)) = K0 (X), and by Lemma 4.15 there exists K E
End~(X) (Jf ® C(X)) "" X ® C(X) = XX (recall Exercise 3.1) such that
158 4. Fredholm Operators on C*-modules

V:= F + K E L(J.f)X is invertible. Clearly the segment t .... F + tK connects


F to V within FredX. On the other hand, Proposition 4.18 (with A = ([)
shows that the group of invertible elements of C(X) ® L(J{) = L(J.f)X
is connected; so there is a path from V to 1 in FredX. In conclusion, the
homomorphism rr0 j is injective.
Ta see that it is surjective, in the light of Theorem 4.19, we need only
show that for any dass [ p]- [q] in Ko ( C (X)), there is some F E Fred X with
indexF = [p]-[q].NowifindexF1 = [p] andindexF2 = [q], thenF = F 1 FJ
will da, so we can suppose q = 0. Using Theorem 3.8, we can also suppose
that p is a projector in Mn(C(X)) for some n. Regard p as a projector-
valued map from X to Mn(<(). If v E Xis an unilateral shift operator,
defined an an orthonormal basis {ukh2:o for J{ by v(uk) := Uk+l• then
v*v = 1 and dimkerv* = 1, because v*(uo) = 0 and v*(uk) := Uk-I for
k ~ 1. We can then define an OperatorF an C(X)n ® J{ by

F(x) := (1n- p(x)) ® v* + p(x) ® lJ-f.

F(x)~ = I (p(x)~k +On- p(x))~k+I) ® ek.


k2:0

The equation F(x)~ = '7 is solved by ~o = p(X)'1o +On- p(x))( for any
( E cn, and ~k = p(X)'1k + (1n- p(x))'1k-l for k ~ 1. Thus, F(x) is
surjective, while kerF(x) ""im(1n- p(x)) = kerp(x); thus FE FredX
and indexF = [ker F] = [p ], as required. D

Putting the last two theorems together, we obtain that [X,Fred] ""K 0 (X)
via the index map, which indeed is Theorem 4.3! Thus Theorem 4.19 mer-
its the name we have given it. The more general result that [X, FredA] ""
K 0 (AX), for X compact and A unital, is proved in [342].
All is well, the reader may conclude, concerning the formal beauty of
the theory of Fredholm operators an C*-modules; but, where are the con-
crete examples? As a matter of fact, there are plenty of them. Vector C*-
bundles over a space M are defined as ordinary vector bundles but with
fibres that arefinite-rank C*-modules over a given C*-algebra A. Elliptic
pseudodifferential operators and their symbols are defined just as in the
classical case (consult Chapter 7); they are seentobe A-Fredholm and their
indices in Ko(A) can be computed by a generalized Atiyah-Singer theorem,
the Mishchenko-Fomenko index theorem [346] -see also the older refer-
ence [345], and [460], where the methods of [243) are generalized for appli-
cation to this index theorem. This is obviously related, in the commutative
case, to the family index theorem [33]. Another index theorem for families
of Fredholm operatorswas given some time ago by Dan Freed [188]; in that
case, the algebra is of the form C""(G), where G can be any Banach-Lie
group.
4.5 Morita equivalence of C*-algebras 159

4.5 Morita equivalence of C*-algebras


The previous work helps with a different circle of ideas, extremely impor-
tant in noncommutative geometry. Let us generalize first the construction
of the tensor product Ecf>('E) at the end of Section 2.5, when a C* A-module
'E and a morphism </>: A - B were given. We shall now start from two
C*-algebras A and B, a right pre-C* A-module 'E, a right pre-C* B-module
:J, and a representation of A by operators on :J, i.e., a given morphism
p: A- EndB(:f). We can form the algebraic tensor product 'E 0 :f, which
is naturally again a right B-module, tagether with a positive sesquilinear
pairing ( · I ·) satisfying
(s1 ® t1 I sz ® tz) := (t1 I p((s1 I sz)) tz) = (p((sz I si)) t1 I tz). (4.9)
This clearly includes the construction of Ecf> ('E), when :F = B, as a parti-
cular case. As before, one can pass to the quotient by the B-submodule of
elements z E 'E 0 :F suchthat (z I z) = 0. We denote by 'E ®p :J, or more
simply 'E ® A :F, the quotient pre-C* B-module. If 'E and :F are C* -modules,
weshall also use the notation 'E ®A :J, usually without further comment,
to denote the C* ~-module obtained by completing that quotient.
Exercise 4.15. Prove, along the lines of the discussion in Section 2.5, that
{ z E 'E 0 :J: (z I z) = 0} is spanned by the elements sa ® t- s ® p(a)t. 0
Such tensor products naturally arise when considering C* -bimodules,
or pre-C* bimodules, under commuting left and right actions of two C*-
algebras. Of course, a representation p: A - EndE ('E) is a left action of A
on 'E by any other name, so we have entered that context already. However,
a symmetrical notation for bimodules is very convenient, so we now bring
in some notation for left (pre)-C*-modules; these are obvious analogues of
the right (pre)-C*-modules treated so far.
Definition 4.6. A left pre-C* -module over a C* -algebra B is a complex vec-
tor space :F that is also a left B-module equipped with a sesquilinear pairing
:F x :F - B, the obvious analogue of (2.5):
{r I s + t} = {r I s} + {r I t},
{brls}=b{rls},
{rls}={slr}*,
{sls}>O for S*-0,
for r,s, t E :F and b E B. Notice that these requirements demand that
this pairing be linear (indeed, B-linear) in the first variable and antilinear
in the second variable, in cantrast to our Standard convention of linearity
in the second variable. (This trick of switching conventions, which greatly
simplifies subsequent formulas, is due to Daniel Kastler [387].) A left pre-
C* B-module :F is full if {:f I :f} is dense in B. A left C* -module over B is
obtained by completing :F in the norm s .... II {s I s} 11 1 /Z.
160 4. Fredholm Operators on C* -modules

If 'Eis any right A-module, its conjugate space Xis a left A-module: using
the obvious notation X = { s : s E 'E } , we can define
as := (sa*)-.

For instance, if 'E = pAm, we get X== mAp. If 'Eisa right pre-C*-module
over A, there is an obvious pairing that makes X a left pre-C* A-module,
namely, {f I .5} := (r I s).
In this case, X is isomorphic to the subspace of compact elements of the
dual 'E~ := HomA ('E, A) of 'E. This is a left A-module, with the operation
a~: s .... a(~(s)), for ~ E 'E~, a E A, s E 'E -which is morenatural than
the right module operation (2.13). Recall that there is an injective module
map X- P given by f .... (r I·); its imagelies in Hom~('E,A). Indeed, if
f = as, then (sa* I·)= a(s I·)= la)(sl, so AT maps onto Hom~0 ('E,A).
Now
lllaslllt := lllsa*lllx = sup ll(r I sa*)ll = sup ll(r I s)a*ll
lllrlllsl lllrlllsl
= sup lla(s I r)ll = llla)(slll.
lllrlll sl

andin the sameway, the T-normofa finite sum LJ a 1 s1 equals the operator
norm of L.1 1a1 )(s1 1. Since the closure of 'EA is all of 'E, the closure of AT
is X, so that X maps isometrically onto Horn~ ('E, A). In particular, we have
established a sort of "Riesz theorem" for C* -modules, namely that every
element of Horn~ ('E, A) is of the form s .... (r I s) for some r E 'E.
Definition 4.7. A pre-C* B-A-bimodule is a complex vector space 'E that is
both a left pre-C* B-module and a right pre-C* A-module and, moreover,
satisfies
r(s I t) = {r I s}t for all r,s, t E 'E. (4.10)
We say that 'Eis right-full if ('E I 'E) is dense in A, or left-full if {'E I 'E} is
dense in B. We call it simply full if both conditions hold.
Lemma 4.21. The two norms naturally defined on a pre-C* B-A-bimodule
coincide.

Proof. We must show that if s E 'E, then II {s I s} IIB = II (s I s) IlA· This fol-
lows from the algebraic properties of both pairings, and from the Schwarz
inequality of Lemma 2.14, by the following calculation [389]:
ll(s I s)lli = ll(s I s)(s I s)ll~ = ll(s I s(s I s))ll~
= ll(s I {s I s}s)ll~ :5 ll(s I s)IIA ll({s I s}s I {s I s}s)IIA
:5 II (s I s) II~ llt .... {s I s}tll 2 = II (s I s) II~ II {s I s} II~.
Thus ll(s I s)IIA :5 ll{s I s}IIB; the opposite inequality is obtained by inter-
changing A and B and the two pairings as well. D
4.5 Morita equivalence of C*-algebras 161

'E is called a C* B-A-bimodule if it is complete in the norm lllslll 2 :=


ll(s I s)ll = ll{s I s}ll .
.,. Although C* B-A-bimodules have a formidably rich structure, there is a
plentiful supply of them. The simplest examples are afforded by the C*-
algebras themselves, with the pairings (a I b) := a*b and {a I b} :=ab*.
Then indeed c(a I b) = ca*b = {c I a}b. Note that they are full: the ideal
A 2 is densein A, since A has an approximate unit.
Now, take any full right C*-module 'E over A, and consider the C*-
algebra B := End~ ('E); it acts on 'E on the left. Moreover, we claim that
'E becomes a C* B-A-bimodule, the B-valued pairing being given simply by

{r I s} := lr}(sl.
To see this, we recall that Tlr}(sl = ITr}(sl forT E EndA('E), andin par-
ticular forT E B. Since {'EI 'E} = End~0 ('E) is by definitiondensein B, 'Eis
left-full. The compatibility (4.10) of the pairings comes from the definition
of the ketbra operators: r(s I t) = lr}(slt.
In general, even if A is commutative, the algebra End~ ('E) will not be.
Thus we glimpse the main role that C*-modules play in noncommutative
geometry, as mediating structures to allow the emergence of new algebras
related, but not isomorphic, to A. That role is hidden in commutative geo-
metry.
.,. The relation between A and End~ ('E) has the following key feature of
reciprocity.
Proposition 4.22. Let 'E be a full right C* A-module. lf B =End~ ('E), then
End~(T) ""A.

Proof. The algebra End~0 ('E) has a right action on 'E, namely

f lr}(sl := s(r I t).


The map i ..... s(r I t) is bounded by lllrllllllslll = II Ir} (slll from the Schwarz
inequality; by continuity, this map extends to a right action of B on 'E. The
pairing (f I ü) := lt}(ul makes 'E a right C* B-module, since if b = lr}(sl,
then
(f I ü)b = lt}(ullr}(sl = lt}(s(r I u)l = (f I s(r I u)) = (f I üb).
Clearly, 'Eis left-full since (TI T) = End~0 ('E) is densein B.
Define a morphism a: A - End~(T) by a(a) : i ..... ta*. Each a(a) is
indeed B-linear since
a(a)(i lr}(sl) = s(r I t)a* = s(r I ta*} = ta* lr}(sl.

On the other hand,


a((r I s))(f) = t(s Ir)= f ls}(tl = f (s I f) = lf)(.SI f,
162 4. Fredholm Operators on C*-modules

so a((r I s)) = lf)(.SI E End~0 (E). Since 'Eis right-full, we conclude that
a maps A into, indeed onto, End~(E). Now a is injective, since a(a) = 0
implies (s I t)a* = (s I ta*) = 0 for s, t E 'E, so ca* = 0 for all c E A, thus
a = 0. Hence a is a C*-algebra isomorphism. D
Exercise 4.16. Show in detail, exhibiting all pairings, that the correspond-
ence r®s- Ir) (sl extends to a C* B-module isomorphism 'E®AT"" B, and
that f ® s - (r Is) extends to a C* A-module isomorphism T ®s 'E ""A. 0
Definition 4.8. If :J is a C* A-B bimodule and (j is a C* B-C bimodule, the
tensor product :J ®s (j becomes a C* A-C-bimodule, with the pairings on
simple tensors given by (4.9):

(r1 ® s1 I rz ® Sz) := ((rz I rdss1 I sz)c


{r1 ® s1 I rz ® Sz} := {r1 I rz {sz I s1} B} A. (4.11)

The compatibility (4.10) of the left and right pairings is an easy calculation
that should by now be routine.
Exercise 4.17. Check the compatibility of the pairings (4.11). 0

Exercise 4.18. If the C* A-B-bimodule :J and the C* B-C-bimodule (j are


full, show that :J ®s (j is a full A-C-bimodule. 0
Exercise 4.19. Check the associativity of C*-bimodule tensor products. o
.,. With these C* -bimodule techniques in hand, we can now introduce a
most important equivalence relation between C* -algebras, weaker than iso-
morphism but stronger than equality of K-theory groups.
Definition 4.9. We say that two C*-algebras A, B are (strongly) Morita-
equivalent, and we write A ~ B, when there is a C* B-A-bimodule 'E and a
C* A-B-bimodule :J suchthat

'E ®A :J "" B and (4.12)

as B- and A-bimodules, respectively. We refer to 'E and :J as equivalence


bimodules.
For instance, by taking 'E =Am and :J = mA, we see that any full matrix
algebra Mm (A) is Morita-equivalent to A.
To say that :J ®s 'E "" A means in particular that the vector space gene-
rated by the pairings ((rz I YI)sSI I sz)A, with r1. rz E :J and s1.s2 E 'E, is
matched via an A-bimodule isomorphism to the dense subspace A 2 of A
generated by products ataz. In particular, A2 !;;; ('EI 'E), so 'Eis right-full.
The same argument, applied to 'E ® A :J "" B and the second set of pairings
in (4.11), shows that 'Eis left-full. In synthesis, the C*-bimodules 'E and :J
that implement a strong Morita equivalence are automatically full.
Proposition 4.23. Morita equivalence is an equivalence relation.
4.5 Morita equivalence of C*-algebras 163

Proof. Reflexivity is trivial (if B = A, take 'E = :F = A, too) and symme-


try is obvious. If we are given a C*-bimodules 'E, 'E', :F and :F' (over the
appropriate algebras), suchthat (4.12) holds and also 'E' ®B :f' == C and
:f' ®c 'E' == B, then the tensor products 'E' ®B 'E, a C* C-A-bimodule, and
:F ®B :f', a C* A-C-bimodule, satisfy
'E' ®B 'E ®A :F ®B :J' :::: 'E' ®B B ®B :J' :::: 'E' ®B :J' :::: C,
and similarly :F ®B :F' ®c 'E' ®B 'E == A, since bimodule tensor products are
associative. 0
In many cases, one can construct pre-C*-bimodules 'E and :F satisfying
the analogue of (4.12) where A and B are replaced by pre-C*-algebras .Jl
and 'B. With the obvious changes, (4.12) then defines the relation of Morita
equivalence between pre-C* -algebras .
.,. Morita equivalence is important for the following reason: if A ~ B, there
is a natural way to match right C* A-modules with right C* B-modules,
or left C* A-modules with left C* B-modules, or C* A-bimodules with C*
B-bimodules. To be more precise, suppose that equivalence bimodules 'E
and :J, satisfying (4.12), are given. Then G' ..... G' ®A :F takes any right
A-module to a right B-module, the inverse map is G ..... G ®B 'E, and this
correspondence is clearly functorial. Likewise, 'E ® A ( · ) takes left A -modules
to left B-modules and is inverted by :F ®B ( · ); finally, 'E ®A ( ·) ®A :F takes
A-bimodules to B-bimodules and is inverted by :F ® 8 ( ·) ®B 'E.
Let us look more closely at the transformation of endomorphisms under
the first of these correspondences. If G is a right B-module and 'E is aB-
A-bimodule, then we can form the right A-module G ® 8 'E. Any adjointable
operator TE EndB(G) can be amplified to an operator T®1:r: r®s ..... Tr®s
in EndA(G ® 8 'E). lt turnsout [171] that if T is B-compact, then T ® 1:r is
A-compact, provided only that 'Eis left-full, by the following argument.
Lemma 4.24. If G is a C* B-module and 'E is a left-full C* B-A-bimodule,
then T ..... T ® 1:r maps End~(G) into End~ (G ®B 'E).
Proof. First take T E End~0 {{i) of the form T = Iub) (v I where u, v E G
and b = { r I s} E B with r, s E 'E. Since 'E is left-full, these cases span a
dense subspace of End~(<;j). If w E G, t E 'E, then
T ® 1:r(w ® t) = ub(v I w) ® t = u ® b(v I w)t
=u ® {r I s}(v I w)t =u ® {r I (w I v)s}t
=u®r((wlv)s I t)=u®r(v®slw®t),
so that lu{r I s}) (v I® 1:r = lu®r)(v ®sl. lt is clear that T ..... T® 1:r is norm-
decreasing on these examples, so it extends continuously to a morphism
between their closed linear spans, i.e., from End~(<;j) to End~ (<;j ® 8 'E). 0
This lemma has an obvious but important consequence.
164 4. Fredholm Operators on C* -modules

Corollary 4.25. If FE HomB(G, G') is a B-Fredholm operator, and if'E is a


left-full C* B-A-bimodule, then F ® 1::t is an A-Fredholm operator.
Proof. lt is clear that 1g ® 1::t is the identity operator on G ®B 'E. If G E
HomB(G', y) is suchthat 1g- GF and 1g· - FG are B-compact, then 1 -
( G ® 1:E )(F ® 1::t) and 1 - (F ® 1:E) ( G ® 1:E) are A-compact, by the previous
lemma. o

~ It is useful to have a characterization of Morita equivalence of C* -alge-


bras that involves only one bimodule, not two. The next theorem gives
such a criterion; it amounts to an alternative definition of (strong) Morita
equivalence, which is in fact the original version introduced by Rieffel [38 7].
This is also the approach taken by Lance [303] and Skandalis [440], and may
be preferred once its symmetry is established.
Theorem 4.26. Two C* -algebras A and B are Morita-equivalent if and only
ifthere is a full right A-module 'E such thatEnd1('E) ""B.

Proof. If B "" End1 ('E), then 'f, with the structure described in Proposi-
tion 4.22, is actually a C* B-A module. Then, by Exercise 4.16, A ~ B.
Conversely, suppose A ~ B via appropriate C*-bimodules 'E and ;:. Since
B = End~(B) "" End~('E ®A :J), it follows from Lemma 4.24 that the mor-
phism S - S ® 1:r takes End~ ('E) into B and is inverted by T - T ® 1'E, which
takes B = End~(B) into End~ (B ® 8 'E) =End~ ('E). Notice that Lemma 4.24
can be applied since both 'E and;: are left-full. 0

Exercise 4.20. Prove that any C* B-A bimodule 'E implementing a Morita
equivalence between unital C*-algebras A and B is finitely generared and
projective as an A-module. 0

Proposition 4.2 7. If p is a projector in a C* -algebra A such that the ideal


ApA is densein A, Iet B denote the C* -algebra pAp; then A ~ B.

Proof. Take 'E := pA, which is clearly a pre-C* B-A-bimodule under the
compatible pairings (pa I pc) := a*pc E A and {pa I pc} := pac*p E
B. Since lllpalll = llpaiiA and 'Eisa closed right ideal in A, it is in fact
a C*-bimodule. Right-fullness of 'E follows from the assumed density of
('E 1 'E) = ApA in A. (Left-fullness is clear since {'EI 'E} = pA 2 p is dense
in pAp.) Since lpa}(pcl is just the operator pe- pac*pe, it is clear that
End~0 ('E) ""pA 2 p, and by completing we get End1 ('E) ""pAp = B.
The Morita equivalence A ~ B follows from Theorem 4.26. In this case
;: = Ap, since pA ®A Ap "" pAp, while Ap ®pAp pA is the completion of
ApA, which is A. 0

An algebra like pAp is sometimes called a corner of the algebra A. If A


is, say, a matrix algebra and p is a projector of the form h E9 On-k. then
pAp consists of matrices whose only nonzero entries occur in the top left
4.5 Morita equivalence of C*-algebras 165

corner. Any (two-sided) ideal of A that includes pAp also includes ApAp
and pApA, so if ApA is densein A, then any closed ideal of A that includes
pAp also includes Ap and pA and their product ApA, and so is ail of A.
Thus the only closed ideal including the subalgebra pAp is A itself: pAp
is then cailed a full corner of A. It can be shown [57] that A ~ B if and
only if there is a C* -algebra C in which A and B may be embedded as
complementary fuil corners.
Proposition 4.28. Any C* -algebra A is Morita-equivalent to its stabilization
As = X®A.

Proof. If Aisunital and p is a rank-one projector in X, then P1 := p ® 1A E


1'(As) is a projector in As suchthat AsP1As = XpX ® A 2 =X® A2 = A~,
whichis dense inAs, while P1AsP1 = pXp ®A = Cp ®A === A. Thus As ~ A
by Proposition 4.27, via the C*-bimodules 'E = P1As === pX ® A === J{ ® A =
J{A and :J = AsP1 === Xp ® A === J{ ® A = J{A· In the nonunital case, one
can directly choose 'E and :J to be J{A with the algebras As and A acting
on the right and left in the obvious manner, with A acting "diagonaily". 0

Corollary 4.29. lf A, B are stably equivalent C* -algebras, then A ~ B. E3

This coroilary strongly suggests that Morita-equivalent C*-algebras have


the same K-theory. Indeed, if its converse were true, that is, if A ~ B were
to imply As === Bs, then Ko(A) === Ko(B) would follow, since Ko(As) = Ko(A).
Now, it happens that Morita-equivalent and <T-unital C*-algebras are in fact
stably equivalent as weil [57]. lt turns out, however, that there are pairs of
C* -algebras, not both <T-unital, which are Morita-equivalent but not stably
isomorphic; an example is also given in [57].
The Morita invariance of the K0 -functor can, however, be shown directly,
by building on the theory of Fredholm operators for C* -modules. The proof
is due to Exel [171].
Theorem 4.30 (Exel). Jf A ~ B, then Ko(A) === Ko(B).

Proof. Let 'E be a left-fuil C* B-A-bimodule. By Coroilary 4.25, we know


that the correspondence F ..... F ® 1:r takes a B-Fredholm operator F E
HomB(§,§') to an A-Fredholm operator in HomA(§ ®B 'E,§' ®B 'E). This
suggests that we can define a map p('E) from Ko(B) to Ko(A) by

p('E): indexF ..... index(F ® 1:r ).

To see that this is weil defined, suppose there are two operators F1, F2 in
FredB(§, §') with the same index. By Coroilary 4.16, we can find n E ~. U
and K suchthat F1 $Fi $1A" = U +Kin FredB(§ $ §' $ Bn, §' $ § $ Bn),
where U is invertible and K is B-compact. But then
166 4. Fredholm Operators on C*-modules

Now U ®1:r is invertible and K ®bis A-compact; therefore, index(F1 ®1:r) =


index(F2 ® 1:r ).
The map p('E) is clearly a group homomorphism, on account of Proposi-
tion 4.13. Also, if :Fis a left-full C* C-B-bimodule, then 1.r ® 1:r is the iden-
tity operator on the C-A-bimodule :F ®B 'E, so that p('E) o p(:f) = p(:f ®B'E).
Finally, if A ~ B, we can find left-full equivalence bimodules 'E and :F sat-
isfying (4.12). The corresponding homomorphisms p('E): Ko(B) - Ko(A)
and p (:f): Ko (A) - Ko (B) invert each other, so they are isomorphisms. D

~ Morita equivalence makes sense for arbitrary unital rings [348]: just re-
move the prefix "C*" from the objects of Definition 4.9. At the purely alge-
braic level, it establishes an equivalence between the categories of (right)
B-modules and (right) A-modules, given by the functors (j - (j ®B 'E and
(j' - (j' ®A :F; see the discussion prior to Lemma 4.24. Similar correspon-
dences hold for left modules and bimodules. For unital C* -algebras, the
relations of "strong" and "ordinary" Morita equivalence coincide [27, 391];
so we choose to omit the adjective "strong". An intermediate notion of
"formal" Morita equivalence of *-algebras, where the equivalence bimod-
ules carry positive semidefinite pairings, has been developed by Bursztyn
and Waldmann [59], who show that two C*-algebras are strongly Morita
equivalent iff their minimal dense ideals are formally Morita equivalent.
For nonunital C*-algebras, there is also a way to implement (strong)
Morita equivalence by functors, but it is necessary to replace the category of
C* -modules by that of "operator modules". A (left) operator module over a
C*-algebra Ais a closed subspace X of L(Jf) suchthat rr(A)X ~X, where
rr is a nondegenerate representation of A on Jf. Any C* -module is an ope-
rator module, and the (inner) tensor product of C*-modules corresponds to
the Haagerup tensor product of operator modules. The morphisms in this
category are the completely bounded module maps </>: X - Y, i.e., those
for which the norms of the corresponding maps <J>(n): Mn (X) - Mn ( Y) are
uniformly bounded. Blecher [37] has shown that A ~ B if and only if there
are functorial equivalences between the categories of left operator modules
over A and B respectively.
Much of the importance of Morita equivalence stems from the corres-
pondence between the representation theories of A and B that is imple-
mented by an equivalence bimodule. Let rr be a nondegenerate *-represen-
tation of A (that is: the only vector ~ E J-frr for which rr (A) ~ = 0 is the zero
vector), and let 'E be a right C* A-module. Weshallshow how rr induces an-
other representation rr of the C*-algebra EndA ('E). Before doing so, we ask
the reader to check the following improved version of the inequality (3.2).
Exercise 4.21. Prove that if s E 'E and TE EndA ('E), then

0::;; (Ts I Ts)::;; IIT*TII(s I s) (4.13)

in the C* -algebra A. 0
4.5 Morita equivalence of C*-algebras 167

On the algebraic tensor product 'E 0 Jfrr there is a seillidefinite scalar


product, given on elementary tensors by
(r ® 111 s ® ~) := (rylrr((r I s))~).
We build a representation of EndA ('E) with the GNS construction of Sec-
tion LA. Namely, if N = {<P E 'E 0 Jfrr : (<P I <P) = 0}, then ('E 0 Jfrr) IN
is a prehilbert space; its completion is a Hilbert space that we denote by
'E®AJfrr. If TE EndA('E), then the linear map s®~ .... Ts®~ preserves Non
account of (4.13), so it passes to the quotient as a linear map fr(T). Again
by (4.13), we obtain llfr(T) II :s; IITII. so thatir is a representation ofEndA ('E)
on 'E ®A Jfrr. By restriction, we also get a representation of End~ ('E). One
may come back with 'E.
In condusion, if A ~ B via the equivalence bimodule 'E, then rr .... fr
is a bijective correspondence between nondegenerate representations of
the C*-algebras A and B. Moreover, fr is irreducible if and only if rr is ir-
reducible. Even more, it turns out that if we denote by A the dual space
of A, i.e., the set of equivalence dasses of its involutive irreducible repre-
sentations, this procedure implements a homeomorphism between A and
B(with the Jacobson topology).
In particular, if a C*-algebra is Morita-equivalent to a commutative C*-
algebra, then its dual space is locally compact and Hausdorff. A dass of
C*-algebras Morita-equivalent to commutative C*-algebras are some con-
tinuous trace C*-algebras [137, 151]. If Ais a C*-algebra in this dass, then
(as anticipated in general terms in Section 2.5), it is isomorphic to the sec-
tion algebra of a continuous field of elementary C*-algebras over A, where
"elementary" means that each fibre is isomorphic to X or to Mn((.) for
some n. In particular, two commutative C*-algebras are Morita-equivalent
if and only if they have homeomorphic spectra, if and only if they are iso-
morphic.
In many other instances, Morita invariance mediates between the com-
mutative and the noncommutative worlds. If G is a locally compact group,
acting on a locally compact space M, and the space of orbits MI Gis a "good"
space (that will happen if the action is proper and free), then C0 (M IG) and
the crossed product C*-algebra C0 (M) ~ G (defined precisely in Chapter 12)
are Morita-equivalent, and they store essentially the same information. The
door cannot open on all noncommutative algebras, by far: for instance, a
factor can be Morita-equivalent to a commutative algebra if and only if it
is of type I. Usually MI G is a "bad" space, and its quotient topology is too
coarse for C0 (M IG) to be of any use. Nevertheless, Co(M) ~ G will still
carry the geometrical information. Similar statements can be made for fo-
liations [91]. This method for dealing with quotient spaces by means of
functional analysis is one of the grand themes of Alain Connes' thinking.
Regrettably, we cannot go into all these applications of Morita invariance.
At least, in this book, it will be decisively used to obtain a noncommutative
understanding of spin structures on manifolds.
Part II

CALCULUS AND LINEAR


ALGEBRA

Alles sollte so einfach


wie möglich gemacht sein,
aber nicht einfacher

- Albert Einstein
5
Finite-dimensional Clifford Algebras
and Spinars

From the noncommutative geometric Standpoint of Part III, commutative


geometry and Clifford geometry are one and the same thing. So here we deal
with their linear-algebraic and Ue-theoretic underpinnings, namely Clifford
algebra. We chosenot to dispense with it in this book, despite the existence
of many excellent treatments, mainly for ease of reference. In particular,
the infinitesimal spin representation is needed to deal with the spin con-
nection, which in turn serves the construction of "unbounded K-cycles" in
Chapter 9. For the present, we limit ourselves to the case of finite dimen-
sions, so as to illuminate more clearly the algebraic nature of many of our
tools. However, we introduce that representation in a way that affords an
easy transition to the infinite-dimensional case, to be developed in Chap-
ter 6. In cantrast with the previous (and succeeding) chapters, here there
are no "big theorems" and we trust that the readerwill find the going easier.
Familiarity with basic properties of the classical orthogonal and unitary
groups and their representations, as presented for instance in [54] or [439],
is assumed. Also, the reader should be conversant with the terminology of
superalgebras, which is reviewed in Section S.A, serving as an appendix.

5.1 The eightfold way


A Clifford algebra is an associative algebra generated by a vector subspace
equipped with a symmetric bilinear form g, in such a way that the square of
any vector is a scalar. For the zerobilinear form, the corresponding Clifford
172 5. Finite-dimensional Clifford Algebras and Spinors

algebra is just the exterior algebra on the given vector space; otherwise, it
has the same underlying vector space as the exterior algebra, but with a
modified product operation.
In the infinite-dimensional case, taken up in the next chapter, the Clif-
ford algebra will be completed to a C* -algebra, and continuity arguments
must then be handled carefully. When the generating vector space is finite-
dimensional, we need only concern ourselves with the (fairly intricate) al-
gebra that arises from the defining anticommutation relations.
We begin, then, with a finite-dimensional real vector space V. Form its
exterior algebra NV, whose elements arefinite sums of wedge products
u1 " · · · "uk with u1, ... , Uk E V and k :5 dim!Rt V, subject only to the
anticommutation relation ui "UJ + UJ "ui = 0. Foreach v in V, there is
an obvious linear map E ( v) on N V given by exterior multiplication:

(5.1)

Clearly E(v) 2 = 0.
A Clifford algebra over V is manufactured by "quantizing" NV by means
of g so as to get nontrivial commutation relations. For that, we add to E ( v)
a contraction L( v) given by
k
L(V)(Ul 1\ • • • 1\ Uk) := 2: (-1)J-lg(V, Uj) U1 1\ • • • 1\ UJ 1\ • · · 1\ Uk.
j=l
(5.2)

(The circumflex on u J means, as usual, that this term is omitted.) To fore-


stall trivial cases, we shall always assume that the symmetric bilinear form
g is nondegenerate. Now L(v) 2 = 0 and E(V)L(v) + L(V)E(v) is the operator
of multiplication by g(v, v); indeed,

[E(Uo)L(Uo) + L(Uo)E(Uo)](ul 1\ • • • 1\ Uk)


k
'
= L ~ 1\ · · • 1\ Uk
(-1) 1.- 1g(uo, Uj) Uo 1\ • • • 1\ Uj
j=l
k
+ 2:<-1)Jg(uo,UJ)uo" · · · "ii] 1\ • • • "uk.
j=O

It follows that the sum of these operators c(v) := E(v) + L(v) satisfies
c(v ) 2 = g(v, v) for v E V. More generally, E(v )L(u) + L(U)E(v) = g(u, v ),
so that
c(v)c(u) + c(u)c(v) = 2g(u, v) for u, v E V.
This anticommutation relation is often written {c (u), c (v)} = 2g ( u, v); in
conformity with physics tradition, the curly brackets are used to denote an
anticommutator. Ordinary brackets [ ·, ·] denote a commutator; but since
5.1 The eightfold way 173

most algebras considered in this chapter are actually superalgebras, i.e.,


they are iEz-graded, the ordinary bracket will occasionally also denote a
supercommutator; we shall usually say so explicitly. For convenience, we
shall call g a "quadratic form" even though this term usually refers to the
function v .... g(v, v) rather than its polarized version g(u, v).
Definition 5.1. Let V be a real vector space on which is given a nonde-
generate quadratic form g. The (real) Clifford algebra Cl( V, g) is the Sub-
algebra of End~~t (NV) generated by the Clifford multiplication operators
{ c (v) : v E V }. This algebra has the same dimension as N V itself, namely
2n if dim~~t V= n. To see that, choose a basis {ei, ... ,en} for (V,g) such
that g (ek, e1) = 0 for k * l and g (ek, ek) = ± 1 for each k; then the operators
c(ek) anticommute, so Cl(V,g) is linearly generated by the ordered prod-
ucts c(ek 1 )c(ek 2 ) ••• c(ekr), for K = {ki < · · · < kr} s;; {1, ... , n}, including
the operator of multiplication by 1 for the case K = 0.
We shall more often be concerned with complex Clifford algebras. One
can complexify the exterior algebra to get NV ®~~t I( "" NVc, where vc
denotes the complexification V ®~~t I( = V Eil iV of V, real-linear operators on
V being amplified to complex-linear ones on vc (with no change of names)
by the rule A(u + iv) := Au+ iAv, and similarly for forms. Regard the
u 1 in (5.1) and (5.2), then, as elements of vc, so each E(v), t(v) and c(v),
with the same defining formulae, on using the amplified quadratic form g,
are !(-linear operators on NVc. The complex Clifford algebra G(V) is the
subalgebra of End(A"Vc) generated by all the operators c(v). (The cute
notation Gis borrowed from the book by Lawson and Michelsohn [314],
where much information on Clifford algebras may be found.)
To lighten the notational burden, we shall describe the basic properties
mainly within the complex case, and ask the reader to check the analogous
statements for real Clifford algebras.
The Clifford algebra G(V) can be thought of as the exterior algebra NVc
with a new product. In this way, the scalars I( and the vectors in V lie in
l(l(V), onidentifying v E V with c(v)l. We write simply VI v2 ... Vk for the
element corresponding to the operator c (VI )c ( v2) ... c ( Vk). Concretely, the
vector space isomorphism O": l(l(V) - NVc (the symbol map) is given by
O"(a) := c(a) l. The product in G(V) satisfies the fundamental relation

uv + vu = 2g(u,v) for u,v E V, (5.3)

equivalent to v 2 = g(v, v).


Theinverse isomorphism Q: NVc - l(l(V) is a quantization map for
the exterior algebra [33]. In particular, since O"(uv) = u 1\ v + g(u, v), we
get Q(u 1\ v) = uv- g(u, v) = t<uv- vu). The general formula is

Q(vi 1\ • • · 1\ Vk) = ~1 L (-1)T VT(l) ... VT(kl· (5.4)


. TESk
174 5. Finite-dimensional Clifford Algebras and Spinors

where the sum ranges over the set Sk of permutations of k indices.


Exercise 5.1. lf u, v, w are three vectors in V, show that

uvw = Q(u A v A w) + g(v, w)u- g(u, w)v + g(u, v)w.

Then check that a- indeed inverts Q. 0

Proposition 5.1. The algebra G(V) satisfies the following universal prop-
erty: if .Jl is a unital complex algebra and f: V - .Jl is a real-linear map
satisfying j(v) 2 = g(v, v) 1.Jt for all v E V, then there is a unique algebra
homomorphism j: G(V) - .Jl suchthat f = ilv.
Proof. The uniqueness is clear, since j(vi ... Vk) := j(vi) .. . j(vk) must
hold in all cases, and then j can be extended by linearity to all of G(V).
The question is whether it is thereby well defined. For that, it is enough
that j(uv + vu- 2 g(u, v)) = 0 for all u, v E V; but this is clear, since

j(uv + vu) = j((u + v) 2 - u 2 - v 2)


:= j(u + v) 2 - j(u) 2 - j(v) 2 = 2g(u, v) 1.Jt. D

The analogaus statement for real Clifford algebras is proved similarly.


Example 5.1. Suppose that g,g' are two nondegenerate quadratic forms
on V that are congruent, i.e., there is an invertible real-linear map h: V - V
suchthat g(u,v) = g'(hu,hv) for all u,v; then h extends to a real alge-
bra homomorphism from Cl(V,g) onto Cl(V,g') that is actually an iso-
morphism, since h- 1 extends to its inverse. Thus Cl(V,g) depends, up to
isomorphism, only on the signature of g. In the complex case, all nonde-
generate quadratic forms are congruent; this is why we omit mention of g
in the notation G(V).
Example 5.2. If h: V - V is a (pseudo)orthogonal linear map, that is, if
g(hu, hv) = g(u, v) for all u, v, then by taking .Jl = Cl(V,g) we can ex-
tend h to an automorphism (}h of Cl(V,g): this is the Bogoliubov auto-
morphism determined by h. The uniqueness implies that (}hk = (}h o (}k
whenever h, k are orthogonal. If we take .Jl = Cl( V) instead, we likewise
obtain a Bogoliubov automorphism of Cl(V), also denoted by lh.
Example 5.3. In particular, if h = -1, then x := (}_ 1 is an automorphism
whose square is the identity, so it gives a ::Z:2·grading of the Clifford alge-
bra into an even subalgebra and an odd subspace: Cl(V,g) = Cl+(V,g) EB
Cl-(V,g), or else G(V) = n+(V) EB n-(V), as the case may be, where
Cl±(V,g) [respectively, G±(V)] denotes the (±1)-eigenspace of x. It is
clear that V c Cl- (V, g) c n- (V); more generally, since x(v 1 v 2 ... Vk) =
( -1 )k v 1 v2 ... Vk, the even subalgebra consists of linear combinations of
products of an even number of elements of V.
lt is clear that a-, Q are even maps and that Gk(V) := Q(EBJ= 0 AJVC)
defines a filtration of G(V).
5.1 The eightfold way 175

Exercise 5.2. Prove that [G(V),G(V)] = on-I(V), where the bracket de-
notes the supercommutator. 0

Example 5.4. Next, let J2l be the opposite algebra <Cl(V) and j(v) := vo;
0
,

thenj: <Cl( V) - <Cl(V)o can be regarded as an antiautomorphism! of G(V)


given by (VI ... Vk)! := Vk ... VI.
Example 5.5. Even the standard complex conjugation on <Cl( V) fits into this
context: we let J2l be the conjugate algebra of <Cl(V) and let f be the iden-
tity map on V; the linear homomorphism j can then be regarded as an
antilinear conjugation a ..... ä of <Cl(V). Composing this with the product
reversal gives an involution a ..... a* := (ä) 1 of the Clifford algebra; notice,
in particular, that (VI ... Vk)* := Vk •.. VI.
Example 5.6. If we combine complex conjugation on <Cl( V) with the grad-
ing x. we getan important antilinear isomorphism K(a) := X(ä), called the
charge conjugation.

.,.. We examine the real Clifford algebras more closely now. Since the iso-
morphism dass of the algebra depends only on the signature of the qua-
dratic form, we need only consider the standard quadratic form of signa-
ture (p, q) on ~p+q:

(x, y) ~ XIYI + · · · + XpYp - Xp+IYp+I - • · · - Xp+qYp+q· (5.5)

The corresponding Clifford algebra is denoted Clp,q· (Same sources, such


as [314] and the seminal paper [15], where the classification is worked out
in full, use the opposite sign convention to (5.3), by taking uv + vu =
-2 g (u, v), so that our Clp,q would be their Clq,p .)

Exercise 5.3. Work out the low-dimensional cases: Cli,o "" ~ES~. Clo,I "" <C,
Clz,o ""M2(~), Ch,o ""M2(<C), Cli,I ""M2(~) and Clo,2 "" IHI, constructing
explicit isomorphisms; e.g., ~ES~ 3 (a, b) ..... ~(a + b) + ~(a- b)ei E
Ch,o. 0

Exercise 5.4. Determine the even subalgebras in low-dimensional cases:


Clt_ 0 "" ~. Cl2,o "" <C, Cl3, 0 "" IHI and Cll.o "" IHI ES IHI, constructing explicit
isomorphisms. 0

Lemma 5.2. Foreach pair of integers p, q E N, there are isomorphisms:

Clp+I,q+I ""Clp,q ®tRM2(~). (5.6a)


Clp+4,q ""Clp,q ®RM2(1Hl) ""Clp,q+4. (5.6b)

Proof. We denote by {ei, ... ,ep,EI, ... ,Eq} an orthogonal basis for ~p+q
with respect to the quadratic form (5.5), where each g(er, er) = + 1 and
each g (Es, Es) = -1; these vectors anticommute in Clp,q. For the first iso-
176 5. Finite-dimensional Clifford Algebras and Spinors

morphism, define a linear map f: ~p+q+ 2 - Clp,q ®111.M2 (~) by

j(er):=er®(~ ~1 ) j(ep+r) := 1 ® (~ ~)'


j(Es):=Es®(~ ~1 ), j(Eq+r) := 1 ® (~ -1)
0 .

These elements anticommute and their squares are ± 1 with the expected
signs, so f extends by Proposition 5.1 to an algebra homomorphism. A
dimension count reveals that this homomorphism is bijective.
Both algebras appearing in (5.6b) have dimension 2P+q+ 4 . Thus we need
only define real-linear maps g,h: ~p+q+ 4 - Clp,q ®111.M2(1Hl), which take a
suitable orthogonal basis into anticommuting elements whose squares are
± 1 with the correct signs. Let i, j, k be the standard unit quaternions of
square -1 suchthat ij = k = - ji. Then we can try

-k)
0 ' g(Es) = h(Es) :=Es® (~ -k)
0 '

for r = 1, ... , p and s = 1, ... , q. On the remaining four generators, we


define respectively

g(ep+r) := 1 ® (~ ~i)' h(Eq+r) := 1 ® (~


~)'
g(ep+2) := 1 ® (~ -j)
0 ' h(Eq+2) := 1 ® (~ _ok),
g(ep+3) := 1 ® G~)' h(Eq+3) := 1 ® (~ ~)'
g(ep+4) := 1 ® (~ ~1)' h(Eq+4) := 1 ® (~ ~).
Once more, Proposition 5.1 yields the desired isomorphisms. D

Exercise 5.5. Refine (5.6b) by finding isomorphisms by which Clp+ 2,q ""
Clq,p ®111.M2(~) and Clp,q+2 ""Clq,p ®IJ!.IHJ. <>

Note that IHl ®IJ!.IHJ ""M4 (~) ""End111. (<C 2). Indeed, by Exercises 5.3 and 5.5,
Cl2,2 "" Clo,2 ®111.1Hl = IHJ®IJ!.IHI; while, fromLemma 5.2, Cl2,2 "" Cl1,1 ®111.M2 (~) =
M2(~) ®111. M2(~) = M4(~). lt is then also true that M2(1Hl) ®111. M2(1Hl) ""
M 1 5(~). Finally, we take note of the following consequence of Lemma 5.2.

Corollary 5.3. Foreach pair of integers p, q E ~. there are isomorphisms:


E3

Now any real algebra .J\. is Morita equivalent, in the purely algebraic sense,
to Mn(.Jl.) = .J\. ®111. Mn(~) -via the bimodules .J\.n and n.J\.- so that the
5.1 The eightfold way 177

Morita equivalence classes of the real Clifford algebras can be read off from
the 8 x 8 table of the lowest-dimensional cases (provided we have the cour-
tesy to put Clo,o := IRI). This "spinorial chessboard", which provides the title
of a manual on the subject [58), also has Morita invariance on lines parallel
to the main diagonal, by (5.6a), which is sometimes referred to as (1, 1)-
periodicity. Therefore, provided we can keep track of the matrix sizes, we
can read off the whole classification of these algebras [118) from the cases
Cln,o. n = 0, 1, ... , 7. The full 8 x 8 table may be found in [314, p. 29).
However, a more attractive picture, which incorporates both periodicities,
is the so-called spinorial clock on the last page of [58). We exhibit it here
in a slightly modified form:

Theorem 5.4. All finite-dimensional real Clifford algebras are classified by


the spinorial clock, by following these operating instructions. To determine
Clp,q. first compute m := (p - q) mod 8, and locate the arrow A ~ B on
the clock. Next, compute the integer N for which dimiR MN (B) = zp+q. Then
Clp,q = MN(B). Moreover, the even subalgebra is given by either Cl;,q =
MN(A) or c1;,q = MN;2(A), depending on its dimension.

Proof. Suppose p ~ q, so that p = Bk + q + m with k ~ 0. Then (5.6a) and


Corollary 5.3 show that

and similarly, Cl;,q ::::: Cl~.o ®!RM2;k+q(IR1.). If p < q, say p + Bl = q + m


with l ~ 1, then Clp,q ::::: Clm,SI ®!RM2q-sl(IR1.) ::::: Clm,o ®!RM2HI(IRI) provided
p ~ m, and similarly for c1;,q; when p < m, one can say instead that
Clp,q ®!RM16 (IR() ::::: Clm,o ®iRM2H<l-o (IRI). In all cases, then, Clp,q ::::: Mr (Clm,o)
and Cl;,q "" Mr(Cl~. 0 ), where r is determined by the requirement that
dimClp,q = zp+q. lt therefore suffices to verify the workings of the clock
for the case q = 0, p = 1, ... ,8.
Hereis the table of the Clp,o and their even subalgebras, for p = 1, ... , 8:
178 5. Finite-dimensional Clifford Algebras and Spinors

p Clp,O Cl;,o
1 IREBIR IR
2 Mz(IR) c
3 Mz(C) IHI
4 Mz(IHI) IHIEBIHI
5 Mz(IHI EB IHI) Mz(IHI)
6 M4(1Hl) M4(C)
7 Ms(C) Ms(IR)
8 M16(IR) Ms(IR EB IR)
The top half of the table summarizes Exercises 5.3 and 5.4. The bottarn
half comes from (5.6b) on tensoring by Mz(IHI), using the isomorphisms
IHI ®IR C == Mz(C), given by Exercise 2.4, and IHI ®IR IHI == M4(IR). 0

Exercise 5.6. From the table, it appears that Clp+l,q ""Cl;,q ®M2 (1R) in ge-
neral. Prove this, and prove also that c1;+l,q "" Clq,p· 0

In conclusion, the algebras Clp,q are simple, except when p -q = 1, 5 mod


8, in which case they possess two nonisomorphic irreducible real represen-
tations. The cases p-q = 0, 2 mod 8, forwhich Clp,q isareal matrixalgebra,
are said to be of the Majorana type.
~ We come back now to the complex case. Here the periodicity is much
simpler. First of all, we may and shall assume, from now on, that g is pos-
itive definite. This amounts to replacing the basis vectors Es in ~RP+q by
ep+s := iE5 , so that {ei, ... ,ep+q} becomes an orthonormal basis for CP+q.
One can complexify the tables for Cln,o to get each Cl (IR n), bearing in mind
that IHI ®IR C ""M2 (C). But it is even simpler to compute Cl(IRn) directly.
Lemma 5.5. If N = 2m, then

Proof. lt is clear that Cl(IR) "" C EB C, and also Cl(IR 2 ) "" M2 (C) -either
by direct calculation, or using that Mz ( C) is the only 4-dimensional com-
plex associative algebra with a I-dimensional centre. The general case is
obtained from the periodicity isomorphism

that is determined, by Proposition 5.1, from the real-linear map f: IRn+Z -


Cl( ~Rn) ®c Mz(C) given, for r = 1, ... , n, by

f(er) :=er® 1(0 1)0 , f(en+d := 1® i


(0 -i)
0
(1
, f(en+Z) := 1® 0

since all the j(e1 ) anticommute and satisfy j(e1 )2 = 1. 0


5.1 The eightfold way 179

Proposition 5.6. There is a unique trace T on Cl (V) such that T ( 1) = 1 and


T ( a) = 0 whenever a is odd.

Proof. Choose an orthonormal basis {e1, ... ,en} of (V,g), and let eK
ek 1 ek 2 . . . ekr for K = {k1 < · · · < kr} !; {1, ... , n}, and e0 := 1. The
several eK form a basis for Cl(V). Suppose that T is a trace satisfying the
stated conditions, and let K have an odd number of indices; then T (eK) = 0.
Moreover, if l rt K, then since the vectors ek anticommute in Cl(V), we find
that ezeK = -eKel and so T(ezeK) = 0; but of course eKul = ±ezeK, so that
T(eKud = 0. Thus a = LK aK eK implies T(a) = a0; this establishes the
uniqueness of T. (The scalar component a 0 is the same for all bases.) lt
also establishes existence, since a ...... a 0 is clearly even and normalized,
and it is tracial since (ab)0 = LK(-l)r(r-ll1 2 aK bK = (ba)0. D

The Clifford algebra Cl(V) becomes a Hilbert algebraunder the positive


definite scalar product:

(a I b) := T(a* b) = L. äK bK, (5.7)


K

the expression on the right being independent of the chosen basis.


Exercise 5.7. Show that any real Clifford algebra Cl(V,g) also carries a
unique normalized even trace, butthat the scalar product defined as in (5.7)
is positive definite if and only if g is itself positive definite. 0

.,. Now suppose that an orientation on the real Hilbert space (V,g) is given.
From now on, let {e1, ... , en} be an orthonormal basis that is compatible
with the given orientation.
Definition 5.2. Write n = Zm or n = Zm + 1 according as n is even or odd.
The chirality element of Cl( V) is

(5.8)

If {ei, ... , e~} isanother oriented orthonormal basis, then ej = Lk=l h 1kek
where h is an orthogonal matrix of determinant + 1, so ( -i)m ei e; ... e~ =
det(h)y = y; therefore, y is independent of the chosen basis.
Since y* = (-l)m(-l)n(n-1li 2 y = y (because n(n-1)/2 = m(Zm + 1)),
and y* y = en ... e1 e1 ... en = 1, the chirality element satisfies y 2 = 1. Also,
y either anticommutes or commutes with each eJ and therefore with each
v E V, according as n is even or odd. Therefore, if n is even, yvy = -v
for v E V, so by Proposition 5.1 y( · )y equals x. the grading operator,
which is then an inner automorphism; whereas if n is odd, yv y = v for
v E V, and therefore y ( ·) y is the identity. It should be clear in general that,
when n is odd, inner automorphisms give rise only to special orthogonal
transformations of V.
180 5. Finite-dimensional Clifford Algebras and Spinars

Suppose that n is even. Then the centre of Cl( V) is just the scalars C. This
follows from Lemma 5.5, but let us show it directly. Indeed, if a is a central
element, then x(a) = ;ya;y = a;y 2 = a, so a is even. If a = IKeven aK eK,
then a = e1ae1 = IKeven aK e1eKel and e1eKe1 = +eK according as l E Kor
not. Therefore, aK = 0 unless K is empty; this means that a = T(a) 1.
On the other hand, if n is odd, then the centre of Cl( V) contains 1 and ;y.
In fact, the centreis exactly two-dimensional, by Lemma 5.5.

5.2 Spin groups


The group of invertible elements of a complex Clifford algebra contains
several interesting subgroups that are of crucial importance in this book.
Definition 5.3. Let v be a unit vector in the real Hilbert space (V, g), thatis,
g(v, v) = 1. Then v*v = v 2 = 1 and therefore (va I vb) = T(a*v*vb) =
T(a*b) = (a I b) for all a,b E Cl(V), so that left multiplication by v is a
unitary operator on Cl(V). More generally, if u E Cl(V) satisfies u*u = 1,
then a - ua is unitary; conversely, if a - ua is a unitary operator, then
T(a*(1- u*u)a) = 0 for all a, which implies u*u = 1. We therefore
call u E Cl(V) a unitary element of the Clifford algebra if u*u = 1 (or
equivalently, if uu* = 1).
If a E Cl(V)x is an invertible element, we can conjugate other elements
by a; this operation Ad(a): b - aba- 1 is not particularly useful per se.
An interesting variant uses the grading automorphism x to form "twisted
conjugates" cp(a): b - x(a)ba- 1 • For instance, if v E Visa unit vector
and if x E V is any vector, then
cp(v)x = (-v)xv- 1 = (-v)xv = (xv- 2g(v,x))v = x- 2g(v,x) v,
so that cp(v) preserves the subspace V of Cl( V) and its action on V is just
the reflection in the hyperplane orthogonal to v.
Any unitary vector in vc is of the form w = .\ v where v is a unit vector
in V and l.\1 = 1; then cp(w) = cp(v) also. These reflections generate the
orthogonal group O(V,g) of (V,g); when no ambiguity is likely, weshall
just write O(V). The orthogonal group has two connected components;
the products of an even number of reflections form the rotation subgroup
SO (V), which is the connected component of the identity.
Definition 5.4. If w 1 , w 2 E vc are unitary vectors, then w 1 w 2 is a unitary
element of o+ (V). The group generated by allsuch even unitaries (i.e., all
products w1 wz ... w 2 k of an even number of unitary vectors) is denoted
Spinc(V), the so-called spinc group of V. Foreach u E Spinc(V), the map
cp ( u): x - uxu - 1 (no x is needed since u is even) is a rotation of V, so we
may regard cp as a homomorphism from Spine (V) to SO (V). Any u E ker cp
commutes with all x E V, so it is an even element in the centre of Cl( V) and
5.2 Spin groups 181

thus is a scalar. Therefore, ker cf> "" lr, the circle group of unitary scalars.
This gives a short exact sequence of groups

1 - l r - Spine(V) J:. SO(V) -1, (5.9)

which says that Spine(V) is a central extension of SO (V) by lr.


Note that if u E Spine(V), then a - uau- 1 , for a E G(V), is just the
Bogoliubov automorphism eq,<u>; indeed, it is clearly an automorphism of
((](V) that preserves V, so it is the unique automorphism that extends the
orthogonal transformation cf>(u) of V.
~ There isanother important homomorphism v: Spine(V) - lr given by
v(u) := u!u. Concretely, if u = w 1 w 2 ... Wzk. where the Wj are unitary
vectors, then

v(u) = Wzk .•• wzwrwz ... Wzk = .\1 Wzk .•. w~ ... Wzk = · · · = .\1.\2 ... .\2k,
where Aj := w] E lr. Notice that v restricts to the squaring map .\ - .\ 2 on
the scalar subgroup lr of Spine (V). Thus (c/>, v): Spine (V) - SO(V) x lr has
kernel {±1}.
Exercise 5.8. Prove by a direct argument that the sequence (5.9) is not split.
0

Definition 5.5. The spin group Spin(V) is defined as the kernel of v in


Spine(V). By the previous remark, this yields a short exact sequence of
groups:
1 - {±1}- Spin(V) J:. SO(V)- 1,
which says that Spin(V) isadouble cover of SO(V). Now u E Spin(V) if
and only if u*u = 1 and also u!u = 1, if and only if u is even, unitary and
"real", i.e., u = ü. 1t is important to realize that Spin(V) is contained in
Cl( V, g). Since complex conjugation equals charge conjugation on n+ (V),
we conclude that Spin(V) is the charge conjugation invariant subgroup of
Spine(V).
We summarize the properties of these groups in the following diagram,
taken from [190]:

1~ :\-,V /1
lr~ ~lr
V
Spine(V) <P
~<P~
Spin(V) SO(V)

1
/ ~
1.
182 5. Finite-dimensional Clifford Algebrasand Spinors

For dim V > 1, we exclude the possibility that Spin(V) have two con-
nected components, since exp(te1e2) = cos t + e 1e2 sint, for e 1, e 2 ortho-
gonal unit vectors, is a continuous arc in Spin(V) connecting 1 and -1.
Moreover, for dim V> 2, Spin(V) is the universal covering group of SO(V),
as rr1 (SO (V)) = lL2 -which can be derived from the well-known fact that
rri(S0(3)) = 7L 2 and the long exact sequence in homotopy.
lt is interesting to note that there are parallel constructions for covering
groups of the symplectic groups. If (V, ß) is a symplectic vector space, that
is, a real finite-dimensional vector space with a nondegenerate alternating
bilinear form ß, its symmetry group is the symplectic group Sp(V, ß). There
is a central extension of Sp(V, ß) by "U", analogous to (5.9), denoted Mpc (V),
that unfortunately does not have finite-dimensional irreducible representa-
tions. There is also a distinguished homomorphism v: Mpc (V) - "U" that re-
stricts to the squaring map on the scalar subgroup "U", and its kernel Mp(V),
called the "metaplectic group", provides a double covering of Sp(V, ß). This
approach to the metaplectic representation of Sp(V, ß) is worked out in full
in the monograph by Robinson and Rawnsley [400], which we recommend
for studying the bosonic analogue to the theory of spin groups .
.,. The Lie algebras of Spin(V) and SO (V) are isomorphic; we describe them
more explicitly. The Lie algebra of the rotation group SO (V) consists of the
skewsymmetric operators

so(V) := { A E Endnt V: g(y,Ax) =-g(Ay, x) }.


On the other hand, recall that the space A2 V of bivectors is included in
<Cl+(V) via the quantization map Q: u 1\ v .... i<uv- vu).
Lemma 5.7. Ifada denotes the operator x .... (a,x], then b .... adb is a Lie
algebra isomorphism from Q(A 2 V) ontoso(V).
Proof. First notice that commutators of (quantized) bivectors and vectors
are vectors: on using the famous relation

[AB,C] = A{B,C}- {A,C}B, (5.10)

we obtain

[ ~(uv- vu),x] = [uv,x] = 2g(v, x) u- 2g(u, x) v. (5.11)


A similar calculation shows that a commutator of two bivectors is a linear
combination of bivectors, so that Q(A 2 V) is indeed a Lie algebra. lf b E
Q(A 2V) and adb(x) = [b,x] = 0 for all x E V, then bis central and even
(so it is a scalar) and moreover T(b) = 0; therefore, b = 0. This means
that b ..... ad b is an injective Lie algebra homomorphism from Q (A 2 V) into
Endnt V. In the case b = i (uv- vu), ad bis skewsymmetric, since

g(y, [b,x]) = 2g(v,x)g(y,u)- 2g(u,x)g(y,v) = -g([b,y],x),


5.2 Spin groups 183

so the image of this homomorphism lies in so(V). lt is also bijective since


dimso(V) = !n(n- 1) = dimA 2V. 0
We need a formula for the inverse isomorphism. For A E so(V), the
unique element A E Q(A2V) suchthat adA =Ais given by

(5.12)

where {e 1, ... , en} is any orthorrormal basis of (V, g). By (5.11), [ekez, e5 ] =
2 8zsek - 2 Dksez. Now, if b derrotes the right hand side of (5.12), then
1 n 1 n
[b,es] =4 2:: g(ekrAez) [ekez,es] =2 2::
g(ekrAez) (8zsek- Dksez)
k,l=l k,l=l
1 n n
= 2 2::
g(er, Aes) er- g(es, Aer) er = 2::
g(er, Aes) er = Ae 5 ,
r=l r=l
by skewsymmetry of A.
Proposition 5.8. The Lie algebra ofthe spin group Spin(V) is Q(A 2V).
Proof. Since Q (A 2V) is a subspace of Cl(V), it generates a Lie subgroup
of Cl(V)x by exponentiation. This Lie group consists of the logarithmic
elements exp b := 1 + Lhl bk I k! and their products (actually, the logarith-
mic elements alone are enough, since the resulting Lie group is compact).
lf u = exp b with b E Q(A 2V) and if x E V, then

uxu- 1 = 2:: - 1- bkx(-b) 1 =


1 1 2::
I1 r 2:: (r)
bkx(-b)r-k
k.l~O k.l. nO r. k=O k

= 2:: ~[b,[b, ... ,[b,x] ... ]] = 2:: ~(adb)r(x) E V, (5.13)


r~o r. r~O r.
and v(u) = exp(-b) expb = 1 since b' = -b for b E Q(A 2V); thus
u E Spin(V). The equation (5.13) can thus be abbreviated as cp(exp b) =
exp (ad b). Since any rotation in S 0 (V) is of the form exp A for some A E
so( V), as is seen, for instance, by expressing its matrix in canonical form,
Lemma 5.7 shows that { cp(exp b): b E Q(A 2V)} = SO(V).
In fact, one sees that { exp b : b E Q (A 2V) } = Spin(V), since the logarith-
mic elements doubly cover the rotation group: if b = Afor a skewsymmetric
operator A, take for e 1, e 2 the leading vectors of an orthorrormal basis for
whichg(ek,Aez) = 0 unless l = k ± l. Thus, by (5.12),

b = !g(e1,Ae2) e1e2 + · · · + !g(e2r-1.Ae2r) ezr-le2r with 2r :s; n.


lt is now clear that b commutes with rre1e2, so - exp b = exp(rre1e2 + b),
as claimed. 0
184 5. Finite-dimensional Clifford Algebras and Spinors

5.3 Fock-space representations


The representation theory of finite-dimensional complex Clifford algebras
l(l(V) is fairly straightforward. It follows already from Lemma 5.5 that if
dim V is even, then all irreducible representations of l(l(V) are equivalent,
whereas if dim V is odd, there are exactly two inequivalent irreducible rep-
resentations of the same dimension.
Let us concentrate for a while on the even-dimensional case dim V = n =
2m. Fixa positive definite quadratic form g on V. All irreducible represen-
tations of l(l(V) are equivalent and are 2m-dimensional; this uniqueness is
evident from Lemma 5.5, since G(V) isasimple algebra.
By definition, the exterior algebra NVc is a left module for <Cl( V), with
the action c (v) = E ( v) + L( v). However, this module is not irreducible since
its dimension is 22m. In fact, if F is any finite-dimensional representation
space for <Cl( V), then F = S 1 $ · · ·$Sr. where l(l(V) acts irreducibly on each
Sk; in particular, dimF = 2mr. Another way to express this is to say that
F =Fa® S, where S is an irreducible module and <Cl( V) acts trivially on Fa,
that is, c(a)(w ® s) = w ® c(a)s. (Notice that in the case F = Nvc, the
vector space Fa is also zm·dimensional.) Now zm is exactly the dimension
of the exterior algebra over an rn-dimensional complex vector space. Since
dim~~t V = 2m, an obvious stratagern to get an irreducible module is to take
this vector space to be V itself, after imposing a suitable complexification.
However, it is a most important circumstance that there is no canonical
way to do so.
Definition 5.6. The 2m-dimensional real vector space V can be made into
a complex Hilbert space by choosing an orthogonal complex structure, that
is, areal-linear operator] E End~~t V satisfying

]2 = -1 and g(]u,]v) = g(u, v) for u, v E V. (5.14)

We regard V as a complex vector space via the rule iv := ]v. It carries a


positive definite hermitian scalar product,

(u I v) 1 := g(u, v) + ig(]u, v). (5.15)

Note that (]u lv) 1 = -i(u lv) 1 , as it should be. The rn-dimensional complex
Hilbert space thereby obtained will be denoted by V1.
The defining relations (5.14) can be rephrased as ] 2 = -1, P] = 1, where
the transpose is taken with respect to g. Thus] is also skewsymmetric:
]t = -].In other words, the operator] liesbothin the orthogonal group
and in its Lie algebra. If { u 1 , ••. , Um} is an orthonormal basis for V1, then
{u 1 ,]u 1 , ... , Um,fum} is an orthonormal basis for (V,g). We say that]
is compatible with a given orientation on V if the latter is an oriented or-
thonormal basis.
5.3 Fock-space representations 185

Example 5.7. Consider V = ~ 4 with the standard Euclidean quadratic form.


The general orthogonal skewsymmetric matrix is of the form

( cos
0 C(
- cos C(
0
- sin C<COS ß - sln" sln
- sin cx sinß sin C<COS ß
ß)
1 = sin cx cos ß sin cx sinß or
0 - cos C( '

sin cx sinß - sincxcosß cos C( 0


- cos C( - sin cxcos ß - slncxslnß)
( cos0 C( 0 sincx sinß - sin cxcos ß
1 = sin cx cos ß - sin cx sinß 0 cos C( •

sin cx sinß sincxcos ß - cos C( 0

One of these is compatible with the usual orientation on ~4 while the other
is compatible with the opposite orientation.
Exercise 5.9. Which is which? 0

It is clear that, if 1 is an orthogonal complex structure and h E 0 (V, g),


then h1h- 1 is also an orthogonal complex structure on V. This action is
transitive: let 1. K be two orthogonal complex structures. There is a unitary
map h between V1 and VK, which are complex Hilbert spaces of the same
dimension; from
(hu I hvh = (u I v) 1 ,
we obtain that hisorthogonal and that K = h1h- 1 • If dimlRt V= 2m, the
manifold of orthogonal complex structures is a submanifold of the Grass-
mannian Gm(Vc), of complex dimension !m(m -1), that can be identified
with 0(2m)/U(m). In generalweshall denote by U1 (V) the isotropy sub-
group of 1. i.e., the unitary group of v1 .
Exercise 5.10. Prove that the manifold of all 1 compatible with a given
orientation on ~ 4 is homeomorphic to § 2 • 0

We can construct a Hilbert space equivalence of v1 with a complex sub-


space W1 of vc, namely, the subspace w1 := { v - i]v E vc : v E V}.
Notice first that w1 is isotropic for the amplified bilinear form g on vc,
since g(u - i1u, v - i1v) = 0 for u, v E V. The conjugate subspace
W 1 := { v + i]v E vc : v E V} is the orthogonal complement of w1
under the scalar product on vc given by

((w I z)) := 2g(w,z). (5.16)

!
The projector in End vc with range w1 , namely P1 := (1 - i]), is then a
unitary isomorphism between v1 and w1 . The projector with range W 1 is
just P _1 = 1- P1 = !( 1 + i]). The subspaces w1 and W1 are the eigenspaces
for the respective eigenvalues + i and - i of 1 on vc. In general, an isotropic
complex subspace W :s: vc, satisfying W n W = {0} and WEB W = vc, is
called a complex polarization for vc; notice that W n V = W n iV = {0}.
186 5. Finite-dimensional Clifford Algebras and Spinors

There is in fact a one-to-one correspondence between polarizations and


orthogonal complex structures. If w E W, then w = u - iv for unique
elements u, v E V, and Jw: u .... v is real-linear; thus W = {u- iJwu : u E
V}. Now 2tg(w 1 ,w 2 ) = 0 for w 1 ,w2 E W shows that Jw is orthogonal,
and 5 g ( w 1 , wz) = 0 gives 1a,
= -I. The orthogonal group acts on the
set of polarizations simply by W .... hW. The maps W .... Jw and] .... W1
are mutually inverse, and it is moreover clear that these correspondences
intertwine the adjoint action Jw .... h]wh- 1 of O(V) with the action W ....
hW; therefore the latter action is also transitive.

Exercise 5.11. Explain how orthogonal transformations in O(V,g) can be


identified with unitary operators on vc -for the scalar product (5.16)-
that commute with the operator of complex conjugation. 0

Definition 5.7. Fix an orthogonal complex structure J on (V,g). The Fock


space ,]'1 (V) is defined as the (complex) exterior algebra A*V1 over VJ. This
is a complex Hilbert space, whose scalar product is determined by

(5.17a)

We choose a unit vector 0 (unique up to a phase factor) in the one-dimen-


sional subspace A 0 V1 == C, and call it the vacuum vector of .J'J(V).
The maps P1 : v1 - w1 extend to unitary isomorphisms from A*V1 onto
N w1 , where the latter carries the scalar product

((ZI/\ · · · 1\ Zk I W1/\ · · · "WL)) := 8k1 det[((zi I wj))]. (5.17b)

Weshall refer to A*W1 with this scalar product as the polarized Fock space
associated to ]. (We identify A 0 V1 with A 0 W1 , so that 0 is also a vacuum
vector for the polarized Fock space.) The elements of both spaces are called
spinors, in the present context.

Warning. If we consider, naturally enough, the scalar product on Vjk given


by
(ul ® · • · ® Uk I v1 ® · · · ® Vk) := (ui I v1) 1 · · · (Uk I Vk)J,
then we must define

in order to have (5.17a). In other words, the 1/ JnT normalization factor is


required so that, if u 1 , •.. , Uk are orthonormal, then u 1 1\ • • • 1\ Uk be of
norm one. Similarly for (5.17b).
The Fock space .J'J(V) is more Standard in quantum field theory, and so
will be generally preferred in this book. The polarized presentation also has
an illustrious pedigree; we shall use it soon, in the proofs of some Iemmata.
5.3 Fock-space representations 187

Definition 5.8. Each w E W1 deterrnines two operators on NW1 : exterior


product E ( w) : 0< ...... w 1\ 0< and contraction L( w) : w1 1\ · • • 1\ Wk ......
L:J=1 (-1)1-l ((w lw1 )) w 1 1\ · · • 1\ Wj 1\ · • · 1\ Wk. Notice that this contraction
depends antilinearly on w, hence the notation. The Clifford action of V on
the polarized Fock space is then defined by

TTJ(V) := E(PjV) + L(P-JV). (5.18)

It is easily checked that TTJ(V ) 2 = (v lv) 1 = g(v, v ), therefore the real-linear


map TTJ: V- L(NWJ) extends to a representation of G(V) on NW1 .
Exercise 5.12. Checkthat E(PJV) is the adjoint operator to t(P_ 1 v), so that
TTJ ( v) is selfadjoint for each v E V. 0

It follows that rr1 is an involutive representation, i.e., rr1 (a*) = rr1 (a)t
for all a E G(V).
The representation rr1 on the (unpolarized) Fock space is defined sirni·
larly, using the "creation" operators a j (v) and the "annihilation" operators
a 1 (v) of quantum field theory, defined on NV1 by

and
k
a1(v)(u1 1\ · • · 1\ uk) := L (-l)i- 1 (v I u 1 ) 1 u1 1\ • • • 1\ Uj 1\ • · · 1\ Uk.
j=l

Thus {a1 (u),aj(v)} = (u I v) 1 = ((PJU I PJV)), while {aj(u),aj(v)}


{aJ(U),a 1 (v)} = 0. It follows that

TTJ(V) := a 1 (v) + aj(v).

Exercise 5.13. The Clifford action rr1 extends in particular to the complex-
ification vc. Show that rr1 (w) = E(w) for w E w1 and that rr1 (z) = t(z)
forzE w1 . 0

Lemma 5.9. The Clifford action (5.18) on Fock space is irreducible.

Proof. This follows from Schur's lemma: we show that any operator T on
NW1 commuting with all { rr1 (v) : v E vc} is a scalar. Indeed, T must
commute with each E ( w) and each L(.i), for w, z E w1 . Then L(.i) TQ =
T(t(.i)Q) = 0 for all z E w1 , so TQ E A 0 W1 , i.e., TQ = tQ for some t E C
Thus, T(wl 1\ • · • 1\ Wr) = T(E(Wl) ... E(Wr)Ol = E(Wl) ... E(Wr)(tQ) =
t W1 1\ · • • 1\ Wr, SO T = tl. 0

Any Fock space is therefore a carrier of the irreducible representation of


G(V), i.e., it is an irreducible Clifford module.
188 5. Finite-dimensional Clifford Algebras and Spinors

Lemma 5.10. The operator TTJ(;y) is the grading operator on the exterior
algebra N W1.
Proof. The claim is that the ( ± 1) -eigenspaces of rr1 ( ;y) are the even and odd
subspaces A±W1 ; in other words, that rr1 (;y)a = (-1)ka for all a E AkW1 .
Given an oriented orthonormal basis {e 1 , ••• ,e 2 m} of (V,g) where e 2j =
Jezj-I. the vectors Zj := P1e2j-I and Zj := P-1e2j-I (j = 1, ... , m) form
orthonormal bases for Wand W respectively. Now ZjZj-ZjZj = -i ezj- 1 ezj.
so that
)' = (ZIZI- ZIZJ}(izZz- Zziz) ... (imZm- ZmZm).

Moreover, rr1 (ijZ j - z j i j) = [rr1 (i j ), TTJ (z j)] = [L(i j ), E(z j)]. A direct cal-
culation now shows that [L(ij), E(Zj)] a = ±a for a = Zr1 1\ • • • 1\ Zrk• with
a negative sign if and only if jE {r1 , •.. , rd; thus, rr1 (;y)a = ( -1)ka. D

~ We resume consideration of the odd-dimensional case, dim V = 2m+ 1.


If V has an oriented orthonormal basis {ei, ... , ezm+I}, Iet U be the sub-
space spanned by {ei, ... ,ezm} and Iet .J"J(U) be a Fock space for Cl(U).
Now Cl(U) ""Cl+(V), via the algebra isomorphism that extends the linear
map v- ivezm+I from U into Cl+(V); this makes .J"J(U) an (irreducible)
representation space for Cl+ (V) also. The algebra Cl( V) is generated by
o+ (V) tagether with the central element ;y (which is odd), so we can ex-
tend the action of Cl+ (V) on .J"1 (U) to all of Cl(V) injust two ways, namely,
by letting ;y act on JJ (U) as either + 1 or -1. We have thus constructed
two irreducible representations for Cl( V), both of dimension zm, which
are inequivalent (no nonzero operator can intertwine +1 and -1), whose
restrictions to the even subalgebra Cl+ (V) coincide.
Since Cl(V) is il 2 -graded, we may also consider "superrepresentations"
that respect the gradings; see Definition 5.21 in Section 5.A. It is convenient
to introduce the terminology of graded modules [282].
Definition 5.9. A (left) graded module for the Clifford algebra Cl(V) is a
complex Hilbert superspace F = F+ el F- with an algebra homomorphism
c: Cl( V) - End F that respects the ilz-grading; that is, c (a)P s;; F± if
a E Cl+ (V) and c(b)P s;; P if b E Cl- (V). In other words, c(a) E End+ F
and c(b) E End- F.
Each superrepresentation of Cl(V) determines an (ungraded) represen-
tation of o+ (V) by restriction; conversely, each representation of Cl+ (V)
on F+ can be amplified to a graded representation of Cl(V) on F by se-
lecting, say, a unit vector v and a unitary isomorphism ß: F+ - F-, and
by defining c(v) := (~ ß~I) E End-F. (It is easily checked that, up to
equivalence, the amplified representation is independent of these choices.)
Symbolically, we can write
5.3 Fock-space representations 189

In particular, there is a unique ::l2-graded representation if dim V is odd,


and there are two inequivalent ::l2-graded representations if dim V is even.
In the even case, the Fock representation is ::l2-graded in the obvious way, by
parity ofthe degree of elements of the exterior algebra; each rr1 (v) in (5.18)
is clearly an odd operator. The unitary isomorphism P1: A"VJ - NW1 is
even, and we may take F+ := A+VJ, F- := A-vJ. The other superrepre-
sentation is given by taking F+ := A-vJ and F- := A+VJ, i.e., by using the
opposite grading. For instance, if v is a fixed unit vector, the operator TrJ ( v)
intertwines these two superrepresentations, takes 0 E A+VJ to v E A-VJ,
and anticommutes with TrJ(y).
These superrepresentations are related to Bott periodicity, in the follow-
ing way [15]. Denote by M~ the free abelian group of finite-dimensional
(possibly reducible) graded modules for (1( ~ n); then M~ == 1: if n is odd,
M~ == 1: Eil 1: if n is even. To deal with the relations between modules for
consecutive dimensions, let i: G(~n) ..... (l(~n+ 1 ) be the standard inclu-
sion and i*: M~+ 1 - M~ the corresponding homomorphism. The cokernel
A~ : = M~ I i* (M~ +1 ) is 0 if n is odd, or 1: if n is even: this is the basic peri-
odicity. Now (l(~n) ® (l(~k) == (l(~n+k), where ® denotes the ::l2·graded
tensor product (see Section 5.A); this gives rise to ring structures on direct
sums M~ = EBn M~ and its quotient A~ = EBn A~. It can then be shown
that A~ == 1: [ 17], the polynomial ring generated by the image 17 E A2 of the
irreducible representation of (1( ~ 2).
The relation between this result and Bott periodicity is found in [15] by
using a topological apparatus to construct a natural ring isomorphism from
A~ onto EBn K(§n), where the generator of the latter ring is then identified
as [L]- [Od E K(§ 2), namely, the Bottelement of Definition 3.23.
The representations of Cl( V) also give rise to left C* G(V)-modules. To
define the G(V)-valued pairing, we can use the recipe

{~ 117} := L (c(eK)I71 ~) eK.


K~{1, ... ,n}

It is easily verified that this definition does not depend on the chosen or-
thonormal basis of V, and that b{~ 117} := {c(b)~ 117} for b E G(V) .
.., Since the (l(V) are complexifications of real algebras Cl(V,g), the un-
derlying real structure shines through in its representations.
Definition 5.10. Suppose that c is an irreducible representation of the Clif-
ford algebra G(V) on a finite-dimensional Hilbert space J{, and that there
is an antiunitary operator C: J{ - Jf, i.e., an antilinear bijection such
that (C~ I C17) = (17 I ~) for all ~. 17 E J{, and also that C 2 = ±1. Then
c(a) := Cc(ä)C- 1 for a E (l(V) is also a representation. Moreover, c is
said to be of complex type if the representation c is inequivalent to c; oth-
erwise, it said tobe ofreal type or of quaternionie type according as C2 = + 1
or C2 = -1.
190 5. Finite-dimensional Clifford Algebrasand Spinors

This classification into types is usually associated with group represen-


tations. lf dim V = n, we shall also use the notation c for the restriction
of this representation to a certain finite subgroup CL(n) of its invertible
elements. Choose any orthonormal basis {e1, ... , en} for (V, g), withg posi-
tive definite. Because of the basic commutation relations (5.3), these vectors
satisfy

(5.19)

when regarded as elements of Cl( V); moreover, the algebra is generated by


these vectors and, up to an automorphism of the algebra, it is immaterial
which orthonormal basis is chosen. If c: Cl (V) - End(W) is a representa-
tion of G(V) on a complex vector space W, then the operators YJ := c(e1 )
satisfy

(5.20)

Moreover, the algebra generated by the YJ is precisely c(G(V)). In fine,


each representation of the Clifford algebra is determined by a representa-
tion (5.20) of the commutation relations (5.19). If K = {k1 < · · · < kr} ~
{1, ... , n} and L = {ll < · · · < l5 } ~ {1, ... , n}, then

eKeL = ±eK.;.L, where K + L := (Ku L) \ (K n L), (5.21)

and the sign isthat of the shuffle permutation that puts k1, ... , kr, l1, ... , l 5
in increasing order. Therefore, the 2n+l elements

CL(n) := { ±eK: K ~ {1, ... , n}}

form a group, called the Clifford group of order 2n+l. Notice that e_K 1
(-l)r<r-lll 2 eK. Each orthonormal basis of (V,g) generates an isomorphic
copy of CL(n) inside (l(V). The finite Clifford groups CL(n) already mark
a distinction between the even- and odd-dimensional cases, since their
centres differ according to parity; indeed, Z(CL(n)) = {±1} if n is even,
whereas Z(CL(n)) = {±1, ±e1...nL a four-element group, if n is odd.
Since eKere_K 1ei 1 = ±1 by (5.21), the commutator subgroup of CL(n) is
just {± 1}, and the abelian quotient CL ( n) I { ± 1} "" 7L~ yields 2n inequiv-
alent irreducible one-dimensional representations of CL(n). Now the di-
mension dp of any irreducible representation p of a finite group divides
the order of the group, and the counting formula Lp d~ = 2n+l, where
the sum runs over all classes of inequivalent irreducible representations,
easily leads precisely to 2m for the dimension of the unique representa-
tion of dimension greater than one (for n = 2m), or of the only two of
dimension greater than one (for n = 2m+ 1). Boya and Byrd have pointed
out [46] that the types of the irreducible representations of G(V) can also
be determined from the types of their irreducible restrictions to CL(n).
5.3 Fock-space representations 191

Indeed, the Frobenius-Schur theorem [54, II.6] or [439, III.S] says that, for
any irreducible representation p of a finite group G,
+ 1 if p is real,
l~l I tr p(g 2 ) {
= 0 if p is complex,
BEG -1 if p is quaternionic.

For the case G = CL(n), p = c, using that e,k = ±1 for each K, one can
compute that
1 , 2 {cos(2rrn/8) + sin(2rrn/8) if n is even,
--L..2trc(e ) =
2n+1 K K 2- 1i 2 (cos(2TTn/8) + sin(2TTn/8)) if n is odd,
(5.22)
for each of the irreducible representations of CL(n) of dimension greater
than one. (For odd n, the computation is the same for both such represen-
= =
tations.) This reduces to + 1 for n 0, 1, 2 mod 8; 0 for n 3, 7 mod 8; and
-1 for n = 4, 5, 6 mod 8. (A glance at the spinorial clock confirms the re-
sult.) The eightfold periodicity of the real Clifford algebras is also manifest
from (5.22).
The question left unanswered by this application of the Frobenius-Schur
theorem is where the antiunitary operator C comes from. Of course, the
representation theory of finite groups guarantees the existence of such a C,
but in a nonconstructive way.
It turnsout that Cis simply an implementor, on the representation space,
of the charge conjugation K on Cl(V); this is an antilinear isomorphism of
the algebra. When V has dimension 2m + 1, we may regard the corres-
ponding space as the Carrier of the (unique) irreducible 2m-dimensional
representation of Cl+ (V), which we extend to Cl (V) by setting c ( y) := + 1.
Let us write A := Cl(V) or o+ (V), according as n = 2m or n = 2m+ 1.
By Lemma 5.5, A is a simple algebra, so that the isomorphism c(a) ......
K(c(a)) := c(x(ä)), for a E A, is implemented by an antilinear operator C,
that is,
C C(a) c- 1 = K(C(a)). (5.23)
We can describe the intertwining by choosing an oriented orthonormal
basis {e 1, ... ,en} for V and writing yJ := c(ej). For n even, (5.23) reduces
to
Cyic- 1 = -yi for j = 1, ... ,n. (5.24)
However, ifnis odd, thenej ft A, and we canonly saythat Cyiyic- 1 = yiyi
for i,j = 1, ... ,n.
Definition 5.11. Weshall say that a pair (c, C), consisting of a representa-
tion of Cl(V) or o+ (V) and an antiunitary operator satisfying (5.23), is a
representation of the pair (1[1<+> (V), K).
192 5. Finite-dimensional Clifford Algebrasand Spinors

lt is therefore of interest to classify such representations. Atiyah pointed


out [11] that this is equivalent to classifying the representations of the real
Clifford algebras Clp,q with p + q = n. Indeed, since K is determined by its
restriction to the subspace IRin ®IR ([ of Cl( IRin) and is antilinear, it is enough
to represent a real subspace IR(P EB i!Riq. Notice that if {e 1, ••• ,ep. E1 , ••• , Eq}
is an orthonormal basis for IR(P EB i!Riq, then K(er) = -er and K(E 5 ) = +Es
for each r, s. The representation theory of such "Clifford algebras with
involution" can thus be obtained, in principle, from the spinorial clock of
Section 5.1.
As before, one may also catalogue superrepresentations of the real Clif-
ford algebras Cln,o. obtained by amplifying ordinary representations of
Cl~ 0 . Their equivalence classes generate abelian groups Mn and An :=
Mn/i* (Mn+d satisfying Mn ""Mn+B• An ""An+B. and the ring structure of
A. is much more complicated than in the complex case (there are three gen-
erators, satisfying four relations) [15]. The topological construction identi-
fies An to the reduced KO-group of §n, thereby yielding an eightfold Bott
periodicity in the real case .
.,.. Any representation c of the Clifford algebra Cl(V) gives, by restriction,
a representation of the group of invertible elements Cl(V)x and its sub-
groups Spine (V) and Spin(V). Consider the even-dimensional case again.
If the algebra representation is irreducible, then so is the group represen-
tation of Cl( V) x, since Schur's Iemma implies that any intertwining ope-
rator is a scalar multiple of the identity. The subgroup Spine (V), on the
other hand, lies in the even subalgebra G+(V), so c(y) commutes with
c(Spine(V)); since V is even-dimensional, this implies that c(y) reduces
this representation of Spine(V). Let S generally denote the Hilbert space
carrying the irreducible 2m-dimensional representation c of G(V). Since
c(v) is selfadjoint for each v E V (Exercise 5.12), the Clifford action is
involutive. In particular, c(u- 1 ) = c(u*) = c(u)t if u is an even unitary
element, i.e., u E Spine(V).

Definition 5.12. The restriction of the algebra homomorphism c: Cl( V) -


End S to Spine (V) is a unitary representation of this group, called the spin
representation of Spine(V) on S. Weshall denote it by J1 when we wish to
distinguish it from the algebra representation c.
The chirality element y lies in the centre of Spine(V); let s+ and s- be
the (±1)-eigenspaces for the operator c(y) on S. Then the spin represen-
tation J1 is the direct sum of its two subrepresentations on s+ and s-.

By Lemma 5.10, when S is realized as the polarized Fock space NW1 ,


thens+ = A+w1 ands- = A-w1 . Thus s+ ands- have the same dimension,
2m- I.

Lemma 5.11. The two spin subrepresentations on s+ and s- are irreducible


and inequivalent.
5.3 Fock-space representations 193

Proof. If w1. ... , Wzk are unit vectors in W1 (with 2k 5 m), then by Exer-
cise 5.13,

W1 1\ • • • 1\ Wzk = E(W1) ... E(W2k)0 = J1(W1Wz) ... J1(Wzk-1W2k)Ü,


so the Spine (V)-invariant subspace of S containing 0 is all of s+. On the
other hand, if w 1, w 2 E W1 are unit vectors, then wzw 1 lies in Spine (V) and
J.l(WzÜ/1) W1 = E(Wz)L(Ü/1) W1 = Wz, SO the cyclic subspace of s- generated
by any nonzero vector in W1 contains all of W1 and therefore equals all
of s-. Thus both spin subrepresentations are irreducible.
Moreover, if R: s+ - s- is any Operator that intertwines the two sub-
representations, then R commutes with all operators of the form c (u v) =
c(u)c(v), when u and v are unit vectors in V. Therefore, R must commute
with c(y), and hence Ra= Rc(y)a = c(y)Ra = -Ra for all a Es+, so
that R = 0. Schur's lemma implies that the subrepresentations are inequiv-
alent. 0

By restricting further to the subgroup Spin(V), we again get a spin rep-


resentation of this group that is the direct sum of two inequivalent irre-
ducible subrepresentations on s+ and s-. This happens because Spine (V)
is generated by Spin(V) and the scalar subgroup T; indeed, as we saw,
Spine(V)"" (Spin(V) x T)/{±1}. The two inequivalent irreducible subrep-
resentations of Spin(V) are usually denoted by ß + and ß- .
.- When V is odd-dimensional, the situation of the representations of the
Clifford algebra l(l(V) and the group Spine(V) is reversed: there are two
inequivalent irreducible representations of the algebra, but only one irre-
ducible representation of the group. Indeed, as already pointed out, the
restrictions of these two representations to the even subalgebra n+ (V) co-
incide. In particular, they coincide on the group Spine (V), since Spine (V) c
1(1+ (V). Therefore, the spin representation J1 -usually denoted by ß on
Spin( V)- is obtained by restriction from either irreducible representation
of the algebra Cl(V).
Example 5.8. Consider 1(1(~ 3 ) ""Mz(l() E9 Mz(C). A representation of it is
given by the map ei - ai, i = 1, 2, 3, for {e1, ez, e3} an orthonormal basis
of ~ 3 . Here y is represented by -ia1a2a3 = +1. Then eze3,e3e1.e1e2 E
Spin(3) correspond respectively to ia1, iaz and ia3. Theseareelements of
the Ue algebra su(2) of the group SU(2) of unitary matrices of determi-
nant 1,
u = (_aß !) E SU(2) if aä + ßß = 1.
As in Section 2.6, we write n · cJ := n1 a1 + nzaz + n3a3 where n
(n 1, nz, n3) denotes a point in § 2; then any u E SU(2) may be expressed
in the form
-
u ( n, 1./J ) = exp ( - zi./J
i
n- · a-) = 1
cos zi./J - t. sm
. 1
zi./J n- · a,
- (5.2 5)
194 5. Finite-dimensional Clifford Algebrasand Spinors

with ii E § 2 and 0 ::5 t.fJ ::5 2rr, with the understanding that u(ii, 2rr) = -1
for all ii. This is called the angle-axis parametrization of SU(2). It follows
that Spin(3) ""SU(2). By direct computation we find that

u(ii,t.fJ) (x · ä)u*(ii,t.fJ) = (RIJlnx) · 0', (5.26a)

where the matrix RIJln• given by Euler's formula

RIJln x := cos t.fJ x + sint.fJ ii "x + (1- cos t.fJ}(ii · x) ii, (5.26b)

is indeed a rotation of axis ii and angle 1./J. Moreover, if

(5.27)

then N 3 = - N and, on one hand,

RIJln = exp(t.fJN) = 1 + Nsint.fJ + N 2 (1- cost.fJ),

while on the other, N -in the notation of (5.12)- is identified with- ~ii · ä,
as it should be.
Concerning the topology of the groups involved, we note that the pa-
rameters (ii, t.fJ), with 0 ::5 t.fJ ::5 rr, cover S0(3) once, except that ii and
-ii correspond to the same rotation when t.fJ = rr. That is to say, S0(3)
is topologically a ball with antipodal points on the surface identified. This
makes it easy to understand why it is doubly connected. On the other hand,
the underlying manifold of SU ( 2) is clearly § 3 • To RI/Jn there correspond
u(ii, 1./J) and u( -ii, 2rr- 1./J) = -u(ii, 1./J).
Exercise 5.14. Identify Spinc(3). Prove that Spin(4) ""SU(2) x SU(2) and
that Spin(6) ""SU(4). 0

..,. Since Spinc(V) and Spin(V) are central extensions of the rotation group
SO(V), their spin representations may also be regarded as projective repre-
sentations of SO (V). The latter viewpoint comes to the fore in the infinite-
dimensional context. We adjourn the matter of a concrete exhibition of the
spin representation until Section 6.2, where, after we have learned about
Pfaffians and Gaussians, it is done for the case of infinite dimensions. Nev-
ertheless, in the next section, when we understand better the workings
of Fock space, we describe it for a distinguished subgroup of Spinc(V). Of
course, Spin(V) has other irreducible representations, besides Ll +, Ll- or Ll.
Merely at the Ievel of the fundamental representations, it is known that the
ones associated with fundamental weights w 1, ... , w J for j = 1, ... , m - 1
if dim V = 2m+ 1, respectively for j = 1, ... , m- 2 if dim V = 2m, lift
from representations of SO(V): the "natural" representation and its exte-
rior powers.
5.4 The exterior algebra viewpoint 195

We do address nowl howeverl the matter of the concrete form for the
corresponding infinitesimal representations. We make use of the isomor-
phism between the Ue algebra so(V) of the specialorthogonal group and
that of Spin(V) and therefore regard the infinitesimal version of f.1 as a
1

representation of so(V).
Definition 5.13. The infinitesimal spin representation is the real-linear
map iJ: so( V) - EndS obtained by composing the isomorphism of Lemma
5.7 from so(V) to Q(i\ 2 V) with the restriction of the irreducible represen-
tation of (l(V) to Q(i\ 2 V); namelyl

{l(B) := c(B) for BE so(V). (5.28)

According to Proposition 5.8 f.l(expB) = exp({l(B)). Each {l(B) is skewad-


1

joint. For that it is enough to note that c(uv- vu)t = [c(u) c(v)]t =
1 1

[c(V) c(u)] = c(vu- uv) for U v E V.


1 1

The basic property of the infinitesimal spin representation is the follow-


ing commutation identity.
Proposition 5.12. For all B E so( V) and v E V,

[{l(B) c(v)] = c(Bv).


1 (5.29)

Proof. lt is enough to show that [B, v] = Bv as elements of the Clifford


algebra Cl(V). By choosing an orthonormal basis {ei 1... I en} for (V 1g) 1we
can write
~ 1 n
[BI v] =4 2:: g(e11Bek) [e 1ek. v]
j,k=l

1 n
= 2 2:: g(eJIBek) (g(ek1 v) e1- g(e1, v) ek)
j,k=l

1 n 1 n
= 2 2:: g(ek. v)Bek + 2 2:: g(e11v)Be1 = BV 1
k=l }=1

on using (5.11) and the skewsymmetry of B. 0

5.4 The exterior algebra viewpoint


The moral of the storyl so farl is that when the dimension of V is evenl
the Clifford algebra Cl(V) is identified to the algebra of operators on the
complex exterior algebra over V1 for some orthogonal complex structure].
I

This presentation of End(NV1 ) is not the morenatural one in many con-


texts. We now want to learn to use naturally occurring operators on NV11
196 5. Finite-dimensional Clifford Algebrasand Spinors

and to express them in terms of "creation" and "annihilation" operators:


this is a kind of toy model for quantum field theory. Given any (complex
linear) operator A on VJ, we can define the lifted (or second-quantized)
operators ArA and dArA on Arv1 by
(5.30a)
y

dArA(u1 1\ • · • 1\ ur):= L u1 1\ · • • 1\ Uj-1 1\ Aui 1\ ui+1 1\ • • • 1\ ur,


j=1
(5.30b)

and A 0A := 1 and dA 0A := 0 on A 0 V1 ""<C. Note that AmA = detA. We write


AA and dA(A) for the corresponding graded operators on NVJ. Clearly
A(exp tA) = exp(t dAA). If Ais unitary, then AA is unitary; if Ais selfad-
joint or skewadjoint, then dA(A) is respectively selfadjoint or skewadjoint.
In these cases, the definitions (5.30) are especially important.
The traces of the operators ArA appear in an important formula:
m
det(l ±A) = L (±1)rtr(ArA). (5.31)
r=O
Before verifying it, we introduce a notation for bases of N v1 . Given an ori-
ented basis {u 1 , ..• , Un} for v1 (not necessarily orthonormal, except when
indicated), for each increasingly ordered subset K = {k 1 < · · · < kr} ~
{1, ... , m}, we write UK := Uk 1 1\ Uk 2 1\ • • • 1\ Ukr in Arv (the circumflex is
used to distinguish it from the product UK in Clifford algebra introduced
in the proof of Proposition 5.6). We set 11 0 := 0. If A has a triangular
matrix with respect to the basis {u i}, then the matrix of ArA for the basis
{uK : IK I = r } is also triangular. Now Iet A1, ... , Am be the eigenvalues of A.
The left hand side of (5.31) -with the plus sign, say- is then n~= 1 (1 +Ar),
while
tr(ArA)= L (uKIArAuK)J= L Ak1 ... Akr·
IKI=r 1sk1 < .. ·<krsm

Summing over r gives (5.31).


Note that A(AB) = AAAB. lt follows that IIAr(AB)II :5 IIArAIIIIArBII. If
A = UIAI is the polar decomposition of A, the same equality gives AA =
AU AIAI, from which IAAI = AIAI. In particular, ArA and Ar lAI have the
same norm. Since the largest eigenvalue of AYIAI is nJ=1
Sj(A), where Sj(A)
is the jth singular value of A (see Section 7.C for this notation), we conclude
that

n Sj(AB) n Sj(A) Sj(B).


y y

:5 (5.32)
j=1 j=1

Recall from the previous section that the Clifford action of V on Fock
space is written as TTJ(V) := a 1 (v) + a} (v ), using annihilation and creation
5.4 The exterior algebra viewpoint 197

operators. Note that TTJ(]V) = i(aj(v)- a 1 (v)). Since we can then solve
for the at and a operators:
(5.33)
any operator on NV1 can be decomposed as follows. To reduce notational
clutter, we shall often suppress the dependence on the chosen complex
structure ], writing a(v), at(v) and so on.
Proposition 5.13. Let {u 1 , ... , Um} be an orthonormal basis for the Hilbert
space VJ. Any element S o(End(NVJ) can be written in the form

(5.34)

whereK,L ~ {l, ... ,m} anda/Kl := IlkEKat(uk),a<Ll := IlLELa(uL) in the


natural order; with a(0 la< 0 l := 1. E3
It is particularly easy to express in this way the operators dA(A) of (5.30).
From now on, weshall abbreviate at := at(uk), a1 := a(uL), with respect
to a given orthonormal basis of V1.
Proposition 5.14. Write Au1 =: I~ 1 Ak!Uk. Then
m
dA(A) = L Akl ata 1. (5.3 5)
k,l=1

Proof. First of all,


m m m
L Akl ata 1Ur = L Akl ata1ato. = L Ak!Ölr ato. = Aur.
k,l= 1 k,l= 1 k,l= 1
Moreover, if b E APVJ, c E NVJ, the following properties hold:
aL(b 1\ c) = aL(b) 1\ c + (-l)Pb 1\ aL(c),
at(b 1\ c) = at(b) 1\ c = (-l)Pb 1\ at(c),

so in particular a1 is an antiderivation, and thus ata1is an (even) derivation:

ata 1(b 1\ c) = ata 1(b) 1\ c+b 1\ ata 1(c).

The identity (5.35) follows by linearity.


Perhaps it is simpler to observe that dA(Iuk)(u!l) := a!a 1 and A =
_LAk!IUk)(u!l. D

Note that I~ 1 atak = dA(l) = r over Arvl; this is the numberoperator.


Exercise 5.15. If Ais invertible, prove the intertwining properties
0
198 5. Finite-dimensional Clifford Algebrasand Spinors

From the relation

one gets the Reisenberg equation:


d
dt at(etAv) = [di\A,at(etAv)],

and the fundamental commutation relations

which are tobe compared to (5.29).


Exercise 5.16. Work out the commutation relation
[di\(A),a(v)] = -a(Atv). 0

Notice that, on using (5.10), one obtains the identities

which go perfectly with these commutation relations .


.,. We can now present the spin representation in a particular case. We
shall show that UJ(V) Iifts to Spinc(V), i.e., that there is a map l: U1 (V)-
Spine (V) so that the following triangle commutes:

~Spinjc(V)
UJ(V) <1>

~SO(V).
Moreover, there is also another commutative triangle:

Spinc(V)

U1 (V) ~~ 11

~End(J)(V)).
Fora given U E UJ(V), Iet {u1, ... ,um} be an orthonormal basis for V1
with respect to which U has the diagonal matrix

.-J·
5.4 The exterior algebra viewpoint 199

Consider the corresponding basis {e 1, ... , e2m} of (V, g) where e21 -1 := u J


for j = 1, ... , m and e21 = Je21-1· Note that TTJ(e2J-dTTJ(e2J) = i(a1 aj-
aja1 ), and that (a 1aj- aja1 )uK = (1- 2aja1 )uK = uK if j fl K, whereas
(a 1aj - aja1 )uK = (2a1aj - 1)uK = -UK if j E K, as in the proof of
Lemma 5.10. Then

n exp(tj(- f + ~e2j-1e2j)) = n e-itj/ 2(cos ~tj + sin ~tj e2j-1e2j)


m m
l(U) =
J~l j~1

lies in Spinc(V), and cf>(l(U)) is the corresponding rotation:

cos t 1 sin t1
- sin t1 cos t1
cos t2 sin t2
- sin t2 cos t2

cos tm sintm
- sin tm cos tm

Moreover,
J.1(l(U))UK = e-i(tj+···+tm)/2 n
jEK
e-itj/2 n
HK
eitj/2 UK = n
jEK
e-itj UK = AUuK.

It is clear that elements in SU1 (V) lift to real elements in Spinc(V), i.e., to
Spin(V). Also, as remarked by Plymen [375], these formulas allow an easy
computation of the character of the spin representation.
Let cf>- 1(UJ(V)) be the preimage of the unitary group in Spinc(V). lt is
clear that the short exact sequence

splits. Indeed, cf> -l ( U1 (V)) is isomorphic to the direct product of "U" and
U1 (V). Compare Exercise 5.8 and the diagram

~ Spinjc(V)
UJ(V) <cf>.v)
~SO(V) x "U".

~ Next, we are interested in the lifting of an element of the Lie algebra


of the orthogonal group, i.e., in expressing Ji in the present language. For
that, we must handle real-linear operators. Any real-linear operator S on V
that commutes with] defines a (complex-)linear operator on V1 . On the
200 5. Finite-dimensional Clifford Algebras and Spinors

other hand, if T is a real-linear operator such that T] = - ]T, then T is


an antilinear operator on VJ. Evidently, any real-linear operator R can be
written uniquely as a sum R = R+ + R_, where R+ = !<R- ]R]) is linear
and R_ = !<R + ]R]) is antilinear. Transposes of operators are taken with
respect to the real bilinear form g. The transpose of a linear operator S
satisfies (u I stv) 1 = (Su I v) 1 , for all u, v E V; indeed,

(u 1 stv) 1 = g(u,stv) + ig(]u,stv) = g(Su, v) + ig(]Su, v)


= (Su I v) 1 • (5.36a)

By contrast, (u I Ttv) 1 = (v I Tu) 1 , for an antilinear operator T:

(u I Ttv) 1 = g(u, Ttv) + ig(]u, Ttv) = g(Tu, v)- ig(]Tu, v)


= g(Tu, v) + ig(Tu,]v) = g(v, Tu)+ ig(]v, Tu)
= (v I Tu) 1 • (5.36b)

Definition 5.14. Denote by Sk(V) the vector space of all operators on V


that are antilinear and skewsymmetric (i.e., antiskew in the terminology
of [378]). This is a complex Hilbert space under the usual rule (oc + iß) T :=
ocT + ß]T for oc, ß E ~. with the Hilbert-Schmidt norm.
Any element B of the lie algebra of the orthogonal group is a real-linear
operator skewsymmetric with respect to g, so its linear part B+ is skew-
adjoint and its antilinear part B_ is antiskew with respect to the hermitian
scalar product on V1.
After these preliminaries, we finally bring ourselves to do the computa-
tion. From the expression (5.12) for B, the operator p(B) can be rewritten
as

41 2.B(Uk,BUL) TTJ(Uk)TTj(UL) + g(]Uk,ßUL) TTJ(]Uk}TTJ(UL)


k,l
+ g(uk. B]uL) TTJ(Uk)TTJ(]uL) + g(]uk, B]uL) TTJ Uuk)TTJ(]uL)
=~ 2.g(uk.BuL)(ak + ak)(ai + a1) + ig(]uk,BuL)(ak- ak)(ai + aL)
k,l
+ ig(]uk,]B]uL)(ak + ak)(at- aL) + g(uk,]B]uL)(ak- ak)(ai- a1)
=~ 2.<uk I B-U!) 1 a!ai + (Uk I B+ul) 1 a!al + (B+ull Uk) 1 akai
k,l
+ (B_ull Uk) 1 akal
1 L(uk
=2 ' t t + 2(uk I B+u!) 1 akal
I B_ul) 1 aka 1
1
t + (ß_u!l Uk) 1 akal- 2TrB+.
k,l

The last equality follows from

(B+uliuk) 1 akai = -(uliB+uk) 1 akat = (u!IB+uk) 1 aiak-8kl (ukiB+uk)J,


5.4 The exterior algebra viewpoint 201

since B+ is skewadjoint and the commutation relation {akt at} = 8k1 holds.
Notice that the trace of B+ is a purely imaginary number. To abbreviate
theserather unwieldy formulae, we adopt a convenient shorthand notation,
taken from [213].
Definition 5.15. Given two orthonormal bases {uk} and {vz} of VJ, wein-
troduce, for each linear operator S and each antiskew operator Ton V1, the
following quadratic expressions in the creation and annihilation operators:

at Sa := I (uk I Svz} 1 ataz,


k,l

aTa :=I (Tvz I uk}J azakt


k,l

atTat :=I (uk I Tvz}J atat. (5.3 7)


k,l

Notice that at Sa isjust dAS, byProposition 5.14, and that atTat and aTa
are mutually adjoint.
Exercise 5.17. If Q and R are linear operators on v1 , and if T E Sk(V),
show that the following identities hold:

Exercise 5.18. Checkthat the right hand sides of (5.37) are independent
of the orthonormal bases used; in particular, there is no loss of generality
in taking Vk = uk. 0

Wehave arrived at a practical formula for {l(B).


Lemma 5.15. lfV has finite even dimension, then

(5.38)

for any BE so(V}. E3

Observe that, even if B is linear, this does not quite coincide with dAB:
that is the difference between the latter Clifford-algebra and an exterior-
algebra calculation. The summand - ~ Tr B+ is necessary for {1 to be a ho-
momorphism of Ue algebras. This represents also a "fermionic" parallel to
the difference between Moyal and Berezin quantizations on the "bosonic
Fock space" i.e., the symmetric algebra. Nevertheless, (5.38) does not ob-
struct the lifting of special unitary elements in SU1 (V} to Spin(V), as their
Ue algebra generators are traceless.
• If b E A 2 V1 isafinite sum of bivectors, the map
202 5. Finite-dimensional Clifford Algebras and Spinors

defines a skewsymmetric bilinear form on V1; thus there is a bounded real-


linear operator Tb on V satisfying

(5.39)

lt follows that (u I Tbv) 1 = -(v I Tbu) 1 for all u, v E V; in other words,


g(u, Tbv) = -g(v, Tbu) so that Tb is skewsymmetric. Also, g(Ju, Tbv) =
-g(Jv, Tbu) so that g(u, (Tb]+ ]Tb)V) = 0, i.e., Tb is antilinear.

Lemma 5.16. b ..... Tb is an isometry.

Proof. Choose an orthonormal basis {uk} for VJ, so that {uk,]uk} is an


orthonormal basis for (V, g). The square of the Hilbert-Schmidt norm of Tb
is computed as

II Tb II~ := "f.g(TbUk, TbUk) + g(Tb]Uk, Tb]Uk) = 2 "f.g(TbUk. TbUk)


k k

= 2"f.(TbUk I TbUk)J = 2"f.l (Uj I TbUk),l 2 = ~ .2. I (Uj 1\ Uk I b) 1 2


k J,k J,k=O

= 2.1 (uJ "uk I b) 1 = 2 llbll 2 . D


j<k

Theinverse isometry between Sk(V) and A 2 V1 isT ..... HT, where

(5.40)

since it is obvious that

lncidentally, the relation b = HTb shows that the right hand side of (5.40)
is independent of the chosen orthonormal basis.
Exercise 5.19. If Ais a linear operator on v1 and R E Sk{V), then ARAt E
Sk(V) also. Prove that HARN=! Li,J(ui I Ruj) Aui A AuJ. 0
Consider now the grading operator on the exterior algebra, given by
(-l)r on Arv1. By Lemma 5.10 and its proof, this grading equals rr1 (y),
and we can write it as

1Tj(}') = n
m

k=l
(akat- atak) = nm

k=l
(1- zatak). (5.41)

Define a natural supertrace on End(NVJ) by

StrS := trrr1 (y)S.


5.4 The exterior algebra viewpoint 203

This patently vanishes on odd operators. LetS = sorr,(;y) + other terms.


From the proof of Proposition 5.6, it is clear that Str S = 2m so (since the
trace on Cl(V) induced by the spin representation is proportional to T by
the factor trrr,(1) =2m). In the expansion (5.34), clearly Str(a/K)a(L)) = 0
unless K = L = {1, ... , m}, while (5.41) shows that
rr1 (;y) = ( -l)m<m+l)/ 2 2maL .. mal, ... ,m + lower-order terms.
So we conclude that StraL .. mal, ... ,m = (-l)m<m+0/ 2 , and then
StrS = (-l)m(m+l)/ 2 Sl, ... ,m;l, ... ,m·
This is a painless derivation of the so-called Berezin-Patodi formula [127].
The supertrace Str on the Clifford algebra is unique up to a constant multi-
ple -in view of Proposition 5.21 and Exercise 5.2.
~ An exterior algebra over an n-dimensional vector space can be thought
of as the algebra generated by n anticommuting variables (often called
"Grassmann variables"). Grassmann variables are used by physicists par-
ticularly in supersymmetric theories, for which the superspace techniques
mentioned in Section 5.A are especially relevant. In fact, as explained by
Winnberg [487] long ago, superfields are just spinor fields. Berezin [32]
proposed a calculus of derivatives and integrals over the Grassmann vari-
ables, which turned outtobe a fruitful idea, since many aspects of ordinary
integration with commuting variables have interesting analogues. In fact,
Grassmann derivatives and integrals with respect to a given variable coin-
cide, and they can be interpreted as the contractions with respect to that
variable in the polarized Fock space representation. In that context, the
Berezin integral -defined directly below- is naturally seen as a multidi-
mensional integral; and the Pfaffian of a skewsymmetric matrix -which is
a square root of the determinant, computable by a certain polynomial in
the matrix elements- turns out to be the analogue of a Gaussian integral
in Berezin's theory.
Definition 5.16. The top tier Anv of the exterior algebra is one-dimensio-
nal, so we can define a linear form on the full exterior algebra by selecting
the coefficient of h .. n· This linear form is called the Berezin integral J/\,
defined by
f l\, •
ex .=
{1, if K = {1, ...
,n},
0, if IKI < n.
Since AnT = det T, the "change-of-variables formula" for this integral

r r
amounts to

AT(oc) = det T oc for oc E NV, TE End(V).


Therefore the Berezin integral is invariant under special orthogonal trans-
formations.
Note that J/\ is essentially Str o Q.
204 5. Finite-dimensional Clifford Algebras and Spinors

5.5 Pfaffians and Gaussians


Definition 5.17. The even subalgebra A+V is a commutative algebra, so
we may define functions of its elements in the usual way. For instance,
if b E A 2 V, the exponential exp bis just the finite series

expb :=

whose top-degree term is (m!)- 1 b"m. (The right hand side makes sense for
odd n also, but in that case the top-degree term is zero.) The Pfaffian of b

r
is defined as the Berezin integral of this expansion:

Pf b := expb.

We remark that a change of orientation on V also changes the sign of the


Pfaffian. Notice also that Pf( -b) = ( -l)m Pf b.
Exercise 5.20. If b = ! 2tz=I hz ek 1\ ez for a skewsymmetric matrix B =
[bkz], checkthat

Pf b = 1
2 m m! I (-l)rr brr(l)rr(2) brr(3)rr(4) ... brr(2m-l)rr(2m)• (5.42)
TTES2m

where rr runs over all permutations of {1, ... , 2m}. 0

The right hand side of (5.42) conveys the usual definition of the Pfaffian
of the skewsymmetric matrix B and is denoted by Pf B. It makes sense if
the entries of B lie in any commutative algebra. Thus the Pfaffian is the
evaluation, at the entries of B, of a certain universal polynomial of degree m
in m(2m- 1) variables.
!
Proposition 5.17. Let b = I~T=I bkl ek 1\ ez E A 2 V. Foreach subset K s;
{1, ... , n}, denote by BK the skewsymmetric submatrix of B formed by delet·
ing all rows and all columns whose indicesarenot in K. Then the exponential
exp b may be expanded in the given basis of N V as

expb = I PfBKeK, (5.43)


IKieven

using the convention that Pf Be; := 1.

Proof. Let U be the subspace of V spanned by the ek for k E K, and let


P: V- U be the projection for which P(ek) := ek or 0 according as k E K
or not. Then the expansion of AP ( exp b) has only one term in its top de-
gree IKI, namely the eK·term of exp b. Its coefficient is obtained byreplacing
b with A 2P(b) =! Lk,IEK bkz ek 1\ ez and computing f" exp(A 2P(b)) in AU;
the result is Pf BK. D
5.5 Pfaffians and Gaussians 205

Remark. Since each Pf BK is given by the algebraic expression (5.42), we


see that the equality (5.43) is in fact obtained by plugging the entries of the
matrix B and the bivectors ek 1\ e1 into a certain polynomial identity. Since a
polynomial is determined by its values, this identity holds universally. What
this means isthat the relation (5.43) will still hold if we replace the bkl and
the ek 1\ e1 by corresponding elements of another commutative algebra.

11> We need, for later use, an identity that relates these "subpfaffians" Pf BK
of the skewsymmetric matrix B to complementary subpfaffians of B- 1 (as-
suming that B- 1 exists). For a short while, we shall suppose that B is in-
vertible and abbreviate C := B- 1 . Then C is the matrix of coefficients
of an element c = ! 2::;7=
1 Crs e~ 1\ e~ E A 2 V', where V' isanother 2m-
dimensional vector space with an oriented basis {e' 1 , ••. , e' 2 m}. By "com-
pleting the square", we find the following relation in the space A2 (V' E9 V):

b+ I e~ 1\ er = !I bkl ek 1\ e1 + I e~ 1\ er
r k,l r

= !I h1 (ek + Lr Crke~) 1\ (el + Ls Csle~) + i I Crse~ 1\ e~


k,l Y,S

= iI bk1!k 1\11 + c,
k,l

where we have used the skewsymmetry of c and put fk := ek + Lr Crke~ in


V' E9 V. After exponentiating in A+ (V' E9 V), we obtain

exp(b + Lr e~ 1\ er) =I Pf BK fK 1\ exp c, (5.44)


K

since the previous remark allows us to replace eK by fK in (5.43).


Notation. Before stating the basic combinatorial identity (5.45), which is a
version of Cramer's rule for Pfaffians, we need a little more notation. For
eachsubsetK = {k1 < · · · < kr} ~ {l, ... ,n},letK' := {k~ < · · · < k;m-rl
be the complementary set of indicesnot in K, and let 11KK' = ± 1 be the sign
of the shuffle permutation (1, ... , 2m) .... (k1, ... , kr, k' 1•... , k' 2m-r ). Then
h 1\ eK' = 11KK' h .. n·
Lemma 5.18. If B is an invertible skewsymmetric 2m x 2m matrix, then for
each subset K ~ {1, ... , n} of even cardinality, the following identity holds:

(5.45)

Proof. The Berezin integral on NV extends to a linear map from N (V' E9 V)


to NV', which we may denote by J,~v· given by L(:v oc' 1\ ß := (J" ß) oc' for
oc' E NV', ß E A"V. For C := B- 1, write fk := ek + Lr Crke~ as before; since
the fJ above satisfy fJ = e1 + lower-order terms, it follows that L~v fJ = 0
if] * {1, ... ,n}, whereas L~vfi ... n = J"e1...n = 1. Applying this map to
206 5. Finite-dimensional Clifford Algebras and Spinors

both sides of (5.44) gives

f/\ NV
exp(b +Ire~ 1\ er)= Pfb expc = PfB_LPfCre[.
L
(5.46)

On the other hand, it is easy to see that

2m 2m
exp(Ik ei 1\ ek) = 0 exp(ei 1\ ek) = 0 (1 + ei 1\ ek)
k=1 k=1
= "'e'
L k1
1\ ek I 1\ •.• 1\ e'k2r 1\ "'(-1)1KI(IK1+1)/2eK
ek 2r = L 1\ e'K·
K K

When IKI is even, ~IKI(IKI + 1) =~IKI mod 2, so the signisjust (-l)IKI/2.


Thus

f/\
NV
exp(b +Ire~ 1\ er) = L (-1) ILI/ 2 J/\ (exp b 1\ er) 1\ e[
L

L (-1) 1K'I/ 2 1JKK' (Pf BK) e]c· (5.47)


IKieven

Comparing the coefficients of e]c on the right hand sides of (5.46) and (5.47)
yields the desired identity (5.45). D

For the case K = 0, (5.45) specializes to Pf B Pf(B- 1 ) = ( -l)m or, equiv-


alently,

PfB Pf(-B- 1) = 1. (5.48)

Let A be any 2m x 2m matrix. If in Proposition 5.17 we replace ek by


I r arker, then b becomes b' := I arkbkzaszer 1\ es with matrix AB At, and

r
e1...n becomes (detA) el...n· Therefore

Pf(ABAt) = exp b' = (detA) Pf B.

Taking A = B- 1 gives the corollary Pf (- B- 1 ) = (det B) - 1 Pf B; by combining


this with (5.48), we arrive at the fundamental relation

(Pf B) 2 = det B. (5.49)

Tobemore precise: both sides of (5.49) are polynomials of degree 2m in


the entries of B, which coincide when Bis invertible, and therefore coincide
for every skewsymmetric matrix B. We remark now that, by squaring both
sides of (5.45), we obtain detB det(B- 1 )K' = detBK, which is just the usual
Cramer's rufe for subdeterminants of B- 1.
5.5 Pfaffians and Gaussians 207

Exercise 5.21. The relation (Pf B) 2 = detB can be proved by a more con-
ventional algebraic method involving polynomial rings [310, Thm. XV.9.1].
Assuming this property, give an alternative derivation of (5.45) by proving
that detB det(B- 1 )K· = detBK (using, say, the Laplace expansion of deter-
minants) and then paying due regard to the signs in the square root of both
sides. o

.". We come now to the main point. Our aim is to show that if A and B
are two skewsymmetric 2m x 2m matrices, then det( 1 -BA) has a distin-
guished square root, which weshall denote by det 112 (1- BA), depending
polynomially on the entries of A and B, that is positive when A = B. As
(5.49) strongly hints, the definition of this square root uses Pfaffians.
Lemma 5.19. lf A and Bare skewsymmetric 2m x 2m matrices with Bin-
vertible, then

Pf(-B)Pf(B- 1 -A) = L PfAK PfBK. (5.50)


IKI even

Proof. Let a,c E A2 V have coefficient matrices A and C := B- 1 ; then

Pf(B- 1 - A) = f' exp(c- a) = J'' exp c 1\ exp( -a)

L PfCKPf(-A)K·J"eKAeK·= L (-1) 1K'II 2 '7KK'PfCKPfAK·


IKI even IKI even

L PfC PfBK' PfAK' = PfC L PfAK PfBK


IK'Ieven IKI even

on using (5.43) and (5.45), and replacing K' by K in the final sum. Since
Pf(- B) Pf C = 1 by (5.48), the result follows. D

Squaring both sides of (5.50) yields the important identity


2
det(l -BA) = ( L Pf AK Pf BK) , (5.51)
IKieven

valid for invertible B in the first instance. But, once again, this is a polyno-
mial identity in the entries of the matrices A and B, so it is generallyvalid
for all skewsymmetric A and B.
Definition 5.18. Let A and B be skewsymmetric 2m x 2m matrices. Then
we define the preferred square root of det(1- BA) tobe

det 112 (1- BA):= L Pf AK Pf BK. (5.52)


IKI even

Notice that the right hand side does not depend on any orientation. Note
also that det 112 (1- B2 ) ~ 1, with equality only if B = 0.
208 5. Finite-dimensional Clifford Algebras and Spinors

.,.. Let us now choose an orthonormal basis {u 1, ... , Um} for the Hilbert
space VJ.
Definition 5.19. For each T E Sk(V), the corresponding Gaussian is the
quadratic exponential in exterior algebra:

where '1K is the sign of the permutation putting the indices k1, ... , kzr in
increasing order, and we adopt the convention that the r = 0 term is the
vacuum vector 0.
The name "Gaussian" refers to the analogaus construction on the bosonic
Fock space, where the corresponding elements are exponentials of qua-
dratic functions.
Example 5.9. Here are a few simple Gaussians. First, 0 itself is fo. Next,
if Tk E Sk(V) is determined by Tk(uzk-1) := -u2k, Tk(u2k) := U2k-l and
Tk(Uj) := 0 for other j, then HTk = U2k-l 1\ U2k, so !Tk = n + U2k-l 1\ Uzk·
Furthermore, if T = T1 + · · · + Tk. then HT = I.~= I UzJ-1 1\ uz 1 , and !T =
0 + U1 1\ • • • 1\ Uzk + lower-order terms.
The linear span of { fT : T E Sk(V)} is clearly all of J'j (V).
Expanding the Gaussian fT with respect to the orthonormal basis {uK}
yields

h=l:(Pfh)UK= I (Pfh)UK. (5.54)


K IKieven

Herewe have denoted by h the skewsymmetric matrix with entries (uk I


Tut) for k, l E K. By convention, we take Pf T 0 := 1 and Pf h := 0 whenever
IKI is odd.
When dealing with subpfaffians of antilinear operators, there is a some-
what subtle point that deserves attention. If C denotes the complex conju-
gation operator on vc -that is, C(u + iv) := u- iv for v E V- and if T is
an antiskew operator on V, then TC becomes a skewsymmetric linear ope-
rator on vc, and all the previous formulae about Pfaffians are applicable to
matrix elements of TC. Now, if T E Sk(V) is invertible with inverseR, then
(TC)- 1 = CR = C(RC)C, so the matrix elements of RC are complex con-
jugates of the matrix elements of (TC)- 1. Consider the matrix Tr. where
L c; {1, ... , n}; the combinatorial formula (5.45) applied to this matrix can
be rewritten in the current notation as

(PfTr} (PfSL\K)* = (-l)IL\Ki/ZT"/K,L\K Pfh,


5.5 Pfaffians and Gaussians 209

whenever K ~ L; here the asterisk denotes the complex conjugate. In the


proof of Lemma 5.19, the term Pf BK is likewise conjugated. From that, it
is easy to see that the analogue of (5.51) is

det(1-TS)=( I (PfSK)*Pfhf. (5.55)


IKieven

Notice also that, if det(1 - TS) is computed by a Straightforward Laplace


expansion, the matrix elements of S are conjugated by the action of the
antilinear operator T, while the matrix elementsofT arenot conjugated.
Corollary 5.20. Foreach T E Sk(V), the expansion (5.53) defines an ele-
ment h of the even Fock subspace J} (V), whose norm is given by

llhll 2 = I I Pf hl 2 = det 112 (J- T 2 ). (5.56a)


IKieven

Polarization of this squared norm yields


<fs I h) = det 112 U- TS) := I (PfSK)* Pfh (5.56b)
IKI even

for the scalar product of two Gaussians. B

Since Pf TK is a polynomial in the matrix entries of T, the right hand


side of (5.56b) is a holomorphic function ofT; and, by the same token, it
is an antiholomorphic function of S. Thus, we can regard det 1 12 (I - T S)
as a sesquiholomorphic function of two variables S, T E Sk(V), satisfying
det 112 (J- ST) = (det 112 (J- TS))*. lts values are therefore determined,
using analytic continuation, by its restriction (5.56a) to the diagonal.
A most transparent expression for the spin representation is given by
its action on Gaussian elements (a fermionic analogue of "coherent states"
in ordinary quantum mechanics). We need to deal with it in the context
of infinite-dimensional Clifford algebras and Fock spaces, and so, to avoid
useless repetition, we leave the matter for the next chapter.
For more about Pfaffians, we recomrnend the excellent tutorial in [331,
§ 1], and its continuation in [263], whose treatments we have followed here.
We close with an instructive exercise extracted from [33].
Exercise 5.22. Let V be a vector space. Consider the function
. (X)·-d sinh(X/2)
]V .- et X/ 2 .

On so(V), by a theorem of Chevalley, its square root [with j~12 (1) = +1]
can be extended to an analytic function. Now consider, for an element a of

r r
Q (A 2 V), the Clifford exponential exp a. Prove that

O"(expa) = j~12 (ada) exp(O"(a)). 0


210 5. Finite-dimensional Clifford Algebrasand Spinors

5.A Superalgebras
The prefix super is used when an object is 7L 2 -graded. Thus, a (real or com-
plex) vector space V is called a superspace if it has a 7L 2 -grading:

v = v+ ES v-.
Elements of v+ are called even, and elements of v- are called odd. The
parity of such an element v is denoted by #v := 0 or 1 according as v E
v+ or V E v-. In subsequent formulae, whenever we write #v' it is to be
understood that v is either even or odd.
Definition 5.20. A superalgebra is a superspace A = A + ES A-, with a bilin-
ear product that preserves the grading:

When an algebra Ais 7L-graded, that is, A = EBkEZ Ak with AJAk s;; AJ+k
for j, k E 7L, it can also be regarded as a superalgebra by setting

A- := E}1Azk+l·
kEZ

The tensor algebra I' (V), the symmetric algebra S(V), the exterior algebra
N(V), and the Clifford algebras Cl(V,g) and Cl(V) of a vector space V are
all examples of that.
A typical feature of superstructures is that certain plus or minus signs
appear in formulae, obtained from a "rule" that can be schematically ex-
pressed by
( .•. U) * (V ..• ) - ( -1 )#u #v (. •• V ) * (U .•• ) .

In other words, whenever there is an interchange of the order of two el-


ements in a formula across any product operations, one should expect
the extra factor (-1)#u#v. This rule is sometimes credited [319) to Koszul
or [227] to Quillen.
If A and B are superalgebras and A ® B is a suitably defined tensor pro-
duct, we can define a 7Lz-grading an it by

(A ® B)+ := (A+ ® B+) E9 (A- ® B-),


(A ® B)- := (A+ ® B-) E9 (A- ® B+).

For elements of definite parity, we set

(a ® b)(c ® d) := (-1)#b#c(ac ® bd).

The resulting superalgebra, denoted by A ® B, is called the graded tensor


product of A and B.
5.A Superalgebras 211

When V and W are superspaces, Hom(V, W) is also a superspace, by


setting

Hom+(V, W) := Hom(v+, w+) E9Hom(v-, w-),


Horn- (V, W) := Hom(V+, w-) E9 Hom(V-, w+).

The grading is defined so that #(L(v)) = #L+#v whenever v andL(v) have


adefinite parity. In the case V= W, End (V) = Hom(V, V) is a superalgebra.
Definition 5.21. A superrepresentation of a superalgebra Ais an algebra
homomorphism p : A - End V, for some superspace V, that respects the
grading; that is, p(A+) ~End+ (V) and p(A-) ~ End-(V).
Definition 5.22. A Lie Superalgebra is a superspace with abilinear bracket,
denoted [ ·, · ], that satisfies the graded versions of anticommutativity and
the ]acobi identity:

[X, Y] = -(-l)#X#Y[Y,X],
[X, [Y,Z]] =[[X, Y], Z] + (-l)#X#Y[Y, [X, Z]],

whenever X, Y, Z have definite parities. In particular, End(V) is a Ue super-


algebra, where its bracket is the supercommutator

[S, T] := ST- (-l)#S#TTS.

If A is a superalgebra, an (even or odd) endomorphism D E EndA is


called a derivation of parity #D if

D(ab) = D(a) b + ( -l)#D#a aD(b),

whenever a E A has a definite parity. Odd derivations are also called an-
tiderivations. Inner derivations are those of the form b ,_ [a, b] for some
a E A. The supercommutator [D, ß] of two derivations is again a deriva-
tion, with #[D, ß] = #D + #ß. Indeed,

[D,ß](ab) = Dß(ab)- (-l)#D#aßD(ab)


= D(ß(a) b + (-l)#Maaß(b))
- (-l)#D#ßß(D(a) b + (-l)#D#aaD(b))
= Dß(a) b + (-l)(#D+#ß)#aaDß(b)
- (-l)#D#ß(ßD(a) b + (-l)(#D+#ß)#aaßD(b))
= [D, ß](a) b + ( -l)(#D+#ß)#aa [D, ß](b).

For instance, on the Superalgebra .Jl• (M) of differential forms on a mani-


fold M, the Ue derivative by a vector field is an even derivation, while the
exterior derivative is an odd derivation.
212 5. Finite-dimensional Clifford Algebrasand Spinors

Definition 5.23. The grading operator on a superspace V is the linear map


x E End+ V defined by x<v) := (-l)#vv when v is either even or odd. If
Tr denotes the usual trace of an operator on V, the supertrace is the linear
functional on End(V) defined by

Str T := Tr(xT). (5.5 7)

lt follows that Str T = 0 if T is odd. When V is a lz-graded Hilbert space,


this formula for the supertrace makes sense provided xT is traceclass.
lt is convenient to write elements of End(V) in block matrix form:

(5.58)

For instance, x = (~ ~l), and hence Str T = Tr T++- Tr L_, Notice that
even endomorphisms are diagonal and odd endomorphisms are antidiag-
onaL
Proposition 5.21. Supertraces vanish on supercommutators.
Proof. Indeed, if Sand T are both even, then Str[S, T] = Tr[S++• T++]-
Tr[S __ , L_] = 0, whereas if S and T are both odd, then
6
The Spin Representation

We benefit from the treatment of the spin representation in the previous


chapter to introduce the infinite-dimensional case as weiL As it turns out,
it provides a kind of abstract version of quantum field theory, which will
surface in Part IV. The main result, a farnaus theorem by Shale and Stine-
spring, is of foremost importance in physics. Furthermore, it points to the
theory of Fredholm modules of Chapter 8, and thus deserves tobe counted
as a source of noncommutative geometry.

6.1 Infinite-dimensional Clifford algebras


Much of what has already been said about finite-dimensional Clifford al-
gebras carries over with only minor changes. Suppose, then, that we are
given an infinite-dimensional separable real vector space V with a positive
definite symmetric bilinear form g. We lose nothing by supposing V tobe
complete in the metric induced by g, so that (V,g) isareal Hilbert space.
Then the scalar product (5.16) makes vc a (separable, infinite-dimensional)
complex Hilbert space. On the exterior algebra A• vc, the exterior mul-
tiplication, contraction, and the Clifford operators c (v) are well-defined
by the same formulas as in the finite dimensional case. The operators
c(v) generate an algebra Gftn(V) that can be regarded as the union of all
the finite-dimensional Clifford algebras l(l(VI), where V1 runs through all
finite-dimensional subspaces of V.
214 6. The Spin Representation

Exercise 6.1. If V1 :::;; V 2 :::;; V with dim Vz < oo, show that there is an injec-
tive algebra isomorphism G(V1) .... G(Vz) and that such algebra inclusions
compose in the natural way. 0
If vl :5 Vz :5 V with Vz finite-dimensional, the (normalized, even) trace T
on G(Vz) restricts, by the uniqueness in Proposition 5.6, to the correspond-
ing trace on Cl(VI). In this way, we can define a unique trace on Gftn(V)
satisfying T ( 1) = 1 and vanishing on odd elements. The scalar product (5. 7)
makes Gfin (V) a prehilbert space; call its completion J{T· The GNS con-
struction (see Section LA) now allows us to identify a E Gftn(V) with the
operator b .- ab on J{T·
Exercise 6.2. Show that each such operator rrT(a): b - ab is bounded,
and is zero if and only if a = 0 in Gftn(V). 0
Definition 6.1. In this way, Gfin (V) is represented faithfully as an algebra
ofbounded operators on J{T· Its closure is a C*-algebra, which we define to
be the (complex) Clifford algebra Cl( V). The C*-algebra Cl( V) is separable
if and only if V is separable.
This construction nails down the intuitive idea that Cl(V) should be
the "C*-completion" of Gfin (V), since each finite-dimensional subalgebra
G(V1) carries a unique C*-norm. For all the details, and a comprehensive
treatment of the infinite-dimensional algebra, its norm closure Cl(V) and
also its weak closure (a von Neumann factor of type IId the reader should
consult [3 78].
We remark that if v E V, then
llrrT(v)ll 2 := sup{ T((vb)*vb): b E Gfin(V), T(b*b) :51}
= sup{ T(b*v 2 b): b E Gfin(V), T(b*b):::;; 1} = g(v, v),
since v 2 b = g(v,v)b in G(V1) whenever v E V 1 and b E G(V1). Thus
the natural inclusion of V (or of vc) in Cl( V) is an isometry. The Clifford
algebra Cl(V) is in factasimple C*-algebra. This happens because we can
define Gftn(V) as the nested increasing union of the G(Vz), where Vz runs
over the even-dimensional subspaces of V. Thus, any nonzero C*-algebra
morphism rr: G(V) - A restricts to a nonzero morphism G(V2 ) - A
(since rr ( 1) * 0). But each G(Vz) isasimple matrix algebra, by Lemma 5.5,
so these restricted morphisms are isometries; thus rr itself is isometric,
and so ker rr = 0. Therefore, any closed ideal of Cl( V) is trivial (since we
can take rr to be the quotient morphism.)
Exercise 6.3. Prove the C* version of Proposition 5.1: that any real-linear
map f from V into a C*-algebra A satisfying j(v) 2 = g(v, v) 1A extends
uniquely to a morphism j: Cl(V) - A. Conclude that the Bogoliubov au-
tomorphisms Bh are C*-algebra automorphisms of G(V). 0
The operator X = B-1 yields the grading Cl( V) = G+ (V) E9 n- (V) as
before. 1t is easy to see that n+ (V) is the C*-algebra completion of the
6.1 Infinite-dimensional Clifford algebras 215

union of allfinite-dimensional even subalgebras n+ (V1 ), so in fact Cl(V)


and n+ (V) are isomorphic now. However, in theinfinite-dimensional case,
the chirality element y is no Ionger available, so we cannot claim that x
is an inner automorphism. In fact, x is not inner: for a proof, see [378,
Thm. 1.2.14].
~ An important dass of irreducible representations of this Clifford C*-
algebra are carried by Fock spaces.
Definition 6.2. Let] be any orthogonal complex structure on (V,g). The
Fock space :JI(V) is defined tobe the Hilbert space completion of the ex-
terior algebra A"VI under the scalar product (5.17a). Note that JI(V) is
separable if and only if V is separable. Analogously, the polarized Fock
space is defined to be the Hilbert space completion of the exterior algebra
A"WI under the scalar product (5.17b). Then the customary definition of a
Clifford action onA"VI, say, extends to a representation of Cl(V) on :JI(V)
-necessarily isometric, from the simplicity of Cl(V). This is the Fock rep-
resentation TTI corresponding to ].
The proof of Lemma 5.9 carries over and shows that each TTI is an irre-
ducible representation. As a consequence, every nonzero element of JI (V)
is cyclic. In particular, consider the Fock state a-I on Cl (V) given by the va-
cuum vector 0, namely

If u E vc, the vector 0 is killed by rrI ( u) if and only if u E WI (recall


Exercise 5.13). In particular,

It turns out that this property characterizes the Fock state.


Lemma 6.1. If a- is a state of the C* -algebra Cl(V) suchthat a-(u*u) = 0
for all u E W I, then a- = a-I.

Proof. The relation I a- (a * b) 12 :::; a- (a * a) a- ( b * b), which is the Schwarz in-


equality for states (l.A), shows that a-(a*b) = 0 whenever either a or b
lies in W I. Now any element in Clfin (V) can be written as a finite sum
c = .\. 1 +Ir Cr where .\. E ([ and each Cr is of the form w1 ... Wkiz ... Z1, with
wiE WI and Zj E WI. Taking a = w1 or b = z1 shows that each a-(cr) = 0;
thus, a- (c) = .\.. This establishes the uniqueness of a- on Clfin (V), and by
continuity on the whole Cl(V). D

Corollary 6.2. Let TT be a representation of Cl(V) on a Hilbert space :J-f,


having a cyclic (unit) vector <I>, and suppose that rr (v + i] v )<I> = 0 {or any
v E V. Then rr is unitarily equivalent to the Fock representation rrI.
216 6. The Spin Representation

Proof. The vector state cr(a) := (4> I rr(a)cl>) complies with the hypothe-
sis of Lemma 6.1, so it coincides with cr1 . Now, the linear correspondence
TTJ(b)O. ..... rr(b)cl> is isometric, since

Since 0. and <I> are cyclic vectors, this extends by continuity to a unitary
operator U: :J1 (V) - J-f. If a, b E ((](V), then

rr(a)Urr1 (b)0. = rr(ab)cl> = Urr1 (ab)0. = Urr1 (a)rr1 (b)0.,

so that rr(a)U = UTTJ(a) for all a E Cl(V). 0

In particular, ifwe start from the Fock state cr1 and build a representation
of Cl( V) with the GNS construction, we recover the Fock representation rr1.
(There are, in fact, many other irreducible representations of Cl(V), arising
from purestatesthat do not satisfy the hypothesis of Lemma 6.1; they do
not concern us here.)
In the infinite-dimensional case, the various Fock representations are not
all equivalent. Indeed, from the discussion after Definition 5.6, which also
carries over, any other orthogonal complex structure is of the form h]h- 1
with h E O(V); giving a unitary equivalence TTJ - TThJh-1 amounts to imple-
menting the Bogoliubov automorphism eh by unitary operators on :J1 (V).
The theorem of Shale and Stinespring [434, Thm. 2.5], to which we come
in Section 6.3, gives a necessary and sufficient condition: namely, that the
J-antilinear part of h be a Hilbert-Schmidt operator .
.". There is the matter of adapting the calculation methods of Section 5.4 to
theinfinite-dimensional context. For the most part, this is routine. The def-
inition (5.30a) goes through, and when U is unitary it gives another unitary
operator. The definition (5.30b) is trickier, because in general dA(A) is un-
bounded, even if Ais bounded, the number operator providing a simple and
notable example; and because we may want to use it for unbounded opera-
tors. When Ais selfadjoint or skewadjoint, we can use the property that the
exterior algebra A·v1 is a set of analytic vectors for dA(A), together with
Nelson's analytic vector theorem [384], to establish selfadjointness or ske-
wadjointness of dA(A); and the Heisenberg relationsarevalid at least for
analytic vectors. Following [68], [213] and [380], we try to dodge technical
complications, for now, by making heavy use of the presence of Gaussian
elements in the infinite dimensional Fock space.
Definition 5.14 carries over, provided that in computations we ask that
the operators in Sk(V) be of Hilbert-Schmidt class. Definition 5.15 (and Ex-
ercises 5.17 and 5.18) will shortlybe seen to work, and Lemma 5.16 clearly
carries over, establishing an isometry between the closure of A 2 V1 in Fock
space and Sk(V). ForeachT E Sk(V), the corresponding Gaussian is the
6.1 Infinite-dimensional Clifford algebras 217

quadratic exponential in exterior algebra:

(6.1)

Since this sum is convergent (as we shall soon verify), fT lies in the even
subspace J"J (V) of the Fock space. As before, we expand the Gaussian h
with respect to the orthonormal basis {uK}. So, again,

~Hyr
r.
= I
IK1=2r
2; 1'1K(Uk 1
r.
I Tuk2) ... (Uk2r-1l TUk 2r)Uk 1 1\ • • ·I\Uk 2 r·

The coefficient of UK on the right hand side is just the Pfaffian Pf h, where
we keep the notation of the finite-dimensional case. Then (6.1) can be rewrit-
ten as

h = I (Pfh) UK, (6.2)


IK[even

where the sum runs overallfinite even subsets K of ~- The linear span of
{h: TE Sk(V)} will be densein J"j (V).
We need a well-known property of (Fredholm) determinants in infinite
dimensions: if R E L(Jf) is a bounded linear operator on a Hilbert space
Jf, then det R is definable if and only if R -I is traceclass. Let us interrupt
our line of reasoning to discuss this; we follow [43 7]. Formula (5.31) is now
used to define the determinant:

I
00

det(l + A) := Tr(ArA),
r=O

for A traceclass. The operator ArA is traceclass in Ar Jf: if s1 (A) denote the
singular values of A, the eigenvalues of IArAl =Ar lAI are SJ 1 (A) · · · SJk (A),
for j1 < · · · < ik· Therefore

and so I det(l + A) I :5 exp IIAII1. Also,


00

ldet(l+A)I :5 nn+sJ(A)).
j=O

All expected properties of determinants flow from these estimates. There


are several good treatments of infinite determinants, besides [437]; for a
clean alternative development using spectral theory, see the appendix to
the article by Araki [8].
218 6. The Spin Representation

Proposition 6.3. ForeachTE Sk(V), the expansion (6.1) defines an element


fT of the even Fock subspace :Fj (V), whose norm is given by
llhll 2 = 2:: 1Pfhl 2 =det 112 (J-T2). (6.3)
IKI even
Proof. Suppose first that T has finite rank. Then we can choose the or-
thonormal basis so that (6.2) isafinite sum; the result follows from (5.56a)
and the definition of det 1' 2.
More generally, T is Hilbert-Schmidt, so T 2 is a traceclass linear operator
on V1 , the determinant det(J- T 2) is defined. Moreover, if T(l), T(2), ... is
a sequence of finite-rank antiskew operators suchthat T(n) - Tin Sk(V),
then det(J- T 2) = limn-oo det(J- Tln)). Thus the series LIKieven I Pf hl 2
converges to the limit det 112 (J- T 2); by Parseval's identity, the series (6.2)
converges absolutely to the Gaussian fT, showing that fT E :J1 (V) and
incidentally that (6.3) is valid. D
Corollary 6.4. The scalar product oftwo Gaussians is given by (5.56b).
Proof. This follows by polarization, since the Schwarz inequality shows
that the series on the right of (5.56b) converges absolutely for S, TE Sk(V).
D

Exercise 6.4. Let exp(b) := L~=o(l/r!) b"r, with b E A2VJ. Verify directly,
by use of a spectral decomposition for Tb, that llb"r11 2 ::; r!llbll 2r and thus

II expbll 2 ::; exp llbll 2.

6.2 The infinitesimal spin representation revisited


The reason we gave for introducing Gaussian elements was to prepare for
a suitable treatment of the spin representation in the infinite-dimensional
case. It is time to justify this expectation. We deal first with the infinitesimal
spin representation f.l; it needs an important modification, as weshall see.
The "global" spin representation J1 will be exhibited in next section, in the
course of proving Shale and Stinespring's theorem. We need the following
lemma, which is just the basic differential equation for Gaussian functions
translated into the language of Grassmann variables.
Lemma 6.5. Ifv E V and SE Sk(V), then

a(v)fs + at(sv)fs = 0. (6.4)

Proof. We may assume that v is a unit vector and construct an orthonormal


basis {uo, u1, Uz, ... } for V1 with uo = v. We expand fs = LK Pf(SK) UK
with respect to the corresponding orthonormal basis in Fock space.
6.2 The infinitesimal spin representation revisited 219

On account of the development (5.42), the Pfaffian of a skewsymmetric


2m x 2m matrix B has a Laplace expansion in complementary minors; for
instance, the expansion along the kth row (and column!) is given by
Pf B = L (±) kl bkt Pf Bk' l',
l*k
where Bk' I' is the principal minor obtained by deleting from B the kth row
and column, and also the lth row and column; and ( ± ht is the sign of the
permutation that reorders (1, ... , 2m) as (k, l, 1, ... , k, ... , f, ... , 2m). In the
present context, this leads to the formula
Pf(SK) = L. (±ht (uk I Sut) 1 Pf SK\!k,l),
LEK
provided that k E K and ( ±) kl is the sign of the shuffle K .... ( k, l, K \ { k, l}).
Consequently, we obtain, for any k E K,
Pf(SK) UK = L. (uk I Sut) 1 Pf(SK\{k,tJ) Uk "Ut "uK\{k,lJ·
lEK\k
(Notice that Uk "Ut "UK\{k,ll = (±ht UK, on reinserting the indices k, l in
their proper places.) Now we get
a(v)fs = a(uo) L,Pf(SK)UK = a(uo) L Pf(Souduo /\ur,
K r~o

because the summands for which UK contains no u 0 give zero contribution.


Therefore
a(v)fs = a(uo) L. L. (uo I Suk) 1 Pf(Sr\k) uo "Uk "ur\k (6.5)
UOker
=- L. L. Pf(SM) (uk I Suo) 1 uk "uM=- L. Pf(SM)Suo "uM.
k>OM~O M~O

Next, we remark that


(a(v) + at(Sv))fs = L. Pf(SN)Suo "UN,
N30

as a direct consequence of (6.5). Thus


(a(v) + at (Sv)) fs = L Pf(Soud Suo 1\ uo /\ur
uo
= L. L. (uo I Suk) 1 Pf(Sr\k) Suo 1\ uo "Uk "ur\k
r~Oker

= L. (uo I Suk) 1 L_Pf(SK) Uk "Suo "uo "uK


k>O K
= L Pf(SK) Suo 1\ Suo /\ uo 1\ UK = 0.
K
The several interchanges of summations are justified since S is Hilbert-
Schmidt, so all these series converge in the norm of ]'1 (V). D
220 6. The Spin Representation

Lemma 6.6. Provided that TE Sk(V) and that R is linear and skewadjoint,
every Gaussian fs lies in the domain of each of the operatorsatT a t, a Ta,
atRa, which are therefore densely defined, and their evaluations are as
follows:

atTat Us) = 2HT 1\ fs,


aTa(fs) = 2HsTs Ais -Tr(ST)fs,
at Ra(fs) = H[R,S] 1\ fs. (6.6)
Proof. The first identity is easy; indeed, the creation operators a t ( v) act as
exterior multiplications on the exterior algebra i\"V1 , so that atTat(~) =
2HT 1\ ~ for any ~ E A"V1 and consequently for any ~in :Jj(V).
Next, if R is skewadjoint and S is antiskew, then for u, v E V,
(Rv I Su) 1 - (Ru I Sv) 1 = -(u I SRv) 1 + (u I RSv) 1 = (u I [R,S]v) 1 •
An application of Lemma 6.5 shows that at Ra(j5 ) equals

.l:<uk I Rui) 1 a}<uk)a1 (uL)fs = - .l:<uk I Rui) 1 a}(uk)aj(Sui)fs


k,l k,l
= .l:<Ruk I UI) 1 Uk 1\SUI 1\fs = L (Ruk I UI) 1 (Ur I Sui) 1 Uk 1\ Ur 1\fs
k,l k,l,r
=- L (Ruk I UI) 1 (UI I Sur) 1 Uk 1\ Ur 1\ fs
k,l,r
=- L (Ruk I Sur)J Uk 1\ Ur 1\ fs
k,r
= L ~ ((Rur I Suk) 1 - (Ruk I Sur) 1 ) Uk 1\ Ur 1\ fs
k,r
= L ~(Uk I [R,S]ur) 1 Uk 1\ Ur 1\fs = H[R,Sl 1\fs.
k,r
lfboth T and S lie in Sk(V), then aTa(fs) equals
.l:<Tull Uk) 1 aJ(uL)aJ(Uk)fs = .l:<Tuk I UI) 1 aJ(UI)aj(Suk)fs
k,l k,l
=- L (Tuk I u1) 1 a}<Suk)aJ(UI) fs + L (Tuk I UI) 1 (UI I Suk) 1 fs
k,l k,l
= .l:<Tuk I UI) 1 a}(Suk)aj(SuL)fs + .l:<Tuk I Suk) 1 fs
k,l k
=- L (up I Suk) 1 (Tuk I Suq) 1 Up 1\ Uq Ais- .l:<uk I STuk) 1 fs
k,p,q k
=- L (TSuq I Uk) 1 (uk I Sup) 1 up 1\ Uq 1\fs -Tr(ST)fs
k,p,q
= L (up I STSuq) 1 Up 1\ Uq 1\ fs- Tr(ST} fs = 2HsTs 1\ fs- Tr(ST} fs.
p,q
6.2 The infinitesimal spin representation revisited 221

Since both S and T are antilinear and Hilbert-Schmidt, their product ST


is a linear traceclass operator on V1 , so that Tr(ST) is weil defined and
finite. D
We now try to combine the constituent parts of (6.6) to get the desired
formula expressing the infinitesimal spin representation on Gaussians. But
here a difficulty arises, due to the last term in the expansion (5.38) for {l(B),
namely the trace of the linear part of B. If V1 is an infinite-dimensional
Hilbert space, it may very weil happen that B+ is not a traceclass opera-
tor. Indeed, while the antiskew operator T in the previous Iemma must be
Hilbert-Schmidt in order that HT exist, no such handicap applies to the
skewadjoint linear operator R, because H[R,Sl makes sense if only the in-
dex S of the Gaussian fs is Hilbert-Schmidt. We do not want then to assume
that R be traceclass. The only way out is to discard the trace term in (5.38)
and redefine {1 without it.
Definition 6.3. Let (V, g) be an infinite-dimensional real Hilbert space, on
which an orthogonal complex structure 1 has been chosen. The restricted
orthogonal Lie algebra o1 (V) consists of those skewsymmetric bounded
real-linear operators B on V for which [], B] is Hilbert-Schmidt, or equiv-
alently, the antilinear part ß_ = ~ (B + 1B]) is Hilbert-Schmidt. (We shail
make acquaintance in Section 6.3 with the Banach-Ue group corresponding
to this Banach-Ue algebra.) The infinitesimal spin representation of o1 (V)
is the correspondence B - {l(B), given by (5.38) with the trace of B+ sup-
pressed:
(6.7)
It foilows from Lemma 6.6 that {l(B)fs makes sense for any B E OJ(V)
and SE Sk(V). Indeed, if we put 1/J(B, S) := ß_- SB_S + [B+, S], then
{l(B)fs = H!JI(B,S) 1\ fs + ~ Tr(Sß_) fs. (6.8)
Notice also that the "vacuum expectation functional" (0 I{l(B)O) 1 vanishes.
The foilowing important exercise is now straightforward, and shows how
the basic commutation relations extend to the infinite-dimensional case
(of course, if V is finite-dimensional, the commutator [{l(B), TTJ(V)] is un-
affected by the suppression of the constant - ~ Tr B+ ).
Exercise 6.5. Prove that

and also that


{l(B)TTj(V)(TTJ(U)fs) = TTj(V){l(B)(TTj(U)js) + TTj(BV)TTj(U)fs,
for B E o(V), S E Sk(V) and u, v E V. Deduce that the commutation
relation
[{l(B),TTJ(V)] = TTJ(Bv)
222 6. The Spin Representation

holds on the dense domain in Fock space spanned by all Gaussians fs and
allTTJ(u)fs. 0

Informally speaking, the redefinition of iJ(B) as (6.7) rather than (5.38)


is accomplished by "subtracting an infinite constant" ~ Tr B+. The price we
pay for this manoeuvre isthat B - iJ(B) is no Ionger a homomorphism of
Lie algebras; there now appears an anomalaus commutator or Schwinger
term.
Exercise 6.6. Show that

Try to do the same computation for bosons, instead of fermions. o


Theorem 6.7. If A,B E OJ(V), then

[iJ(A), iJ(B)]- iJ([A,B]) = -~ Tr[A-,B-]. (6.9)

Proof. Recall that [A,B]+ = [A+,B+] + [A_,ß_] and [A,ß]_ = [A+,B-] +


[A_, B+]. The simplest way to see the result is to reconsider the finite-
dimensional case, with the new definition of iJ; the redefinition does not
affect [iJ(A),iJ(B)], so the difference is -~Tr([A,B]+) = -~Tr([A_,ß_]),
We remark that although [A_, ß_] is a traceclass commutator, its trace need
not vanish, because it is the commutator of antilinear operators.
Fora direct proof, we may recall Exercise 5.17, that yielded the identities
[atA+a,atB+a] = at[A+,B+]a and [atA+a,atB_at] = at[A+,B-]at. By
taking adjoints, the second equality yields the formula [aA_a, at B+a] =
a[A-,B+]a also. Thus the only contributions to the left hand side of (6.7)
come from terms involving both A- and B-. Explicitly, we compute

[iJ(A),iJ(B)]- iJ([A,B])
= -![atA_at,aB_a]- ![aA_a,atß_at]- at[A-,B-]a

= 41 .L ((uk I A_uz) 1 (B_uq I up) 1 - (uk I ß_uz) 1 (A_uq I up) 1 )


k,l,p,q
x [a!af,apaq]- _L<ur I [A_,ß_]us) 1 atas.
r,s

Using Exercise 6.6 and the antiskewness of A_ and ß_, so that, for example,
(Uk I A_uz) 1 = -(uzl A-uk) 1 , the quadratic terms in the at and as cancel
out and we are left with

a purely imaginary number; this is an absolutely convergent sum since A-


and ß_ are Hilbert-Schmidt, but need not cancel to zero. D
6.2 The infinitesimal spin representation revisited 223

~ The simplest interpretation of this result is in terms of Lie algebra coho-


mology [250). We can construct a !-dimensional central extension of o1 (V)
by i~. with commutator

[(A,ir),(B,is)J := ([A,B], a(A,B))

where

(6.10)

To see that indeed this defines a Lie algebra, we need to verify its Ja-
cobi identity, i.e., that a is a 2-cocycle for the real Lie algebra cohomology
of o1 (V). The coboundary operator for this cohomology is

8a(A,B,C) := a([A,B],C)+a([B,C],A)+a([C,A],B) = L a([A,B],C),


cycl!c

where Lcyclic denotes a sumover the three cyclic permutations of (A, B, C).
That 8 a = 0 is easily checked:

-2 8a(A, B, C)

=Tr( L [[A,B]_,c]) =Tr( L. [[A-,B+],C]- [[B-,A+],c])


cycl!c cychc

= Tr( L_ [[A-,B+], C] + [[A+, C],B_])


cychc

= Tr( L_ [[A-,B+], C] + [[B+,A-l. C]) = 0.


cychc

Thus, a acts as a generator for the cohomology space H 2 (o1 (V), ~) = ~­


When V is finite-dimensional, it simplifies to

which isatrivial 2-cocycle. In other words, the Lie algebra formed from a
is then isomorphic to the trivial extension. In theinfinite-dimensional case,
this is no Ionger true, since [A+, B+] is in general not traceclass. In sum-
mary, the Schwinger term comes directly from the obstruction to linearity
of the infinitesimal spin representation.
Proposition 6.8. The Schwinger terms can also be rewritten as
i
a(A,B) = BTr(J[],A][],B]), (6.11)

where the trace is taken on the Hilbert space vc (and], A, B are the amplified
operators on vc ).
224 6. The Spin Representation

Proof. Note thatTr(J[],A][],B]) = Tr([J,B]J[J,A]) = -Tr(J[],B][],A])


and also that [J,A] = 2]A_. Therefore, the right hand side of (6.11) is
unchanged under skewsymmetrization: -~Tr(A_Jß_) = tTr(J[A_,ß_]).
1t remains only to observe that
Tr(J[A_,ß_]) = Tr(iPJ[A-,B-]P1 ) + Tr(-iP_ 1 [A_,ß_]P_ 1 )
= 2iTr(PJA-P-JB-PJ- P1ß_p_1 A_P1 ) = -4i a(A,B).
For the final equality, we have used the unitary isomorphism P1 : v1 - w1
to compute the trace of [A_, ß_] over v1 . D
The form (6.11) of the Schwingerterm will become pertinent when we
first deal with cyclic cocycles, in Section 8.4.

6.3 The Shale-Stinespring theorem


Two problems concerning Fock representations turn out to be closely re-
lated, so much so, in fact, that the solution to one gives a solution to the
other. Consider, on one hand, the equivalence problem: let ], K be two
different complex structures; we wonder what are the conditions for the
Fock representations rr1 , TTK tobe unitarily equivalent, i.e., that a unitary
T: J"J(V) - J"K(V) exist suchthat TTK(a) = TTTJ(a)T* for a E Cl(V). Then,
there is the implementation problem: for a fixed choice of ], find conditions
on h E 0 (V) suchthat there exists a unitary operator U on ]"1 (V) satisfying
TTJ(Bha) = UTTJ(a)U*.
v
Let K := h]h- 1 . Clearly h: 1 - VK is unitary, and then a unitary isomor-
phism between J"J(V) and J"K(V) is given by Ah. Let v E V and 1;;" E ]"1 (V).
Then
Ah(aj(v)l;;") = Ah(v A 1;;") = hv A Ahl;;" = a"k(hv)(Ah(l;;")).
Therefore,
a"k(hv) = Ahaj(v) (Ah)*.
By the same token, aK(hv) = Ah aJ(V) (Ah)*, so
TTK(hv) = Ah TTJ(V) (Ah)*. (6.12)

From the C* version of Proposition 5.1 (namely, Exercise 6.3) we conclude


that
TTK((}ha) = Ah TTJ(a) (Ah)*,
for any a E Cl(V). lt is immediate that (JK 0 eh = Uj, for the Fock states of
the algebra.
Moreover, if T: J"J(V) - J"K(V) is a unitarymap, intertwining TTK and TTJ,
then we consider the unitary operator U := T* (Ah) on JJ (V), for which
UTTJ(a)U* = T*(Ah) TTJ(a) (Ah)*T = T*TTK((}ha)T = TTJ((}ha).
6.3 The Shale-Stinespring theorem 225

Therefore, a solution to the equivalence problern gives a solution to the im-


plementation problem. Reciprocally, if U solves the implementation prob-
lern, then T := (Ah)U* is a solution to the equivalence problem.
We concentrate now on the implementation problem. lt is relatively easy
to find sufficient conditions for h E 0 (V) to be implementable. Note, to
begin with, that if h E U1 (V) c O(V), then it is certainly implementable,
as K := h]h- 1 = ], and so, in view of (6.12), Ah will do.
Consider now a general element h E O(V). We decompose it into its
linear and antilinear parts: h = Ph + qh where Ph. qh are given by
Ph := !(h- ]h]), qh := ! (h + ]h]).
We find that Ph-1 = !(h- 1 - ]h- 1 ]) = !(ht- ]ht]) = p~ and likewise
qh-1 = q~.
Taking linear and antilinear parts of hh -l = h -l h = 1, we get the rela-
tions
PhP~ + qhq~ = P~Ph + q~qh = 1,
phq~ = -qhp~. p~qh = -q~ph. (6.13)
Definition 6.4. The restricted orthogonal group 0 1 (V) is the subgroup of
O(V) consisting of those h for which (], h] is a Hilbert-Schmidt operator,
or equivalently, for which qh is Hilbert-Schmidt.
The group 0 1 (V) includes the union O(oo) of finite-dimensional ortho-
gonal subgroups, defined just like U(oo) in Chapter 3. Similarly, U(oo) c
UJ(V); in fact, the seminal paper by Jordan and Wigner [265] on the canon-
ical anticommutation relationswas framed in terms of U(2"").
The rest of this section involves the structure of 0 1 (V}. This is a topo-
logical group, when provided with one of several natural topologies: for
instance, we may decree that hi - h if and only if qh, -%in the Hilbert-
Schmidt norm and Ph, - Ph in the strong operator topology, or alterna-
tively in the operator norm. The first option is natural if one considers
unbounded generators for curves on OJ(V); the second alternative makes
0 1 (V) aBanach-Ue group, whose Banach-Ue algebrao1 (V) we have already
met. Verification of the continuity of the operations is routine, on using the
formulae
t t t t
Phk-1 = PhPk + qhqk, qhk-1 = phqk + qhpk.
Alreadyin the finite-dimensional case, when OJ(V) = O(V), the short exact
sequences of Section 5.2 do not tell the whole story. We already mentioned
that 0 (V) has two connected pieces, that can be distinguished, for instance,
by the value of the determinant; SO(V) is the connected component of the
identity, and, by looking back at Section 5.2 and not excluding now the
product of odd numbers of unitaries, we obtain the covering group Pin(V)
of O(V) and the short exact sequence
1 - {±1}- Pin(V)- O(V)- 1.
226 6. The Spin Representation

It turnsout that (in either of the topologies considered) 0 1 (V) also has two
connected components! In order to show this, in the spirit of Chapter 4,
all we need is to define a two-valued index map on 0 1 (V), i.e., a homo-
morphism i of 0 1 (V) onto 71. 2 suchthat two elements h, k of 0 1 (V) can be
continuously connected if and only if i(h) = i(k). This weshall do later.
Definition 6.5. When h E OJ(V) and Phis invertible, we put Th := qhph 1 •
This is antilinear and Hilbert-Schmidt, and is skewsymmetric on account
of (6.13); thus Th E Sk(V). lt also follows from (6.13) that Ph (1 - T~ )ph =
1. This indicates that the pair of operators (ph, Th) parametrizes the set
SOj(V) := { h E 0 1 (V): ph 1 exists }. (This is not a subgroup of 0 1 (V), but
neighbours of the identity in 0 1 (V) do belang to SOj(V).)
Proposition 6.9. LethE SOj(V). Then lh is implementable.

Proof. According to Corollary 6.2, it is enough to find a unit vector 4> E


:fJ(V) suchthat TTJ(ßh(v + i]v))4> = 0 for all v E V -that is, a vacuum
vector for TTJ 0 eh. From (5.33) we obtain

~TTJ(eh(v + i]v)) = ~TTJ(hv) + ~rr1 (h]v) = a(phv) + at(qhv). (6.14)

It is then clear that the element we are looking for is a multiple of frh, as
the equation

(6.15)

follows from Lemma 6.5, on replacing v by Ph v and S by Th. We can normal-


ize it by taking 4> := det- 114 (1- T~) frh, on account of Proposition 6.3. 0

We now consider the case in which Phis not invertible. This demands a
deeper study of the orthogonal groups. If K = h]h- 1 is the new complex
structure, then ] - K has operator norm at most 2; but if IIJ - Kll < 2,
then the operator ~(1- ]K) = 1 + ~1(]- K) would be invertible, since
II~J(J- K)ll < 1, and hence Ph = ~(1- ]K)h would be invertible. On
the other hand, it is clear that noninvertibility of Ph will occur even in the
finite-dimensional cases. We restriet the set J(V) of complex structures
by introducing J1 (V) := { K E J(V) :] - K is Hilbert-Schmidt }. Also, we
call "restricted polarizations" those W for which]- Jw is Hilbert-Schmidt;
theseform the respective orbits of] and of w1 under OJ(V). Clearly U1 (V)
is again the isotropy subgroup of 1 or W1 under the respective actions
of 0 1 (V).
Let W = hW1 be a restricted polarization, and consider G := ~(1- Jw]).
This is a Fredholm operator on the complex Hilbert space vc, since ct G -
1 = cct - 1 = ~ Uw - ]) 2 is compact, indeed traceclass (recall Proposi-
tion 4.1). Moreover,

ker G = ker ct = { z E vc : 1z = - ]wz }.


6.3 The Shale-Stinespring theorem 227

In particular, its index is dim(ker G) - dim(ker ct) = 0. Since G = hph, it


follows that kerph = kerG and dimc(kerph) = dim(kerGt), so that Phis
also a Fredholm operator of index zero.
The argument establishes, inter alia, that Pi/ exists if and only if] + Jw
is invertible.
Exercise 6.7. Prove that, if W = hW1 and Phis invertible, then

h = U-Jw)U+Jw)- 1 • <>

A few remarks on the structure of Ph and qh are in order. We find that

in view of (5.36). lt follows that

(6.16)

Reciprocally, if (6.16) is an orthogonal decomposition, then (6.13) holds,


and thus h is orthogonal. We can decompose VJ as ker Ph E9 (ker Ph).L and
as ker Ph E9 (ker Ph) .L, where .L can be understood with respect to either g
or ( · I ·) 1 , as ker Ph is a complex subspace. Moreover, im Ph = (ker Ph) .L, so
Ph can be written in the form

(6.17)

as an operator from ker Ph E9 (ker Ph) .L to ker Ph E9 (ker Ph) .L. We recall that
dimker Ph = dimker Phis finite. On the other hand, on again using Phqh =
-qhPh. we obtain %(kerph) ~ kerph. Ukewise, qh(kerph) ~ kerph. lt
follows that h (ker Ph) ~ ker Ph, and then h ( (ker Ph) .L) ~ (ker Ph) .L by
orthogonality. We conclude that %(kerph).L ~ (kerph).L, i.e., qh also has
the block form

Moreover, this B is a (finite-dimensional) antiisometry, since qh qh v = v


for v E ker Ph. and qhqh u = u for u E ker Ph; C is obviously still Hilbert-
Schmidt.
Now consider h E ÜJ(V) with dim(kerph) = dim(kerph) = n > 0. Let
{ u 1, ... , Un} be an orthonormal basis (in V1) for the subspace ker Ph. Let Yk
be the reflection of (V,g) for which Yk(Uk) = -]uk. rk(]uk) = -uk while
rk(v) = v if g(v, Uk) = g(v,]uk) = 0. This operator has the identity
restricted to the subspace orthogonal to Uk as its linear part, and it is
antilinearon the spanofuk and]uk. Let r := r1 ... Yn. The operatorr is also
in block form, its lower right corner being just the identity on (ker Ph) .L,
228 6. The Spin Representation

and rh E SOj(V). We claim that t;;" /\ frrh• where t;;" is a nonzero element of
An (ker PU. is proportional to the vacuum vector for TTJ 0 eh. i.e.,
(aJ(PhV) + aj(%V))[t;;" /\ fr,.h] = 0. (6.18)
One can take t;;" = u1 /\ · · · /\ Un, for the sake of the argument. It is enough
to verify (6.18) for v E kerph and v E (kerph)\ respectively. In the first
case qhv E kerp~. therefore aj(%v)t;;" = 0, and (6.18) holds. Next let
v E (ker Ph).t.. As in the proof of Proposition 5.14,

aJ(PhV)[t;;" /\ fr,-h] = aJ(PhV)t;;" /\ frrh + (-l)nt;;" /\ aJ(PhV)frrh"


But aJ(PhV)t;;" = 0, since PhV E (kerp~).~.. On the other hand,

aj(%V)[t;;" Ajyrh] = (-l)nt;;" /\ aj(qhv)frrh"


Now, if v E (ker p~).t., then PrhV = PhV and qrhV = %V. This reduces
(6.18) to (6.15) with h replaced by rh, and our claim is proved. Wehave
arrived at the following result.
Proposition 6.10. LethE OJ(V). Then eh is implementable. E3

Note that in this case the operator A in (6.17) is an isomorphism. Also,


we see that the "out-vacuum" vector u 1 /\ · · · /\ Un /\ !Trh lies in :FJ
(V) or
:Jj (V) according as n is even or odd .
.,. The foregoing discussion makes it highly plausible that h E OJ(V) is
also a necessary condition for eh to be implementable. Before proving this
second (harder) half of the Shale-Stinespring theorem, however, we want
to exhibit the implementors -by describing their action on Gaussians in a
fully explicit form.lt is clear that the implementor U(h) of a given Bogoliu-
bov transformation is uniquely defined up to a phase factor. It is also clear
that it generalizes to the infinite-dimensional case the (global) spin repre-
sentation, when the latter is thought of as of a projective representation of
the orthogonal group; thus we shall write Jl(h) for it. We concentrate on
the case h E SOj(V), which is the more pertinent one for the applications
in Part IV.
Let us abbreviate fh := Th-1· The antilinear part of the equation 1 =
hh- 1 = (1 + Th)Ph(1 + Th)P~ gives 0 = ThPh + PhTh, so that

(6.19)

It is convenient to write Pht := (p~)- 1 = (ph 1 )t for the contragredient


operator to Ph when h E SOj(V). Then Ph + qhTh = (PhP~ + qhq~)pht =
Pht· Whenever h- 1 ,k,hk E SOj(V), this gives the product formula

Thk = qhkPhl = (qh + PhTd(Ph + qhTk)- 1 = (qh + PhTk)(1- ThTk)- 1 ph 1


= qhph 1 + (% + PhTk- qh(l- ThTk))(l- ThTk)- 1 ph 1
(6.20a)
6.3 The Shale-Stinespring theorem 229

The first step toward the construction of the spin representation is to


note that there is a local action of SOj(V) on Sk(V), given by

h ·S := (qh + PhS)(ph + %5)- 1 .

(lf Phis invertible, so is Ph +%S, for S small enough in the Hilbert-Schmidt


norm.) The relation

follows from (6.13) since st = -S. lt follows that h · S is also skew, and it
is clearly antilinear and Hilbert-Schmidt; thus h · S E Sk(V). 1t is readily
checked that hk · S = h · (k · S) whenever k ·Sand hk · S are defined; the
group action is local since they may be undefined in particular cases, this
being of course a signal that Jl is not equivalent to a true representation.
From (6.20a) with Tk = S, we obtain the useful alternative form

(6.20b)

Fora given h E SOj(V), different from the identity, those S for which
Ph + %S is invertible form an open neighbourhood of zero in Sk(V), so
{ fs : h · S exists} spans a dense subset of J"J (V). Thus we may define a
unitary operator on J"J (V) by the prescription

where eh isapositive number and cl>h (S) is a complex number, chosen so


as to make Jl(h) unitary. More precisely: first let S = 0. Then the definition
becomes
Jl(h)O. := Ch c/>h(O) frh,

so unitarity is met by eh:= llhh 11- 1 = deC 114 (1- T~), c/>h(O) = 1. Notice
that

(0. I Jl(h)D.) = Ch. (6.21)

Lemma 6.11. A suitable choice for the scalar factor is

(6.22)

Proof. To establish unitarity, we require that

(6.23)

The right hand side is det 112 (1- TS), by (5.56b). The left hand side, also
from (5.56b), equals

cPh(S) c/>h(T) det 112 (php~) det 112 (1- (h · T)(h · S)),
230 60 The Spin Representation

since c~ = dec 112 (1- T~) = det 112 (PhPUo Using (6o19) and (6o2Gb), we find
that

Ph (1- (h T)(h S))ph


0 0

= Ph (1- T~)ph- T(1- fhn- 1Pi/Thph- pftThp;;ts(l- fhs)- 1


- T(l- fhn-1p;;1p;;ts(l- fhs)-1
= 1 + (1- TTh)- 1Tfh + fhs(l- fhs)- 1
1 1
TTh)- T(1- Th)S(1- Ths)-
A A2 A
- (1-
= (1- Tfh)- 1(1- TS)(1- fhs)- 1,
and therefore the left hand side of (6023) reduces to

The formula (6022) simplifies this expression to <fs I fr)o 0

Finally, on the odd subspace :Jj (V) the spin representation is given by
the prescription

J.1(h)(TTJ(V)fs) := TTJ(hv) J.1(h)fs,

for S E Sk(V) suchthat h 0S is definedo For the general case, when Ph is


not invertible, we refer to [213]0
~ We seek to factorize J.1 in a convenient mannero For h E SOj(V), we
introduce the following Operators s1. Sz, S3:

SI(h) = exp(~atThat),
Sz(h) = :exp(at(p;;t -l)a):,
1
S3(h) = exp( 2 aTha)o (6o24)
A

The colons in the second operator denote a Wick-ordered or normally or-


dered product: each monomial aL a 11 oooaLa 1r obtained by exponentiating
at (p;;t -l)a is replaced by aL 000aLa 1r 000a 1,, with all creation operators
on the lefto lt will be enough to compute the effect of these operators on
Gaussianso
Lemma 6.12. S1 (h) is a bounded operator on JJ (V), with S1 (h)fR = fn+R
for any R E Sk(V)o

Proof. We already noted, in the proof of Lemma 606, that (atThat)m; =


zmH;hm 1\; for any; E :J'J(V)o Moreover, (5.40) gives the norm estimate
IIHihm /\;II ~ IIHTh um 11;11 ~2-m IIThllr u;u, with llollz denoting the Hilbert-
Schmidt normo The series exp( ~atThat) := I:=o(2mm!)- 1(atThat)m
6.3 The Shale-Stinespring theorem 231

convergcs in the bounded operator norm on :Jj(V) and satisfies the es-
timate II exp( ~atThat) II ~ exp( ~ IlTh ll2 ).
We now see that S1(h)~ = hh A ~ for any ~ E :JJ(V). In particular, for
~ = fR, we obtain SI(h)fR = hh A fR = hh+R· 0

Lemma 6.13. lfR ESk(V) and (1- RTh) is invertible, then


1/2
S3(h)fR = det (1- RTh)fRo-fhRJ+
A

Proof. Notice that S3 (h) restricts to the even subspace :Jj (V), which is
generated by the Gaussians fs. In view of the previous lemma, it is enough
to compute
(fs I S3(h)fR} = (S3(h)tfs I fR} = (SI(h- 1)fs I fR}
= Us+fh I fR} = det 112 (1- R(S + Th))
= det 112 (1- RTh) det 112 (1- (1- RTh}- 1 RS),

because det 112 (1 - (1- RTh)- 1RS) = det 112 (1- R(l - fhR)- 1S), which
equals (fs I fRo-fhRJ-1 }. 0

Lemma 6.14. lfR E Sk(V), then S2(h)fR = fph'Rphi.


Proof. First, let us note that if Ais any bounded linear operator on v1 , then
ARAt E Sk(V), and HARA' = Li,J(ui I Ru 1 }Aui A AuJ by Exercise 5.19.
Therefore,
fARA' = L Pf(RK)AUk 1 1\ • • • 1\ AUk 2 m;
IKI even

Secondly, we must check that, if Km:= {1, ... , 2m}, then


:exp(atca):uKm = (1 + C)u1 A • • · A (1 + C)uzm· (6.26)

For then, :exp(atca):uK = (1 + C)Uk 1 A • • • A (1 + C)uk 2 m by a change of


orthonormal basis. Tagether with (6.25), the substitution C = ph_t- 1 gives
the desired result.
To verify (6.26), note that the left hand side is a finite series, since the
terms azi with z1 > 2m give no contribution. Thus
2m
n(Ukj I Cuzj}at ... atazn ... azi UKm
n
:exp(atca):uKm = L ~! L
n=O kJ ... kn J=l
IJ ... In

= L I~I .1 11L,Km\Lat(Cuz 1 ) ... at(Cuz")UKm\L


Ls:;Km
= L f1 1\ • • • 1\ fzm,
Ls:;Km
where fJ = Cu J if j E L, fJ = u J otherwise. But the latter sum is just an
expansion of the right hand side of (6.26). 0
232 6. The Spin Representation

It is worth noting that since

fp-;;tRp-;;1 = I Pf((phtRph_ 1 )K) uK = I det((ph_ 1 )K) Pf(RK) UK,


IKI even IKI even

and ldet((ph_ 1 )K)I = det 112 ((ph_t)K(Ph_ 1 )K) ::5 det 112 (ph_tph_ 1 ) = c;;_ 2 , then
S2 extends to .Jj (V) as a bounded operator with norm at most c;;_ 2 .
Now we see that, on applying S1, S2, S3 in turn to JR, the index of the
Gaussian transforms as R .... Th + ph_t R(l- ThR)- 1 ph_ 1 = h · R, and so

(6.27)

whenever ph_ 1 and h · R exist. Thus p(h) = chS 1S 2S 3 holds on .Jj (V) when-
ever h E SOj(V). For the factorization in the general case, when Phis not
invertible, we refer again to [213] .
.,. Now for the necessity result. Webegin with a lemma about creation ope-
rators.
Lemma 6.15. Let Y be a closed complex subspace o(V1 . Then, ifY is infinite-
dimensional, nvEY ker a t (V) = {0}; however, if Y iS finite-dimensional,

n kerat(v) = JJy1. 1\ AdimYy_


VEY

Proof. Since Y is closed, we can assume that an orthonormal basis {uk}


for VJ has been chosen so that some subset { Uk : k E M } is an orthonormal
basis for Y.
Let~ E nvEYkerat(v). Suppose first that dimY = oo; then for each
finite subset K c f':::l, we can find l E M \ K, so that

and thus ~ = 0. On the other hand, if dim Y is finite, this calculation goes
through unless K 2M, in which case we can write UK = ±ur A UM, with L
disjoint fromM. Therefore, ~ * 0 only if ~ = 17 A uM with 1J E ;:1 P; and it
is clear that every such 1J A UM lies in ker a t (Uk) for each k E M. D
Theorem 6.16 (Shale-Stinespring). Let h be an orthogonal operator on V.
The corresponding Bogoliubov automorphism (}h o(Cl(V) is unitarily imple-
mentable in the Fock representation TTJ if and only if qh is Hilbert-Schmidt.

Proof. We have already seen that (}h is implementable if qh is Hilbert-


Schrnidt; only the necessity of this condition remains to be proven. This
requires a more thorough study of the full orthogonal group O(V). We
follow [3 78] closely here. Our remarks on the structure of Ph and qh for
h E 0 1 (V) remain valid for h E O(V), except that now kerph need not be
finite-dimensional, and im Ph is no Ionger necessarily closed, although its
closure is still (ker p~) J..
6.3 The Shale-Stinespring theorem 233

When Phis selfadjoint, so that ker Ph = ker Ph· then Ph· qh and h pre-
serve the decomposition ker Ph E9 (ker Ph) j_. In the polar decomposition
Ph = wh IPhI of a general Ph. the partial isometry Wh maps (ker Ph) j_ onto
(ker Ph)j_ itself, and we can construct a unitary operator u by adding to it
a partial isometry w' mapping ker Ph onto ker Ph· The relation Ph = u IPhI
holds, so that h = u IPhI+% with u unitary. Now it will be enough to exam-
ine the implementability condition for orthogonal operators whose linear
part is selfadjoint. For suppose h E O(V). Then lh = lhu•eu. and eu is
automatically implementable, while the operator h u * has selfadjoint lin-
ear part ulphlu* and antilinear part qhu*, which will be Hilbert-Schrnidt
if and only if qh is Hilbert-Schrnidt.
Let us then assume that Ph is selfadjoint. Let U be a unitary operator
on F1 (V) implementing eh, and let <I> := UD. be the out-vacuum vector.
Because of (6.14), the vacuum condition on <I> implies that at(qhv)<l> = 0
for all v E ker Ph. and then, since qh is antiunitary on ker Ph. it follows
that at (v )<I> = 0 for all v E ker Ph· Now, ker Ph must be finite-dimensional,
since otherwise Lemma 6.15 would imply <I> = 0.
Let m = dim ker Ph and write <I> = ~ A 1,; with 1,; E Am ker Ph and ~ E
_r1 (ker Ph)j_. For v E (ker Ph)j_ the vacuum condition can be written as
0 = (a(phv) + at(%V))(~ A 1,;) = (a(phv) + at(qhv))~ A 1,;.
Denote by h' the restriction of h to (ker Ph)j_; we conclude that ~ is a cyclic
vacuum vector for the representation TrJ o eh' of G((ker Ph)j_ ). The range
of Ph is a dense subspace of (ker Ph) j_.
The Fock space .fJ(ker Ph)j_ is the Hilbert-space direct sum of the sub-
spaces Ak (ker Ph)j_, so we can write ~ =: .Lk ~k with respect to this decom-
position. The odd components vanish, since the vacuum condition partic-
ularizes to a(phv)~ 1 = 0 for all v E (kerph)j_, implying ~ 1 = 0, and then
again to
(6.28)
for n = 1, 3, ... , forcing ~ 3 = ~ 5 = · · · = 0 in turn. Of course, the same
argument would prove that <I>= 0 if ~o were zero; thus ~o * 0. We can even
suppose that ~ 0 is the vacuum vector of (ker Ph)j_, after dividing ~ and
multiplying 1,; by a suitable constant, if necessary. The case n = 0 of (6.28)
then gives
a(phv)~2 + qhv = 0.

This can be restated as saying that (phv Au I ~2) + (u I %V) 1 = 0 for


all u E (kerph)j_; therefore on (kerph)j_ the relation qh = 2T"E; 2 Ph holds,
where T"E; 2 E Sk( (ker Ph)j_) is associated to the bivector ~2 by (5.39). Now T"E; 2
is Hilbert-Schrnidt, therefore so is the restriction of% to (ker Ph) j_. Also, its
restriction to the finite-dimensional ker Ph is trivially Hilbert-Schrnidt. D
Corollary 6.17. Let 1, K be two orthogonal complex structures. Then rr1 and
TrK are unitarily equivalent if and only if1- K is Hilbert-Schmidt.
234 6. The Spin Representation

Proof. Choose h E O(V) so that K = h]h- 1 • Both Fock representations


will be equivalent if and only if ]h] + h is Hilbert-Schmidt, by the previous
theorem. Equivalently, (Jh] + h)]h- 1 = K-] must be Hilbert-Schmidt. D

Wehave actually proved that 0 1 (V) acts transitively an the set J1 (V) of
restricted complex structures that differ from] by a Hilbert-Schmidt ope-
rator. We can remark here that ÜJ(V) - JJ(V) is then a principal bundle,
with contractible fibre U1 (V). Therefore, Statements about the topology of
the base translate into statements about the topology of the bundle, and
reciprocally .
.,. Weshall now produce the promised index map an 0 1 (V). Denote by x1
the grading operator on :fJ(V). It is clear that if Jl(h) implements eh on
:fJ(V), then so does XJJl(h)XJ· Therefore, since x] = 1, either XJJl(h)XJ =
Jl(h) or XJJ1(h)x1 = -Jl(h). To decide whether Jl(h) is even or odd in this
sense, it is enough to look at the transformed Fock vacuum, which is known
tobe of the form u1 1\ · • · 1\ Un "fT, where n is the dimension of ker ph:
the parity of Jl(h) must be the parity of n. But this parity is a property of
the (restricted) orthogonal transformation itself! The argument, moreover,
shows that the sign map i: ÜJ(V) - 7Lz given by XJJl(h)XJ =: i(h)Jl(h)
is a homomorphism of groups in the abstract sense. We shall denote by
SOJ(V) the preimage of +1 by i, justifying some previous notation.
Exercise 6.8. If K = h]h- 1 , with h E 0 1 (V), then i(h) = +1 or i(h) = -1
if and only if ~ dimker(J + K) is respectively even or odd. 0

Theorem 6.18. The sign homomorphism i: 0 1 (V)- 7L 2 is an index map.

Proof. The only remaining issue is the continuity of the sign map i. Con-
sider an element h E 0 1 (V). We must prove that if dimkerph jumps, then
it does so in steps of even complex dimensions. We argue, for simplic-
ity, in the Banach-lie topology of ÜJ(V). Because the restriction of Ph to
(kerph).L is an isomorphism, the spectrum of PhPh excludes an interval
(0, E) for some E > 0. Choose 8 with 0 < 8 < ~E; and then choose a neigh-
bourhood N of h such that, for all k E N, sp(PkPk) excludes (8, E- 8) and
moreover IIPkPk- PhPhll < 8. Let P be the orthogonal projector of V1 with
range ker Ph. and let TI be the spectral projector of PkPk for the interval
[0, 8].
Und er those conditions, P maps TI (V) bijectively onto ker Ph· Indeed, any
unit vector v E kerph n TI(V).L would satisfy (v I (PkPk- PhPh)v) 1 > 8,
violating the norm bound for elements of N; thus P: TI (V) - ker p h is
surjective. On the other hand, any unit vector u E (ker Ph).L n TI(V) would
satisfy (u I (PhPh- PkPk)v) 1 <::: E- 8 > 8, overshooting the norm bound
again; thus P: TI(V)- kerph is one-to-one.
ClearlyTI(V) 2 kerpk;ifthisisanequality, thendimkerpk = dimkerph.
Otherwise, consider the antiskew operator qipk on (ker Pk).L n TI (V). Since
6.3 The Shale-Stinespring theorem 235

qipk commutes with PiPk by (6.13), it preserves this spectral subspace for
PiPk and, in particular, has trivial kerne! there. If its polar decomposition
is Klqipkl• then K is nonsingular, isometric and antiskew (with respect
to ]), so it is a complex structure anticommuting with ]. Together,] and
K generate a representation of the real Clifford algebra Clo,2 "" !HI acting an
(ker Pk).l n Il(V), which must therefore have even complex dimension. D

Carey and O'Brien, whose treatment [67] we have followed an this matter,
point out that the Atiyah-Singer mod 2 index theorem for real skewadjoint
Fredholm operators [17] can be derived from Theorem 6.18 .
.,.. Let us denote the group of the implementors by CurrJ(V); the notation
stands for "current group": see [8]. An outcome of the discussion in this
section is the existence of nonsplit short exact sequences:

1 - "U"- Curr1 (V)- 0 1 (V) - 1

and

1 - "U"- Currj(V)- S01 (V) -1, (6.29)

where CurrJ (V) denotes the component of the even implementors sit-
ting over S01 (V). The group Currj(V) is the strict analogue of the Spine
group in the infinite-dimensional case. But there is in infinite dimensions
no twofold covering of S01 (V), i.e., no spin group. Were it to exist, the
sequence would be trivial, since S01 (V), just like UJ(V), is known tobe
simply connected [8]. One can manage an exact sequence of the form

1-{±1}- Spin1(V)- S01 (V) - 1

by further restricting to the group S01 (V) of specialorthogonal transfor-


mations h for which h -1 is traceclass and ker(h + 1) is even-dimensional.
Theseare "universally implementable", i.e., SOI(V) c SOJ(V) for all ]. We
shall not dwell an this question any further .
.,.. In the rest of this section, we reconsider the relation between the global
and infinitesimal spin representations, that is not so Straightforward since
neither {1 nor J1 is a representation strictu sensu. lt follows from (6.22),
using (6.19) and (6.20), that

cf>hk(S) =det 112 (1- S(Tk + Pk" 1 Th(1- TkTh)- 1 Pk"t))


= det 112 (Pk"t(l- sfk>Pi- Pk"tSPk" 1 Th(l- Tdh)- 1))
= deC 112 (1- Tdh) det 112 ((1- STk)Pi(l- Tdh)Pk"t- Spk" 1 ThPk"t)
= deC 112 (1 - Tdh) cf>k (S) det 112 (1 - Tdh- Pk"t (1 - STk)- 1Spk" 1 Th)

= der- 112 (1- TkTh) cf>k(S) cf>h(k · S),


236 6. The Spin Representation

whenever h 1 k 1 hk E SOj(V)I provided k·S and hk·S bothexist. This yields


the (group) cocycle equation

Jl(h)Jl(k) = c(h 1 k)Jl(hk)l (6.30)

where the cocycle c(h 1 k) is given by

c(hl k) := ChCkCj:,~ det 112 (1- TkTh) = exp(iargdet 112 (1- TkTh))

= exp(iargdet 112 (pj:, 1 PhkPk" 1 )). (6.31)

The second equality follows from (6.23) on taking S = T = Tk and noting


det 112 (1- TkTh) = cPh(Tk); the third one is clear from Phk = PhPk + qhqk
and (6.19). This confirms that h .... Jl(h) is a projective representation of
the restricted orthogonal group 01 (V). The last expression in (6.31) makes
clear that if the determinants of the Ph operators were individually definedl
the cocycle would be trivial. (For the nonregular casesl which are actually
easierl we refer to [213].)
Exercise 6.9. Check the cocycle properties of c:

c(ll k) = c(hl 1) = 1; c(h k)c(hk g) = c(hl kg)c(k g).


1 1 1

Definition 6.6. The extended orthogonal group 81 (V) is the one-dimen-


sional central extension of 01(V) by -u- whose elements can be written as
(h where h E 0 1 (V) E -u- with group law
1 .\) 1 1 ,\ 1

(h 1 ,\) • (k 1 ,\') = (hk 1 ,\,\' c(h k) )


1 1

so that (h t\Jl(h) isalinear unitary representation of the extended


1 ,\) ....

group. lts Ue algebra ö1 (V) is a one-dimensional central extension of o1 (V)


by i1Rt with commutator
1

[(Al ir)l (BI is)] := ([A~B] 1 a(A B))I


1

where a(AI B) is defined as

d I
-d2
d d I
c(expsAiexptB)- -d2
d c(exptA expsB). 1 (6.32)
t S t=s=O t S t=s=O
Next we produce the promised relation between iJ and Jl. One considers
one-parameter groups with values in 0 1 (V). Let t .... exptB be one such;
then PexptB is invertible for small enough tl and

ddt It=O TexptB = B_.

In particularl ß_ is Hilbert-Schmidt; this is why the antilinear part of o1 (v)


is just Sk(V). Write f.JB (t) := Jl (exp tB) and exp( i~B (t Is)) := CB (t Is) :=
6.3 The Shale-Stinespring theorem 237

c(exp tB, exp 5B). There is an old trick [230] to redefine 1.1 for a one-parame-
ter group so that it becomes a homomorphism. On the dense set of analytic
vectors in i\ • v1 given by states of finitely many particles, we may differen-
tiate the relation
/JB(t)/JB(5) = CB(t,5) /.IB(t + 5)
with respect tot. Writing hB(t) := exp tB for small t, we obtain

:t 't=O/J(exptß)j5 = :t lt=OChB(tl<J>hB(t)(S)fh 8 (t)·5

= :t lt=O<PhB<n(S)fs + ~ :t lt= 0 Pf((hB(t) · S)K)EK

= !Tr(SB_)f5 +H{/J(B,Sl Afs = /1(B)fs,


according to (6.8). Wehave used the invertibility of (PexptB + qexptBS), for
small t and any SE Sk(V). Therefore

/1(B) /JB(5) = ifB(5) /.IB(5) + d/.1:: 5 ),

where fB(s) := (dfdt)it=O~B(t,5). By redefining

/JB{t) := eiOB(t)/JB(t) := exp(i s: fB(T) dT )/JB(t),

we obtain
'(B) ' ( ) - dl.lß(5) (6.33)
/.1 I.IB 5 - d5 .
In conclusion, l.lß(t) = exp(t/1(B)), and thus
/Jß(t}/Jß(5) = l.lß(t + 5).
Note that ÖB(O) = 0 by Exercise 6.9. Hence it It=ol.lß(t) = it It=o/J(exp tB).
The vacuum expectation value (0 I exp(t/1(B))0) 1 is the so-called vacuum
functional.
Exercise 6.10. Using the Campbell-Baker-Hausdorff formula, check that
the Ue algebra cocycle defined by (6.32) satisfies
a(A,B) = [/1(A),/1(B)] -li([A,B]),
for all A,B E o1 (V), just as in (6.9).

~ As well as the anomalous commutators, there are anomalous transfor-


mation laws. The group 01 (V) acts on ö1 (V) by the adjoint action of the
central extension; this action is of the form
(B, ir)- (Ad(h)B, ir + y(h,B)),
where the anomaly y(h, B) E i!Rt depends linearly on B. In fact,
y(h,B) = /.l(h)/1(B)/.l(h)- 1 -/1(Ad(h)B).
238 6. The Spin Representation

Exercise 6.11. Prove that y(h, [A,B]) = a(Ad(h)A,Ad(h)B)- a(A,B). 0


Exercise 6.12. Prove that the anomaly is given by
y(h, B) = - 21 Tr[B_ - ~ ~ ~2
ThB+, Th (1- Th )- ]
1
(6.34)
whenever h E SOj(V) and BE o1 (V). 0

6.4 Charged fields


Both in physics and in noncommutative geometry, most pertinent exam-
ples of the spin representation correspond to charged fields. In that case,
there exist two relevant complex structures, and one of them commutes
with allrelevant operators, including the other. We may as well baptize the
formeras i and regard V as the realification of a complex Hilbert space J{.
The second complex structure is necessarily of the form] = i(E+ - E_) for
suitable projectors E± with E+ + E- = lJ-{. The original complex structure
Q := i is precisely the charge, the generator of "global" gauge transforma-
tions, in physicists' parlance, which are simply multiplications by a phase
factor.
To repeat: J{ and V denote the same space, regarded as complex or
real, respectively. The complex structure defining the Fock space structure
on V is not Q but ]; it should be clear that both i and] are orthogonal
with respect to g, the real part of the "natural" or original inner product
( · I · )Q of J{ (denoted simply by ( · I ·) from now on). In what follows, the
structures J{ and V1 on V should be carefully distinguished. The complex
space J{ is graded by E+- E_, and weshall write J{± := E±J{· Operators
on J{ = J{+ Eil J{- can be presented in block form,

A = (A++ A+-)
A-+ A __ '
all corners being linear for the original complex Hilbert space structure;
the diagonal terms represent the linear part with respect to ] and the off-
diagonal terms the antilinear part. Note the simple formula
(6.3 5)
where, for instance, l/J± := E± 1/J. The overall trace on v1 (rather than on J{)
is therefore given by
A++ A+-) t (6.36)
Tr ( A-+ A__ = Tr A++ + Tr A--·

For instance, since (A+_)t = -A-+ when A E o1 (V), the Schwingerterms


reduce to
a(A, B) = - ~ Tr(A+-B-+ - B+-A-+) - ~ Tr(B-+A+- - A-+B+-)
=- Tr(A+-B-+ - B+-A-+) = -2i~ Tr(A+-B-+ ).
6.4 Charged fields 239

In charged field theory, only orthogonal transformations on V commut-


ing with Q, i.e., unitary with respect to the original complex structure, are
considered. They define a subgroup U(J{) c O(V). In particular, we can
write SE U(J{) = U(J{+ EB J{-) in block form:

s = (s++ S+-)
s_+ s__ , thus Ps = (s++
0 s~_), qs =( 0
S-+
S+-)
0 .

Similarly,

s-1 = st = (sf+ s!+)


S+- s!_ ,
Pst = (sL
0 s~J, t
qs
0
= ( sL
s!+)
0 '

where, in our notation, sL means (S+_)t, not (St)+- = s!+, and so on.
The formulae (6.13), with S instead of h, give the identities

s++sL + s+-sL = s!+s++ + s!+s-+ = E+,


s__ s!_ + s_+s!+ = s!_s__ + sLs+- = E-,
s++s!+ + s+_s!_ = sLs+- + s!+s__ = o,
s__ sL + s_+sL = s!_s_+ + sLs++ = o. (6.37)

It is clear that p5 1 exists if and only if S++ and s__ are invertible (as ope-
rators on J{ + and on J{-, respectively). It is immediate that

Ts = (s os-1 (6.38)
-+ ++
We pause a moment to pander what "antiskew" means in the present con-
text, namely,
T- ( 0 T+-)
0 ' .h
Wlt T-+ = - Tt+-·
- L+

From (6.37), we verify again that Ts, Ts are antiskew. It is also clear that
SE 0 1 (V) if and only if S+- and S-+ are Hilbert-Schmidt.
Example 6.1. Although it is by now well known that the Schwinger term
allows one to compute "from first principles" the anomalaus term of the
Virasoro Ue algebra, this example keeps all its instructive value. Let V be
(the realification of) the complex Hilbert space J{ = L2 (§ 1 )/C of square-
integrable functions of zero mean on the circle with the standard measure
-one may consider instead the real Hilbert space of real square integrable
functions in L2 (§ 1 , ~)/~. but we choose the charged field context. Explic-
itly, the scalar product is given by
240 6. The Spin Representation

where j(n) denotes the nth Fourier coefficient of f. The group Diff+ (§ 1)
of orientation-preserving diffeomorphisms is represented on J{ by

(6.39)

Plainlyl the representation is unitary. The complex structure ] on the real-


ification V of J{ is given by
](eike) := isignkeikO_

(This makes a lot of sense in the real context as ] restricted to real func-
1

tions is the unique rotation-invariant complex structure compatible with


the inner product.)
Lemma 6.19. Each Scf> is implementable in :J1 (V).

Proof. Since Lk>oeikO- Lk>oe-ike = icot~B~] is given by the principal-


value integral

]j(B) = -21 PJTT cot ~(a- B)j(a) da.


1T -TT

We compute the integral kernel K(8 1 8 2 ) of 1 [Scf>~JJ~ setting (/J := cp- 1:

2 ~ J~rr (8((/)(81) - 8)~(/J' (BI) cot ~ (8 2 - 8)

- cot ~(8- 8 1)8((/J(B)- Bz)~t/J'(B)) dB


= cot ~(Bz- (/J(8I))~t/J'(81)- cot ~(</J{Bz)- 81)~</J'(Bz)~
which is continuous except perhaps when 8 1 = </J{Bz). Since cotx- ~
vanishes at x = 0 and is smooth we need only observe that
1

2~ - 2..j(/l(82) - -2__!_!_(~</J'(Bz)) = cp"(Bz)


Bz- t/J(BI) <jJ(Bz)- 81 dBz (</J'(Bz)) 3 i 2
as 8 1 - <P ( 8 2 ) I to conclude that [Scf> I]] has a continuous (indeed smooth)
-hence square-integrable- kernell and thus is Hilbert-Schmidt. D
(This proof is found in [336 §7.2] and is a bit different from that by
1

G. Segal in [425] for the boson case).


1

~ Denote by Ul (J{) the group of ]-restricted unitary operators U(J{) n


OJ(V); whereas 0 J(V) has two connected piecesl it will soon become clear
that the group U1 (Jf) has an infinite nurober of connected piecesl naturally
indexed by 7L. But let us look first at the quantization rule that corresponds
to {11 in the present context. We veer closer to physics notationl where one
wants the second-quantized operators to be selfadjoint rather than skew- 1

adjoint; in factl as only infinitesimally orthogonal transformations on V


6.4 Charged fields 241

commuting with Q (i.e., skewadjoint with respect to the original complex


structure) are considered, we can construe the infinitesimal spin represen-
tation as a map from (restricted) selfadjoint operators on J{ to selfadjoint
operators on F1V; this is accomplished by the rule A ..- -i[l1 (iA), for A
selfadjoint.
Note that -ifJQ (iA) would be simply dA(A) on JQ V, but -i[l1 (iA) is not
dA(A) on :F1V. Recall that

flJ(B) = ajB+aJ + ~(ajB_aj- a 1 B_a 1 ), (6.40)

for any BE OJ(V). In our present context, when BE u(Jf) n OJ(V),

B+ = (B~+ B~-)' ß_ = (B~+ B~-)'


where B++ = -BL, ß __ = -B!_, B+- = -B!+, and B+-• B-+ are Hilbert-
Schmidt.
Notation. In what follows, we distinguish the "particle" and "antiparticle"
sectors by writing b( cp) := a 1 ( <P) and bt (<P) := a j (cp) when cp E J{+,
and d(t/1) := a 1 (t1J), dt(t/J) := a}(t/J) when t/J E J{-. Let {cf>k} and {t/Jk}
denote arbitrary orthonormal bases for J{+ and J{-, respectively; weshall
abbreviate bk := b(cf>k), dk := d(tiJk), and so on.
We now compute in turn each term in (6.40); from (5.37) we find that
l
2(a 1t B-a 1t ) = 2
1"'"t"' "'I B+-tiJJ} 1 a 1t (tiJJ)
L.a1 (.,...k)(.,...k
j,k
+ aj(tiJJ){tiJJ I B-+cPk} 1 aj(cf>k)

= ~ Lbk(cf>k I B+-t/Jj}dJ + dJ(B-+cf>k I t/lj}bk


j,k

= ~ Lbk(cf>k I B+-t/Jj}dJ- dJ(cf>k I B+-t/lj}bk


j,k
= Lbk(cf>k I B+-tiJJ}dJ =btB+_dt, (6.41a)
j,k

using the canonical anticommutation relations {b j, dk} = 0 and (6.3 5). By


the same token,

- 2l (aJB-aJ) -z- j,k


= 1"'"
L.aJ(cf>k)(B-+cf>k I t/lj} 1 aJ(t/Jj)

+ aJ(lfJJ)(B+-tiJJ I cf>k} 1 aJ(cf>k)


= -z- j,kL bk(t/Jj I B-+cf>k}dj- dj(t/lj I B-+cf>k}bk
1

= Ldj(t/lj I B-+cf>k}bk = dB-+b. (6.41b)


j,k
242 6. The Spin Representation

Finally,

aJB+aJ =I bt(cJ>k I B++ci>J} bJ + dt(!JJk I B--!JJJ} dJ


1 1
j,k
= Ibt(cJ>k I B++ci>J} 1 bJ- dt(B--1/Jk I!JJJ} 1 dJ
j,k
=I bt(cJ>k I B++c/>j}bj- dk(l/Jj I B--1/Jk}dj
j,k
=btB++b + :dß__ dt:. (6.4lc)

Recall that for the two first operators to make sense (in that the vacuum is
in their domain of definition), the Shale-Stinespring criterion must be met.
But, taking that into account, nothing prevents us from making the result
of the calculation a rule for general operators:

dA(A) := bt A++b + bt A+_dt + dA-+b + :dA __ dt:.

We just showed that dA(A) = -iiJJ(iA) whenever Ais selfadjoint.


Definition 6.7. We can now construct a few "currents", namely, the (quan-
tized) charge

and the number operator

Vectors of the form bJ1 ••• bJ1 dt ...


dLn form a basis of Fock space;
they are immediately seen to be eigenvectors both of N and Q with respec-
tive eigenvalues l + m and l- m. lt follows that sp(Q) = 71.. We shall call
charge sectors, and denote by :!k (V), the eigenspaces corresponding to the
eigenvalues k E 7L of Q. The matter of charge sectors is intimately related
to the topology of Ul (Jf). Relabel the vacuum vector Q as 10} or IOin} and,
for SE Ul (Jf), write IOout} := J.I(S) 10} ("out-vacuum") for the new vacuum
vector. From the theory of Section 6.3, we know that if dim(ker Ps) = n > 0,
the out-vacuum is written

IOout} oc U1 1\ • • • 1\ Un 1\ hrs = at(ur) ... at(un) exp(~atTrsat) 10},


(6.42)

where {u 1, ... , Un} is an orthonormal basis of ker Pi and Prs is invertible.


lt should be clear from (6.4la) that all states of the form fy are charge zero
states (for a higher-brow argument, consult [213]). Since Psis a Fredholm
Operator of Order 0, it follows that index S++ = - index 5 __ .
Exercise 6.13. Checkthat S(kerS±±) = kerst.. o
6.4 Charged fields 243

Thus, indexS±± = dimkerst.- dimkersl±· For out-vacua of the form


(6.42), we can choose orthonormal bases {c/> 1 , •.. , cpz} and {tJ1 1 , .•• , tJim}
for ker sL and ker s!_ respectively, so that l + m = n. We get, for the
expectation value of the charge in the out-vacuum \0 0 ut) = J.i(S)\0),

(Üout I Q Üout) = l- m = dimker sL - dimker s!_


= indexs __ = -indexS++•

andthen
J.i(S):fk{V) = Jk+indexs __ (V),
which can be rewritten as

[J.i(S),Q] = indexS++· (6.43)

.,. Formula (6.43) is an anomalaus identity, since S commutes with Q; we


will see in a minute that the charge anomaly is "topological" in nature. By
using the Calder6n formula (4.3), it is possible to give an interesting formula
for [J.I (S), Q] in terms of] and S. Taking into account (6.3 7), we obtain

indexS++ = Tr(S!+S-+)m- Tr(S+-SL)m for m ~ 1.

Since

it follows that

indexS++ = 4 ~ Tr(l ([J,S][],st])m). (6.44a)

Since ][], S][], st] = -S[], St] + ]S[], st]], this can alternatively be writ-

:!
ten as

indexS++ =- Tr(S[J,st] ([J,S][J,stnm- 1 ). (6.44b)

Similarly,

indexs __ = indexsL = 4~ Tr(l ([J,st][J,S])m),


from which the relation index s++ = - index s __ follows at once.
For a superficially different proof of (6.44), see [293]. The expression
exhibits a truly interesting periodicity. lt is actually an important index
formula in noncommutative geometry. lt gives an index map K1 (A) - 7L
for suitable algebras [91, IV.l].
Exercise 6.14. Check formula (6.44) for the simple unitaries {eine} of the
Virasoro example. 0
244 6. The Spin Representation

Let us write i(S) := indexS++· The outcome of the discussion preced-


ing (6.43) isthat J.l(S) in general sends the original vacuum into a charged
vacuum, with charge -i(S). Two unitaries S1 , S2 can be continuously path-
connected in Ul (Jf) if and only if i(Sd = i(Sz), with i being constant on
the curve connecting S1 and Sz. The connected components uf of Ul (Jf),
interchanging the charged vacua, are the sets i- 1 (k) for k E 71., and the
connected component of the identity is ker i, the map Ul (J{) I ker i - 7l
induced by i being an isomorphism. (Concerning the topology of Ul (Jf),
we know that the connected component of the identity in the norm topo-
logy is algebraically generated by logarithmic elements, by Exercise 1.13.
But the same is true in the topology induced by the Hilbert-Schmidt norm
on the J-antilinear parts of elements of Ul (Jf), introduced in this chapter,
because the same Banach algebra arguments of Chapter 1 will work. Note
also that the distance between elements of different connected pieces of
uJ (Jf) is the generic 2.) All this follows at once from the noncommutative
Atiyah-Jänich theorem 4.19; for a bare-hands proof, we refer to [65]. (We
can even use the strong topology for convergence on the ] -linear parts, as
established in [66].)
Note that uf belongs to the neutral component of 0 1 (V) for k even and
to the other for k odd. The distinction between "neutral" fields (called Ma-
jorana fields by physicists, and most naturally described by the original
formalism of Section 6.3) and charged fields (where we distinguish between
a particle and an antiparticle sector) may largely be ascribed to the topolo-
gical differences between the pertinent groups. When a Majorana particle
undergoes electromagnetic interactions (a possibility still not ruled out in
neutrino physics) charge is therefore not conserved; but it is conserved
modulo 2.
Be that as it may, it has conclusively been shown by Matsui [332, 333],
for nonchiral fermions and quite general gauge fields, that the "classical"
scattering matrix S always belongs to the neutral component u6 of the
group; the contrary is generally true for chiral fermion fields. Our dabbling
in physics in this book is limited to the external field problern of quantum
electrodynamics (discussed in Chapter 13), where Q is conserved, and va-
cuum polarization in that sense does not occur -only pair creation arises.
Of course, is always possible that ker S++ * 0, dubbed the "spontaneaus
pair creation" case in the literature. The name seems inapt, as pair creation
occurs in essentially the same way whether Ps is invertible or not -see be-
low and Chapter 13. (We remark that charged vacua in principle do occur
in QED of strong static fields; but that is not a scattering Situation. Consult
the definitive discussion in [414].)
~ We turn to the task of giving the explicit form of J.l(S) for the charged
fermion field. In view of the previous remarks, we consider only the case in
which indexS++ = 0, looking in turn at the "regular" case in which ker ps =
0 and then the "singular" one in which ker ps * 0; we shall see that there
6.4 Charged fields 245

is little that is singular about the latter. We think of any S E Uf (J-f) as


a classical scattering matrix, and of J.i(S) as the corresponding quantum
scattering matrix.
In the light of (6.21), we can regard es as the (absolute value of) the
vacuum persistence amplitude (Om I Oout}. Namely,
es= I(Oin I Oom)l = deC 114 (1- Tg) = dec 112 (1 + (S+_s_::-~)ts+-S.::::-~)
= dec 112 ((s!_)- 1(s!_s __ + sLs+_)s::::-~)
= dec 112 ((s!_)- 1 S::::-~) = det 112 (s __ s!_),
using s!_s __ + sLs+- = E-. On the other hand, since p}(l- Tg)p 5 = 1,

I(Oin I Oom)l = det 114 (p5 p}) = det 114 (S++s!+) det 114 (S __ s!_), (6.45a)
and therefore both factors on the right hand side of (6.45a) are equal. We
thus arrive at
I(Oin I Oom)l = det 112 (S __ s!_) = det 112 (S++SL)
(6.45b)

Exercise 6.15. Checkthat the Operators s+_sL and s_+s!+ have the same
spectrum up to multiplicity. 0
Recall now the factorization of the quantum scattering matrix in Sec-
tion 6.3. Combining (6.38) and (6.41), we see that, in the notation of (6.24),
S1 = exp(bts+_s.::-~dt), S3 = exp(ds::::-~s-+b).
(We abbreviate S1(S) to S1 and so on, to avoid a clash of notations.) Now
write I+-:= s+_s=~ = (Ts)+- and L+ := s=~s-+ = -(Ts)-+. These expo-
nentials are explicitly developed thus:

sl = f
n=O
~! . I.
J 1 , ... ,J 11
(cf>kl I 1+-1/Jjl) ... (cf>k" I 1+-1/Jj") btdJ! ... btdJ"
k! .... ,k"

= f
n=O
~! . I.
J 1 , ... ,J 11
(c/>k 1 I 1+-1/J}I) ... (cf>k" I h-1/Jj") bt ... btdJ" ... dJ 1

k! .... ,k"

and
246 6. The Spin Representation

No ordering prescription is necessary for 5 1 and 53. For the 5 2 factor, the
Wick-ordered product :exp(at <Pst -I)a): can be written as 5 2b5 2d by sepa-
rating the band dterms. Since (c/>k I <Pst -1)c/>j) 1 = (c/>k 1 ((5L)- 1 -1)cp1),
we obtain
52b = :exp(bt ((5L)- 1 - 1)b):.
The remaining factor is

52d = :exp(dk(I/Jk I ((5!_)- 1 - l}I/JJ) 1 dJ):


= :exp(dk(((5!_)- 1 - 1)1/JJ I f./Jk)d1 ): = :exp(dk(I/JJ I (5:::-~- 1)f./Jk)d1 ):
= :exp(d1 (1/JJ I (1- 5:::-~)1/Jk)dJ): = :exp(d(l- 5:::-~)dt):.

Put h+ := (5L )- 1 -1 and [__ := 1- 5.:::-~. The fully developed exponentials


are

and

Putting everything together, we arrive at the exact regular 5-matrix for


the charged fermion field:

S = ewJ.1(5) = (Oin I Oout) exp(bt5+_5:::-~dt) (6.46)


x :exp(bt((5L)- 1 - 1)b- d(5:::~- 1)dt): exp(dS.::::~5-+b).

lt is customary in mathematical treatments to take the phase factor equal


to zero, but in Part IV, by methods generalizing the discussion after Def-
inition 6.6, we shall determine it; this is necessary in view of the physical
interpretation.
By construction,

Srr1 (v) st = rr1 (5v) for any v E V. (6.47a)


6.4 Charged fields 247

In other words 1

Sa1 (v) st= a 1 (psv) + aj(qsv)l


saj(v) st = aj(psv) + a,(qsv)l (6.47b)

yielding an one hand

sb(c/>) st = b(S++cl>) + dt<s-+c/>) 1


sbt(cp)st = bt(s++cl>) +d<S-+c/>). (6.47c)

and an the other

Sd(f./1) st = d(S--f./1) + bt (S+-f./1)1


sdt (f./1) st = dt (S __ f./1) + b(S+-f./1). (6.47d)

Howeverl these "global" formulae are not very useful in practice. This is be-
causel as weshall find in Chapter 13 1different physical processes are asso-
ciated to the various factors of the quantum scattering matrix. Ta compute
theml we need instead the following commutation relations:

[exp(bt Adt )1b(cp)] = -dt (At cp) exp(bt Adt ) 1


[exp(bt Adt )1d(f./1)] = bt (Af./1) exp(bt Adt )1
:exp(bt (A- I)b): bt (cp) = bt (Ac/>) :exp(bt (A- I)b): I

:exp(d(l- A)dt ): dt (f./1) = dt (At f./1) :exp(d(J- A)dt ): I (6.48a)

and their adjointsl

[exp(dAb) 1bt (cp)] = d(Acp) exp(dAb) 1


[exp(dAb) 1dt (f./1)] = -b(A t f./1) exp(dAb) 1
:exp(bt(A- I)b):b(cp) = b((At)- 1 cp) :exp(bt(A- I)b): I

:exp(d(l- A)dt):d(f./1) = d(A- 1 1./1) :exp(d(l- A)dt):. (6.48b)

Other combinations commute. For instancel [exp(bt Adt) 1bt (cp)] = 01and
so an.
Exercise 6.16. Prove formulae (6.48) and then verify equations (6.47) the
hard wayl by slogging through (6.46) with (6.48). Then verify that SQ = QS
by the same method. 0
We summarize our results: the regular quantum scattering matrix is of
the form
S = (Üin I Üom} :expdA(l): 1
248 6. The Spin Representation

where (Oin I :expdA(l):Oin) = 1. For reference:


_ (<sL>- 1 - 1 s+_s~~) (6.49)
I - s~~s-+ 1- s~~ ·
This operator is the soul of the quantum field theory of external fields; for
it we use, except for an unimportant imaginary unit factor, the notation of
Bellissard [28], who was the first to recognize its importance (in the boson
case). Notice that I satisfies the following equation:
I = S - 1 - (S - UE-I. (6.50)
Exercise 6.17. Verify that I is the only solution of (6.50). 0

~ The preceding computations are valid in arbitrary orthonormal bases.


Now, following the excellent paper [431], we introduce "canonical" ortho-
normal bases on the Hilbert space J.f. In effect, as recalled in Section 7.C,
the Hilbert-Schmidt operator S+- has the following canonical expansion:
S+- = 2:sklcPk)(«J..kl.
k
where {<J>d, {l/.ld are respective orthonormal bases in J{+ and J{-, and
the singular values Sk are such that Lk sf < oo and 0 =:; Sk =:; 1, taking into
account the unitarity of S. It follows that
s!_s__ = E-- sLs+- = 2:o- sf>I«J..k)(«J..kl.
k

from which we infer that


s __ = - 2: ~1- sf I«J..k)(«J..kl
k

for a suitable orthonormal basis {«J..k} for J{-. By working a bit more
with (6.3 7), we find one more orthonormal basis {<Pk} for J{ +, so that

S++ = 2: ~1- sf lcPk)(<J>kl and S-+ = 2:ski«J..k)(<J>kl·


k k
The canonical bases are uniquely deterrnined if there are no degenerate
eigenvalues. From (6.45b), we find that the vacuum persistence amplitude
equals I(Oin I 0 0 m) I = fh ~1- sf. For the quantum scattering matrix we
then get

s= ei 8 n~1- sf
k
6.4 Charged fields 249

where we have used the freedom to sandwich operators between different


orthonormal bases in (5.37). It should be clear that the exponential cor-
responding to each mode has only two nonvanishing terms and that this
equation can be rewritten as

Nowit is clearwhat the trouble maybe with invertibility of s±±: the Operator
Psis noninvertible if and only if the eigenvalue 1 occurs inS+- and S-+•
say n 0 times. But nothing really singular happens then, as the apparent
singularities cancel neatly with the zero value of the determinant, yielding
for the contribution of such modes the simpler formula

n (b<cf>k>t- d(IJ!k>) (b<cf>IJ- d(IJ!k>t)=


no
=
k=l

and moreover, that situation is clearly nongeneric .


.,. Finally, we introduce what is usually called the fermion field (and its
conjugate), to wit, the pair of charge-conserving Operators

'I'(IJ!) := aJ(IJI+) + aj(IJI_) = b(IJI+) + dt(IJ!-),


'l't(IJ!) := a}<IJ!+) + a 1 (1J!_) = bt(IJ!+) + d(IJI_). (6.51)

The commutation relations are simply

{'l'(cf>),'I'(IJ!)} = 0,

and of course TTJ(IJ!) = 'I'(IJ!) + 'l't (IJ!) always holds. Clearly,

S'l'(j) st = 'I'(Sj), (6.52)

directly from (6.47).


In some treatments (see [453], for instance), people try to work the quan-
tization formulas rigorously from 'I', forgetting the delicate foundation of
the theory. This is equivalent to the usual rule of thumb: quantize na'ively
using dA and the standard scalar product, then exchange the roles of d
and dt. That still requires a tricky subtraction of infinity for "normal or-
dering" (reflecting the fact that the Fock representations corresponding to
Q and ] are inequivalent!). Our approach (which goes back to I. E. Segal),
profiting fully from Clifford algebra, has many advantages: it allows a uni-
fied treatment of the Majorana and the charged case, and makes for clean,
streamlined proofs. lt transpires, then, that we da not have much use for
250 6. The Spin Representation

the field 'I'; but it is a bridge to the formal version of the theory that appears
in all physics texts. We make contact with that in Chapter 13.
Let us finish by pointing out that a bosonic quantization theory exists,
running parallel to the fermionic one developed in this chapter; especially,
the analogue of Theorem 6.16, due to Shale [433], applies. Consult, for
instance, [400] for the finite-dimensional theory and [212] for the infinite-
dimensional one. See also [91, V.11.ß] for a delightful application of bosonic
quantization, that in turn introduces Connes' approach to Riemann's con-
jecture on the zeta function [100].
7
The Noncommutative Integral

The road to integral calculus on noncommutative manifolds passes through


spectral analysis. In fact, inasmuch as it tries to discover the geometry of
manifolds from the analysis of operators strategically associated to them,
spectral analysis is noncommutative geometry avant Ia lettre.
lt stands to reason that the noncommutative integral be an operator trace
of some sort. Since the sixties, thanks to the work of Dixmier, certain non-
normal, abstract traces on the space of operators on a Hilbert space have
been known to exist. In the early eighties, Wodzicki uneavered the only ex-
tant trace on the algebra of (complete symbols of classical) pseudodiffer-
ential operators; it also became clear that the residues of the zeta functions
associated to those operators had tracial properties. Later on, Connes real-
ized that, for an important dass of operators, the Dixmier trace(s) and the
Wodzicki (and zeta-theoretic) residues were one and the same thing; this
automatically yielded a Dixmier-type formula for ordinary integrals. The
details form an interesting story, though by no means a short one.

7.1 A rapid course in Riemannian geometry


We consider a connected manifold M without boundary, of dimension n, on
which we impose a Riemannian metric. (For convenience, we take M to be
compact, although the reader will notice that much of what follows applies
to noncompact manifolds with minor modifications; we reserve the right
to specialize to M = IRl.n without further apology.)
252 7. The Noncommutative Integral

Definition 7.1. A Riemannian metric on M isapositive definite pairing on


smooth real vector fields, i.e.,
g: 1€(M, IR{) X 1€(M, IR{)- C"''(M, IR{),
which is C"" (M, IR{)-bilinear and satisfies g(X, X) > 0 for any nonzero X in
1€(M, IR{). We also denote by g the obvious extension to a symmetric C"" (M)-
bilinear form on the complex-valued vector fields 1€(M). The pair (M,g) is
a Riemannian manifold. If g is only assumed to be nondegenerate instead
of positive definite, (M, g) is then a pseudo-Riemannian manifold. This is
clearly the globalization of the bilinear form g in Chapter 5.
The metric g induces two bijective co (M)-module maps, the so-called
"musical isomorphisms" X - X~ from 1€(M) to . :·V(M), and its inverse
Of.- Of.~ from .J\ 1 (M) to 1€(M), by the reciprocal recipes

x•(Y) := g(X, Y), g(Of.~, Y) := Of.(Y). (7.1)


There is then an associated pairing of 1-forms, given by
B- 1 (0f.,ß) := g(Of.~,ß~).
to which we shall also refer as the metric.
On a chart domain U of M with local coordinate system {x 1 , ... , xn}, we
write the basic vector fields as o1 = o/oxJ.
Exercise 7.1. In local coordinates, the metric g is described by the matrix
[giJ], where BiJ := g(oi,OJ); let [grs] be the inverse matrix. Show that
(dxr)~ = 2. 1 grJa1 and that g- 1 (dxr,dx 5 ) = grs, thereby justifying the
notation g- 1 • o
When using local coordinates, we shall write X = xJ o1 , Of. = Of.r dxr,
using the Einstein summation convention of summing over each index that
appears twice, as both an upper and a lower index; thus
g(X,Y) =giJXiyJ and g- 1 (0f.,ß) =gr5 0f.rßs-
The metric also yields obvious hermitian pairings on 1€(M) and .J\ 1 (M),
which we shall denote, in both cases, with the ( · I ·) notation; namely,
(X I Y) := gCX, Y) = BiJXiYJ, (7.2a)
(Of. I ß) := B- 1 (iX,ß) = Br 5 Ö<rßs· (7.2b)
The isomorphism X - x~ is of the form["" g, for a certain bundle iso-
morphism g: TM- T* M, in view of Theorem 2.10 (Serre-Swan). Namely,
9x: TxM - TiM is the linear isomorphism Bx(v) := (w - Bx(V, w) =
BiJ(x)viwJ).
The gradient grad f of a scalar f E C"" (M) is the vector field defined by
g(gradj, Y) = dj(Y),

i.e., gradf := (dj)~; locally, (gradj)i = giJ o1j.


7.1 A rapidcoursein Riemannian geometry 253

Definition 7.2. If E M is any vector bundle, we use the handy nota-


tion [33]

.Jtk(M,E) := f"'(M,E) ®c"'<Ml .Jtk(M) ""f"'(M,AkT*M ®E).

Then .Jt • (M, E) is a graded C"' (M)-module of "E-valued differential forms


onM".
A connection \7 on E is a linear mapping

which satisfies the Leibniz rule

\?(sa) = (\?s)a + s ®da, (7.3)

It extends uniquely as a C-linear operator of degree 1 on .Jt • (M, E) by re-


quiring that

\?(s ® w) = \?s ® w +s ® dw, s E f"'(M,E), w E .Jt"(M). (7.4)

This can be rewritten as a graded Leibniz rule:

\?(a" w) = \?a" w + (-1)ka "dw, (7.5)

for a E .Jtk(M,E), w E .Jt"(M).


Definition 7.3. Any smooth vector field X on M gives rise to an operator
on .Jt"(M,E), of degree -1, namely the contraction tx, defined on simple
tensors s ® w by tx(s ® w) := s ® txw, where txw is the usual contracted
form
txw(YI·····Yk-1) := w(X,YJ, ... ,Yk-1),
if w E .Jt k (M), Y~o ... , Yk-1 E X(M). In particular, txs = 0 for s E roo (M, E).
Any two contraction operators tx, Ly anticommute.
Exercise 7.2. Show that tx is an odd derivation on .Jt • (M, E).
Recall that the Lie derivative Lx with respect to the vector field X satisfies
the Cartan identity Lx = tx o d + d o tx on the algebra .Jt • (M) [1]. Given a
connection \7 on E, we use the analogaus composition

\?x := tx o \7 + \7 o tx,

which is a zero-degree operator on .Jt • (M, E) for X E *(M).


Exercise 7.3. Show that \7 x satisfies the (ungraded) Leibniz rule

\?x(a" w) = \?xa" w + a "Lxw, a E .Jt"(M,E), w E .Jt"(M). (7.6)

In the particular case of \7 = d, L x is an even derivation on .Jt • (M). 0


254 7. The Noncommutative Integral

Exercise 7.4. Using the known identity [Lx, Ly] = L[X,YJ on forms, show
that
'Vx(Lya) = Ly('Vxa) + L[x.na for a E Jt•(M,E). 0
Exercise 7.5. Check that V' fX = f'V x, for f smooth. 0
The last exercise teils us that V' x depends only on the value of X at a given
point of a manifold (and not on its derivative); that allows us to consider
operators of the form V' v, for v E TM.
~ Any affine connection V' (that is, a connection on TM) has a torsion tensor
V'(), where () E 5\. 1 (M, TM) is the fundamenta/1-(orm, satisfying LxO :=X
for all X. Now Exercise 7.4 yields the familiar expression for the torsion:

'VB(X, Y) = Ly(Lx'VO) = Ly('VxO- 'V(Lx0)) = Ly('VxO- V' X)


= 'Vx(Ly0)- L[x,YJO- 'VyX = 'VxY- 'VyX- [X, Y]. (7.7)

Definition 7.4. Given a Riemannian manifold (M,g), the Levi-Civita con-


nection 'VB on TM is the unique affine connection that is torsion-free and
satisfies the metric compatibility relation

g('VBX,Y)+g(X,'VBY)=d(g(X,Y)), for X,YEJ€(M,IR). (7.8a)

This condition can also be stated with complex-valued vector fields X, Y E


Jt(M), provided we also demand that V'~X be a real vector field whenever
X and Z are real. Substituting X- X, (7.8a) can then be rewritten as

('VB X I Y) +(X I V'BY) = d(X I Y). (7.8b)

Alternatively, after contracting with another vector field, metric compati-


bility means

g('V~X,Y)+g(X,'V~Y)=Z(g(X,Y)) forall X,Y,Z. (7.8c)

The lack of torsion of 'VB can be expressed, using (7.7), by the identity
'V~Y- 'V~X =[X, Y].
The relation (7.8c) may be written in several equivalent ways:

g('V~X. Y) + g(X, 'V~Z) = Z(g(X, Y)) + g(X, [Y, Z]),


g('V~Z,X) + g(Z, 'V~Y) = Y(g(Z,X)) + g(Z, [X, Y]),
g('V~Y. Z) + g(Y, 'V~X) = X(g(Y, Z)) + g(Y, [Z,X]).

Adding and subtracting gives

2g('V~Y. Z) = X(g(Y, Z)) + Y(g(Z,X))- Z(g(X, Y))


+ g(Y, [Z,X]) + g(Z, [X, Y])- g(X, [Y, Z]). (7.9)
7.1 A rapidcoursein Riemannian geometry 255

lt is easily checked that the right hand side is C"" (M)-linear in Z and X;
therefore, (7.9) uniquely determines an element VBY E .J\. 1 (M, TM) that
satisfies V~(jY)- f V~Y = X(j) Y, and thus VB(jY)- f VBY = Y ® df
for f E C""(M). This shows that (7.9), taken now as a definition of VB,
determines a unique affine connection, as claimed. The same argument
holds for pseudo-Riemannian manifolds as weH.
We can now extend the Levi-Civita connection to the whole tensor bundle,
using recursively the tensor product recipe for connections:

We can define on .J\. 1 (M) = Homc"'(Ml (X(M), C"" (M)) an associated con-
nection, also denoted by VB and referred to as the Levi-Civita connection
on 1-forms. For scalar fields, we settle for VB f := df. When f = a(X), a
formal application of the Leibniz rule provides the definition

VBa(X) := d(a(X))- a(VBX) for a E .Jl. 1 (M), XE X(M), (7.10)

or V~a(X) = Z(a(X))- a(V~X) after contracting with Z. The tensor pro-


duct prescription (7.4), followed by skewsymmetrization, allows us to ex-
tend this to a Levi-Civita connection on all of .J\. • (M), satisfying the graded
Leibniz rule (7.5) -with (]" E .J\. k (M) in the present case.
Exercise 7.6. Show that (VB) 2 : .Jl."(M, TM)- .Jt•+ 2 (M, TM) is an .J\. "(M}-
module map:
(VB) 2 ((J" 1\ W) = ((VB) 2 (J") 1\ w
for (]" E .Jl."(M, TM), w E .Jl."(M). <>

The Riemannian curvature is the element R E .J\. 2 (M, End TM) given by

R(X, Y)Z := LyLx(VB) 2 Z = Ly(V~(VBZ)- VB(V~Z))

= v~v~z- v~v~z- vfx.Y 1z


forZE X(M) = f""(TM), on using Exercise 7.4. If g(Z, Z) =: h, it follows
that 2 g(Z, V~Z) = X(g(Z, Z)) = X(h) by the metric compatibility (7.8c),
and then

g(Z,R(X, Y)Z) = g(Z, V~V~Z)- g(Z, V~V~Z)- g(Z, Vfx.nZ)


= X(g(Z, V~Z))- Y(g(Z, V~Z))- g(Z, Vfx.nZ)
= ~X(Y(h))- ~Y(X(h))- ~[X, Y](h) = 0, (7.11)

so that g(W,R(X, Y)Z) is skewsymmetric in Wand Z (and also has the


obvious skewsymmetry in X and Y). Thus R(X, Y) can be regarded as a
section of the bundle so(TM) of g-skewsymmetric endomorphisms of TM.
256 7. The Noncommutative Integral

Exercise 7.7. Verify the first Bianchi identity,

R(X, Y)Z + R(Z,X)Y + R(Y, Z)X = 0,

by assuming (as you may, since R is a tensor) that the Ue brackets of X, Y


and Z all vanish at a given point of M. Then show that

g(W,R(X, Y)Z) = g(Y,R(Z, W)X) for X, Y, Z, W E X(M). 0

.,.. Most calculations involving VB and R can be done efficiently in local


coordinates. Locally, VB = d + oc, where oc E 5\ 1 (U, End TM) is the connec-
tion 1-form on the chart domain U c M. The Christoffel symbols ri~ of the
Levi-Civita connection are the functions in C"" (U) defined by

VBo1 =: ri~ dxi ® ak.

r:
Thus 1 give the matrix components ocj of oc. From the torsion-freedom
and [oi, a1 ] = 0, one easily derives rt1 = fji· Since V~ia1 = rt ak. the metric
compatibility equation (7.8c) becomes

BJmf{[' + Bimf{j = ozBii•

and (7.9) reads Oi9il +o1ga-ozBii = 2BzmftJ, which yields the fundamental
formula
(7.12)

The Levi-Civita connection on the cotangent bundle is determined locally


by (7.10):
(7.13)

Equivalently, V~i dxk = -ri~ dxi. Also, ri~ = dxk (V~i a1 ) = -(V~i dxk)a1.
The symmetries of the Riemannian curvature tensor are also easily ex-
pressible in local coordinates. Writing

RJkl oi := R(ok. oz) oi = V~k V~1 oi- V~1 V~k Oj,


we obtain
:= akr}1 - ozrJk + rJir~k - f}/:f~ 1 .
R)kz
Two other important quantities may be obtained by contraction of indices,
namely the Ricci tensor
(7.14)

and the scalar curvature


(7.15)
7.1 A rapidcoursein Riemannian geometry 257

We also use other forms of the Riemannian curvature tensor, obtained


by raising or lowering indices, in particular

RiJkl + RaJk + RiklJ = 0; RiJkl = RkliJ·


(7.16)

Exercise 7.8. Deduce that the Ricci tensor is symmetric: R 1z = R!j. 0


Finally, we shall assume the following result [401, Prop. 1.14] obtained
from the Taylor expansion of the metric in (geodesic) normal Coordinates
at a point of M.
Lemma 7.1. On a suitable neighbourhood of any point xo of a Riemannian
manifold it is possible to define normal coordinates so that

E3

.,. Integration of functions over orientable Riemannian manifolds is facil-


itated by the presence of a distinguished volume form. However, integra-
bility does not in principle require this: the element of integration is the
canonical density determined by the metric. The subject of densities is of
the utmost importance in this chapter, so we start from scratch.
Definition 7.5. Let V be a real vector space of dimension n and let c:x E ~­
A continuous mapping d: vn - ~ is an c:x-density if it verifies

d(AVI, ... ,Avn) = I detAioc d(VI, ... , Vn),

for all VI, ... , Vn in V and all A E End V. The vector space of c:x-densities is
denoted I.Jlloc(V); 1-densities are calledjust densities, and we write I.Jll (V)
for I.JI.II(V).
The space I.Jlloc(V) is one-dimensional; for let d~t d2 E I.Jlloc(V), and let
{ui, ... , Un} be a basis for V. Let a = d2(ult ... , Un)fddui, ... , Un) -we
assume that d I is nontrivial. Now, if VI, ... , Vn E V, let A be defined by
Aui := Vi for i = 1, ... , n. Then

lt follows that d2 = adi.


Densities can be constructed from volume elements: if w E .Jln(V), then
lwloc is defined by lwloc(vi, ... ,Vn) := lw(vi, ... ,Vn)loc. This correspond-
ence matches positive volume forms in .Jl n (V) with positive densities in
I.Jlloc(V), for any c:x. Suppose now that g is a nondegenerate symmetric bi-
linear form on V, and that an orientation on V is given. Then there is a
unique volume element v9 such that v9 ( WI, ... , Wn) = 1 for all positively
258 7. The Noncommutative Integral

oriented g-orthonormal bases {w 1 , ••• , Wn} of V; if {u 1, ... , Un} is an arbi-


trary positively oriented basis with dual basis {v 1, ... , Vn}, then
v9 = Idet[g(ui, uj)] l 112 v1 1\ · · · A Vn.

Indeed, if Uj = Lr arjWr = Awj, then g(ui, Uj) = Lr(±l)rariarj. so


det[g(ui,UJ)] = ±(detA) 2 ,whereasuiA· · ·1\Un = (detA)WIA· · ·1\Wn,
so that v9 = I det Al VI A · · · A Vn.
The corresponding oc-densities Iv9 Ioc are defined by

lvgloc(VI, ... , Vn) := Idet[g(vi, Vj)] Ioc/ 2 .


Wehave followed [1] for the algebra. The picture can be easily globalized
in Riemannian manifolds (or pseudo-Riemannian ones, for that matter).
Definition 7.6. For any oc E ~. on any Riemannian manifold (M,g) there
exists a unique Riemannian oc-density lv9 1oc that takes the value 1 on all
orthorrormal bases of the tangent spaces TxM. If VIx •... , Vnx E TxM, then
lvgloc(VJx, · · ·, Vnx) = Idet[gx(Vix. Vjx)] Ioc/ 2 •
Indeed, for any manifold M there is a real line bundle I.Jl.l oc (M) of oc-
densities (whose transition functions are expressed by octh powers of ab-
solute values of Jacobians). These bundles are trivial, as there is no ob-
struction to the construction of nowhere-vanishing sections (whereas the
bundle of volume elements is trivial if and only if M is orientable). In the
Riemannian case, the Riemannian oc-density provides a canonical basis ele-
ment, and so a canonical identification of I.Jl.l oc (M) with Mx~- The density
bundle (for oc = 1) is most important, as the classical change of variables
formula for integrals tells us that there is a unique linear form fM on the
space of sections of I.Jl.l (M), which is invariant und er diffeomorphisms and
agrees with the Lebesgue integral on local charts. We shall write, in local
coordinates (XI, ... ,Xn),

~detg ldx 1 1\ · · · 1\ dxnl =: p ldx 1 1\ · · · 1\ dxnl =: p ldnxl


for this distinguished Riemannian density.
Densities allow for more intrinsic definitions of scalar products on spaces
of sections in some contexts; for instance, the Hilbert space used in Sec-
tion 6.4 to represent the Virasoro group could be replaced by a Hilbert
space of half-densities.

7.2 Laplacians
We want to consider the various Laplacians associated to the Riemannian
manifold (M,g). For simplicity, we consider for the time being only scalar
Laplacians, i.e., operators acting on sections of the trivial bundle Mx I[ - M.
7.2 Laplacians 259

A natural generalization to (M,g) for the Laplacian - L.1 oJ on an Eu-


clidean vector space (the minus sign makes it a positive operator) would be
"simply" given by -giJ'VBo; 'VBo· = -giJ'VBo; 0J..
We may understand this formula thus: on the trivialline bundle L :=
Mx ([ - M, the Levi-Civita connection reduces to d: po (L) - .Jl 1 (M, L) =
=
f""(T*M®L);then 'VB mapsf""(T*M) f""(T*M®L) to.Jl 1 (M, T*M®L) =
f"" (T* M ® T* M ® L), so it can be composed with d. Contraction with the
metric g- 1 on T* M gives a C"" (M)-linear map TrB: f(T* M ® T* M ® L) -
f(L). The composition of these three maps yields the following operator
on the space of sections f(L):

Ll := - TrB o 'iJB o d,

called the connection Laplacian associated to d.


On a chart domain, using (7.13) to compute 'VB df, we get immediately
the local expression for the Laplacian:

(7.17)

Before continuing, we make the trivial remark that, just as the metric de-
termines the Laplacian, reciprocally the Laplacian determines the metric:
just apply it to a function of the form xi xJ on some local chart. lt is only
tobe expected, then, that the analysis of Ll and the geometry of (M,g) are
intimately related.
Another natural (and time-honoured) generalization to (M,g) for the
Laplacian on an Euclidean vector space is the Laplace-Beltrami operator,
which we consider next.
Definition 7-1. The divergence of a vector field X on an oriented mani-
fold M, relative to the volume form v, is the scalar field divv X E C"" (M)
determined by
(divv X)v := Lxv.
Now, Lxv = d(lxv) by Cartan's identity Lx = lx d + d lx. Hence, forM
without boundary, the divergence theorem

is an immediate consequence of Stokes' theorem. If f E C"" (M) never van-


ishes, then fv isanother volume form, with (diVJvX)fv = .Lx<fv) =
(Xj)v + f .Lxv = (Xj + f divv X)v, so that

If X = Xi Oi then divdxli\···Adxn X = OjXJ, since Lx(dx 1 1\ · · · 1\ dxn)


L,1 dx 1 1\ · · ·A.Lx(dxl) 1\ · · ·1\dxn = L,1 dx 1 A · · ·1\d(Xl) 1\ · · ·Adxn.
260 7. The Noncommutative Integral

We deduce a local formula for the Riemannian divergence (denoted by div X


without subscript):

Another local formula for the Riemannian divergence is

(7.18)

The second equality follows from the Leibniz rule for 'iJB, the first one
follows if we can show that

Now, if [Gii] denotes the matrix of cofactors of [BiJ], then by Cramer's


rule gii = Gii I det g. Also, gi 1( aJBkz - ozBkJ) = 0 since the matrix [BiJ] is
symmetric, and so, from (7.12),

J 1 ·z 1 ·z 1 op 2
f. = -g1 okg·z = - - G1 okg·z = - - o k g · z
Jk 2 1 2detg 1 2p 2 OBJZ 1

= p- 1 akP = ok(logp), (7.19)

on using the Cramer expansion detg = LJ g 1zGi 1 (no summation over l).
Definition 7.8. On an Euclidean vector space, (gradj)i = a1j and div X=
a1xJ. This leads us to define the Laplace-Beltrami operator ßLB by
1 "k
(7.20)
ßLBf :=- divgradf = --oJ[pg 1 Ok]f.
p

To see that this leads back to the formula (7.17) for the Laplacian, we
need the local formula

(7.21)

For this, we compute, from (7.12),

gikrJk = gikglm(Ok9mj- ~OmBJd·

Also, from 0 = Ok(g 1mBmJ) we obtain gikglm OkBmJ = -gik((Jkglm)BmJ =


-okg 1k. On the other hand,

.!.gikglm 0 B .k = _1_ cikglm 0 g .k = .!_ 9 zm 0 p


2 m J 2 det B m J P m '

because of (7.19); and (7.21) follows. We conclude that, on scalars, the


Laplace-Beltrami operator ßLB and the connection Laplacian ß coincide.
7.2 Laplacians 261

.,.. In Section 9.B, we shall consider the Hodge-de Rham Laplacian ßH =


d dt + dt d on forms. Its restriction to scalars can be envisaged here: de-
fine dt: .:.t 1 (M)- C""(M) by dtTJ := -diVTJ~ (and df(C""(M)) := 0); then
!:iHf := dt dj = ßj.
The complex vector space Jt 1 (M) becomes a prehilbert space under the
positive definite Hermitian form

(oc I ß) := JM(oc I ß) lv9 1. (7.22)

(When M is noncompact, we define this scalar product on the space Jt~ (M)
of 1-forms with comf.act support.) Its completion is a Hilbert space that
may be denoted by Li (M). Notice that the complex conjugation of forms
is an antiunitary operator on this Hilbert space. The inner product is ex-
tended to Jt 0 (M) $ Jt 1 (M) by declaring that forms of different degree be
orthogonal; the completion L~· 0 (M) $ L~· 1 (M) is a Hilbert space direct sum.
In fact, in Chapter 9, by globalizing the Clifford algebra construction of
Chapter 5, the whole space of forms Jt • (M), or Jt~ (M) in the noncompact
case, is made into a prehilbert space, and in the completed Hilbert space
the operators d, dt and d dt + dt d are densely defined.
If M is compact with two Riemannian metrics g, h, the corresponding
Hilbert spaces are naturally isomorphic (the identity operator from one
to the other being bounded). In the compact case we can thus omit the
subscript g; also, if M is compact, we normalize the density so that ( 111) =
1. (However, if M is not compact, there are always metrics g, h suchthat
L~· 0 , say, contains smooth elementsnot contained in L~· 0 .) Of course, one
can decide that the Hilbert spaces are made of sections of a bundle of half-
forms (i.e., forms tensored with half-densities), rather than forms; that is a
more natural viewpoint for some applications.
The notation dt is well justified, in view of the relation
-(divXIJ) = (X'Idf), for XE]t(M), jE.:.t 0 (M), (7.23)
which implies (dt TJ I f) = (TJ I df) for TJ E Jt 1 (M), i.e., dt is the (at least
formal) adjoint of d, and moreover (df I df) = (j I ßj), so ß is a for-
mally positive selfadjoint operator. lt is enough to check (7.23) when f is
supported in a chart domain U and X is real, and this is accomplished by
computing

L (X' I dj) lvgl = LgijXj okf(dxi I dxk) p ldnxl

= fugiJXiokfgikpldnxl = fupxioifldnxl

=- Lp- 1 oj(pXi)J p ldnxl = - JM(divX I j) lv9 1.

In fact, the Laplacian ß on a compact manifold is a bona fide selfadjoint


operator, but in this book weshall be content with proving selfadjointness
of its close relative, the Dirac operator, in Chapter 9.
262 7. The Noncommutative Integral

Proposition 7.2. The adjoint of the operator 'V{ is -'V{ - div X.


Proof. lf f E C"" (M), then

L (Xf + f div X) vB = L Lx(fvB) = L dtx(fvB) = 0.

Taking f = ( a Iß) for any 1-forms a, ß, the metric compatibility of 'VB gives

L[(Y'{a I ß) + (a I Y'{ß) + (a I ß) divX)] vB = 0,

so that ('V{a I ß} = - (a I Y'{ß) - (a I (div X) ß). D

Exercise 7.9. Prove that


1 0 ( n-1 0 ) 1 0
ß~n = yn-1 or r or + y2 ßsn-1.

~ The connection Laplacian generalizes more easily to the nonscalar case.


Consider again the line bundle L - M (which we no Ionger assume to be
trivial). On L one can define a more general connection 'VL = d + a, where
the connection 1-form a = iA = iA i dxi has only one matrix component. If
we ask that the connection be hermitian, then Ais real; this corresponds, in
the language of physics, to a U ( 1) gauge field. Let VL be the tensor product
of 'VL with the Levi-Civita connection 'VB on T* M. Then, as before, t'L maps
[(T* M ®L) to .5\ 1 (M, T* M ®L) = [(T* M ®T* M ®L), so it can be composed
with 'V L. Contraction with the metric again yields the following operator on
the space of sections f(L):
(7.24)

This is called the Laplacian associated to the connection vr, which in the
literature, such as [197], is often (and confusingly) written & = -gii'V a; 'V ai'
where both the Levi-Civita connection and the line bundle connection are
understood. Proceeding as before, we obtain its local expression:
(7.2 5)

In a completely explicit way, we can write

ßL = gii (DiDi + (ifi~Dk + 2AiDj) + (AiAj + ifi~Ak + DiAj)), (7.26a)

where we have taken the opportunity to introduce the customary notation


Dj := -ioj, familiar from the theory of distributions, in order to absorb the
negative signs.
just as before, this is synthetically written as
L 1 ..
~ = -Wi + Ai)(pg 11 (Di + Aj)), (7.26b)
p
7.2 Laplacians 263

which can be obtained from (7.20) by application of the "minimal coupling


recipe" D i .... D i +Ai. In conclusion, connection Laplacians on line bundles
have a physical-geometrical interpretation, corresponding to the presence
of a gravitational field and an Abelian gauge field.
Now, as donein [33], we can decide that a (generalized) Laplacian is any
differential operator with principal symbol gii(x)~i~j; then the most ge-
neral scalar Laplacian is locally written as

H == gik(x) Dj Dk + b 1(x) Dz + c(x),


where the b 1, c are smooth. In cantrast to [33], we shall demand formal
selfadjointness from the outset:

(f I Hh} == (Hf I h}. (7.27)

Exercise 7.10. Checkthat the condition (7.27) is equivalent to

(7.28)

using integration by parts. 0

Note that ~ bi and ~ c are left undetermined by the formal selfadjoint-


ness condition. The diligent readerwill immediately see that the first equal-
ity of (7.28) for the connection Laplacian is precisely (7.21) and, writing
Ak :== ~Bki ~ bi, will check in earnest the second equality of (7.28) in the
expression (7.26a). Now let V :== ~ c - gik A jAk. We arrive at

H==t::.L+V.

Wehave thus verified, by local computations as in [193], that any gener-


alized Laplacian Hismade from three ingredients: (a) the metric g on M;
(b) a (hermitian) connection on a line bundle L; (c) a section V of the bundle
of endomorphisms of L, determining the zeroth-order piece.
Exercise 7.11. Checkthat the definition of V is not affected by local coor-
dinate changes. 0

.,.. Our His elliptic because the principal symbol a 2 (H) (x, ~) == g- 1 (~. ~) ==
gii (x)~i~j > 0 for all ~ * 0 -we refer to Section 7.A for the definition
of ellipticity and all nomenclature and background on (pseudo)differential
operators used in this book. The utility of the form (7.26b) for the Laplacian
is that we immediately find

(J I Hf}== fu ((D + A)f I (D + A)j) lv 9 1+ (f I Vf}.

Therefore, it is enough that V be bounded below for H tobe semibounded


and then, according to general theory, to have selfadjoint extensions. We
264 7. The Noncommutative Integral

spell out in geometric language the complete symbol associated to (7.26a):

a(ßL)(x, ~)

= gii(~i~J + (ifi~~k + 2Ai~J) + (AiAJ + i(fi~Ak- oiAJ))) + V(x). (7.29)


In the pure Levi-Civita case, we get simply PH(X, ~) = giJ (~i~J + ifi~~k ).
At this point, we realize that the formula for the Laplacian on a general
vector bundle Eis exactly the same, provided that we reinterpret each A 1(x)
as a matrix acting on Ex -that is, A 1 E f""(U,EndE). Note that the terms
gii~i~J + igiJ[i~~k will act diagonally. Given a connection '\JE on E, the
connection Laplacian ßE is obtained directly from (7.24) or (7.25).

7.3 The Wodzicki residue


Integration on ordinary manifolds may be recast into a noncommutative
mold due to the existence of an important functional on pseudodifferential
operators, called the residue of Wodzicki. In the one-dimensional case, this
was discovered previously by Adler (3] and Manin [325], and studied closely
by Guillemin [226]. The higher-dimensional case was developed by Wodz-
icki [490,491], who realized its role as the unique trace (up to multiples) on
the algebra of classical pseudodifferential operators. We also recommend
the notes by Kassel [277], a nice summary of the whole Wresidue matter.
Our discussion is restricted to the case of compact boundaryless mani-
folds; for an excellent introduction to noncommutative residues on mani-
folds with boundary, see the paper by Schrohe [420].
Symplectic cones constitute the proper geometrical ground for the defi-
nition of the Wodzicki residue.
Definition 7.9. A symplectic cone is a symplectic manifold (Y, ß) that is
a principal bundle with fibre IRl.+ over a compact base X. Let p denote the
action of IRl. +. We also require that

Pt ß = t ß for all t > 0. (7.30)

The most pertinent example is Y = T* M minus the zero section, for M


compact, with the standard symplectic form and the dilations pr(x, ~) :=
(x, t~). We can take the cosphere bundle §* M as the base. If t =: e5 , the Lie
algebra generator on IRl,+ ist d/dt = d/ds. The corresponding fundamental
vector field (t d/ dt)y is the so-called Euler vector field X, given by

From the definition, or by direct computation, DRß = ß. Indeed, s ..... Pes is


the flow of X, so L:Rß = (d/ds) 1s=oe 5 ß = ß in view of (7.30). The symplectic
7.3 The Wodzicki residue 265

form may be expressed in Darboux coordinates as ß = d~1 1\ dxJ where


(x 1 , ... , xn) are local coordinates on a chart domain in M; we abbreviate,
as before, dnx := dx 1 1\ • • • 1\ dxn. Let w := ß"n be the volume form on Y;
note that L'Rw = n w. Consider the (Zn- 1)-form lRW:
lRW = (-l)n(n-1)/2l'R(d~1 1\ ••• 1\ d~n 1\ dnx) =: (-l)n(n-1)/20"!; 1\ dnx,

where
n
0"!; := 2:: (-1)1- 1 ~jd~1 1\ • • • 1\ ~ 1\ • • • 1\ d~n· (7.31)
j=1

Of course, <T!; restricts to the volume form on the unit sphere { ~ : I~ I = 1 }


in each TiM. (The notation I~ I means the Euclidean norm of the coordinate
vector (~1 •... , ~n) in ~n.) Clearly d(lRW) = ( -l)n<n- 1)1 2d<T!; 1\ dnx = n w
(which, in view of Cartan's identity, follows also from LRW = n w).
We restriet to a single cotangent space TiMfora short while, in order to
examine <T!; more closely. We regard it as an (n-1)-form on ~n \ {0}; that is,
<T!; = tRdn~ with 'R = 2.']= 1 ~J o/o~1 . We need to rely on the connectedness
of the sphere §n- 1 , so weshall suppose until further notice that n ~ 2.
Definition 7.10. A smooth function f defined on ~n \ {0} is homogeneaus
of degree ,\ if j(t~) = t?-. j(~) for all t > 0 -or, equivalently, if 'Rf = .\f.
A derivative of a function homogeneaus of degree ,\ is homogeneaus of
degree ,\ - 1.
Proposition 7.3. For any function a_n(~) homogeneaus of degree -n, the
form a_n<T!; on ~n \ {0} is closed.

Proof. Since lR is a graded derivation,

da-n 1\ 0"!; = da_n 1\ l'R dn~ = l'R da_n 1\ dn~ = -na-n 1\ dn~,

so d (a-n <T!;) = da_n 1\ <T!; +a-n d<T!; = -na-n 1\ dn ~ + na_n 1\ dn ~ = 0. 0

Weshall consider integrals of the form f§n-1 a_n(~)I<T!;I· From Propo-


sition 7.3 it follows that they can be calculated using any section of the
~+-principal bundle ~n \ {0}. In particular (remember that n ~ 2), we can
"blow up" the sphere §n- 1 into a cylinder §n- 2 x ~. which isahomologaus
cycle in ~n \ {0}, and obtain the same value for the integral.
Now, homogeneaus functions are generically sums of derivatives, on ac-
count of Euler's theorem: if f is homogeneaus of degree .\,

_1_
n + ,\
±
i= 1
o(~d)
o~i
= _1_(nf + 'Rf)
n + ,\
= f.

This argument fails, however, for ,\ = -n. Instead, we get a more restricted
result.
266 7. The Noncommutative Integral

Lemma 7.4. fsn-1 a-n(~) 10'~ I = 0 if and only if a_n is a sum of derivatives.

Proof. Since !:lj = 0 if and only if df = 0, the kernel of the Laplacian on


functions on a compact space consists of the constants only. Assurne that
the integral is zero. Given a function h on §n- 1 , the equation

must therefore have a solution. We may extend h to !Rn \ {0} by defining


h(~) := l~l-<n- 2 lh(~/l~l), applying the result ofExercise 7.9 to this func-
tion, we get

Conversely, takea-n := ob1-nlo~1. say, with b1-n homogeneaus of order


1 - n. Replacing the cycle §n- 1 by §n- 2 x IR and writing 17 = (~2 •.•• , ~n),
we compute

since b1-n vanishes at infinity, due to its homogeneity. D

Before proceeding, we make a few remarks on the general symplectic


case -see [226] or [491]. Nothing really changes: using the Euler vector
field 'R = (tdjdt)y, and introducing oc := L'Rß, we see that pioc = toc, and
we get ß = doc, in view of L'Rß = ß. Then we consider J1 := oc 1\ ßn- 1 , and
clearly PiJ.l = tnJ.l. lt also follows that L'Roc = oc and L'RJ.l = nJ.l. A smooth
function f defined on a symplectic cone Y .!!.. X is homogeneaus of degree
,\ when L'Rf = Af. Let f be homogeneaus of degree -non Y. Then !11 is
invariant under the action of [R+; moreover, it is "horizontal" with respect
to the fibration, since L'R (j J1) = f L'RJ.l = 0; hence it is of the form rr* J1f for
some form J.lJ (of top degree) on the base space X. The symplectic residue
of f is then defined as the integral fx I J.lf 1 .
.,. lt is now time to consider spaces of pseudodifferential operators. For
the reader's convenience, we give in Section 7.A abrief overview of these
operators and outline the symbol calculus. The (scalar, for now) pseudodif-
ferential operators of order at most d on a manifold M form a vector space
'I'd(M). These spaces are increasingly nested, as d - oo, the intersection
being the space of smoothing operators 'I'-oo (M). The symbol a determines
the operator A = Op(a) up to a smoothing operator. Properly supported
pseudodifferential operators form an algebra under composition; under
our standing hypothesis that M is compact, proper support is automatic
and 'I'"" (M) is an algebra. It contains an important subalgebra 'I'cl' (M) of
"classical pseudodifferential operators". The smoothing operators form an
ideal in both these algebras.
7.3 The Wodzicki residue 267

Definition 7.11. The quotient algebra 'P(M) := 'I'cl (M) /'I'-co (M) is called
the algebra of classical symbols on the compact manifold M. Elements of
'P(M), restricted to a local chart of M, can be identified to asymptotic ex-
pansions of the form
CO

a(x, ~)- I ad-J(X, ~). (7.32)


j=O

where each ar(x, ~) is r-homogeneous in the variable~ (this is the "classi-


cality" condition); moreover, any such formal series gives rise to a symbol
a(x, ~) that is uniquely determined modulo smoothing operators. In other
words, we get a short exact sequence
0 - 'I'-co (M) - 'I'ct (M) !!.. 'P(M) - 0, (7.33)
where a- denotes the symbol map.
The term of order-n of the expansion (7.32) has a special significance. It
is coordinate-dependent, but we can always fix a coordinate domain U c M
over which the cotangent bundle is trivial, and consider a_n (x, ~) as a
smooth function on T* U \ U.
We come now to the key results. We shall follow rather closely the beau-
tiful elementary proofs by Fedosov et al. [17 4].
Theorem 7.5 (Wodzicki). Given a classical pseudodifferential operator A,
there exists (independently of local representation) a l-density on M, denoted
wres A, whose local expression on any coordinate chart is

wresxA := (J1~1=1
a_n(X, ~) la-~1) ldx 1 1\ • • • 1\ dxnl. (7.34)

This is the Wodzicki residue density. By integrating this l-density over M,


we get the Wodzicki residue or noncommutative residue functional

WresA := JM wresxA. (7.35)

(In the literature, this residue is commonly written res A; we adjoin the W
-and the w- to help to distinguish the density from the functional.)
Proof. In order to prove independence of local representation, we need to
know the behaviour of symbols under diffeomorphic changes of Coordi-
nates. This is dealt with in Section 7.A, where it is shown that under a dif-
feomorphism 4J : U - V between open subsets U and V of ~n, with inverse
ljJ : V - U, any operator A E 'I'd(U) with symbol a(x, 17) can be pushed
forward to an operator 4J*A E 'I'd(V) by setting 4J*A(j) := A(4J* j) o 4J- 1
when f is a test function on V. The complete symbol has, by taking ~ =
ljJ' (x)tlJ in Theorem 7.29, the following transformation rule:

ä(x, ljJ'(x)tlJ) =I Ccx(X, 1]) a~a(ljJ(X), 1]),


IX
268 7. The Noncommutative Integral

where co(x, fl) = 1 and the other Coc are polynomials in fl. This means that
ä-n (x, 1/J' (x) t '7) differs from a_n ( ljJ (x), '7) by sums of derivatives in the
r]-variable.
Let us now examine the behaviour of J1!; 1=1 a_n(x, ~) Ia!; I. for x fixed,
under a nonsingular linear transformation h in the ~ variable. First of all,
h*a!; = (deth) ah!; from (7.31). Also, if S is the sphere 1~1 = 1, then

f a_n(X,~)
S
la!;l = ±f
h(S)
a_n(X,~) la!;l

by Proposition 7.3, since the ellipsoid h(S) is homologaus to S if h pre-


serves orientation; the minus sign applies if h reverses orientation. Thus,

f 1!;1=1
a_n(X, ~) Ia!; I = ± J1!;1=1 h* (a-n(X, ~) Ia!; I)
= ldethiJ a_n(X,h~) lah!;l·
1!;1=1
Finally, bywriting y = ljJ(x), ~ = IJJ'(x)tfl, we obtain

f1!;1=1
ä_n (x, ~) Ia!; lldnxl = I det ljJ' (x) I J1111=1 ä_n (x, ljJ' (x)t r]) la11 lldn xl

= JIIJI=1 ä_n(x,ljJ'(x)tr]) la11 11dnyl


= JIIJI=1 a_n(y, 17) la11 lldnyl, (7.36)

as the cycles 1~1 = 1 and 1171 = 1 arehomologaus for fixed x, and the de-
rivative terms do not contribute to the last integral. Therefore, the density
(7.34) is weH defined and independent of local coordinates. D
Theorem 7.6 (Wodzicki). The noncommutative residue (7.35) is a trace on
'P(M). IfdimM > 1, it is the only such trace, up to multiplication by a con-
stant.
Proof. Assurne that the symbols a, b are supported on a compact subset of
a chart domain U (since, in viewofTheorem 7.5, we canlater patch tagether
with a partition of unity). The commutator [A, B] = C of pseudodifferential
operators corresponds to the composition of the respective symbols a, b, c
given by the expansion (7.88)

Each term in this expansion is a finite sum of derivatives, either of the form
op/oxi or of the form oqjo~i· For instance, the leading term is

. n oa ob ob oa n 0 (. ob ) 0 (. ob )
-t i~ o~i oxi - o~i oxi = i~ oxi ta o~i - o~i ta oxi .
7.3 The Wodzicki residue 269

In particular, C-n(X, ~) = LJ=1 apj/axJ + aqjja~j. where PJ and qj are


bilinear combinations of the symbols a and b and their derivatives. Thus,
J1 ~ 1 = 1 apJ;axJ Ia~ I = aP1 ;axJ, where P1 is a smooth function of compact
SUpport within U; therefore the integral over u of aPj I axJ vanishes. At the
same time, J1 ~H aq 1 ;a~J Ia~ I= ObyLemma 7.4. Therefore, Wres[A,B] = 0.
To verify the uniqueness property, let T be any trace on 'P(M). The sym-
bol calculus shows that derivatives are commutators in 'P(M), since the
composition formula particularizes to

[~J.a] = -iaaa.
xJ

Hence T must vanish on derivatives, and thus T(a) depends only on the
( -n)-homogeneous term a_n(X, ~). To apply Lemma 7.4, we first average
this term over spheres:

tLn(X) :=}
~Ln
J
1~1=1
a_n(X,~) la~l.

where On is the standard volume of the sphere §n- 1 • Therefore the centered
( -n)-homogeneous term a_n (x, ~) -ä-n (x) I~~-n isafinite sum of deriva-
tives. Thus, T(a) = T(ä-n(X) 1~1-n). This means that h .... T(h 1~1-n) is a
linear functional on coo (U) that kills derivatives (with respect to the local
coordinates x 1 , .•. , xn), so it must be proportional to the Lebesgue integral.
In summary,
T(a) =C fu ä_n(X) ldnxl = C WresA
for some constant C. D

The hypothesis that n > 1 was used in appealing to Lemma 7.4; in the
case n = 1, the cotangent bundle T* M is disconnected, so there are actually
two residues, which may be linearly combined [491, 2.14) .
.,. We now look at some examples.
Proposition 7.7. Let ß be the scalar Laplacian on an n-dimensional compact
manifold. Then
2 rrn/2
Wresß-n/ 2 =On= f(I) (7.37)

is the standard volume of the sphere §n- 1 •

Proof. Note first that the inverse powers of ß are defined, modulo smooth-
ing operators. Since ß is second-order, the operator ß -n/ 2 has order-n and
its principal symbol is (giJ (x)~i~J )-n/ 2 , from (7.29). Since g(x) is positive
definite, we can make a local coordinate change y = 1JJ (x), ~ = IJJ' (x) t 11
270 7. The Noncommutative Integral

with the property that I.J.I'(x) := g(x) 1i 2; in the new coordinates, the prin-
cipal symbol is 1111-n, and (7.36) shows that

wresxA =On ldnyl = Ondetl/J'(x) ldnxl = Onp(x) ldnxl =On lv9 1.


Thus Wres~-n/ 2 := fMOnp(x) ldnxl =On, because we have assumed the
Riemannian density on M to be normalized. D

In particular, for the n-dimensional sphere with the standard volume,


-n/2 2(2rr)n
Wres~§" = On0n+1 = (n _ 1 )!'

on using the Legendre duplication formula f( I )f( n; 1 ) = .JIT(n -1)!/2n-1.


Naturally, things start to get interesting when we compute the Wresidues
of powers of the Laplacians for which the a_n symbol term is no Ionger
principal. In this context there appears a very important quantity in non-
commutative geometry, the residue of (the subprincipal symbol in) ~- ~ +1,
that we may call the action functional; its role in a variational principle will
emerge in Chapter 11.
Theorem 7.8 (Kastler-Kalau-Walze). In dimension n ~ 3, the action func-
tional on a Riemannian manifold is proportional to the integral ofthe scalar
curvature by a constant depending on n:

(7.38)

This is obviously a Riemannian analogue of the Einstein-Hilbert action


functional of general relativity for Lorentzian 4-manifolds. A proof is given
in the next section, where we introduce different incarnations of the action
functional. The relation (7.38) was first announced by Connes in the early
nineties; the result was proved independently by Kastler [281] and by Kalau
and Walze [268]. We recommend a Iook at the original papers.
For more general classical 'l'DOs operating on spaces of sections of vector
bundles over M, i.e., with matrix-valued symbols, the same arguments are
applicable, except that wresx A is now a matrix-valued density. Its matrix
trace is a scalar-valued density whose integral

WresA := JJ
M 1~1=1
tra-n(X, ~) la~lldx 1 A • • • A dxnl

defines a trace, unique up to multiples, on the algebra of classical symbols


of pseudodifferential operators whose coefficients are endomorphisms of
a given vector bundle E over M .
.,.. The Wresidue of a classical pseudodifferential operator A can alterna-
tively be obtained from the asymptotics of its distributional kernel near the
7.3 The Wodzicki residue 271

diagonal. This kernel kA (in local coordinates) is defined in Section 7.A:

(7.39)

where a(x, ~) is again the symbol of A. lt is of the form lA (x, x- y) for y


near x, where l(x, ·) is the inverse Fourier transform of a(x, ·) -see the
notationforthat in Section 7.B- and kA is smooth off the diagonal y = x.
The coincidence Iimit of the kerne} on the diagonal incorporates the follow-
ing logarithmic divergence (if the order is -n or greater), whose coefficient,
as Connes has remarked [92, 97], is precisely the noncommutative residue.
Theorem 7.9. The kernel of a classical pseudodifferential operator A of in-
teger order d has the following Iocal form near the diagonal:
d+n
kA(x,y) = L W-k(x,x- y)- uo(x) log lx- Yl + 0(1) as y - x,
k=l
(7.40)

where each w -k (x, z) is homogeneaus of degree - k in z and u 0 (x) Idn x I


is the Iocal expression of a 1-density on M. In fact, uo (x) Idn x I = wresx A.

Proof. The asymptotic expansion (7.3 2) of the symbol of A may be rewritten


as a finite sum
N
a(x.~) = L ad-J(x.~) +rN(x.~). (7.41)
j=O

for x E U c ~n and ~ E ~n. where the remainder term rN belongs to


the symboldass sd-N (U) of Definition 7.17. By (7.85a), this is integrable
in ~ whenever N > d + n. Applying the inverse Fourier transform in the
~variable to both sides of (7.41), we obtain

N
lA(X,Z) = L Wj-d-n(X,Z) + tN(X,Z),
j=O

where tN(x,z) remains bounded as z - 0.


Each term ad-J(X, ~)in (7.41) is homogeneaus of degree d- j for ~ * 0.
Homogeneaus distributions on ~n are discussed in detail in Section 7.B.
Before taking the inverse Fourier transform, we may regularize ad- J (x, ~)
at ~ = 0 if necessary, and the consequent behaviour of w J-d-n (x, z) at z =
0 is as follows. For j < d +n, the term WJ-d-n(x, z) is homogeneaus in z of
negative degree j - d- n, as required. For j > d + n, write k : = j - d- n > 0;
then, by (7.100), Wk (x, z) is of the form

Wk(X, z) = Vk (x, z) - Pk (x, zj lzl) lzlk log lzl,


272 7. The Noncommutative Integral

where Vk and Pk are continuous and are k-homogeneous in the second vari-
able; in fact, z .... Pk(x, z) is a k-homogeneous polynomial. Consequently,
Wk(X,Z)- 0 as Z - 0.
Finally, for j = d + n, (7.99) shows that wo(x, z) = -uo(x) log lzl where
uo(x) is independent of z. Since kA(x,y) = lA(x,x- y), equation (7.40)
follows.
On changing coordinates by a local diffeomorphism x .... 1J1 (x), kA (x, y)
is replaced by kA(!Jl(x),!J.I(y))L( x,y), where L(x,y)- ldetljJ'(x)l as
y- x. Furthermore, log I!Jl(x) -ljJ(y)l -log lx- yl as y- x, so that the
logarithmically divergent term -uo(x) log lx- yl of (7.40) is replaced by
-uo(!Jl(x)) I det !Jl' (x) I log lx - y 1. Thus, the local 1-density uo(x) ldnxl
is invariant.
To compute this 1-density, it is enough to observe that the regulariza-
tion of 1~1-n at ~ = 0 contributes -(2rr)-n On log lzl to the inverse Fourier
transform, on account of (7.99), and that a function of ~ whose average over
the unit sphere is zero contributes nothing, since the principal-value distri-
bution is homogeneous. In particular, a-n (x, ~)- ä_n (x) I~~-n contributes
nothing of this form. It follows that uo (x) = ä_n (x) and that the corres-
ponding 1-density is just wresx A. The proof of Theorem 7.5 confirms its
invariance under local coordinate changes.
Finally, notice that the argument Ix - y I of the logarithm is the local
expression of a Riemannian distance on M; we are free to change this dis-
tance to another (equivalent) one d(x, y), since log d(x, y) -log lx- yl is
bounded as y - x, so it may be absorbed in the 0 ( 1) term of (7.40). D

7.4 Spectral functions


Wodzicki residues are directly related to the asymptotic expansions of
spectral functions. The most venerable of these is the counting function
of the Laplacian,
N 6 (:\) := #{ /1 E sp~: /1 s; :\},

where the symbol # denotes cardinality. Spectral analysis was probably


born in 1911, when the young Hermann Weyl, under the instigation of
Hilbert, endeavoured to prove the famous estimate:

N6 (:\) _ On volM :\n!Z.


n(2rr)n

(Actually, Weyl considered the Dirichlet Laplacian for bounded domains


in !Rtn, but the result is of course the same for compact manifolds.) A source
of frustration for spectral analysts was the subsequent difficulty in mak-
ing higher-order expansions, due to heavy oscillations of the remainder.
In fact, for manifolds without boundary, the only thing that can be said in
general, in order to improve on this estimate, is that the remainder term
7.4 Spectral functions 273

is of order ,\ <n-ll/ 2 . Carleman proposed to "average out" Nt!. by using the


Laplace transform, i.e., the partition function

where 0 ~ ,\ 1 ~ ,\z ~ ... is the sequence of eigenvalues of ß, counted with


their multiplicity. This was found to possess an asymptotic expansion at
all orders as t l 0. Another popular spectral function is the zeta function

(tJ.(s) :=So"" ,\ 5 dNt!.(,\).

Taking things at their root, the spectral theorem teils us that any (Borel)
function of a selfadjoint operator H can in principle be calculated from the
spectral family of projectors EH(,\) associated to it; more precisely, what
is used is the derivative, which exists in the distributional sense:

d H (t\1 )·=dEH(,\)
• d,\ .

To be concrete, let J{ be the space of square-integrable sections of an Eu-


clidean vector bundle over a compact Riemannian manifold M, and let H
be a positive elliptic pseudodifferential operator of positive order on J{,
with domain DomH. Then H has compact resolvent. We pause to discuss
this matter. Let T be an operator on a Hilbert space that is either bounded
or unbounded selfadjoint (or normal), so that the spectral theorem applies
toT. The Operators R;\(T) := (,\- T)- 1 , for ,\ f/ sp(T), are bounded, and
they are linked by the resolvent equation

R;>..(T) - R 11 (T) = (J.l- ,\) R;\(T)R 11 (T). (7.42)

In particular, R;\(T) and R 11 (T) commute, and if one is compact then so is


the other. To prove (7.42), just write R;>.. (T)- R 11 (T) = R;>.. (T) (J.l- T)R 11 (T)-
R;>.. ( T) (,\ - T)R 11 ( T) and simplify. For more information on resolvents, see,
for instance, [383]. We say that the operator H has compact resolvent if
one, and therefore any, R;>.. (H) is compact. When H is a positive operator,
it suffices to show that (1 + H)- 1 is compact.
In the present context, ( 1 + H) -l is an elliptic pseudodifferential ope-
rator of negative order, whose compactness is guaranteed (see the end of
Section 7.A). To avoid annoying trivial exceptions, we shall assume that
0 is not an eigenvalue of H. Therefore, H has discrete spectrum of finite
multiplicity; if 0 < Al ~ ,\z ~ · · · is the complete set of its eigenvalues,
with the corresponding orthonormal basis of smooth eigenfunctions u i,
the spectral family is given [253] by

EH(,\):= 2: luj)(ujl.
AjSA
274 7. The Noncommutative Integral

and its derivative is

dH(t..) := 2::1uj}(ujl8(.\- Aj). (7.43)


j

This spectral density is a distribution with values in L(DomH, Jf). The


defining properties of EH(.\) are

(7.44)

J:
These are integrals in the weak operator sense, that is,

(y I Hx} = .\d(y I EH(.\)x},

for x E DomH, y E Jf. Then (7.44) becomes, in distributional notation,

The spectral density may be used to construct the functional calculus


for H. Indeed, we can define cp(H), whenever cf> is a distribution, by

where x E Domcp(H) means that the evaluation ((y I dH(A)x},cp(.\));.., is


defined for all y E :Jf. As important special cases, we mention the "zeta
operator"

(7.45a)

the "heat operator"

(7.45b)

and the I-parameter unitary group generated by H, which is just the Fourier
transform of the spectral density,

The symbolic formula

recommends itself, and we shall employ it from now on.


.,. In an ideal world, one would study the asymptotic behaviour of the spec-
tral density, and from there derive the asymptotic expansions for the other
spectral functions. This seems a bit desperate, as 8(.\- H) is a quite sin-
gular object (in the .\-variable). However, it has lately been realized, thanks
mainly to the work of Ricardo Estrada [164,166,168] that local asymptotic
7.4 Spectral functions 275

expansions at all orders do exist for the spectral density and the kernels
associated to it, in a certain generalized Cesaro sense; indeed, these as-
ymptotic expansions give rise to more or less benevolent expansions for
functions of H, according to the smoothness of each such function. In the
case of the counting function, however, we have to compose H with the
Heaviside function, which is not smooth.
Definition 7.12. A distribution f E 1J' (~). is of order ß at infinity in the
Cesaro sense, for ß E ~ \ {-1, -2, ... }, written as
j(t) = O(tß) (C) as t - oo,
if for some NE ~. there exist an Nth orderprimitive fN of fand a polyno-
mial p of degree at most N- 1, suchthat fN is locally integrable for large t
and the relation
fN(t) = p(t) + O(tN+ß) as t - oo
holds in the ordinary sense.
By an Nth order primitive we mean an fN whose Nth distributional de-
rivative equals f. When ß isanegative integer, the Cesara theory becomes
much more complicated, but we do not need that case for our purposes.
An evaluation in the Cesara sense
(j(t),</>(t))r =L (C) (7.46)
means that the distribution g = f<P has a primitive G suchthat G(t)- L =
o(1) (C) as t- oo. It turnsout [168] that the evaluation
('L.~= 1 an8(t-n),</>(t))t =L (C)

holds if and only if L.~=l an </>(n) = L in the Cesara sense of the theory of
summability of series. Likewise, if f is locally integrable and supported in
aninterval[a, oo ), then (7.46) is valid if and onlyif J;
j(t)</>(t) dt = L in the
Cesara sense of the theory of summability of integrals: a merit of Estrada's
Cesara theory is that it works equally well for operators with discrete or
continuous spectrum.
The main space of symbols for Cesara theory is the space X(~) of
Grossmann-Loupias-Stein symbols, already considered in Definition 3.23
in connection with the Moyal asymptotic morphism.
Periodic distributions with zero mean belang to the dual space X' ( ~): if
f is such a distribution, then, for n suitably large, the periodic nth-order
primitive with zero mean fn of f is a continuous function and defines the
evaluation of f at </> E X by a convergent integral:
(j(x),</>(x))x := (-l)n(fn(X),</>(nl(x))x.

In this case all the moments are zero. Distributions that have a moment
asymptotic expansion are characterized by the following crucial result [164,
168].
276 7. The Noncommutative Integral

Theorem 7.10 (Estrada). Let f E V'(~) be a distribution. Then f EX'(~)


if and only if f satisfies

j(x) = o(lxl-oo) (C) as x - ±oo,

if and only if there are constants #Jk for k E ~ such that

j(Tx) _1Jo8(x) _1118'(x) + 1128"(x) _ ... as T - oo


T T2 2!T3

in the weak sense. Moreover, if f E X'(~) and </> E X(~). the evaluation
(j (x), </> (x)) x is Cesaro summable. a

We want to study the asymptotic behaviour of 8(,\- H). Consider the


smooth domain
n Dom(Hk);
00

Dom 00 (H) :=
k=l
when H is selfadjoint, Domoo (H) is dense in J-{ [384, X.6]. The density of
Dom"' (H) has momentaus consequences. The relation

holds in the space of continuous linear maps L(Dom"' (H), J-f). Therefore,
8(,\ -H) belongs to the space X' (~-L(Dom 00 (H), J-f)), the following mo-
ment asymptotic expansion holds:

and so 8(,\- H) vanishes to infinite order at ±oo in the Cesära sense:

8(,\- H) = o(IA.I-oo) (C) as 1,\1- oo. (7.47)

(Of course, (7.47) is trivial when His bounded.)


.,. Our task is to figure out the symbol for a spectral density. We first ex-
amine a simple case: an elliptic operator with constant coefficients on Eu-
clidean space that is essentially selfadjoint. Let H be the selfadjoint exten-
sion of such an operator. The assumption of constant coefficients implies
that (T(Hn) = (T(H)n; we may therefore define

(T(8(,\- H)) := 8(,\- (T(H)).

Indeed, by regarding 8 (,\ - H) as a distribution with values in the space


of operators mapping Domoo (H) into J-{, the functional calculus gives the
identities
kEN,
7.4 Spectral functions 277

while
kEN.

In the general case of nonconstant coefficients, we define a ( 8 (i\ - H))


as a distributional symbol that satisfies the relations

kEN. (7.48)

In the light of Theorem 7.10 and (7.47), we make the following Ansatz for
the symbol:

a(8(i\- H))- 8(i\- a(H))- q 1 8'(i\- a(H)) + q2 8"(i\- a(H))


- q3 8"'(i\- a(H)) +··· (C), (7.49)

where the coefficients qk depend on H, but not on i\; to determine what


they must be, we recompute (7.48) for each k E N. What happensisthat
any power of i\ will pair nontrivially with only a finite number of terms in
the development (7.49), since

L""oo p (i\) Ö(r) (i\ - t) di\ = (-l)r p<r) (t)

for any polynomial p. Thus, for k = 1, 2, 3, ... we get

a(H) = a(H) + qr,


a(H 2) = a(H) 2 + 2qla(H) + 2q2,
a(H 3 ) = a(H) 3 + 3qla(H) 2 + 6q2a(H) + 6q3,
etcetera. This is easily solved for each qk in turn: obviously q 1 = 0, and

and so on.
Exercise 7.12. Work out q 4 .

~ To compute traces of pseudodifferential operators, one needs the coin-


cidence Iimit for their kernels. If Ais a pseudodifferential operator, we can
rewrite its kernel (7.39) as

In general, the coincidence Iimits are not small in the Cesara sense. How-
ever, in our case, ellipticity actually implies that dH(x, y; i\), the kerne} of
8(i\- H), is smooth in (x, y), on account of (7.43). Letting y - x, we get
the coincidence Iimit on the diagonal:

(7.50)
278 7. The Noncommutative Integral

Let us try, then, to compute the first terms of the Cesara asymptotic ex-
pansion, as ,\ - oo, of the coincidence limit for the kernel dH (x, y; ,\)
of 8(,\ - H), where H is a positive elliptic operator. We shall use nota-
tion appropriate to the scalar case, and abbreviate c := O"(H). From (7.49)
and (7.50), we get

dH(x,x;,\)
- (2rr)-n(l, 8(,\- c(x, ~)) + qz(x, ~) 8" (,\- c(x, ~))- · · · )~ (C).

In polar coordinates on the cotangent fibres, ~ = l~lw with Iw I = 1, the


right hand side becomes

(2rr)-n Jlwl=l (l~ln-l, 8(,\- c(x, l~lw))


+ qz(x, l~lw) 8"(,\- c(x, l~lw))- · · · ) 1 ~ 1 dw.

To evaluate this, it is helpful to recall the change-of-variables formula


,8(t-to)
8 (j(t)) = L. lf'(to)l ' (7.51)

where the sum is taken over the (simple) roots of the equation j(t) = 0.
In the present case, t = 1~1 is taken positive, and weshall assume that the
equation c(x, l~lw) = ,\ has a unique positive solution 1~1 = 1~1 (x, w;,\).
We therefore need to compute

(2rr)-nJ , l~ln-l(x,w;,\)
lwl=l c (x, l~l(x,w;,\)w)
+ ~(qz(x, l~l(x,w;,\)w)l~ln-l(x,w;,\)) _ ... dw (7.52)
(),\2 c'(x, l~l(x, w;,\)w) '

where c' denotes the derivative of c with respect to the I~ I variable.


The asymptotic expansion (7.32) of O"(H) may be written as

Solving the equation c (x, I~ Iw) = ,\ amounts to a series reversion. Let us


assume for a short while that the order d of His 1. We then expect
. 1 co(x, w) -1
l~l(x,w,A)- (
Cl X,W
),\- (
Cl X,W
)-c_I(x,w),\ +···.
The integration over Iw I = 1 will give an expansion:

The calculation of ai(x), a 2 (x), and so on, is quite involved. However,


at the first order qz and subsequent terms do not contribute, and then
7.4 Spectral functions 279

ao -which depends on the principal symbol only- equals the coefficient


of .\n-l in the expansion of flwl=ll~ln-l(x,w;.\)/ci(x,w)dw, which is
flwl=l c1 (x, w)-n dw, so that ao(x) ldnxl = wresx H-n.
Returning to the general case, where H is a positive pseudodifferential
operator of order d, observe that A := H 1id isapositive pseudodifferential
operator of first order. Setting J.1 := .\ 1 /d gives

andthus
dH(x,x;.\)- (2TT)-n(a 0 (x).\<n-dl!d + a!(x).\<n-d-1)/d
+ az(x).\<n-d-2)/d + ... ) (C), (7.53)

the leading term in this expansion now being

ao(X) ldnxl = d1 wresx H-n!d.


It turnsout that the development (7.49) gives ever lower powers of .\in
the asymptotic expansion of dH (x, x; .\).
Exercise 7.13. Show that the order in .\ of qz (x, I~ I(x, w; .\) w) is at most
2d -1, so that its higher-arder contribution to the development (7.53) is in
principle to a 1 ; that the order in.\ of q 3 is at most 3d- 2, so it contributes
to az at the earliest, and so on. <>
Note that Weyl's estimate follows from our calculation, at least in the
weak sense: since wresx tl.-n/Z =On lv9 1,

Actually, the first term of the expansion is valid in the ordinary sense, not
merely in the Cesara sense, as the following estimate shows.
Proposition 7.11. Keeping the hypotheses on H, the asymptotic behaviour
of its eigenvalues is
(7.55)

Proof. Choose and fix E > 0. Pick smooth functions g, h such that 0 ~
g(.\) ~ 1, 0 ~ h(.\) ~ 1 and also g(.\) = 1 on [0,1- E], g(.\) = 0 for .\ > 1,
and h(.\) = 1 on [0,1 ], h(.\) = 0 for .\ > 1 + E. The indicator function of the
interval [0,1] is then sandwiched between the positive smooth functions
g and h; therefore
280 7. The Noncommutative Integral

By pairing (7.53) with g and h, we can find constants c and 11 = /-I(E) such
that if ,\ > /-I(E), then

and (7.55) follows immediately. D

For higher-arder terms this control of the indicator function no Ionger


works, so N(,\) possesses no ordinary asymptotic expansion. On the other
hand, because rtt. is a very weil behaved function, (7.45b) yields the bona
fide asymptotic expansion for the kerne! KH of the heat operator rtH. For
the coincidence Iimit we get

2: ak(X) t(-n+k)!d
00

KH(t,x,x)- as t l 0, (7.56a)
k=O

with

(7.56b)

We shun here the case in which negative integer powers of ,\ appear in the
Cesära expansion, leading to logarithrnic terms in the development (with
coefficients always given in terms of the ak), for the reasons indicated in
Section 7.B. For that, consult [168]. When necessary for clarity, we shall
write ak (x; H) for the coefficients. A bit more generally, if P is pseudodif-
ferential of order 0, the coincidence Iimit of the kerne! KH,P of the operator
Pe-tH is

2: lXk(x;H,P) t(-n+k)/d.
00

KH,P(t,x,x)- (7.56c)
k=O

.,. One of the deepest ofWodzicki's discoveries can be rephrased as follows:

(7.57)

not just the first, but all the coefficients of the Cesära development for the
spectral density are Wodzicki residue densities! This particularly means, in
view of (7.56), that one can read off the result of the Kastler-Kalau-Walze
theorem from the known value of a2, a remark probably made first by
Ackermann [2].
To sustain (7.57) by a direct proof would take us too far afield. However,
we can provide an indirect argument. The result of Wodzicki (and thus
Ackermann's remark) was actually framed in the context of zeta functions.
Let us now address them. The dassie paper on zeta-function theory is that
7.4 Spectral functions 281

of Seeley [423]. He considers an elliptic (determined system of) pseudo-


differential operator(s) of positive order; no selfadjointness condition is
required. The idea is to define H 5 , for s complex, as a Dunford integral

H 5 := 2 1 . i ,V(,\-H)- 1 d,\, (7.58)


rrt Jr
where the contour r must wind araund the spectrum of H while avoiding
the branch point of the function ,\5 at ,\ = 0. For this to work, H must
possess a ray arg,\ = e of minimal growth, i.e., a ray on which there lies
no (eigen)value of the principal symbol ad(x,~). One makes r startat oo,
coming toward the origin along the ray arg,\ = e of minimal growth, turn-
ing clockwise in a small circle araund the origin, and going back to oo along
the ray. Naturally, the results depend somewhat on the ray chosen; it was
the study of the resulting "spectral asymmetry" that led Wodzicki to the
discovery of the noncommutative residue.
For positive operators, however, these elaborate precautions in the def-
inition are unnecessary, since r may wind araund the positive real axis
with any angle of aperture, there is no spectral asymmetry and Seeley's
definition coincides with (7.45a), of course. We should remark that H 0 , as
given by (7.58), does not equal1, but rather 1 -Po, where Po is the (finite-
dimensional) projector on ker H. This happens since the contour r does
not enclose the possible eigenvalue ,\ = 0 of H. All the complex powers H 5 ,
in particular H- 1 , also vanish on ker H, though they do satisfy the usual
semigroup law Hs+t = H 5 Ht on (ker H).L. We shall understand inverses,
such as ß - 1 , in this sense.
For d > 0, and when ~ s is negative enough (smaller than -n/ d), H 5 has
a very weil behaved kernel, whose trace gives the numerical zeta function
(H (s). Seeley was able to prove that kHs (x, y) extends to a entire function
of s for x * y, whereas kHs(x,x) extends to a meromorphic function. To
describe this, Seeley used the pseudodifferential calculus to construct a
good approximation to (,\- H)- 1 • Let
00

B;>.. - 2:: b-d-k;J..


k=O
be defined by
a(B;>..) o a(,\- H) = 1;
then he was able to prove that H 5 is a pseudodifferential operator with
symbol

a(H5 ) = ~ -2 1 . i A5 b-d-k;J.. d,\. (7.59)


k=O rrt Jr
Each pole of kHs (x, x) is simple and is related to a particular term in this
expansion; the poles are located at
k-n
S=-d-, k=0,1,2, ... ,
282 7. The Noncommutative Integral

as we thoroughly expect from (7.53), at least for positive operators. Be-


yond the first or second term, the expressions of the b-d-k;J.. become very
complicated; the associated residues are given by

Res
s=(k-n)/d
kHs(X,X)=d( 2 i)
TT n+
J
1 1~1=1 fi\ 5 h-d-k;;>..di\lu~l.
Jr
(7.60)

We know now that, in fact [490],

Res
s=(k-n)/d
kHs(x,x) =- d ( 21 )
TT n
J 1~1=1
CT-n(H(k-n)fd)(x, ~)I er~ I. (7.61)

and the (ordinary) residues at the poles of the zeta function are essentially
our ak. Seeley states that there is never a pole at s = 0 (rather obvious for
us, as the noncommutative residue must vanish), nor at 1, 2, .... This asser-
tion is true for differential operators, whose positive powers have vanishing
Wodzicki densities, but false for pseudodifferential operators in general.
The values of kHs (x, x) at s = 0 can also be computed by similar local for-
mulae -which are instrumental in the proof of the Atiyah-Singer theorem
by zeta functions.
For the firstpole of the zeta function '(,Hat s = -n/d, the equality of the
right hand sides of (7.60) and (7.61) is immediate: the leading term of the
expansion (7.59) of u(H-nfd) is

_21. f i\-nld(i\- ad)-1 di\ = ad(X,~)-nld,


m Jr
and the associated residue is
-d-1(2rr)-n f ad(X,~)-nld lcr~l. (7.62)
Jl~l=1
But, if H is a positive elliptic operator of order d, the power H-n!d is
an elliptic pseudodifferential operator of order -n, and wresx H-n/d
Jl~l=1 ad(X, ~)-nfd lcr~lldnxl.
1t is time to clinch the main result of this section.
Theorem 7.12. Let H be a positive elliptic pseudodifferential operator of
order-n over an n-dimensional compact manifold M without boundary.
Then the zeta residue of H, defined as
ResH :=Res '(,H(s),
s=1
is proportional to the Wodzicki residue
1
ResH = n (2 rr)n WresH.

Proof. Theinverseoperator H- 1 is of positive order, so the previous theory


applies to it. Now, Ress=1 '(,H(s) = - Ress=-1 '(,H-1 (s) = n- 1(2rr)-n WresH,
in view of (7.62). D
7.4 Spectral functions 283

Wodzicki's formula (7.61) can be rewritten as


1
Res kH-s(x,x) ldnxl = wresxH<k-n)/d. (7.63)
s=(n-k)/d d(2rr)n

The relation between the residues at poles of the zeta function and the
coefficients of the heat kernel expansion is derived in many books by a
routine Mellin transform. We review this briefly.
Definition 7.13. Let f E co (0, oo) be a smooth function satisfying lf(t) I s
ce-til. for some .\ > 0 and t large enough, and suppose that

j(t) - ,L bktk 12 + clog t


k;,;-n

for some n as t l 0. lts Mellin transform :Mf is the meromorphic function

:Mj(s) = [(15) fooo ts-1 j(t) dt. (7.64)

Examples are

1 = _1_ foo u-1 dt 1 foo ts-1 dt


[(s) Jo et ' l;;"R(s) = [(s) Jo et=T·
The first formula has already been used, and l;;"R derrotes the Riemann zeta
function.
From the transform formula

(7.65)

it is clear that :M[Pe-tH] = PH-s, for P bounded. Using the asymptotic


development (7.56c) and formula (7.64), by splitting the integration into
the intervals [0, 1] and [1, oo), it follows that

R k ( ) OCk(x;H,P)
es
s=(n-k)/d PH-sX,x =r(( n- k)/d)"

Putting this tagether with (7.56b) and (7.63), we obtain (7.57). Moreover,
from the well-known [197] result oc2(x;ß) = s(x)/6(4rr)ni 2, where s(x)
derrotes the scalar curvature of M, there ensues the Kastler-Kalau-Walze
formula:
Wresß-n/ 2+ 1 = JM 2a2(x) ldnxl,

where, using normal coordinates,

2(2rr)n 01.2 2(2rr)ns


Za 2 = f( ~- 1) = 6(4rr)n/2f( ~- 1)
284 7. The Noncommutative Integral

The perceptive reader will have noticed an apparent contradiction, in


that for even-dimensional manifolds Wres ß -nl 2 +k can be different from
zero only as lang as - ~n + k < 0, whereas the asymptotic development
for the heat kernel has an infinite number of terms. But then, as explained
also in [164, 168], the residues are replaced by "moments". For n = 2, the
coefficient oc2 is already a moment and cannot be computed by a Cesara
development. This strikingly different behaviour of the odd-dimensional
and the even-dimensional cases is concealed in the uniformity of the heat
kernel method, but is reflected in the behaviour of the corresponding zeta
functions: there are infinitely many poles in the odd-dimensional case, and
only a finite number in the even-dimensional case.
Exercise 7.14. Prove the Kastler-Kalau-Walze formula (7.38) from a direct
calculation of the Cesara development coefficient a2 (x) of (7.53), using the
formula (7.52). o

7. 5 The Dixmier trace


An "infinitesimal" operator T on an infinite-dimensional Hilbert space J{
with a countable orthonormal basis is one for which, roughly speaking,
II T II < E for any E > 0. Of course, the only operator satisfying this condi-
tion, as stated, is T = 0. lf, however, we first shave off a finite-dimensional
subspace of J{ before computing the norm of T, then there is an ample
supply of infinitesimals, namely, the space X of compact operators on Jf.
Tobemore precise, an operator T E .L (Jf) could be called "infinitesimal"
if, for each E > 0, there is a finite-dimensional subspace E c J{ such that

IITIPII < E. (7.66)

If h is the orthogonal projector with range E, then PET and TPE are ope-
rators of finite rank, and

IITip II = IIT(l- PE )II = IIT- TPEII < E,

so T is a norm-limit of a family of finite-rank operators, and thus is compact.


If Sk ( T) denotes the k-th singular value of T (see Section 7.C), occurring in
the canonical expansion ITI = Lhosk luk)(ukl. then the condition (7.66)
is satisfied by E := span{ Uk: Sk(T) ~ E} for any E > 0, so the infinitesimals
of .L(Jf) are precisely the compact operators.
On the ideal .L 1 = .L 1 (Jf) of traceclass operators (see Definition 7.26),
the trace Tr T := Lk=o(Uk I Tuk) is absolutely convergent (and indepen-
dent of the orthonormal basis {uk} of Jf). It is a positive linear functional
on .L 1 that satisfies Tr(TS) = Tr(ST) whenever TS and ST lie in .L 1 . This
"trace property" makes Tr Iook like a good candidate for a "noncommuta-
tive integral", and indeed, an analogue of Lebesgue integration theory using
7.5 The Dixmier trace 285

this trace was developed in the fifties by I. E. Segal [427]; for instance, the
Schatten ideal LP is the analogue of the usual LP space. However, this the-
ory turns out tobe unsatisfactory for our purposes, since the infinitesimals
in L 1 are "too small": we want infinitesimals of order 1, i.e., those whose
eigenvalues go to zero like 1/n, tobe integrable. Later, Dixmier [136] dis-
covered another kind of trace functional with a larger domain of compact
operators, which turns out to be just right for noncommutative geometry.
To construct it, we begin with the partial sums of the singular values:

an(T) := so(T) + si(T) + · · · + Sn-dT).

Clearly, T is traceclass if and only if the increasing sequence {an(T)} is


bounded, and IITII1 = limn-oo an(T). Each an is a norm, by Corollary 7.33,
and can be characterized as

an(T) = inf{ IIRII1 + niiSII: R,S EX, R + S = T }.

This is proved in Proposition 7.34. A useful trick [94, 113] is to extend the
sequence of norms {an} to a family of norms on X indexed by .\. E [0, oo),
by defining

a;~.(T) := inf{ IIRII1 +AllS II: R,S EX, R + S = T }.

Since IIRII1 ~ IIRII in general, it follows that a;>.(T) = AIITII in the case
0~.\~1.

Exercise 7.15. If n = L.\J is the integer part of .\., so that .\. = n + t with
0 ~ t < 1, show that
(7.67)

so.\. .... a;~. (T) is piecewise linear. Conclude that each a;>. satisfies the trian-
gle inequality. 0
Indeed, the piecewise linear function .\. .... a;~. ( T) is concave, since the
increments an+ I (T)- an (T) =Sn (T) decrease as n - oo,and so any chord
of its graphlies below the graph. If T is traceclass, the function is bounded
above by IIT11 1. Weshall be interested in the case that the function .\. ....
a;~. ( T) grows logarithmically.

Lemma 7.13. If A, B EX are positive, then am+n (A + B) ~ am (A) +an (B)


forallm,n.

Proof. By Lemma 7.32 and the positivity of A and B, we can find subspaces
FandEwithdimF = manddimE = nsuchthatam(A) = sup{Tr(PFAPF):
dimF = m} and an(B) = sup{Tr(PEBPE): dimE = n }. The subspace E +F
has dimension r ~ m + n, so that if E + F ~ G with dim G = m + n, then

Tr(hAPF) + Tr(PEBPE) ~ Tr(PcAPc) + Tr(PcBPc) = Tr(Pc(A + B)Pc).


286 7. The Noncommutative Integral

Conversely, if Gis an (m+n)-dimensional subspace and if F, E are any sub-


spaces of respective dimensions m and n with E + F ~ G, then Tr(PFAPF) +
Tr(PEBPE) :5 Tr(Pc(A + B)Pc). Taking suprema over all possible E, F, G
yields the desired inequality. D

Exercise 7.16. Extend Lemma 7.13 to nonintegral values of the norm pa-
rameter: if A and B are compact positive operators and if J..l, A ~ 0, then
a 11 +t\ (A + B) ~ a 11 (A) + CT" (B). 0

In particular, we conclude that, for A, B E X positive and A ~ 0,

(7.68)

a"
These inequalities suggest that for large A, (log A) - 1 is almost, though
not quite, an additive functional on the cone of positive compact operators.
If it were actually additive, it would be a trace, since a" (V AVt) I log A =
a" (A) I log A for V unitary, and it could be extended by linearity to a trace
functional on operators that arenot necessarily positive. Weshall now iden-
tify the domain of this putative trace, and obtain the trace itself by suitably
averaging the norms T - a" (
T) I log A.

Definition 7.14. The Dixmier ideal of compact operators is defined by

. . U,\ ( T)
L 1 + ..= { TEX.IITII1+·=sup- } (7.69)
1 1 <oo. t\2:e og t\

Clearly II · ll1+ is a norm, satisfying a"(T) :5 IITII1+ logA for A ~ 0. Since


A- a"(T) is bounded forT E L 1, it is clear that L 1 c L 1+. lt turnsout
(see Section 7.C) that L 1+ c LP for any p > 1, whence the notation for the
Dixmier ideal.

Consider the following Cesaro mean of the function a" (T) I log A:
TA(T) := -1 1 1
ogr\
J"
3
au
--(T)
-- du
-
1ogu u
for A ~ 3. (7.70)

Then T,\ (S + T) :5 T,\ (S)+T" (T)for S, TE L 1+ (see Exercise 7.15). This is still
not an additive functional, but it has an "asymptotic additivity" property:
the following Iemma is due to Connes and Moscovici [113].

Lemma 7.14. If A,B E L 1+ arepositive operators, then

log log
T,\(A + B)- T,\(A)- T,\(B) = 0 ( logA
A) as A- oo. (7.71)
7.5 The Dixmier trace 287

Proof. From(7.68)and(7.69),au(A+B) 5 (IIAIII++IIBIII+) loguforu ~ e.


Using (7.68) again with .\ > 3,

0 5 T,\ (A) + T,\ (B) - T,\ (A + B) 5 --


1 J" O"zu (A + B) - O"u (A + B) -
du
-
log .\ 3 log u u
= _1_ f 2" ( O"u (A + B) _ O"u (A + B)) du
log.\ J6 log ~u log u u

__ 1_
log.\
(J"3 _J6f2") O"u (A + B) du.
logu u
( 7.7 2 )

Using the estimates

f 2" (
J6
logu -1) du= J"(log2u -1) du=
log ~u u 3 logu u
J" log2du
3 ulogu
< log2loglog.\

and

I (J,\- J6r2") O"u(A + B) du I= I (J 6- f2,\) O"u(A + B) du I

u:
3 log u u 3 " log u u

5 (IIAIII+ + IIBIII+) + J:") duu = 2log2 (IIAIII+ + IIBIII+).

the right hand side of (7.72) is majorized by


2 + loglog .\
(IIAIII+ + IIBIII+) log2 log.\ •

and (7.71) follows. 0

Of course, the constant 3 in (7.70) could be replaced by any nurober


greater than e, the point being that on the interval [3, oo ), the function
log log .\/log.\ is bounded and falls to zero at infinity. Thus .\ .... T,\ (A)
lies in Cb([3, oo)), and the left hand side of (7.71) lies in the C*-subalgebra
Co( [3, oo) ).
Definition 7.15. In the quotient C*-algebra Boo := Cb([3, oo))/Co([3, oo)),
let T(A) E Boo be the dass of .\ .... T,\(A), for A a positive element of LI+.
Then T is additive and positive-homogeneous, i.e., T(A + B) = T(A) + T(B)
and T(cA) = cT(A) for c ~ 0.
By Lemma 7.35 of Section 7.C, the ideal LI+ is linearly generated by its
positive elements: if T = yt in LI+, there are spectral projectors P± so
that P+ + p_ = 1 and T = P+ITIP+- P-ITIP- is a difference of positive
elements of LI+. Therefore, T extends to a linear map from LI+ to Boo by
defining T(T) := T(P+ITIP+)- T(P-ITIP-) if T is selfadjoint, and then
T(T) := ~T(T + Tt) + iT(iTt- iT) for a general T.

Exercise 7.17. Show that T(VAVt) = T(A) for any positive A E L 1+ and
unitary V; conclude that T(ST) = T(TS) forT E L 1+ and SE L(H). 0
288 7. The Noncommutative Integral

To define a trace functional with domain L 1+, all we have to do is to


follow the map T: L 1+ - Boo with a state w: Boo - ((, given, say, by a
positive linear form on Cb([3, oo)) that vanishes on C0 ([3, oo) ), normalized
by w(l) = 1. (The commutative C*-algebra Boo has plenty of states; the
Gelfand-Naimark theorem guarantees that, since each character is a state.
The states of Boo correspond to generalized Iimits, as A - oo, of bounded
but not necessarily convergent functions.)
Definition 7.16. To each state w of the commutative C*-algebra Boo there
corresponds a Dixmier trace Trw T := w (T ( T)) whose domain is L 1+.
We say that T E L 1+ is measurable if the function A - TA ( T) converges
as A - oo. Since in that case -and only in that case- Tr w T = lim.\ _"" TA ( T),
independently of w, we define the Dixmier trace of a measurable compact
operator by
Tr+ T := lim T.\ (T) .
.\-oo

Any traceclass operator T is measurable, since A - a" (T) is bounded by


IITII 1 , so that a.\(T)/logA- 0 and thus also TA(T) - 0, as A- oo. This
implies that Trw T = 0 for any traceclass operator T (so the Dixmier traces
are quite different from the usual trace!).
Actually, the Dixmier traces vanish on an operatorideal somewhat larger
than L 1. Since TA(T) ::5 IITII 1+ for all A ~ 3 by construction, it follows that
I Trw Tl ::5 II Tll1+ for any w. That is to say, the Dixmier traces define contin-
uaus linear functionals on the Banachspace L 1 +, that vanish (in particular)
on finite-rank operators. Therefore, they also vanish on the ideal LÖ +, which
is the closure for the 11.11 1+ norm of the ideal of finite rank operators.
The norm of L 1+ is a symmetric norm, i.e., IIATBII1+ ::5 IIAIIIIT111+ IIBII
if TE L 1+ and A,B are any bounded operators. Any such "symmetrically
normed operator ideal" is known to be separable if and only if the finite-
rank operators are dense [200, Thm. 3.6.2]. We shall shortly exhibit exam-
ples of measurable operators on which the Dixmier traces do not vanish;
from that we conclude that the Dixmier ideal L 1 + is not separable, even
though it is sandwiched between Schatten classes which are themselves
separable operator ideals. Symmetrie norms and their operator ideals are
discussed at greater length in Section 7.C.
Proposition 7.15. The measurable compact operators form a closed sub-
space of L 1 + that contains LÖ+ and is invariant under conjugation by boun-
ded invertible operators on J{ (and thus independent of the choice of scalar
product).

Proof. The condition of measurability persists under linear combinations


and norm Iimits, because each Trw is linear and continuous on L 1+. Ele-
ments of LÖ + are clearly measurable since Tr w T = 0 there. If S is invertible
in L(Jf), then Trw (STS- 1) = Trw (T) by Exercise 7.17, so that STS- 1 is
measurable if T is measurable. 0
7.5 The Dixmier trace 289

For future use, we prove the Hölder inequality for the Dixmier trace. Let
us recall the ordinaryHölderinequality: if {ak} and {bk} arefinite or infinite
sequences of positive numbers, p > 1 and q = p I (p - 1), then

(7.73)
The case p = 1, q = oo can be included by interpreting the last term as
supk lhl.
Proposition 7.16. Let T, S E X besuch that TP, Sq E L 1+, with p > 1 and
q = p I ( p - 1); or T E L 1+ and S is bounded. The following inequality holds:
Trw ITSI 5 (Trw ITIP) 11 P(Trw 1Siq) 1iq, (7.74)

where the right hand side, {or p = 1, is Trw ITIIIS 11.


Proof. First of all, we establish Horn's inequality [254]:
n-1
an(TS) 5 L Sk(T)sk(S). (7.75)
k=O

This follows from (5.32):


n-1
n
k=O
Sk(TS) 5 n
n-1

k=O
sk(T) Sk(S).

Taking logarithms gives


:Z::logsk(TS) 5 :Z::log(sk(T)sk(S)),
k k

and the convexity of the exponential function then yields (7.75). In more
detail, let Xk := logsk{TS), Yk := log(sk(T)sk(S)); one can check, case by
case, that 2:r= 1(xk- t)+ 5 2:r= 1(Yk - t)+ for all t E ~; then the integral
representation ex = f:'oo (x- t) +et dt allows to conclude that 2:r= 1 exp Xk 5
2:r= 1 exp Yk, which is (7.75). (Fora Ionger but more elementary argument,
see [34, Il.3 ].)
In particular, for the case p = 1, an(TS) 5 an(T) IISII.
Using now the ordinary Hölder inequality for p > 1, it follows that
CTn(TS) 5 CTn(ITIP) 11 Pan(ISiq) 11 q.

If n 5 ,\ < n + 1, the interpolation (7.67) allows us to replace n by .\:


a;o.(TS) 5 (1- t)an(ITIP) 11 Pan(ISiq) 11 q + tan+diTIP) 11 Pan+diSiq) 11 q
5 ((1- t)an(ITIP) + tan+1(1TIP)) 11 P
· ((1- t)an(ISiq) + tan+diSiq)} 11 q
= a;o. (I TIP ) 11 Pa;o. ( ISiq) 11 q,
290 7. The Noncommutative Integral

where we have used the ordinary Hölder inequality again.


Dividing by log.\ and averaging, the ordinary Hölder inequality for inte-
grals [316] yields

(7. 76)

and so T( ITSI) ~ T( ITIP) 1iPT( ISiq ) 1 iq in the commutative C*-algebra Boo.


Since any state w of Boo is given by an integral, it yields one more Hölder
inequality that we can apply:

Trw ITSI ~ w(T(ITIP) 1iPT(ISiq) 1 iq)


~ w(T(ITIP)) 1 iPw(T(ISiq)) 11 q = (Trw ITIP) 11 P(Tr00 ISiq) 11 q.

Of course, since we only needed the Hölder inequality for measurable ope-
rators, we could have omitted the last step by just taking the limit as .\ - oo
in (7.76). 0
No explicit general formula for Trw (TS) can be given without specifying
the state w, but we can at least rewrite it as a generalized limit of a se-
quence. If {an} is a bounded sequence, we can extend it piecewise-linearly
to a function in on Cb([3,oo)), as in (7.67), and let aoo be the its image
in Boo; we write limn-w an := w(aoo). Clearly, limn-w isapositive linear
functional on the space .eoo of bounded sequences, coinciding with the or-
dinary limit on the subspace of convergent sequences. If TE L 1+, then

Trw T = lim CTn(T).


n-w logn

Lemma 7.17. Let A E L 1+ be a positive operator, and Iet S E L(J-{). Let


En := L:i:,:-J IUk) ( Uk I be the spectral projector of A = L:ho Sk IUk) ( Uk I for
its first n eigenvalues. Then

Trw(AS) = lim - 1 Tr(EnAS). (7.77)


n-w 1ogn

Proof. Since ITr(EnAS) I ~ IISII Tr(EnA), it follows that (log n)- 1 Tr(EnAS)
is a bounded sequence, so the right band side of (7.77) makes sense: call
it 1/J(S). This defines a positive linear functional !fJ on L(J-{), suchthat
111/JII = 1/1(1) = Trw(A). Now S .... Trw(AS) = Tr 00 (A 1 i 2 SA 1 i 2 ) isanother
positive linear functional on L(J-{) with the same norm; as remarked in
Section LA, to establish their equality it is enough to show that their dif-
ference is also positive. Therefore, we need only show that

lim - 1 -Tr(EnA 112 BA 1 i 2 ) ~ Tr 00 (AB) = lim - 1 -an(A 112 BA 112 )


n-w log n n-w log n

when Bis a positive operator. This follows from Lemma 7.32, which shows
that Tr(EnC) ~ an(C) for C positive, since En has rank n. 0
7.5 The Dixmier trace 291

.,.. We want to compute Dixmier traces for some examples, before going fur-
ther. Clearly, if T E Xis suchthat the sequence {crn (T) I log n} converges,
then TE L 1+, T is measurable, and

T r + T = 1·1m T71 (T) = 1·1m -


crll (T)
- = 1·1m -
crn-
(T).
71-oo 71-oo log,\ n-oo log n
An important dass of examples is provided by the (scalar, for simplicity's
sake) Laplacians on compact Riemannian manifolds.
Example 7.1. Take, for instance, the n-dimensional torus 1rn := ~n (ll.n. For
l E &:n, define <J:>z E C 00 (P) by <J:>L(X) := exp(2rri(llx 1 + · · · +lnxn)). These
are the eigenfunctions of ß; indeed, ß<l:>1 = 4rr 21l1 2<t>L. and sp ß = {4rr 21l1 2 :
l E &:n}. The multiplicity m11 of any ,\ = 4rr 21l1 2 is the number of lattice
points in &:n of length lll. The only harmonic functions are the constants.
For any s > 0, the compact operator ß -s has spectrum { ,\ -s : ,\ E sp ß },
with the same finite multiplicities m71.
To compute the Dixmier trace of ß -s, we need to estimate crn (ß -s) I log n
for n large. In fact, weshall use the subsequence {crN, (ß -s) I log Nr}, where
Nr is the total number of lattice points in a ball of radius r centred at
the origin. (If Nq and Nr are two successive values of those numbers, and
Nq ~ n ~ Nr. then CYNq(ß- 5 )/logNr ~ crn(ß- 5 )/logn ~ CYN,(ß- 5 )/logNq,
so the full sequence tends to the same Iimit, since logNrl logNq - 1.) Since
a large ball may be approximated by a union of nonoverlapping unit cubes,
we get N r - vol { x : Ix I ~ r }, and so the shell of radius r and thickness
dr has the volume
Nr+dr- Nr- vol{ X: r ~ lxl ~ r + dr}- On rn-l dr.

We estimate

(JNR(ß- 5 ) = L (4rr 21l1 2)-s- r 1


(4rr 2r 2)-s (Nr+dr- Nr) (7.78a)

s; s;
l5III5R

-On (4rr2r2)-srn-l dr = (2TT)-2son yn-2s-l dr.

JJ
The unit ball has volume Onrn-l dr = Onln, and therefore logNR -
1og(n- 1 0nRn)- nlogR as R- oo. Thus,

CYNR (ß-s)
---:-.:"-'--,-:--'- - (2rr)-2son IR r n-2s-l d r (7.78b)
logNR nlogR 1 ·

If s < nl2, the integral diverges and thus ß-s rt L 1+; we write Tr+ ß-s =
+oo. lf s > nl2, thenRn- 2s /logR- 0, so ß-s is measurable and Tr+ ß-s =
0. Finally, if s = nl2, the right hand side of (7.78b) is independent of R, so
again ß -s is measurable and

Tr+ ß -n/2 = On (7.79)


n(2rr)n
292 7. The Noncommutative Integral

Example 7.2. Consider now the Riemann sphere § 2 (with the standard area
form), where sp ß = { l(l + 1) : l E ~ }, with respective multiplicities m1 =
21 + 1. Let Nr := Ir=o(2l + 1) = (r + 1) 2; then log(Nr+d/log(Nr) - 1 as
r - oo, so that limn-oo O"n(ß- 5 )/logn can be computed as the limit of the
subsequence lTNr (ß -s) /log Nr. We estimate

and thus
lTNr(ß-S)- _1_ ±(l+ !)1-2s as r - oo.
log Nr log r I=O 2

If 5 < 1, the right hand side diverges, so ß -s rt L 1+. If 5 > 1, the series
converges, so ß -s is traceclass and Tr+ ß -s = 0. Lastly, if 5 = 1, then
r r 2 2r+1 2 r 2
L (l + ~ )-1 = L 21 + 1 = L k- L 21 = 2H2r+1- Hr,
1=0 1=0 k=1 1=1
where Hr is the rth partial sum of the harmonic series. Since Hr -log r + ;y
where ;y is Euler's constant, 2H2r+1 - Hr -log r + ;y + 2log 2; therefore

lim lTNr(ß- 1) = lim 2H2r+1 -Hr = 1.


r-oo log Nr r-oo log r

Thus ß - 1 E L 1+ is measurable and Tr+ ß - 1 = 1.


Exercise 7.18. The Laplacian on the n-dimensional sphere §n (with the
standard volume form) has eigenvalues l(l + n - 1) for l E ~. with cor-
responding multiplicities ml = e~n) - (1+~- 2 ). Show that ß-S rt L 1+ for
0 < 5 < n/2, that Tr+ ß-s = 0 for 5 > n/2, and that

Tr+ ß-n/ 2 = ~· 0
n.

~ Before concluding, we add a few words about the problern of choos-


ing the state w to define a particular Dixmier trace. On the (nonseparable)
C*-algebra Boo, exhibiting even one such state involves using the axiom
of choice. This problern can be finessed in various ways; what we did in
Definition 7.16 was to notice that a function f E Cb([3,oo)) has a limit
lim.\-oo j(A) if and onlyif w(j) does not actually depend on w: the states of
Boo separate the points of accumulation at infinity of elements of Cb([3, oo) ).
No "naturally occurring" operator has come to our attention that lies
in L 1+ but is not measurable, although it is easy to construct artificial
7.6 Connes' trace theorem 293

examples of nonmeasurable Operators (without recourse to the axiom of


choice). For pseudodifferential operators, we are assured of measurabil-
ity by Connes' trace theorem below, and the noncommutative geometrical
conditions discussed in Chapter 11 also allow us to work with measurable
operators only.
In a similar way that Solovay's axiomatic model [441], forbidding the exis-
tence of nonmeasurable Lebesgue subsets of the realline (and, in a stronger
sense, of non-Baire subsets of a complete separable metric space), provides
an alternative framework for ordinary functional analysis, sturdier than the
Zermelo-Fraenkel model of set theory plus the axiom of choice, one can
speculate that a natural framework tailored to the needs of noncommuta-
tive geometry would banish nonmeasurable operators from the outset.

7.6 Connes' trace theorem


From the examples at the end of Section 7.5, we see that the ratio between
Wres ~ -n/ 2 and Tr+ ~ -n/ 2 is the same for tori as for spheres! It appears to
depend only on the dimension:
Wres~-n/ 2
Tr+ ~-n/2 = n(Zrr)n.
This is much more than a mere coincidence.
Theorem 7.18 (Connes). Let H be an elliptic pseudodifferential operator of
order -n on a complex vector bundle E on a compact Riemannian mani-
fold M. Then HE L 1 +, indeed His measurable, and

ResH = Tr+ H = n(Z~)n WresH. (7.80)

We consider first the case of positive H. The equality of the first and
the third terms is Theorem 7.12. To complete the proof, it is enough to
establish either the first or the second equality. We do both.
Short Proof. The second equality follows from (7.53): we get
1
dH-dx, x; ,\) - (Zrr)n ao(x) + · · · (C),

where ao(x) ldnxl = n- 1 wresx H. Then, arguing as in Proposition 7.11, we


obtain

NH-1(;\.)-
f
11
Jo f
M dH-dx,x;J.l) ld n xl df.l- nWresH
(Zrr)n ;\. as ;\.- oo,

and, since NH-d,\) = k when ;\. = Sk(H- 1 ), we conclude that Sk(H)- c/k
as k- oo, where c = n- 1 (2rr)-nWresH. (For this reason, Weyl's estimate
(7.54) can be regarded as a particular case of Connes' trace theorem -and
as such, a harbinger of the theory of noncommutative integration.)
294 7. The Noncommutative Integral

Long Proof. The first equality (without the restriction to pseudodifferential


operators) is a consequence of the next two Iemmas.
Lemma 7.19. Let {ak} be a decreasing sequence of positive numbers with
limk ak = 0, suchthat Ir= 1 ak - 1og(11an) as n - oo. Then log(llan) -
log n (and then, of course, Ir= 1 ak - log n).
Proof. Denote l := liminfn log a~ 1 I log n, L := lim supn log a~ 1 /log n. We
prove first that l = 1.
Suppose for a moment that l < 1. Then there is some oc < 1 such that

E := {n: lo~~~l:n) :5 oc} = {n: an~ :a}


is infinite. For each n E E,

a1 + ···+an > ,.....---...:..:..._-


nan n 1-a
> ---
1og(11an) -log(llan)- oclogn'
and then limsupn(a1 + · · · + an)Jloga~ 1 = oo, contrary to hypothesis.
Suppose instead that l > 1. Then there is some ß > 1 suchthat log af: 1 ~
ß log k for large enough k. Therefore ak :s; c I kß for some c and all k E ~­
lt would follow that
a1 +···+an c(l + 112ß + · · · + 11nß) c"((ß)
~-~-~ < < -----
1og(11an) - ßlogn - ßlogn'
and so liminfn (a1 + · · · +an) I log a~ 1 = 0, which is also excluded.
To finish the proof, it remains to show that L > 1 is forbidden. For if
L > 1, there is some ß > 1 suchthat
log(11an) } ß
E1:= {n: logn ~ß ={n:an:51ln}

is infinite; since l = 1 < ß, we can find oc, with 1 < oc < ß, such that
log(llan) } a
Ez:= { n: logn :soc ={n:an~11n}

is infinite. Choose n 1 E E1, nz E Ez with n 2 < n1 but k $ E1 u E2 for


nz < k < n 1 (and so 1 I kß < ak < 1 I ka). Given E > 0, choose no such that

Ia1log(
+···+an I
1 lan) - 1 :5 E for n ~ no;
suppose also, as we may, that nz ~ no, so a 1 + · · · + an 2 :s; ( 1 + E) log a~; :s;
(1 + E)oclognz. Then
a1 + · · · + an 1 + · · · +an?
a1 ______
___;:,_ • + · · · +an I
____;_=· + an?+1
1og(11an 1 ) log(11an 1 ) log(1lan 1 )
< (1 + E)oclognz + Ln 2 <k,;n 1 k-a
- ßlogn1
(1 ) oclognz ((oc)
< +E
ßlognl
+ ßlogn1 '
7.6 Connes' trace theorem 295

sothatlimsupn(a1 + · · · +an)llog(11an) :5 cxlß < 1. Wehavereacheda


contradiction. 0

Lemma 7.20. Suppose that ak l 0, that Ik'~ 1 af is convergent for s > 1, and
that Ik'~ 1 af- 11(s -1) as s l 1. Then Lk~ 1 ak -log(llan) as n- oo.

Proof. We can assume that 0 < ak < 1, for all k. We define, for each x > 0,
a positive measure J.lx on [0, 1] by J.lx(t) := Ik'~ 1 a~+x 8(t- af), so that
1
2: a~+x g(af).
00

( g(t) dJ.lx(t) =
Jo k~1
We claim that dJ.lx (t) - dt I x, as x I 0. For this, it is enough to checkthat

1
limx ( g(t) dJ.lx(t)
xlO Jo = e
Jo g(t) dt (7.81)

for all polynomials g, or even for all monamials g ( t) = t r. But

(
Jo
1
tr dJ.lx(t) = f. a~+X(r+ 1 )
k~ 1
- 1
x(r + 1)
as X I 0,

by hypothesis. Now consider the function h given by

for 1 I e :5 t :5 1,
h(t) := {11t,
0, for0:5t<11e.

JJ
Then h(t) dt = 1, and Lk~ 1 ak can be written as JJ
h(t) dJ.lx(t) for any
x satisfying a~ ~ 11e and a~+ 1 < 11e, or equivalently log(1lan) :5 11x <
log(1lan+1). Thus n- oo and x - 0 together, and (7.81) applied to h(t)
implies that
1 1
2:n ak -
k~ 1
- - log -
X an
as n - oo. 0

Long proof of Theorem 7.18, continued. The Lemmata 7.19 and 7.20, taken
together, establish that whenever His a positiveoperatorsuch that ?,;'H(s)
is defined for ~ s > 1 and has a simple pole at s = 1, then H E L 1+, and,
moreover, that when ResH = 1 then limn-oo CTn(H)jlogn = 1, too. lt fol-
lows at once that limn-oo CTn (H) I log n = Res H. Therefore His measurable
and Tr+ H = Res H. This establishes (7.80) for positive elliptic pseudodif-
ferential operators of order -n.
Now, to conclude both the long and short proofs, any elliptic pseudodif-
ferential operator Hoforder-n is of the form H = H 1 - H2 + iH 3 - iH4
where each H 1 is a positive elliptic operator of order -n. Indeed, the ad-
joint Ht is also pseudodifferential of the same order, so by taking real
and imaginary parts we may suppose that H is selfadjoint; but then H =
296 7. The Noncommutative Integral

!(IHI+H)-!(IHI-H),soitremainsonlytoobservethat IHI = (H- 2)- 112,


as defined by (7.58), is a pseudodifferential operator andin fact is elliptic.
Thus H lies in L 1+ and is measurable, Tr+ His defined, and (7.80) holds by
linearity. 0

Exercise 7.19. Prove the following converse of Lemmata 7.19 and 7.20: if
Ir= 1 ak -logn, then (s -1) Ik'= 1 a~- 1 as s I 1. o
~ The story of the proofs of Theorem 7.18 is interesting. Connes origi-
nally [88] put forward Theorem 7.18 by proving the second equality di-
rectly. That proof was somewhat telegraphic, and we expounded it in [469],
to which the interested reader is referred. Also in [469], relying on the lke-
hara Tauberian theorem, we gave a proof of Theorem 7.18 by checking the
first equality for pseudodifferential operators. For detailed expositians of
this second approach, see [4] and [228].
Example 7.3. Let us note, however, that the conditions log a~ 1 -log n and
an - 1/n arenot equivalent: the second does not follow from the first. A
simple counterexample is obtained by taking the sequence defined by ao :=
4log 2/3, a1 := 2log 2/3 and ak := 4log 2/3. 4n, for 22n-1 ::; k < 22n+1,

r
with k ~ 2 and n ~ 1. Then

f a~ f (~ 1~:;
k=1
=
n=O
(22n+1 - 22n-1)

= ~ ( 4log 2 ) s 1 __1_ as 5 1 1,
2 3 1- 41-s s- 1
whereas nan oscillates between ! and 2. (An even more spectacular ex-
ample of Ir= 1 ak- logn not implying an- 1/n is found in Lemma 7.37
in Section 7.C.) Connes later mentions [91, Prop. IV.2.4] that the Taube-
rian theorem of Hardy and Uttlewood [232] guarantees the first equal-
ity of (7.80). We have just given an elementary version, due to Ricardo
Estrada, with the advantage of showing clearly when and why the property
(logn)- 1 Ir= 1 ak- 1 is tobe expected -so larger classes of measurable
operators are glimpsed.
lt is fair to say again that the funny behaviour in the counterexample
cannot take place when the an are the eigenvalues of a pseudodifferential
operator -it is excluded by the argument of our short proof. The coun-
terexample does not fulfil the hypothesis of the lkehara-Wiener theorem,
as there are other pol es on the line ~ s = 1, and then the lkehara-type proof
of Res A = Tr+ A is correct, provided one cares to restriet it to elliptic pseu-
dodifferential operators.
Let us add, as a final comment to Theorem 7.18, and also in view of The-
orem 7.9, that the selection of a kind of "principal value" for operators that
are "logarithmically divergent", epitomized by the Dixmier trace, has been
in heuristic use for many years by the pioneers of quantum field theory.
7.6 Connes' trace theorem 297

..,. At last, we can bring all the strands together. Connes' trace theorem
allows us to compute the integral of any function on a Riemannian manifold
by an operatorial formula. Thus, it is the starting point of a generalization
of the notion of integral.
Corollary 7.21. For any function a E C""(M),

La(x) lv9 1= n(~:)n Tr+(aß-n 12 ). (7.82)

Proof. Think of a as a bounded multiplication operator on J{ = L~· 0 ; then


aß -n/ 2 is a pseudodifferential operator of order -n. lts principal sym-
bol is just a_n(X, ~) = a(x)(giJ~i~)-nl 2 , and therefore wresx(aß-n1 2 ) =
Ona(x)p(x) ldnxl, by the proof ofProposition 7.7. Thus, the left hand side
of (7.82) is just 0; 1 Wres(aß-n1 2 ). Then apply Theorem 7.18. D
Madore [323, pp. 213-214] remarks that the left hand side of this equa-
tion is essentially the integral of a(x; ß, a) over M.
Compactness of M is not decisive; we can make a more general statement.
Corollary 7.22. Let a be an integrable function on a Riemannian mani-
fold M. Then (7.82) again holds.
Proof. This follows from the previous corollary if a is a smooth function
with compact support. For a positive and integrable, use monotone con-
vergence; the general case then follows at once. D
Corollary 7.21 suggests that we define a noncommutative integral as the
functional given by
n(2rr)n
a ~ J[ aß-nl 2 := On Tr+(aß-n12). (7.83)

This is normalized so that, for elements of a pertinent algebra .J\. (a non-


commutative description of a manifold), the identity 1.Jt has integral 1. It
is to be expected, from the examples we have seen, that the exponent n be
uniquely determined by the requirement that the right hand side be non-
trivial -neither zero nor infinite- when a is a positive element. However,
we cannot at present say in all generality what is the Laplacian on a non-
commutative manifold -because neither can we say in all generality what
is a Riemannian noncommutative manifold. This problern is addressed, and
solved, in Part III by using a first-order operator instead of a second-order
one. What we then obtain is the noncommutative analogue of spin mani-
folds .
..,. To close, we pointout that the Wodzicki residue has found many appli-
cations in quantum field theory: for instance, in calculations of the multi-
plicative anomaly for zeta-function regularized determinants (see, for in-
stance, [155] and references therein), and of Schwinger terms for current
298 7. The Noncommutative Integral

algebras, as in [73 ), in terms of a twisted version of the Kravchenko-Khesin-


Radul cocycle [295, 382).

7.A Pseudodifferential operators


Pseudodifferential operators are a generalization of differential operators
born out of the construction of parametrices (approximate inverses) for
elliptic differential operators [458). Nowadays, they belang in every ana-
lyst's toolkit. They can be considered at several Ievels of generality, de-
pending on one's purpose. In this book, it will be enough to consider only
the simplest case: the standard pseudodifferential operators, also known
as pseudodifferential operators of type (1, 0) in the more technical termi-
nology [253,449,450).
Let U denote an open subset of !Rl.n and C;'(U) the space of smooth
(complex-valued) functions with compact support in U. The space of dis-
tributions on U, i.e., the dual space of c; (U), is denoted by 'D' (U), while
'E' (U) denotes the space of distributions with compact support, which
is the dual of C"" (U). When U = !Rl. n, we let S' ( !Rl. n) denote the space of
tempered distributions on !Rl.n. Weshall write Dj := -io/oxi, and D 01 =
Df 1 •• • Di:" for any multündex oc E ~n. Let P(x,D) = Llal,;d aa(X)D 01 be a
differential operator of order d, where a 01 E C"" ( U) for each oc E ~n. Using
the Fourier transform and its inverse, we can write

P(x,D)j(x) = (2!)n Jeix·~p(x, ~)j(~) d~


= (2!)n ff ei(x-y)·~p(x, ~)j(y) dy d~, (7.84)

where p(x, ~) = Llal,;d aa(X)~ 01 and f E c; (U). A pseudodifferential


operator is an operator of the general form (7.84), where the function
p (x, ~)in the second integral belongs to a suitable dass of functions on U x
IRl.n; it may even be replaced by a function a(x,y, ~) defined on U x U x !Rl.n.
We shall also use, when convenient, the notation D~ to denote derivatives
with respect to the variables ~i·
Definition 7.17. A function p E C"" ( U x IRl.n) is a symbol of order d, written
p E Sd(U), if for any compact subset K c U and any oc, ß E ~n. there is a
constant CKaß suchthat
(7.85a)
for all x E K and ~ E IRl.n. Similarly, a function a E C"" (U x U x !Rl.n) is an
amplitude of order d, written a E Ad(U), if for any compact K c U and
any oc, ß, y E ~n, there is a constant CKaßy suchthat

(7.85b)
7.A Pseudodifferential operators 299

for all (x,y) E K x K and ~ E ~n.

The set Ad(U) is a Frechet space with the topology defined by the semi-
norms PKocßY• whose value at a is the infimum of the constants CKocßy
in (7.85b). As we shall shortly see, the dass of amplitudes does not pro-
duce more general pseudodifferential operators than the dass of symbols,
but it does offer more freedom to construct such operators. However, the
symbolic calculus of amplitudes is somewhat involved.
Definition 7.18. A pseudodifferential operator of order d is an operator P
defined by (7.84), with p E Sd(U) or a E Ad(U) in the integrand. We write
P = Op(p) or Op(a) in 'l'd(U).
The integrals in (7.84) are to be understood as iterated integrals and it
is, therefore, easy to check that Op(a): c; (U) - coo (U) is continuous.
By transposition, Op(a) extends continuously to a map Op(a): 'E' (U) -
'D'(U). When U = ~n. the relation x 01 eix·~ = D~(eix·~) and integration
by parts shows that Op(a) maps S(~n) into itself continuously and, by
transposition, maps S' (~n) into itself also.
The (Schwartz) kernel of Op(a) is the distribution ka E 'D' (U x U) given
by

(7.86)

where the integral is to be regarded as an "oscillatory integral", that is, in


the distributional sense [450]. When U = ~n. the kernellies in S'(~ 2 n).
Since (x- y) 01 ei(x-y)·~ = D~(eHx-yl-~). integration by parts shows that
(x- y) 01 ka(X,y) = (2rr)-n f ei(x-y)·~D~a(x,y, ~) d~ and, by (7.85b), this
integral converges absolutely for Ioc I > d + n. Differentiating r times und er
the integral sign is allowed and shows that (x- y) ()( ka (x. y) is er. provided
that locl > d + n + r. Therefore, ka is smooth off the diagonal in U x U.
This implies that pseudodifferential operators are pseudolocal: for each f E
'E' ( U), Pf is smooth on every open subset of U where f is smooth. We recall
that differential operators P (x, D) are characterized as local operators: if
f vanishes on an open subset of U then also P(x,D)j vanishes ön this
subset.
Definition 7.19. An operator K: c;(U)- 'D'(U) is smoothing (or "regu-
larizing") if its kernel k lies in coo ( U x U) or, equivalently, if it extends as
a continuous linear map from 'E' ( U) into coo ( U).
A smoothing operator K is actually a pseudodifferential operator whose
amplitude is in Ad ( U) for all d E ~.in which case we write a E A -oo ( U) and
K = Op(a) E '1'-oo (U), and we say that K is a "pseudodifferential operator
of order -oo". Indeed, on account of (7.86), we can take
300 7. The Noncommutative Integral

where cp E C~(l~n) is suchthat f cp(~) d~ = (2rr)n. The converse is also


true: all pseudodifferential operators in 'Y-oo ( U) are smoothing operators.
This is related to Sobolev-space continuity of pseudodifferential operators,
as we shall see in a moment.
Definition 7.20. For each s E IRI., the Sobolev space H 5 (IR{ n) is the space of
tempered distributions u E S' ( IRI. n) whose Fourier transform u is a square
integrable function for the measure (1 + 1~1 2 )5 d~. This is a Hilbert space
whose norm is given by

When s = k isapositive integer, this norm is equivalent to

lliulli~ := I IIDaull6.
le<l:5k

and, in particular, H 0 (IR{ n) = L 2 (IR{ n). Moreover, the inclusions H 5 (IR{ n) ....
for s > t, are norm-decreasing, therefore continuous.
Ht(~R~_n),

More generally, if K c ~R~_n is compact, H 5 (K) denotes the subspace of


H (!RI.n) whose elements have support in K. We then denote by Hf(U) the
5

union of all the spaces H 5 (K) over compact subsets K c U (with the induc-
tive limit topology). We also consider the space H{0 c(U) of distributions
u E 'D'(U) suchthat cpu E H 5 (!RI.n) for all cp E C~(U). For pseudodiffer-
ential operators, the following continuity property holds [458].
Proposition 7.23. Let PE 'Yd(U) and s E IRI.. Then P can be extended to a
continuous linear map from Hf ( U) into H{0~d ( U). E3

lf the symbol of P also satisfies (1 + I~Wd SUPxEK ip(x, ~)I - 0 as 1~1 -


oo for any compact K c U, then Pis not only bounded but compact [458].
=
As a corollary, when p(x, ~) 1, we obtain Rellich's theorem, which says
that the natural inclusions Hf (U) .... H~ (U) are compact operators. Rellich's
theorem and Sobolev's lemma [407, Thm. 7.25], which says that Hf(U) c
ck (U) if > s in + k, imply that the operators in 'Y-oo (U) are smoothing
operators, as advertised.
lt is not always possible to compose two pseudodifferential operators,
since they are defined on compactly supported distributions, but their
range consists of distributions that need not have compact support. The
following condition offers a way out.
Definition 7.21. A pseudodifferential operator P is properly supported if
both P and its adjoint pt map 'E' ( U) into itself.
There are several equivalent ways to rephrase this condition [253,449,
458]; for instance, using the pseudolocality of pseudodifferential operators,
one finds that P is properly supported if and only if P and pt map C~ ( U)
7.A Pseudodifferential operators 301

into itself. Moreover, since the kernel is smooth off the diagonal, one can
write any pseudodifferential operator as the sum of a properly supported
one and a smoothing operator [253,458] .
.,.. The symbol calculus is developed by using asymptotic expansions of
symbols. If p i E sd J ( U) where {d i} is a decreasing sequence of real num-
bers with di - -oo, then one can find p E sdo (U), which is unique modulo
s-"'(U), suchthat
k
P-LPd 1 ESdk(U) forall kEN,
j;Q

and we write p - Lj"O PdJ' Such a symbol may be constructed [449] with
an auxiliary function cp E C"'(~n) that vanishes for 1~1 < 1 and equals 1
for 1~1 > 2, by taking p := Lj'P(~/ri)Pd 1 (x.~) with a suitable sequence
Yj- oo.

Definition 7.22. The most important examples are the classical symbols,
which correspond to the case where the Pd1 (x, ~) arehomogeneaus in~ of
degree dj (i.e., Pd1 (x,.\~) = ,\dJpj(X,~) for every ,\ > 0) and the degrees
differ by integers: di- dj+I E N. In that case, it is customary to redefine
d i := d - j and to write the expansion as

p(x, ~)- L Pd-j(X, ~), (7.87)


j"O

where Pd-j(X, ~) is homogeneaus in~ of degree (d- j). The leading term
(that of highest homogeneity) Pd (x, ~) is called the principal symbol for p.
The key observation to obtain a tractable symbol calculus is the associ-
ation of an explicit symbol to each amplitude. Formally,

where p(x, ~) := e-ix·~P(eix·~); one should write P(ei(·H)(x) instead, but


we shalllet that pass. Often one writes u(P) for the symbol p correspond-
ing toP. For instance, if P = Op(a) is given by

P j(x) = (2rr)-n JJ ei(x-y)·'1a(x,y, ry)j(y) dy dry,


then

p(x, ~) = (2rr)-n JJ ei(x-y)·(ry-~la(x,y, ry) dy dry


= eiD~·Dya(x,y, ~) ly;x•
Using a formal expansion eiD~·Dy = 1 + iD~ · Dy- !<D~ · Dy) 2 + · · ·, we
arrive at the following formula, proved in [450].
302 7. The Noncommutative Integral

Proposition 7.24. Let P = Op(a) E 'l'd be a properly supported pseudodif-


ferential operator. Then P = Op(p) where the symbol p E Sd(U) is given
by p(x, ~) = e-ix·1;;p(eix·1;;), and the following asymptotic expansion holds,
in the sense of (7.87):

E3

Corollary 7.25. Let Op(p) E 'l'd be a properly supported operator. Then its
adjoint is also in 'l'd. Moreover, Op(p)t = Op(p* ), where

E3

The symbol for the product of two pseudodifferential operators is com-


puted in the same way, based on the following formula for an amplitude
for the product:

a(x,y, ~) = (2rr)-n JJ ei(z-yl·<IJ-l;;la!(x,z, ~)az(z,y, 1]) dzd1].


Proposition 7.26. Letüp(p) E 'l'd 1 andOp(q) E 'l'd 2 beproperlysupported
Operators. Then Op(p) Op(q) is a properly supported operator in 'l'd 1 +d 2
whose symbol p o q has the following asymptotic expansion:

(7.88)

In particular, the principal symbols of classical pseudodifferential operators


compose as pointwise products:

(7.89)

Proof. The formal expression is easily derived from (7.89). For the proof of
the validity of the asymptotic expansion, we refer to [450,458]. D

Thus, the set of properly supported pseudodifferential operators is an in-


volutive algebra. Moreover, it is clear that the commutator [Op(p), Op(q)]
of two pseudodifferential operators, of orders d 1 and d 2 , is also a pseu-
dodifferential operator of order d 1 + dz - 1 whose leading term in the
asymptotic expansion is -i times the Poisson bracket of p and q:

. . ~ ap aq ap aq
-t{p(x, ~), q(x, ~)} = -t L ~ axi - axi ~·
J=l '=>J '=>J
7.A Pseudodifferential operators 303

~ The formula p(x, ~) := e-ix·~P(eix·~) yields an expression for the prin-


cipal symbol of a classical pseudodifferential operator. Substituting t~ for
~ with t > 0, and dividing by td, we obtain

t-de-itx-~P(eitx·~) = t-dp(x,t~)- 2: t-1Pd-j(X,~),


j<!:O

so that, formally, t-de-itx·~P(eitx·~) = Pd(X, ~) + O(t- 1 ). For differential


operators, the series terminates and this growth estimate is certainly valid.
Also, we can replace x · ~ in the exponent by h (x), where h E C"" (U) is any
function suchthat dh(x) = ~. keeping the estimate t-de-ith(xl P(eith(x)) =
Pd(x, ~) + o(t- 1 ). This yields the following useful formula [33, Prop. 2.1].
Proposition 7.27. Let P be a differential operator of order d > 0 on U. Its
principal symbol is given by

Pd(X,~) = limt-de-ith(xlp(eith(x)) (7.90)


t-oo

whenever h E C"" (U) satisfies dh(x) = ~· 8

Definition 7.23. We say a pseudodifferential operator Op(p) E 'l'd (U) is


elliptic if there exist strictly positive continuous functions c and C on U
suchthat
lp(x,~)l ~ c(x)l~ld for 1~1 ~ C(x), x EU.
Elliptic pseudodifferential operators are characterized by possessing a
parametrix, that is, an inverse modulo smoothing operators. The symbol
of a left inverse Q of an elliptic pseudodifferential operator Pis constructed
by successive approximation based on the product formula. One can define
recursively {q-d-1 }J",O by the relations q_d(x, ~)p(x, ~) = 1, and
·Iai
q-d-J(X, ~)p(x, ~) =- 2: :,.v~q-d-J-Ial (x, ~) D~p(x, ~) (7.91)
l:slal<j OC.

for j ~ 1. We refer to [458, Cor. 1.4.3] for the full proof of the following
proposition.
Proposition 7.28. IfOp(p) E 'l'd(U), then Op(p) is elliptic if and only if
there is some q E s-d (U) such that q o p = 1, or suchthat p o q = 1, or else
suchthat
Op(p) Op(q) =Op(q) Op(p) =1 mod 'I'_"" (V).
In particular, such an Op(q) is elliptic of order -d. 8

Moreover, by Corollary 7.25 and (7.91), the adjoint of an elliptic pseudo-


differential operator is also elliptic.
~ To define pseudodifferential operators on manifolds, we need first to
study the behavior of Op(p) E 'l'd(U) under the action of a diffeomor-
phism. Let cJ>: U - V be a diffeomorphism between open subsets of Rn. lf
304 7. The Noncommutative Integral

P = Op(p) E 'l'd(U), then

cp*P(j) := P(cp* j) o cp- 1

defines an operator cp*P: C;' (V) - C"" (V) that is actually pseudodifferen-
tial. To see that, write 1.fJ := cp- 1 and ]1/1 (x) := det 1./J' (x) for its Jacobian
determinant, and define 'I' (x, y) for x * y by the equation

(1./J(X)- ljJ(y)). ~ =: L ~i'I'ij(X,y)(Xj- YJ)·


ij

Setting 'I' (x, x) := 1./J' (x), we get a matrix-valued function 'I' that is smooth
and invertible in a neighbourhood N of the diagonal in V x V. Let 'I' (x, y) -t
be the contragredient matrix. Then

cp*P(j)(x) = (2rr)-n JJ eiiJ·(IIJ(x)-zlp(ljJ(x), ry)j(cp(z)) dz dry


= (2TT)-n ff eii)·(IIJ(X)-IIJ(y))p(ljJ(X), ry)j(y)jl/l(y) dy dry

= (2TT)-n ff eil)·'l'(x,y)(x-y)p(ljJ(X), ry)j(y)JIIJ(y) dy dry

= (2TT)-n JJ ei~·(x-y)p(ljJ(X) ' 'I'(X ' y)-t~) I detjl/l(y)


'I' (x, y) I
j(y) dy d~.

Let p: V x V - [ 0, 1] be supported on N with p =1 in a smaller neighbour-


hood of the diagonal. If we insert the cutoff p (x, y) in the integrand, we
change cp*P by a smoothing operator (since 1- p(x,y) vanishes near the
diagonal), and thereby obtain the following amplitude of order d:

b (x,y, "")
'!: · - ]1/l(y) ) -tJ:)
.- p(x,y) I det'l'(x,y)l p(!.fJ(x ''l'(x,y) "" ·

Therefore, cp*P = Op(b) mod '1'-""(U), which proves that cp*P E 'l'd(U).
Nowwe can apply Proposition 7.24 to compute a symbol q(x, ~) for cp*P
from the amplitude b. We remark that b(x, x, ~) = p(!.fJ(x), !.fJ' (x)-t~), and
that 'l'(x,y)-t~ -ljJ'(x)-t~ vanishes to first order as y - x. We arrive at
the following result.
Theorem 7.29. If P is a properly supported operator in 'l'd ( U), and cp: U -
V a diffeomorphism, then cp*P E 'l'd(V) is properly supported. Moreover,
cp*P = Op(p<l>), where

p<l>(x, ~) - L ~qa(X, ~)Dgp(ljJ(x), !.fJ' (x)-tl;). (7.92)


lal<oO a.

where q 0 (x, 1;) = 1 and, in general, qa(x, ~) is a polynomial in~ of degree


: :; !Iai. a
7.A Pseudodifferential operators 305

A more explicit description of the terms qcx(X, ~). and a proof of the
estimate on the degree in ~. can be found in [253, Thm. 18.1.17). Ob-
serve that if p is a classical pseudodifferential operator, then the a-term
of the sum (7.92) is also polynomial in~ of degree at most d, so that, after
some rearrangement, one finds that p<l> is also a classical symbol. Using
t.JI'(x)- 1 = c:J>'(t.J!(x)), equation (7.92) also shows that

(7.93)

so the principal syrnbol transforms as P!<x.~) = Pd(t.JI(x),c:/>'(t.JI(x))t~),


that is,
(7.94)

.".. Now we can describe pseudodifferential operators on manifolds; weshall


consider only compact manifolds, on which the condition of proper support
is automatic -such operators form an involutive algebra.
Definition 7.24. Let M be a compact manifold. An operator P: C%' (M) -
coo (M) is called a pseudodifferential operator of order d, if the kernel of P
is smooth off the diagonal in M x M and, for every coordinate chart ( U, c:J>)
on M, the operator f ..... P(c:J>* j) o c:J>- 1 from C%'(c:J>(U)) into C 00 (c:/>(U))
belongs to 'I'd(c:J>(U)). We say that P is classical if this local expression
is a classical pseudodifferential operator, for each coordinate chart. More
generally, given vector bundles E and F over M, a linear map P: rt (M, f) -
roo (M, F) is a pseudodifferential operator of order d, if the kernel of P
is smooth off the diagonal in M x M and the local expressions of P are
pseudodifferential Operators with matrix-valued symbols, in the obvious
way.
Formula (7.94) shows that the principal symbol of a classical pseudodif-
ferential operator is invariantly defined as a function on the cotangent bun-
dle T* M .!!.. M (more generally, as a bundle morphism from the pullback
bundle rr* E to the pullback bundle rr* F). By (7.87) and the multiplicative
property (7.89), the mapping sending a classical pseudodifferential opera-
tor to its principal symbol is a homomorphism from '!' 00 (M) to coo (T* M). In
particular, if Pisa differential operator of order d on M, the formula (7.90)
for the principal symbol remains valid; we need only take h E coo (M) such
that dh(x) = ~ for (x, ~) E TiM. In fact, one can define a principal symbol
for any pseudodifferential operator on M by exploiting (7.93) on overlap-
ping charts and patehing with a partition of unity; the result [450) is a
well-defined dass in the quotient space sd (T* M) I sd- 1 ( T* M).
If c:J>: U - V is a diffeomorphism between open subsets of ~n, and
P E 'I'd ( U) is elliptic, then, by (7.93) c:/>*P is also elliptic of order d. Thus, we
say that an operator P E 'I'd (M) is elliptic if its local expression in each co-
ordinate chart is elliptic. Using a partition of unity, one can patch together
local parametrices in each chart to construct a parametrix for P [458).
306 7. The Noncommutative Integral

Sobolev spaces H~(M) and Hf0 c(M) can be defined for any manifold M:
a distribution u on M lies in Hf0 c(M) if (cp- 1 )*(xu) E H 5 (~n) for every
local chart (U, cp) and every x E C~ ( cp (U)); the subspace of compactly
supported elements is H~ (M). When M is compact, these two spaces agree
and will be denoted by H 5 (M). Proposition 7.23 implies that any P E 'l'd (M)
extends to a continuous linear map P: H 5 (M) - Hs-d(M), and Rellich's
theorem can be restated as follows.
Theorem 7.30 (Rellich). If M is a compact manifold and if 8 > 0, then the
natural inclusion H 5 (M) ._ Hs-t5 (M) is a compact operator. E3

In particular, if P is a pseudodifferential operator of negative order -d


on a compact manifold M, the extended operator P: L2 (M) - Hd(M) is
continuous; composing it with the inclusion Hd (M) ._ L2 (M) yields a com-
pact operator P: L2 (M)- L2 (M).

7.B Homogeneausdistributions
A basic problern in analysis and quantum field theory is that of regulariza-
tion of divergences at a point. Thus, if V is a neighbourhood of a point xo E
~n and u is a (locally integrable, say) function on V\ {xo}, one wishes to
extend u to a distribution on V. This is not always possible, if the diver-
gence of u at x 0 is too wild; indeed, it is known [165] that u is regularizable
at x 0 if and only if u(x) = O(lx- xol-a) as x - Xo, in the Cesära sense
of Section 7.4, for some oc E ~ \ {n- 2, n- 3, ... }.
Here we consider the following special case, where the existence of a reg-
ularization is guaranteed. Suppose that u is a smooth function on ~ n \ { 0},
that is homogeneaus of degree ,\: how can it be extended to a distribution
in S'(~n), and does the extended distribution remain homogeneous? (We
shall first define, by duality, what a homogeneaus distribution is; it turns
out that homogeneity is equivalent to the Euler formula of Definition 7.10
-but in the distributional sense.) Since u and its derivatives have polyno-
mial growth at infinity by homogeneity, the regularized distribution will
be tempered, and it remains to determine the nature of the singularity, if
any, at the origin. This matter is a fairly standard aspect of distribution
theory: we follow [142], [450, §3.8] and [170, Chap. 2], for the most part.
See also [25, §15] foravariant which is suitable for the analysis of some
hypoelliptic operators.
If we denote dilations of test functions by <Pd~) := cp(t~) for t > 0, the
identity

shows that elements of S' (~n) may be dilated by defining


7.B Homogeneaus distributions 307

In particular, the Dirac delta 8 is homogeneaus of degree -n. Since 8 ex-


tends the zero function on !Rn \ {0}, we see that the required extension
need not be unique if ,\ s -n. Indeed, for ,\ = -n- j with jE ~. any linear
combination Llal=j CaDa8 extends the zero function.
In what follows, we assume n ~ 2; the one-dimensional case is somewhat
different, since IR\ {0} has two connected components. Write ~ = rw, with
r = 1~1 and Iw I = 1; then a ,\-homogeneous smooth function on !Rn\ {0}
may be written as
u(~)=r'v(w), where veC""(§n- 1 ).
If ,\ > 0, this is trivially extended by setting u(O) := 0. If 0 ~ ,\ > -n,
then u is still a locally integrable function on !Rn, and the integral (u, cf>) :=
f u(~)c/>(~) dn~ defines the desired distribution.
Suppose, then, that ,\ = -n and assume also that

fsn-1
v(w) a(w) = 0. (7.95)

We may extend u to a principal value distribution Pu, as follows. Choose


a function f: [0, oo) - IR such that f (t) = 1 for 0 s t s f decreases !,
!
smoothly from 1 to 0 on the interval s t s 1, and j(t) = 0 for t ~ 1.
With the cutoff ~- j(l~l), we now define

(Pu, c/>) := f u(~)(c/>(~)- c/>(O)j(l~l)) dn~. (7.96)


JDl"
If f, g yield any two cutoffs, the corresponding expressions differ by

c/>(0) l
0
oo dr
(j(r)- g(r))-
r
f §n-1
v(w) a(w) = 0,

on account of (7.95), and thus Pu is well defined by (7.96).


Let u(~) = 1~1-n v(~/1~1); assuming that (7.95) holds, then

umf
EIO I~I>E
u(~)f(l~l)dn~ = 0,
and thus the principal value distribution is also given by the familiar for-
mula
(Pu, c/>) = limJ u(~)c/>(~) dn~.
EIO I~I>E

Therefore, in this case, Pu is homogeneaus of degree -n.


~ Next consider u(~) := 1~1-n. The right hand side of (7.96) now depends
on the cutoff f (I~ I), but still defines an extension or regularization of u, call
it R JU, inS' (IR n). (Several alternative ways to regularize I~ 1-n are discussed
and compared in [169].) However, RJu is notahomogeneaus distribution.
Indeed, for cf> E S(!Rn),
((RjU)t- t-nRJU, cf>) = t-n (RjU, cf>l!t- c/>).
308 7. The Noncommutative Integral

Since R JU is clearly rotation-invariant, we can evaluate the right hand side


by first integrating over 1~1 = 1, and thus we can as weil assume that cf> is
radial: cf>(~) = !JJ(I~I) for a suitable !JJ. In that case,

(Rju,cf>) =On 1oo


o
dr
(IJI(r) -ljJ(O)j(r))-
r
= On 1oo
o r
dr
( ljJ ( r I t) - ljJ ( 0) f (r I t)) -

for any t > 0; therefore,

((RJuh- t-nRJu.cf>) = On!JJ(O) t-n 1oo


o
dr
{J(rlt)- j ( r ) ) -
r
=Oncf>(O)t-n fooJs dr j'(s)ds
Jo st r
= Onc/>(0) t-n logt.
In other words,
(7.97)

The regularization of UJ(~) := 1~1-n-J, for j = 1, 2, ... , proceeds along


similar lines. We use a cutoff to replace the test function cf> by another
which vanishes to order j at ~ = 0:

Note that two different cutoffs yield regularizations which differ by a finite
sum of delta derivatives:

where Ca:= Ha (j- g) ( 1~1) 1~1-n- i dn~. This procedure is associated with
the names of Epstein and Glaser [162] in quantum field theory computa-
tions.
Exercise 7.20. Show that the dilations of RJ,Ju i are given by
(RJ·U·)t=t-n-JR_f.U·+
,]} ,]} L "'c t-n-i(1-t-i-lai)Da8
IX
lal<i
+ 2: Cat-n- j logt Da8. 0
lal=i

lt is convenient to adjust the regularization by absorbing the terms of


homogeneity higher than -n- j:
7.B Homogeneous distributions 309

The inhomogeneity of Rf,Ju J then has the simpler form:

(Rj,jUj)t = t-n-j RJ,jUj + t-n-Jlogt I CocDoc8. (7.98)


locl=j

.,. Let ;: denote the Fourier trans(ormation, defined on test functions cp E


S(~n) by

One defines the Fourier transform of a tempered distribution by duality;


homogenaus distributions are always tempered. The Fourier transform of
a ,\-homogeneous distribution on ~n is homogeneaus of degree -n- ,\
since, for test functions, :Jcp(tx) = t-n-1l :fcp(x) by the obvious change of
variables, and this formula extends toS' (~n) by duality. Indeed,

((:Ju)t, c/J) = t-n (:Ju, c/Jl/t) := t-n (u,:J(c/Jl!tl) = t-n-1l (Ut,:f(c/Jl/tl)


= t-n-11 (Ut, tn(:Jcp)t) = t-n-11 (u,:Jcp) = t-n-11 (:Ju, c/J).

Now, if u E S'(~n) is both ,\-homogeneous and smooth off the origin,


we can write :Ju = :J(gu) + :J((l- g)u) where g(l~l) is any standard
cutoff. The first summand is analytic since gu has compact support. Also,
v := (1- g)u is a smooth function that vanishes near 0, suchthat if lßl >
,\ + n + Iex I, then Doc(xß :Jv) = :J(~ocvßv) is continuous since ~ocDßv is
integrable. It follows that :Jv is smooth except possibly at x = 0, and so
also is ;:u.
The same argument yields smoothness off 0 of :Ju, when u is an inho-
mogeneaus distribution satisfying either (7.97) or (7.98), since those condi-
tions ensure that such distributions (and their derivatives) have polynomial
growth at infinity.
By Fourier-transforming (7.97), we find that wo(x) := :J(RJI~I-n) is a
rotation-invariant function, smooth except at the origin, which satisfies
t-n(wo(xjt)- wo(x)) =On t-n logt, so that

wo(x) = wo(xft) -Onlogt for t > 0.

Since wo is radial, C := wo(x/lxl) is constant. lt maybe suppressed by sub-


tracting C 8 from the regularization R f I~ 1-n, so we may and shall assume
that C = 0. By taking t = lxl, we thereby arrive at

Wo(x) =-On log lxl for x * 0. (7.99)

In the same way, w 1 (x) := :J(RJ.Ji~l-n-J) is radial, smooth off 0, and


satisfies
t-nw1 (xjt) = t-n-J(w1 (x) +logt I c~xoc),
locl=j
310 7. The Noncommutative Integral

or, more simply,


WJ(xft) = t-JwJ(x) + PJ(x) t-Jlogt,

where PJ(x) is ahomogeneous polynomial of degreej.Again taking t = lxl,


we get
(7.100)
where VJ(X) := lxiJwJ(X I lxl) is also homogeneaus of degree j. Therefore,
for j = 1, 2, ... , WJ(x) remains bounded as x - 0.

7.C Ideals of compact operators


We assemble here, for ease of reference, the definitions and properties of
operator ideals that we refer to at several places in this book. Two excel-
lent general references are the books by Gohberg and Krein [200] and Si-
mon [438]. The particular ideals needed for the Dixmier trace are treated
in [91, IV.C], [469] and [475].
We consider compact operators on a Hilbert space J{; we remind the
reader that we assume this space tobe separable and infinite-dimensional,
so its orthonormal bases are countably infinite.
Let A be a positive compact operator on Jf. The spectrum of A consists
of the number 0 and countably many (positive) eigenvalues of finite mul-
tiplicity (that may be arranged in decreasing order). We can then assemble
an orthonormal basis {Uk} for J{ by choosing a finite orthonormal basis
for each eigenspace, concatenating these, and appending, if necessary, an
orthonormal basis for the kerne! of A. This gives us an expansion
A = I Sk luk)(ukl. (7.101)
b:O
where the coefficients Sk =: Sk (A) are the eigenvalues of A (with multi-
ple eigenvalues repeated), so that so ~ s1 ~ sz ~ . . . and Sk I 0. It is
convergent in the norm of L(Jf): indeed, for each E > 0 there is an inte-
ger N(E) suchthat s0 , ... , SN-I are the coefficients greater than E, and so
IIA- L.~:J Sk luk)(uklll :5 E.
Definition 7.25. If T E X is any compact operator, its "absolute value"
ITl := (TfT)l/2 is a compact positive operator satisfying ker ITl = ker T, so
it admits an expansion (7.101) where the Uk are eigenvectors of ITl, forming
an orthonormal basis for the closure of T(Jf). The polar decomposition
T = UITI is obtained by defining U(ITix) := Tx and Uy := 0 for y E ker T;
thus U isapartial isometry uniquely determined by T. Now let Vk := Uuk;
then T has the norm-convergent expansion
T = I Sk lvk)(ukl. (7.102)
b:O
7.C Ideals of compact operators 311

where {Vk} can also be completed to an orthonormal basis of J{ by ap-


pending a basis for ker T, if necessary. We refer to (7.102) as a "canonical
expansion" ofT; it is unique if and only if allpositive eigenvalues of ITl
are nondegenerate.
The coefficients Sk =: Sk(T), namely, the eigenvalues of ITI, are called
singular values of T; we always list them in decreasing order, repeated
according to multiplicity. Since Tt = Lk Sk IUk} ( Vk I. the basis {Vk} consists
of eigenvectors of (TTt ) 112 • Also, if ul. u2 are unitary Operators on Jf, then

u1 TU2 = :l:Sk IU1 vk}<U1 ukl.


k

and therefore sk<U1TU2) = sk(T).


Compact operators are those that can be uniformly approximated by
operators of finite rank. Any finite-rank operator R an J{ may be written
as R = I}:Ö lvJ} (wJI. where {vo, ... , Vk-d is an orthonormal basis of the
range space R(Jf).
Exercise 7.21. Show that the adjoint operator Rt has the samerank as R,
and that Rt(Jf) c E implies Rlp = 0. 0
Lemma 7.31. 1fT is a compact operator, then

Sk(T) = inf{ IIT- Rll: rankR ::5 k} = inf{ IITip II: dimE = k }.

Proof. If Eisa finite-dimensional subspace of Jf,let h denote the orthogo-


nalprojectorwithrangeE.Notethat IITIPII = liT-TPEll andrank(TPE) ::5
dimE; this, tagether with Exercise 7.21, implies that these infima are equal.
If {uk} is an orthonormal basis of J{ for which (7.102) holds, taking R :=
I}:ö s1 (T) u 1 or E := span{uo, ... , Uk-d shows that the infima are at-
tained. If Fis a k-dimensional subspace other than E = span {u 0 , .•• , Uk-l},
then IITxll ~ Sk(T)IIxll for some nonzero x E E n P·, and so sk(T) ::5
IITIPII. 0

Definition 7.26. Several subclasses of compact operators may be deter-


mined by imposing suitable conditions on the singular values. If 1 ::5 p < oo,
we may define the Schatten p-class LP by

2:: Sk(T)P < oo.


00

TE LP if and only if (7.103)


k=O

This is anidealin X and hence also in L(Jf). lnparticular, L 2 is the Hilbert-


Schmidt dass, and L 1 is the trace dass, whose norm is the trace of ITI,
00 00

IITII1 :=TrI Tl = L (uk IITiuk) = L Sk(T).


k=O k=O
312 7. The Noncommutative Integral

For p > 1, LP is a Banachspace [415], whose norm is the .fP-norm of the


sequence {Sk ( T)}:

(:L Sk(T)P)
00 1/p
IITIIp := = (TriTIP) 11P.
k;O

1t is immediate that T E LP if and only if ITI P E L 1 •


The trace Tr T := .l:k';o (Uk I Tuk) is absolutely convergent if and only
if T E L 1 , and it defines a linear functional on this Banach space (that is
independent of the orthonormal basis used). If A is a compact positive ope-
ratornot in L 1 , we write Tr A := + oo, and thus regard Tr as a positive linear
functional on all of X, whose domain of finiteness is just L 1 • The trace
satisfies Tr(TS) = Tr(ST) whenever TS and ST lie in L 1 , or equivalently,
Tr{U AUt) = Tr A whenever Ais positive and U is unitary.
If so(T) = IITII :::;; 1, the sum (7.103) decreases as p increases, so that
LP c Lr if p < r (the singular value sequence Sk := k-I/p produces an
operator in Lr but nor in LP). lt is also easy to see that if II Tll = 1, then
IITIIp - 1 as p- oo; for this reason, the convention L"" :=Xis sometimes
used, since it allows certain propositionstobe stated for 1 :::;; p :::;; oo.
Recall the Hölder inequality (7.73) for sequences; by taking ak := Sk(T)
and bk := Sk(S) for compact operators T,S, we obtain
1 1
Tr ITSI :::;; IITIIp IISIIq whenever - +- = 1.
p q

In particular, TS is traceclass whenever TE LP and SE Lq. The case p = 1,


namely IITSIII :::;; IITIII 11511, remains validforT E L 1 and SE L(.J{).
Exercise 7.22. Generalize the previous inequality to the case of compact
operators To, ... , Tn. with Ti E LPi and p 01 + P1 1 + · · · + p;; 1 = 1:

(7.104)

provided that each Ti E LPi and p 01 + P1 1 + . · · + p;; 1 = 1. 0

.,. lt is often useful to work not with the singular values individually, but
with their partial sums:

Un(T) := So(T) + si(T) + · · · + Sn-dT). (7.105)

Lemma 7.32. IfT is compact, then Un(T) = sup{ IITPEIII: dimE = n }. IfT
is also positive, then Un(T) = sup{ Tr(PETh): dimE = n }.

Proof. If T and S are compact Operators such that yt T ~ st S ~ 0, then


(1- PE)TtT(1- PE) ~ (1- PE)sts(l- PE) ~ o, so that IITIP II ~ IISIP II
for any finite-dimensional subspace E of J{; from Lemma 7.31 it follows
7.C Ideals of compact operators 313

that Sk(T) ~ Sk(S) for each k ~ 0. In particular, if dimE = n, then Sk(T) ~


sk(TPE) for k = 0, 1, ... , n- 1, and thus

n-1 n-1
O"n(T) = I Sk(T) ~ I sk(TPE) = IITPEII1.
k=O k=O

Thus, O"n(T) ~ sup{ IITPEII1 : dimE = n}. On the other hand, taking
F := span { uo, .. . , Un-1 } shows that II T PF ll1 = O"n ( T), so the supremum is
attained.
If T is also positive, then Tr(PETPE) = Tr(TPE) ::; IITPEII1 ::; O"n(T);
moreover, O"n(T) = IIThll1 = Tr(TPF) = Tr(PFTh), because now PF =
.Li::J luk) (uk I commutes with T and hence TPF = PFTPF is positive. D
Corollary 7.33. Each O"n obeys the triangle inequality: O"n (S + T) ::; O"n (S) +
o-n ( T), and thus defines a norm on X. This norm satisfies o-n ( T) ::; n II Tl!.

Proof. lt is enough to note that II(S + T)PE111 ::; IIShll1 + IITPEII1 for each
subspace E with dimE = n. Since sk(T) ::; s0 (T) for k = 0, ... , n- 1, it
follows that O"n(T)::; nso(T) = niiTII. D

The following characterization of O"n ( T) is useful; see [34, Prop. IV.2.3]


or [91, IV.2.oc].
Proposition 7.34. O"n(T) = inf{ IIRII1 + niiSII: R,S EX, R + S = T }.
Proof. Both sides of this equation are invariant under the replacement of
T by VT with V unitary, so, by polar decomposition, we can assume that T
isapositive operator.
If T = R + S, then o-n(T) ::; o-n(R) + o-n(S) ::; IIRII1 + niiSII by Corol-
lary 7.33. To see that the infimum is attained, take F := span {uo, ... , Un- 1}
where uo, ... , Un-1 are the leading eigenvectors of the positive operator T,
and split T as Rn+ Sn where Rn:= (T- Sn(T)l)h and Sn:= Sn(T)PF +
T(l- PF); clearly, IISnll = Sn(T) and Rn= .Li::J<sk(T)- Sn(T))Iuk)(ukl,
so that II Rn ll1 = .Li::J (Sk (T)- Sn (T)) = O"n (T)- nsn (T), and thus O"n (T) =
IIRnll1 + niiSnll. D

~ Many other ideals of compact operators can be defined using the follow-
ing device. (We need not concern ourselves with noncompact operators: a
theorem of Calkin guarantees that an ideal of L(J-f) containing even one
noncompact operator must be all of L(J-f) [438, Prop. 2.1].)
Definition 7.27. A norm 111·111 on an ideal J 5; Xis called symmetric if

IIIRTSIII::; IIRIIIIITIIIIISII for all TE J; R,S E L(J-f).

We say that J is a symmetrically normed ideal if J is complete in the


norm 111·111; it is then a Banach space.
314 7. The Noncommutative Integral

Exercise 7.23. Show that 5k(ST) ::s; IISII 5k(T) and 5k(ST) ::s; IITII 5k(S) for
S, TE X. Conclude that all the norms G"n and II · llp are symmetric. 0
A symmetric norm is unitarily invariant, that is, IIIUTVIII = IIITIII when-
ever U and V are unitary. 1t follows from the canonical expansion (7.102)
that a unitarily invariant norm depends only an the singular values ofT,
and the previous exercise then shows that it is a symmetric norm. Thus, an
arbitrary symmetric norm is given by

IIITIII = <I>(5o(T),5I(T),52(T), ... ), (7.106)

for a suitable function <I>.


Definition 7.28. Let coo be the space of complex sequences with only finite-
ly many nonzero terms. A symmetric gauge function is a function <I> : c 0 o -
[0, oo) suchthat
(a) <I> is a norm an coo,
(b) lf>(x) = <I>(a) if ak = lxk I for each k,
(c) <I>(x) = <I>(y) if Yk = Xrr<k> for some finite permutation rr, and
(d) <1>(1,0,0, ... ) = 1.
Symmetrie gauge functions have the following monotonicity property
[438, Thm. 1.9]: if Ir,:-J Xk ::5 Ir,:-J
Yk for all n, then <I>(x) ::5 <I>(y). Prop-
erty (c) means that if x is a positive sequence, we may rearrange the en-
tries of x in decreasing order when computing <I>(x). We can then ex-
tend <I> to infinite sequences (converging to zero) provided that <I>(x) :=
SUPn <I> (XI, ... , Xn, 0, 0, ... ) remains finite.
Before going to the examples, we take note of the following algebraic
property of symmetric normed ideals.
Lemma 7.35. Any symmetric normed ideal J on J{ is linearly generated by
its positive elements.
!( i
Proof. If T E J, then T = T + yt) + (i yt - i T), so J is linearly generated
by its selfadjoint elements.
Suppose, then, that T E J is selfadjoint. On account of (7 .1 06), ITI lies
in J also. If T = UITI is the polar decomposition ofT, we can define a
symmetry F commuting with ITI by setting F := 1 ankerT and F := U
!
an (ker T) .L. The projectors P ± := (1 ± F) also commute with ITl; indeed,
P+ is the spectral projector ofT for the interval [O,oo), so that P±ITI =
P±ITIP± arepositiveoperatorslyinginJ,and T = (P+-P-)ITI = P+ITIP+-
P-ITIP-- D
Definition 7.29. The operator ideal

L<l> := {TE X: <l>(5o(T), 51 (T), 52 (T), ... ) < oo}


7.C Ideals of compact operators 315

is a symmetrically normed ideal, with the norm given by (7.106). Examples


are X with the norms CTn or the usual operator norm (by taking ct>(x) :=
maxk lxk I) and the Schattenideals LP.
If 'Y is another symmetric gauge function, then L4> = L'"~ if and only if ct>
and 'Y are equivalent in thesensethat C1'Y::; ct>::; C2'Y, for some constants
satisfying 0 < C1 ::; C2 < oo.
Exercise 7.24. Show that maxk lxkl ::; ct>(x) ::; Ik lxkl for any symmetric
gauge function ct> and any x E c 00 . Conclude that L 1 !;;; L4> !;;; X. 0
There are symmetrically normed ideals that are not separable, because
the finite-rank operators are not dense. Given T E L4> with canonical ex-
pansion (7.102), write Tm:= Ik=osk(T) lvk)(ukl; then Tm E L4> also. The
finite-rank operators are denseifand only if IIIT- Tmlll - 0 as m - oo.
Thus, L4> is separable if and only if

lim ct>(xm+l•Xm+2· ... ) = 0 whenever ct>(x) <


m-oo
oo. (7.107)

When this condition is not fulfilled, we denote by L~ the closure of the


finite rank operators.
Any ct> has a dual symmetric gauge function, given by the dual norm
on c 00 , which is easily verified to satisfy the requirements of Definition 7.28:

ct>'(y) := sup{ l(x I y)l: ct>(x) = 1},

where (x I y) := Ik XkYk· To examine the Banach-space duality of the sym-


metric ideals, we first consider the sequence spaces c4> = { x : ct>(x) < oo},
normed by ct> itself, and its subspace c~, defined as the closure of coo. A
standard sequence-space argument [438, Thm. 1.17] shows that the dual
space to c~ is isometrically isomorphic to c4>'. This has the following con-
sequence.
Proposition 7.36. If L~ =t- L 1 , the dual Banach space to L~ is isometrically
isomorphic to L4>', and the duality is given by tunetionals of the form T ,_
Tr(ST), for S E L4>'. E3

The duality of the Schatten ideals LP and Lq when p > 1, q = p I ( p - 1),


is a special case of that. The case p = 1 is exceptional since the dual space
of L 1 is L(Jf) rather than X.
There is a device that produces many interesting examples of nonsepara-
ble ideals [200,475]. For any sequence x, write CTn (x) := xo+X1 + · · · +Xn-1;
for instance, if x is the sequence of singular values of a compact operator T,
then CTn(x) = CTn(T) by (7.105). Now suppose z is a decreasing sequence
of positive numbers with z 0 = l. It yields two symmetric gauge functions,

ct>z(x)
n
CTn (x)
:= sup --(-),
CTn Z
ct>;(y) := L YkZk.
hO
316 7. The Noncommutative Integral

that are dual to each other. By [200, 111.14), .[<f>z = .L 1 if the sequence z
is summable; .[<f>z = X if lim1 z 1 > 0; andin the intermediate case that
limJ ZJ = 0 but the series 2.1 z 1 diverges, (7.107) does not hold, so .[<f>z is
not separable. The dual ideal .[<~>;, however, is always separable.
Example 7.4. The Dixmier ideal.L 1+ falls und er this heading, by taking Zk : =
11(k + 1); the gauge function 'Y(x) := supn O"n(X)/logn that defines the
Dixmier ideal by (7.69) is equivalent to <l>z. The subspace .Lb+ is in fact the
common kerne} of all the Dixmier traces. lts dual space x- := { T E X :
Lho Sk (T) I (k + 1) < oo} is called the Ma<;aev ideal of .L(Jf) [322). Hölder's
inequality shows that x- includes every Schatten ideal .Lq, and by duality
.L 1+ c .LP for each p > 1.
Example 7.5. If 1 < p < oo, Iet Zk := (k + u-
1/P, q := pl(p - 1). The
notations .[P+ := .[<f>z and .Lq- := .[<~>; denote the corresponding operator
ideals, with the respective norms
O"n(T) ' Sk(T)
IITIIp+ := s~p n(P-lllp, IITIIq- := k7o (k + 1)1/p.
(7.108)

Then .Lq- is separable with dual space .[P+; the latter is not separable, but
the dual of .Lg+ is .Lq-. The notation comes from the following inclusions.
Exercise 7.25. Show that .Lr c .Lq- c .Lq if 1 :5 r < q. Conclude that
.[P C .[P+ C .L5 if 5 > p. 0

Exercise 7.26. If 1 < p < oo, a compact operator T lies in .LP+ when
O"n (T) = O(n(P- 1 llp) as n - oo. Show that TE .[P+ if and only if NITI (,\) =
0(,\P) as ,\- oo, where NITI is the counting function of ITl. o
Lemma 7.37. If 1 < p < oo and A E .[P+ is positive, then AP E .L 1+. How-
ever, the converse need not hold.
Proof. Suppose that A E .[P+ with IIAIIp+ = C; then O"n(A) :5 Cn(P- 1 llp for
JJ
n ~ 1. Since t(P- 1 liP = (p- 1)1p s- 1 1P ds, we can find C' ~ C(p- 1)/p
so that Sk(A) :5 C'(k + 1)- 1/P for each k. Then Sk(AP) :5 C'P /(k + 1), and
so O"n (A) :5 C" log n for all n, with C" ~ C'P.
To give a counterexample [479) of a positive operator BE .L 1+ for which
A = B 1 1P does not lie in .[P+, it suffices to construct a sequence x ofpositive
numbers suchthat O"n (x) I log n is bounded, but k Xk is not. (Then take A :=
Lkxt 1Piuk)(uki.) Let xo := 1, and define Xk := (logm)lm! for (m -1)! :5
k < m!. It is easily checked by induction that O"m! (x) :5 log m! and therefore
O"n(X) :5 logn for all n; however, limsupk-oo kxk = limm-oo m!Xm! = oo.
0

We mention another instructive example, also taken from [479).


Exercise 7.27. Define a sequence x by log(11xk) := logk- ~ogk. Show
that the positive operator Lk Xk iuk) (Uk I belongs to each .LP for p > 1, but
not to .L 1+. 0
7.C Ideals of compact operators 317

For other examples of symmetrically normed ideals, we refer to [91,


IV.2.oc]. These examples form a two-parameter family .f.(p,q), with 1 < p < oo
and 1 5 q 5 oo, constructed by Banach-space interpolation starting from
the pair of spaces L 1 and X. They can be matched to points (p- 1 , q- 1 ) of
the unit square (except for the left and right edges). Same of these interpo-
lation spaces are examples we have already seen: each .f.(p,oo) on the lower
edge of the square is our .f.P+, each .f.(P. 1 l on the upper edge is our .f.P-,
and on the diagonallie the Schatten classes .f.(p,p) = .f.P. The corners are
X, L 1 , the Dixmier ideal .f. 1+ and its dual, the Ma<;aev ideal x-.
~ A feature of Banach-space interpolation theory is that the norms for
intermediate spaces may be estimated precisely in terms of the norms of
the extremes. We shall need later one such estimate [91, IV.2.8]: if 1 < p <
oo, there is a constant ßp > 0 suchthat for any traceclass operator T, the
following inequality holds:

(7.109)
8
Noncommutative Differential Calculi

Until now, our noncommutative callisthenics have involved generalizations


of topology and linear algebra, and a new integral. We are now ready to cross
the Rubicon into differential calculus. The first, and crucial, step is to intro-
duce first-order differential forms on a noncommutative space defined by
a (complex) pre-C*-algebra .Jt. We say crucial because, in most developed
differential calculi (e.g., the usual de Rham complex of differential forms
on a manifold), specification of what is to be understood by a space of 1-
forms is effectively enough to introduce the full calculus. Weshallstart by
indicating the simplest thing that one can do barehanded withjust the alge-
bra, that is, introduce the module of universall-forms. The construction is
actually simpler if .Jl is assumed to be noncommutative. lt has a rather ab-
stract appearance, but, as we shall eventually see, has something to do with
noncommutative geometry proper. Along this road, near the end, we arrive
at the Hochschild-Kostant-Rosenberg-Connes theorem, which amounts to
a homological construction of differential forms. This is one of the key re-
sults in this book, and in the whole of noncommutative geometry. Along
the way, we prove the Chern isomorphism theorem.
There are several differential algebras in this chapter, and we trust that
the reader will not get too confused by the many differentials: d, 4_, d, 8
and so on. Then there is the question of the topological setting of the cal-
culi we use. lt has no general answer, so we are often explicit only on the
abstract algebraic aspects. At any rate, we tacitly assume that the maps be-
tween topological algebras are continuous; and, of course, the homological
constructions will require topological versions of the tensor product.
320 8. Noncommutative Differential Calculi

8.1 Urüversalfornns
Let 'E be a bimodule over a (complex) algebra .Jt that weshall suppose unital.
Definition 8.1. By a derivation D: .Jt - 'E, we understand a linear map that
satisfies the Leibniz rule
D(ab) =Dab+aDb.
Note that D kills the constants: D(1) = 2D(l), so D(1) = 0. Weshalllet
Der(5l, 'E) denote the complex vector space of all5l-bimodule derivations
with values in 'E. Any element m of 'E defines a derivation:
ad(m)a := ma- am,
called an inner derivation; a bimodule is called symmetric if allinner deriva-
tions are trivial. Weshall denote by Der' (.Jt, 'E) the subspace ofinner deriva-
tions.
We remark that the module Der(.Jt, .Jt) has a complex Lie algebra struc-
ture, since the commutator of two derivations is a derivation.
We pander the following "universal" problem: find a derivation d from .Jt
into a bimodule Q 1.Jt, such that, given any derivation D of the unital algebra
.Jt into a bimodule 'E, there is a unique bimodule morphism LD: 0 1.Jt - 'E
with D = LD 0 d,

y
Q1.Jt

',~:
5l D 'E.
The assignment cp - cp o d defines a linear map
HomA (0 1.Jt, 'E) - Der(5l, 'E);
the universal property is the assertion that this arrow is an isomorphism. It
is clear that such an universal derivation (0 1.Jt, d) is uniquely determined,
up to isomorphism.
The linear map d: .Jt- .Jt ® .Jt given by da:= 1 ® a- a ® 1 obeys
d(ab) = 1 ®ab- ab® 1 = a ® b- ab 181 1 + 1 ®ab- a 181 b
= adb + dab, (8.1)

so d is a derivation. Let Q 1.Jt be the subbimodule of .Jt ® .Jt generated by


elements of the form a db. Then
0 1.Jt = ker(m: .Jt ® 5l- .Jt),
where m: a ® b - ab is the multiplication map. Indeed, if I.J aJ ® bJ E
ker m, then I.J a J b J = 0, so we can write
I.J a J ® bJ = I.J a J ( 1 181 b J - bJ 181 1) = I.J a J db J.
8.1 Universal forms 321

(Here and from now on, a simple Lj denotes a finite sum.)


The left and right module structures on 0 1 5l. are respectively given by
a'(adb) = a'adb, (adb)a" = ad(ba")- ab da". (8.2)

Thus, our construct yields what may be called a first-order differential cal-
culus for 5l..
Note also that 0 1 5l. is generallynot symmetric even if 5l. is commutative.
Weshall call 0 1 5l. the bimodule of universall-forms over 5l.. Suppose now
that 'E is any 5l.-bimodule and let D: 5l. - 'E be a derivation. We define
tv: 0 15l. - 'E by its action on simple tensors:
tv(a ® b) := aDb, (8.3)
restricted to 0 15l.. Then tv is a bimodule morphism, because of (8.2) and
the derivation property of D; also, tv(da) =Da.
~ Assurne now that 5l. is commutative, and consider derivations of 5l.-
modules that verify the Leibniz rule
D(ab) = bDa + aDb.
(We shall write the module action on the left, as is usual in homological
algebra books.)
Exercise 8.1. Show that in the commutative case Der(5l., 5l.) possesses, in
addition to its Ue algebra structure, an 5l.-module structure. 0

The construction of a module of universal 1-forms Iooks paradoxically


more complicated in the commutative case [319], for which it makes sense
not to distinguish between the "universal1-forms" a db and db a. Thus we
pander again a universal problem: find a derivation 4 of the commutative
unital5l. into an 5l.-module n~b5l., such that, given any derivation D of 5l.
into a module 'E, there is a unique module map «/Jv: n~b5l. - 'E suchthat
D = «/Jv o 4_. In this case, the isomorphism
1
Hom_:a. (Oab5l., 'E) -- Der(5l., 'E) (8.4)

will be a module map.


As any module on a commutative algebra is trivially a symmetric bimod-
ule, it stands to reason that n~b5l. will be a quotient of 0 1 5l.. Weshallshow
that n!b5l. = 0 15l./ (0 15l.) 2 . Here (0 15l.) 2 is a subbimodule of 5l. ® 5l., ob-
viously included in 0 1 5l., and
adb-db a = a®b-ab® 1-1 ®ba+b®a = -(1 ®a-a® 1)(1 ®b-b® 1),
since 5l. is commutative. Thus, the quotient 0 15l./ (0 15l.) 2 is a symmetric
5l.-bimodule, and therefore can be regarded as an 5l.-module by identifying
the left and right actions of 5l.. The differential is defined to be
4a := (1 ® a- a ® 1) mod (0 15l.) 2 •
322 8. Noncornmutative Differential Calculi

The computation in (8.1) shows again that 4 is a derivation. To verify the


universal property, consider a derivation D: 5'1. - 'E, and define !fJD by
!fJv(a ® b mod (0 1 5'1.) 2 ) := aDb. This is well defined: whenever LJ a 1 ® b1
belongs to 0 1 5'1., then
L.1 (aj®bj)(1®c-c®1) ~ L.1 ajD(b1c>-L.1 a 1cD(b1 ) = L.1 a 1b1 Dc = 0,
and it follows that !fJv((0 15'1.) 2 ) = 0. Clearly, !fJD o 4 = D, and !fJD is unique
since 0 15'1. and 0 15'1./ (0 15'1.) 2 are generated respectively by the images of
d and4.
The elements of n!b5'1. are sometimes called Kähler differentials. Note
that the presentation of n!b5'1., under the form
n!b5'1. := 0 15'1. I { L.1 a 1 db1 - db1 a 1 : a 1, b1 E 5'1.},
makes sense also for noncommutative algebras, but it no Ionger equals
Ql.Jl/ (Ql.Jl)2.

11> Returning to the general case where 5'1. is a unital but not necessary
commutative algebra, we can now construct a universal graded differential
algebra over 5'1..
Definition 8.2. A graded differential algebra (R", 8) is a (complex) asso-
ciative algebra R" = EBk'=o Rk whose product is graded (in other words,
Rk R 1 ~ Rk+ 1), tagether with a differential 8, namely, a linear map of de-
gree + 1 such that 8 2 = 0 is an odd derivation:
8(Wk1J) = (8Wk) 1} + (-1)kWk 817 when Wk E Rk. (8.5)
We shall use the notational convention, when dealing with graded differen-
tial algebras, that Wk denotes a homogeneaus element of degree k.
What we want, then, is a graded differential algebra n· 5'1. = EBk'=o Qk 5'1.
with 0° 5'1. = 5'1. and 0 15'1. as already defined, endowed with a differential d
that extends the derivation from 5'1. into 0 15'1.. Moreover, if (R", 8) is an-
other graded differential algebra, any algebra homomorphism !fJ of 5'1. into
the degree-zero subalgebra R 0 should be lifted to an algebra homomor-
phism of degree zero !fJ: n· 5'1. - R" intertwining the differentials d and 8.
The algebra product in 5'1. and the derivation d must determine the product
in n· 5'1..
Denote 5'1. := 5'1./1(; for brevity, we write ä E 5'1. for the image of a E
5'1. under the quotient map. Then we remark that 5'1. ® 5'1. = 0 1 5'1. by the
identification ao ® ä1 ,.._ ao da1, which is well defined since d1 = 0. If
c E 5'1., then c(ao ® ä1) ,.._ cao da1, while
(ao ® ä!)c := ao ® a1c- aoa1 ® c~ ao d(a1c)- aoa1 dc = ao da1 c,
so this correspondence is a bimodule isomorphism. The reader will have
also noted the direct decomposition of 5'1.-bimodules:
(8.6)
8.1 Universal forms 323

Put 0 2 .J\ := 0 1 5\ ®.Jt 0 15\ = (.Jl ® .Jl) ®.Jt (.Jl ® .Jl) = .J\ ® .J\ ® .J\. More
generally, define nn .J\ := 0 15\ ®.Jl 0 15\ ®.Jl •.• ®.Jl 0 15\ (n times), so that
nn .J\ = .J\ ® -®n
.J\ ; in other words, we take the tensor algebra over .J\, but
we quotient out the scalar terms except in degree zero. The differential
d: .J\ ® .J\ ®n - .J\ ® .J\ ®(n+ 1l is given simply by the shift

d(ao ® ä1 ® • · • ® än) := 1 ® äo ® ä1 ® · • · ® än.

Since I = 0, we get d 2 = 0 at once. Starting from an in degree zero, multi-


plying on the left and applying d repeatedly gives

We make n· .J\ an .J\-bimodule. The left module property is immediate:


a' (ao da1 ... dan) = (a' ao) da1 ... dan.

To get the right module property, we use the postulated derivation property
da b = d (ab) - a db to pull the elements of A through to the left:

(ao da1 ... dan)a"


= ao da1 ... dan-1 d(ana")- ao da1 ... dan-1 an da"
= · · · = ( -l)naoa1 daz ... dan da"
n-1
+ L (-l)n-Jaoda1···d(aJaJ+1) ... danda"
j=1

+ aoda1 ... dan-1 d(ana").


Lastly, we define

so that n· .J\ becomes a graded algebra; weshall call it the universal graded
differential algebra over .J\.
Notice that d(ao da1 ... dan) = dao da1 ... dan. There is also the useful
formula
ao [d, ad ... [d, anll = ao da1 ... dan,
where the a 1 on the left hand side are regarded as left multiplication ope-
rators. This is easily verified by induction: [d, an] 1 = dan - an d1 = dan,
and [d, an-d dan = d(an-1 dan)- an-1 d(dan) = dan-1 dan.
Now, suppose we are given the differential graded algebra (R", c5), and an
algebra homomorphism lJ.I: .J\ - R0 • The universality of n· .J\ is now clear,
since we can extend this map to the homomorphism lJ.I: n· .J\ - R" given
by
324 8. Noncommutative Differential Calculi

Exercise 8.2. Check that this ljJ is a graded algebra homomorphism. 0


Note the following generalization of (8.6): for each n ;e: 0, the bimodule
sequence

where i(wda) := w ® a- wa ® 1, is exact.


Dubois-Violette [148] has pointed that the action an .Jt oftheUe algebra
of derivations Der(.Jt, .Jt) can be extended to an action an n· .Jt, in the
following way.
Exercise 8.3. Given D E Der(.Jt, .Jt), show that the bimodule morphism
LD: 0 1 .Jt - .Jt extends to an odd derivation of n· .Jt of degree -1 (which
we also denote by Lv). Checkthat Lv := Lv o d + d o Lv is an even derivation
of n· .Jt and establish the supercommutator relations:
[Lv, Lß] = 0, [Lv, Lß] = L[D,ß], [Lv, Lß] = L[D,ß],
for all D,!!. E Der(.Jt, .Jt). 0

Before continuing, Iet us note that, for a nonunital algebra .Jt, the preced-
ing constructions can be performed an the unitization .Jt +. Many authors,
including Connes [86, 91], construct the universal differential algebra an
.Jt +, whether .Jt is unital or not to begin with. In that case, the degree-
zero term is declared tobe .Jt rather than .Jt +. Also, Kastler [280] has dealt
at length with the case in which .Jt is a superalgebra. The "purely alge-
braic" construction of n· .Jt given here is a simplification of the "tangent
algebra" construction of Arvesan [10], which Iifts bounded derivations to
morphisms of C* -algebras.
~ Hereis an important example. ForA = C(M), with M compact, we can
identify A®n with C(M x · · · x M). lndeed, C(M) 18> C(N) ""C(M x N) under
the obvious identification (g ® h) (x, y) := g(x)h(y), since the subalgebra
of C(M x N) generated by such simple tensors is dense (Stone-Weierstrass
again). Also, for .Jt = C"' (M), we can identify .Jt ®n with C"' (Mx · · · x M),
as we shall see in Section 8.5. The multiplication maps m(g ® h)(x) :=
(gh)(x) = g(x)h(x) come from restriction to the diagonal in Mx M. If
h E A, then dh(xo,XI) = (1 ® h- h ® 1)(xo,xd = h(xi)- h(xo). Thus
0 1 A (or 0 1 .Jt, as the case may be) is identified with the set of functions of
two variables vanishing an the diagonal. lt is instructive to check the left
and right actions of A on 0 1A given by
(gj)(xo,xd := g(xo) j(xo,x!),
(jg)(xo,xd := j(xo,x!)g(x!).
More generally, nn Ais identified with the set of functions of n + 1 variables
vanishing an contiguous diagonals. The differential is given by
n
dj(xo, ... ,Xn) := L(-l)kj(xo, ... ,Xk-I,Xk+l•····Xn),
k=O
8.1 Universal forms 325

and d 2f = 0 is immediate. The left and right actions of A on nn A are given


by

(gj)(xo, ... , Xn) := g(xo) j(xo, ... , Xn),


(jg)(xo, ... ,Xn) := j(xo, ... , Xn) g(Xn).

The product of an m-form fandann-form h is

fh(xo, ... ,Xm+n) := j(xo, ... ,Xm) h(Xm, ... ,Xm+n).

Even after restricting to the smooth subalgebra 5\ = coo (M), this is obvi-
ously much larger than the space of differential forms in one "variable".
(There is an obvious surjective map from n· 5\ onto the de Rham complex
5\ • (M) of differential forms on M, resulting from universality.)
In fact, the universal complex (0" (M), d) has a cohomology that is trivial
except in dimension 0. To get a nontrivial cohomology, one can restriet to
neighbourhoods of the diagonal, as follows. Call f E nn A "locally zero"
if there is an open cover {Uj} of M suchthat j(xo, ... ,Xn) = 0 when-
ever x 0 , ..• , Xn E u1 for some j; it is clear that df is also locally zero.
After factaring out the subcomplex of locally zero elements, the quotient
(0" (M), d) is known as the Alexander-Spanier complex of M, and its coho-
mology is no langer trivial: it is the Alexander-Spanier cohomology of M
(with complex coefficients). This is known to be isomorphic to the Cech
cohomology of M [442, Chap. 6], and hence to the de Rham cohomology if
M is a smooth manifold [41]. See [111] and also [350], which treats the dual
homology theories.
In the commutative case, it is usual to form simply the exterior algebra
i\5t n!b.A of n!b.A over 5\, denoted n;b.A; this is "supercommutative"' that
is to say, WkWZ = (-1)k 1wzwk if Wk. wz have respective degrees k and l.
The graded differential algebra n;b.A is a cochain complex, where 4 acts
as the coboundary operator.
Exercise 8.4. Formulate and prove a universal property of n;b.A for mor-
phisms of 5\ into the degree-0 subalgebra of a supercommutative graded
algebra. 0
Proposition 8.1. The exterior algebra n;bcoo (M) may be identified to the
de Rham complex 5\ • (M) of differential forms on M.

Proof. The Lie algebra of derivations of 5\ = C"" (M) is just the space Jt(M)
of vector fields on M (with complex coefficients). Therefore, on account
of (8.4), n!b.A is indeed identified to the 5\-module 5\ 1(M) of first-order
differential forms on M; and then, on account of Proposition 2.6, n;b.A is
identified to the 5\-module 5\ • (M) of all differential forms on M. D
The equality 5\ 1 (M) = 0 1 C"" (M) 1(0 1 Coo (M) )2 conveys nontrivial infor-
mation; it has already the flavour of the locality of continuous Hochschild
326 8. Noncommutative Differential Calculi

homology of C"" (M) that will be instrumental in understanding the Hoch-


schild-Kostant-Rosenberg-Connes theorem.
The approach to "noncommutative differential calculus", based on uni-
versal differential forms, looks at first sight unsatisfactory. Sometimes
n· .Jt is so huge that an integral may not be defined on it. We now turn
to an apparently more demanding approach, trying to serrer de plus pres
the spirit of the de Rham complex, in the noncommutative geometric con-
text; we shall return to n· .Jt afterwards.

8.2 Cycles and Fredholm modules


Definition 8.3. An n-dimensional cycle is a (complex) graded differential
algebra n· = EBr=o nk, i.e., nk = o for k > n, tagether with an integral f,
i.e., a linear map f: n· - C, suchthat f Wk = 0 for k < n,

f WkWL = (-l)kl f WtWk and f dWn-l =0


for homogeneaus elements of the indicated degrees. Given an algebra .Jt, a
cycle over .Jt is a cycle (n•, d, f) tagether with a homomorphism from .Jt
to 0°.
The simplest examples are afforded by de Rham complexes. Let 0° be
an algebra of smooth functions on ~n, with values in Mm ( 0 and appropri-
ate growth conditions. Then the usual exterior differentiation constructs a
cycle of dimension n, with integral

J Wn := JIR" tr Wn.
Or take a compact smooth manifold M of dimension n without boundary,
and consider again the space of smooth differential forms. Then, more
generally, any closed de Rham current C on M defines a cycle, with di-
mension 5 n and integral w - fc w, since fc dw = fac w = 0 by Stokes'
theorem.
~ A very interesting dass of examples comes from Fredholm modules over
a given algebra.
Definition 8.4. Let .Jt be an algebra (often a C*-algebra). An odd Fredholm
module over .Jt is given by an involutive representation a of .Jt and a sym-
metry F (i.e., an operator satisfying F = pt and F 2 = 1) on a (usually sepa-
rable) Hilbert space Jf, such that
[F,a(a)]EX(Jf) forall aE.J\.. (8.8)
An even Fredholm module involves an (even) representation a = a 0 e a 1 of
the algebra .Jt on a l 2 -graded Hilbert space Jf 0 e Jf 1 , and F is now an odd
8.2 Cycles and Fredholm modules 327

symmetry on Jf satisfying (8.8) -in other words, F essentially intertwines


0" 0 and 0" 1.

We have already met this structure in Section 6.4, in a slightly different


guise: if] and Q are commuting orthogonal complex structures on a real
Hilbert space (V,g), and Jf = VQ, then F := -i] is a symmetry on Jf.
An algebra satisfying (8.8) is, for example, the algebra generated by the
restricted orthogonal group 0 1 (V).
A pre-Fredholm module is defined similarly to a Fredholm module, except
that, instead of demanding that F be a symmetry, we ask only that each
O"(a)(F- Ff) and O"(a)(F 2 - 1) be compact. A Fredholm module can be
associated to each pre-Fredholm module, as follows; we consider only the
(unital) even case, as the odd case can be recovered by using the trivial
grading. Suppose, then, that F is an operator on Jf 0 EB Jf 1 of the form

F:=(~ ~). (8.9)

where P E Fred (Jf 0 , Jf 1), Q E Fred (Jf 1, Jf 0 ), and 1 - PQ and 1 - QP are


compact. We double the graded Hilbert space by setting !it 0 := J{ 0 EB Jf1,
j:{1 := J{1 EB Jf 0 , put a-o(a) := 0" 0 (a) EB 0 and 6" 1(a) := 0" 1(a) EB 0, and
. - (0p Q)0 , where
define F :=

- ( P 1- PQ ) - ._ (2Q- QPQ
p := 1 - QP QPQ- 2Q ' Q .- 1-PQ

Exercise8.5. Checkthat Q = .P- 1, so that F2 = 1, and that [F,G"(a)] is


compact for each a E .Jl. Explain why Fis selfadjoint. 0
Let M be a smooth compact manifold, .Jl = C"" (M) and let Jf 0 , Jf 1 be
the respective Hilbert spaces of sections of Hermitian vector bundles E0 ,
E 1 over M; let P be an elliptic pseudodifferential operator of order 0 from
E0 to E1 that willlie in Fred(Jf0 , Jf 1), and consider F given by (8.9) with
Q a parametrix for P. This setup clearly defines a pre-Fredholm module.
This classical example comes from Atiyah's work on index theory [14] and
justifies thinking of Fredholm modules as "abstract" or "noncommutative"
elliptic operators.
The main point for us hereisthat a Fredholm module (.Jl,Jf,F) gives
rise to a cycle with 0° = .Jl. Observe first that F in fact defines a grading
on L(Jf): for any operator T, we can write

T = i<T + FTF) + i<T- FTF) =: T+ + L.

Also, (TS)+ = T+S+ + LS_ and (TS)_ = T+S- + LS+. Here T = T+ if


T = FTF and T = T+ if T = -FTF; this Z2·grading of operators should
328 8. Noncommutative Differential Calculi

be distinguished from the one coming from J{ = J{ 0 E9 J-{ 1, in the case of


even Fredholm modules.
To eventually define an integral, we need to impose (relatively weak)
summability conditions on the algebra. We assume again that the reader
is familiar with the main properties of Schatten classes, as expounded, for
instance, in [3831 or [4381; a summary treatment is given in Section 7.C, to
which we refer for notation. For all a E .Jl and for some chosen nonnega-
tive integer n, we shall assume [F, a(a) 1 E Ln+ I (J{). Moreover, we shall
suppose that n is odd when dealing with odd Fredholm modules, and that
n is even in the case of even Fredholm modules. Of course, this can always
be arranged by adding 1 to n if necessary, since Ln+l c L»+ 2 ; but in in-
teresting examples, the parity of the minimal summability degree and that
of the Fredholm module do indeed coincide, for reasons that will emerge
later.
For instance, if [F, a(a) 1 is required tobe Hilbert-Schmidt, we obtain a
1-summable Fredholm module -a charged field by any other name. His-
torically, the Shale-Stinespring version of quantum field theory acted as a
springboard for noncommutative geometry.
From now on, we shall suppress the representation a in the notation,
writing [F,a] rather than [F,a(a)], say. The graded differential algebra
structure is introduced by defining, for any operator a in .Jl,

da:= i[F,a] = 2iFa_. (8.10)

There is a certain inevitability in the definition of the operator differential


as a commutator. After all, as mentioned in Section LA, any normal opera-
tor on a Hilbert space is unitarily equivalent to a multiplication operator. If
Mf is the multiplication operator by a function f on the space of square in-
tegrable functions on a manifold, then, as noted in [4281, the multiplication
operator associated to oif is given by [oj,MJ].
The factor of i is introduced in (8.10) in orderthat d commute with the
involution, i.e.,
d (a *) = (da)*,
where the right hand side denotes the operator adjoint to da. Same authors
prefer to define da as [F, a], but this would lead to (da)* = -d(a* ), which
clashes with the more natural involution on ordinary differential forms.
The differential (8.10) in fact selects the F-odd part of a and the sum-
mability condition can be rewritten as da E Ln+l. Consider the vector sub-
space of "1-forms" in L(J{), spanned by operators of the form ao da 1, with
ao, a 1 E .Jl. We can define the second-order differential by the rule

d(ao da!) := i [F, ao da!] = i[F, ao] i[F, ad = dao da1.

The bracket here denotes a supercommutator; for instance, [F, ao da!] is


an anticommutator. Since d(ao da 1) = 2iF(ao da!)+, the differential now
8.2 Cycles and Fredholm modules 329

selects the even part of the 1-forms, belanging to L(n+ll/ 2, whereas the
odd part still belongs to Ln+l. Proceeding in the same vein, we consider
the space of 2-forms, spanned by elements of the type a 0 da 1 da2 with
a 0 , a 1 , a2 E .Jl. Then the differential selects the odd part, which now be-
longs to L(n+ll/ 3 , whereas the even part still belongs in L(n+ll/ 2 ; and so
on. It is natural then to define o.k as the space of operators spanned by
forms aoda1 ... dak with ao, ... ,ak E A. If a E o. 2r, then a+ E L(n+ll/ 2r
and a_ E L(n+l)/( 2r+O, whereas if a E o. 2r-l, then a+ E L(n+ll/ 2r and
a_ E L(n+IJ/(Zr-IJ. The algebra multiplication is just the operator product.
The basic identity d 2 = 0 follows from
[F, [F, T]] = F 2 T + FTF ± FTF- TF 2 = 0,

for T either even or odd.


From the identity
[F,ad· · · [F,ak]F = (-1)kF[F,ad· · · [F,ak],

we check the equivalence of both ways to define differentiation:

d(aoda1 ... dak) := i[F,ao[iF,ad· · · [iF,ak]] = daoda1 ... dak. (8.11)

This yields the simple rule: dw = i [F, w] for any w E o.·.


Exercise 8.6. Verify that if Wk E o.k and wz E 0. 1, then WkW! E o.k+l. o
To summarize, the algebra of the cycle can be considered as the direct
sum of operators of different degrees on n + 1 copies of the Hilbert space
Jf. The usual argument for the ordinary exterior differential shows that d
is an odd derivation.
We adopt the block matrix notation of Section 6.4 for the Fredholm
module, denoting by J{± the ±1-eigenspace of F. (Note, please, that for
even Fredholm modules, this is generally a very different grading from
J{ = J.[O $ J{l ). Then we may write

F= (1 0)
0 -1

whereby
0
a~_)' a - = ( a_+

and
i [F, a] 0
= da = 2i ( -a-+
Now for the integral part. Given an operator T on Jf, we introduce its
conditional trace
Tr' T := TrT+. (8.12)
330 8. Noncommutative Differential Calculi

Note that Tr' T = Tr T if T E .f.I, by cyclicity of the (usual) trace.


Assurnefirst that n is odd. Then (wn)+ E .f.I, so it makes sense to define
the integral by

J Wn := Tr' Wn = -~Tr(Fdwn), (8.13)

and f Wk := 0 for k '* n. Since f dw = - ~ Tr(F d 2 w) 0, closedness


is automatic, and it remains to check the graded trace property. For that,
consider Wk, Wt with k + l = n and, say, k odd and l even. Then

f WkW! = -~Tr(Fd(WkWt)) = -~Tr(FdwkWt-FWkdwt)


= -~ Tr(-dWtFWk + WtFdwk) = -~ Tr(FdWt Wk + Fw1 dwk)
= -~Tr(Fd(W!Wk)) = f W!Wk. (8.14)

as it should be. The commutation under the trace in the third equality is
allowed since kJ(n + 1) + (l + 1)/(n + 1) = 1; here we are using the relation
Tr(AB) = Tr(BA), valid whenever A E .f.P, BE .Lq and 1/p + 1/q = 1.
Assurne now that n = k + l is even and Iet x denote the original grad-
ing operator on :Jf, whose (±1)-eigenspaces are J{ 0 and :Jfi. In this case,
(wn)- E .f.I. We define the integral by

J Wn := Tr'(XWn) = Tr(X(Wn)-) = -~Tr(xFdwn), (8.15)

and f Wk = 0 for k < n. We may also write

J Wn = Tr(x(wn)_) =: Str(wn)-,

where Str is the supertrace (5.57), i.e., Str A := Tr(xA), which -always- in-
volves the original grading X· Closedness is again automatic and f WkW! =
( -1 )kl f WtWk is worked out just as before, because our conventions imply
that the parity of elements of n· with respect to x coincide with the parity
induced by the degree, that is, XWk = (-1)kWkX·
Exercise 8.7. Checkthis graded trace property by reworking (8.14) for the
other parities of k and l. 0

The important matter of groups of transformations of Fredholm mod-


ules [91, IV.4.ß] -also without the restriction of compatibility with a given
orthogonal or symplectic structure- was taken up in [39] .
..,. A distinguished Fredholm module is given by the Hilbert transform.
Definition 8.5. Given a function h on ~. its Hilbert transform Fh is the
Cauchy principal-value integral

Fh(x) := .!:._ limJ h(x- t) dt =: .!:._ pf h(x- t) dt.


1T E10 lti>E t 1T J t
8.2 Cycles and Fredholm modules 331

Exercise 8.8. If Q(t) is a harmonic polynomial on ~n. homogeneaus of


degree k > 0, and if :F denotes the Fourier transformation, show that

p Q(t) - -l(rrn/2[{~) Q(u)) (8.16a)


ltin+k - : f ik[(n;k) iuik

in the tempered distribution sense. In other words, show that

PJ
,f Q(t) A rrn/ 2 [{~)
ltin+k<J>(t)dt = ik[(n;k)
J!Uik<J>(u)du.
Q(u)
(8.16b)

for all <PE S(~n). 0

By taking k = n = 1, Q(t) := t, so that Q(u)/lul = signu, it follows that


(:fF<t>)(u) = (sign u) <f,(u) (8.17)
for <t> E S( ~), and therefore also for <t> E L 2 ( ~). Thus the eigenfunctions
for the Hilbert transform have Fourier transforms that vanish on the nega-
tive or positive halfline; such functions satisfy dispersion relations. Notice
also that this formula shows that the Hilbert transform is in fact a pseudo-
differential operator of order 0.
Hereisa quickerandnot very dirty argument for (8.17). Denote gE(t) :=
1/t for !tl ~ E andgE(t) := 0 otherwise. This is a square summable function
for E > 0, and thus

f
But
BE(u)
oo
= -2i E usinc(tu)dt, where
.
smcx :=---x-;
sinx

since fooo sinc t dt = rr /2, the formula (8.17) follows.


It appears that F defines a Fredholm module. The (commutative) algebra
could be defined as the set of multiplication operators by f E L oo (~) such
that [F, f] E X. What this means is not entirely known, but there is at least
the classical result by Kronecker [379], which says that df is of finite rank
if and only if f is almost everywhere a rational function (with poles outside
the realline).
~ Also, one can pander anew the "Virasoro" example in Section 6.4, to
wit, the space of periodic square-integrable functions on the unit interval
0 :5 fJ :5 1 of zero mean, with F given by
F (e2rrikO) = ( sign k) e2rrikO. (8.18)
(Actually, in Section 6.4, we kept to the usual convention that the angle fJ
run from 0 to 2rr. However, in computations with tori it is handier to nor-
malize the measure on the torus and write 2rrfJ for the angle. We trust the
reader will have no difficulty in sorting out the factors of 2rr .)
332 8. Noncommutative Differential Calculi

The operator (8.18) was shown to be representable by a Hilbert trans-


form:
Fh(e) = PJ: ih(e- 9') cotrrO' d9'.
More precise conditions are known in this case for [F, h] tobe compact, or
p-summable, as indicated in [91].
Exercise 8.9. Verify that Ph = h- f~ h(O) d9 when hisnot of zero mean.
0

Exercise 8.10. Using a fractionallinear transformation from ~ to lr, such as


t ..... ( t- i) I ( t + i), establish a relation between both Fredholm modules. 0
The Hilbert transform is distinguished by its invariance properties: it is
the only symmetry on L 2 (1Rl.) that commutes both with translations and
homotheties. We recall the argument here: any operator Ton L 2 (1Rl.) com-
muting with translations is of the form
J"Th(u) = a(u) :fh(u)
for a E L"'(!Rl.). Then a(i\u) = a(u) for all i\ > 0 forces a(u) oc signu, if
we leave aside the trivial constant solution. It is also clear that F anticom-
mutes with reflections.
Definition 8.6. The Hilbert transform can be generalized to the higher-di-
mensional case in several ways. The most interesting one, from our point
of view, is given by the Riesz operators R1 on L 2 (1Rl.n), defined by

R1h(x)
.
.= ~
2i
hm
Un+l EIO
. ilti>E
t1h(x- t)
It In +l dt.

We recall that 2i/On+l = i[( n;l) ;rr<n+ll/ 2 • It follows from the same dis-
tributional formula (8.16) that

:FR1 h(u) = 1
;
1
:fh(u),

from which it is immediate that the R1 are selfadjoint operators of norm 1


that commute with translations and homotheties. Moreover, 2.}= 1 R] = 1.
Exercise 8.11. Work out the properties of the Riesz operators with respect
to rotations. 0
Therefore, one method to make a Fredholm module out of the Riesz
operators is to find N x N matrices :n .... , Yn. where N depends on the
dimension n, such that

and to define
(8.19)
8.2 Cycles and Fredholm modules 333

on L 2(l~n) ® cN. A Fredholm module can be defined on tori by the same


trick. Such matrices can certainly be found: they are simply the generators
of a representation of the Clifford algebra G(~n). We mayasweH use an
irreducible representation of this complex Clifford algebra, by taking N :=
2ln/ 2 J. For definiteness, we use the following convention: for n = 1, take
Y1 := 1; for all odd n, define inductively

Yj(n) .-
·- (
Yj
0
(n-2) v(n)
"n-1 ·
·= (0 -oi) ,
i
v<n)
.tn
·=
·
(1 0)
0 -1 ·
(8.20)

In particular, for n = 3, we get the Pauli matrices (2.19). For all even n,
d efi ne Yj(n) := Yj(n+ 1) f or J· = 1 , ... , n. The matnx
· Yn+
(n)1, 1mp
· licn
· ly d efi ne d
by this procedure, is simply the grading operator x of the representation
in the even case .
.,.. Now consider matrix functions a in Mm (coo (lfn)). They can be regarded
as multiplication Operators on J{ : = L 2co·n) ® (( N ® cm, say. We may extend
F of (8.18) to J{ in the obvious way, using (8.19).
Exercise 8.12. Prove by a Fourier series calculation that [F, a] E LP for
p > n, and that there are no nonconstant functions a suchthat [F, a] E LP
for some p ~ n. o
The calculus that we have constructed on Fredholm modules has several
analogies to the usual de Rham calculus. Those are more than formal simi-
larities. Exercise 8.12 indicates that it makes sense to think of the elements
a as belonging both to the cycle defined by integration of differential forms
and to the cycle defined by the Fredholm module operator F. Surprisingly,
both integrals coincide! We now state a theorem by Connes and Langmann
in that respect, which lifts the hem of the curtain on one of the basic results
of noncommutative geometry.
Theorem 8.2. If On is the volume of the sphere §n- 1, and

C ·= (2i)ln/2J On (8.21a)
n· n(2rr)n'

then, foranyao, ... ,an E Mm(C 00 (lfn)),

J ao da1 ... dan = Cn J-u-n tr(ao da1 1\ • • • 1\ dan). (8.21b)

where the integral on the left is the abstractintegral (8.13) or (8.15) associ-
ated to the Fredholm module. a
The theorem is proved (actually, for ~n rather than -n-n) by Langmann
[312] by a direct Fourier-led assault. We do not prove it here, as it will
be superseded by a foundational result of noncommutative geometry (see
334 8. Noncommutative Differential Calculi

Chapter 10). However, to check the case n = 1 is something that every


analyst (and theoretical physicist) should do at least once. Here we do it
for the case of the circle.
We can assume, without lass of generality, that m = 1 and that a 0 and
a1 are complex conjugates of each other, so (8.21) boils down to

~ Tr(F[F, ä][F, a]) = l_ ,[ ä da, (8.22)


7T jlr
for a E C"" (lr), where ä and a on the left hand side are to be regarded as
multiplication operators on L 2 (lr). (We extend the operator F of (8.18) to
L 2 (lr) by declaring that F = 1 on the constant functions.) The right hand
side of (8.22) is equal to

where an is the nth Fourier coefficient of a(e) = L~=-oo an e 2 rrine. Also,

([F, a]hh = 2:: (signk- signq) ak-q hq,


qEZ

where we temporarily adopt the convention sign 0 1, to simplify the


calculation. Therefore,

(F[F, ä][F, a]hh = 2:: sign k(sign k- sign p) (sign p- signq)äp-kap-q hq,
q,p

and (8.22) is established by computing

~Tr(F[F,ä][F,a]) = -~ 2::signk(signk-signp) 2 iap-kl 2


k,p
= -~ 2:: signk(signk- sign(n + k)) 2 1an1 2
k,n
= ~ 2:: 2:: 4lan1 2 - ~ 2:: 2:: 4lanl 2 = 2i 2:: nlanl 2 •
n>O -n:Sk<O n<O O:Sk<-n nEZ

~ A similar result for the 2-torus lr 2 was first obtained by Connes [86,
pp. 274-275], by a different method. The Fredholm module used there
was obtained by considering lr 2 as the complex manifold C/ (2rr;l + 2rri;l),
where
F ( 0 (ä + E)- 1 )
= (ä + E) 0 '
with E rt ;r + i;l (ensuring that ä + E is invertible). In order that such an F
be a bounded operator, the graded Hilbert space J-{ is chosen by setting
J-{ 1 := L 2 (lr 2 ) and then J-{ 0 := { ~ E L 2 (lr 2 ) : ä~ E L 2 (lr 2 ) }, a Sobolev
space. 1t is then enough to establish (8.21) for multiplication operators of
8.3 Connections and the Chern homomorphism 335

the form a(cJ>I, cl>2) = exp(2rri(ml cl>1 +m2cl>2)) with m E 71. 2. The left hand
side of (8.21) gives

Jaoda1da2 = -iTr(xF[F,ao][F,ad[F,a2]).

When this is explicitly computed on the vectors of an orthonormal basis


of J-(, a family of Eisenstein series is obtained; these can be summed to
yield (i/2rr) f-a-z ao da11\ da2. Notice that c2 = 2i0 2/8rr 2 = i/2rr. We refer
to the original paper for the details.

8.3 Connections and the Chern homomorphism


In commutative geometry, linear connections send tensor fields (spaces of
sections of certain vector bundles) to 1-form-valued tensor fields: see Defi-
nition 7.2. In noncommutative geometry, we need not change the definition
at all!
Definition 8.7. Let 'E be a right .Jl-module. Consider the right .Jl-module
'E ®5\ 0 15\, where, despite our use of the universal form notation, in this
section 0 1.Jl now denotes a suitable module of 1-forms, like any of those
examined in this chapter. A connection on 'Eisa linear mapping

that satisfies the Leibniz rule, generalizing (7.3):


V'(sa) = (\i's)a+s®da (8.23)

for s E '.E, a E .Jl. There is often a graded algebra 'E ®5\ n· .Jl of "'.E-valued
forms", to which V' extends uniquely as an operator of degree 1 by requiring
\i'(s®w)=V's®w+s®dw, forall sE'E, wEO".Jl.
Hereweare identifying ('.E ®5\ 0 1.Jl) ®...:<\ nn .Jl to 'E ®5\ nn+l.Jl. We still write
V': 'E ®5\ n· .Jl - 'E ®5\ n•+l.Jl for the extension thus defined.
Regarding 'E ®Jl n· .Jl as a right n· .Jl module, we find that
\i'(uw) = (\i'u)w + (-1)kudw (8.24)

for u E 'E ®5\ ok .Jl, w E n· .Jl, as is readily checked -compare (7.5)- by


considering u = s ® 11 with 11 E ok .Jl:
V'((s ® 17)w) = \i's ® 11W + s ® d(17w) = V'(s ® 17) w + (-1)ks ® 11dw .

.".. lt is gratifying that only projective modules admit (universal) connec-


tions [126].
336 8. Noncommutative Differential Calculi

Proposition 8.3. A right module admits a universal connection only if it is


projective.
Proof. Let 0 1 .Jl be the universal.Jl-bimodule of Section 8.1, with differen-
tial da:= 1 ® a + a ® 1. Define right .Jl-module homomorphisms

o- 'E ®5\ 0 1..Jt..i... 'E ®c .Jl ~ 'E- o


by j(s ®da) := s ® a- sa ® 1 and m(s ® a) := sa. This yields a short
exact sequence of right .Jl·modules (think of 'E ®c .Jl as a free .Jl-module
generated by a vector-space basis of 'E). Any linear map \7: 'E - 'E ®Jt 0 1 .Jl
gives a linear section of m by j(s) := s ® 1 + j(\?s). Then j(sa)- f(s)a =
j(\?(sa)- \ls a- s ®da), so f is an .Jl-module homomorphisms precisely
when \7 satisfies the Leibniz rule (8.23). If that happens, f splits the exact
sequence and embeds 'E as a direct summand of the free .Jl-module 'E ®c.Jl,
so 'E is projective. D

Cuntz and Quillen pointout that this theorem can be viewed as a noncom-
mutative analogue of the Narasimhan-Ramanan theorem on (commutative)
universal connections [354].
One should realize that the universal1·form bimodule is not projective in
general. Cuntz and Quillen called quasifree, or formally smooth, an algebra
suchthat this bimodule is projective. A criterion for formal smoothness has
been found by Kontsevich and Rosenberg [289].
When the algebra .Jl is commutative, we can form the tensor product
'E ®5\ :J of two .Jl-modules 'E and :f. If these carry respective connections
vx and \7 :r, the tensor product connection on 'E ®5l :f is given by
v'E®~:r := v'E ® 1:r + 1x ® v:r, (8.2 5)
since the Leibniz rule (8.23) is easily checked .
.. Let us look at examples. The simplest one is d itself on the free module
.Jln. We identify .Jln ®Jt 0 1.Jl to (0 1.Jl)n and write

for this connection. If V' is another connection on .Jl n, then V' - d is an


.Jl-linear map from .Jl n to (0 1.Jl)n, i.e., we can write
V'= d + oc,
where oc is an n x n matrix with entries in 0 1.Jl. If { u J} is the standard
basis in .J\n, then du1 = 0, so V'u 1 = Ir=l UiOCiJ· If b = [biJ] is the matrix
of a change of basis u1 =Ir= I uibiJ, the new 1-form matrix is given by
(8.26)
8.3 Connections and the Chern homomorphism 337

Exercise 8.13. Check that the formula 'V s = ds + as still applies for s E
(0" . .:<\}n. 0

On a finitely generated projective module 'E with associated idempo-


tent e, the connection given by ed plays an important role.
Definition 8.8. If 'E = e.JI. n is a finitely generated projective .JI.-module, let
i: 'E .... .JI. n denote the inclusion, and let 'V be the composition

'E J... .JI. n ~ .JI. n ® .Jl 0 1.JI. .!!_ 'E ® .Jl 0 1.JI..

Regarding 'E as a submodule of .JI. n, we simply write 'V s = e ds. The connec-
tion thus defined is called the Levi-Civita connection or the "Grassmannian
connection" on 'E.
This terminology incorporates the usual Levi-Civita connection on the
tangent bundle of a Riemannian manifold M, given in Definition 7.4, which
can be described algebraically as follows. The manifold can be smoothly em-
bedded in IR(N for a suitably large N. Let f: M .... IR(N be such an embedding.
Any tangent vector at j(x) is of the form ak(x)fk(x), where fk := of /oxk,
so we may identify TxM with span{fl (x), ... .fn (x)} c IR(N; the vector field
X = ak Ok is identified with ak fk E C"" (M, IR(N). The standard scalar pro-
duct ( · I ·) on IR(N induces a metric on M by ßij = g(Oi, Oj) := (ji I fj ). In
fact, the embedding f can always be chosen so that this restricted met-
ric coincides with an intrinsically given metric on M -this is the Nash
embedding theorem [450, Thm 14.5.1]. If p(x) now denotes the orthogo-
nal projector on !Rtn with range TxM. then pfk = fk for each k, so that
the space 'X(M, !Rt) of smooth real vector fields on M is identified with
{ u E C"" (M, IR(N) : pu = u }. In short, the embedding f determines the
projector-valued function p (and vice versa, provided p is orthogonal).
Since we work with complex vector fields, we like to complexify this pic-
ture. One can regard f as embedding M in f[N, so that the vectors fk(x)
span the complexified tangent space T~ M and take ( · I ·) to be the usual
Hermitian scalar product on cN; the relation 9ij = (ji I fj) is unaffected
since each fk (x) is real. We redefine p (x) as the orthogonal projector on f[N
with range T~M and identify the (complex) smooth vector fields 'X(M) with
{ u E C"" (M, f[N) : pu = u }; it is a module over the algebra C"" (M).
Recall from Section 7.1 that the torsion of a connection 'V on 'X(M) is
'V() E .JI. 2 (M, TM), where () is the fundamental 1-form in .J\. 1 (M, TM), de-
termined uniquely by the requirement that Lxe = X for each X E 'X(M).
The above identification says that () = df. For 'V = pd the torsion van-
ishes, because 'V() = p d(dj) = 0. Metric compatibility of 'V = pd follows
easily:
('Voi I oj) + (oi I 'Voj) = (pdfi I h> + (fi I pdh)
= (dfi I pfj) + (Pfi I dfj) = (dfi I h> + (fi I dh>
= d(fi I fj) = dgij.
338 8. Noncommutative Differential Calculi

using the orthogonality of p and the relations p fk = fk· Thus the Levi-Civita
connection induced by the embedding is precisely \7 = pd.
Definition 8.9. When working in the pre-C*-module framework, we shall
consider only Hermitian connections, i.e., those that are compatible with
the inner product structure:
(\7r I s) + (r I \7s) = d(r I s), for all r,s E 'E. (8.27)
The summands on the left hand side are obtained by extending the Her-
mitian pairing on 'E, in the obvious way, to sesquilinear pairings of 'E with
'E ®5\ 0 1.Jt, taking values in 0 1.Jt; e.g., (r I s ® adb) := (r I s)adb.
In the case of the Levi-Civita connection, this is equivalent to selfadjoint-
ness of the idempotent p. If \71, \72 are two Hermitian connections on 'E,
then \7 1 - \7 2 =: ß E Hom5l ('E, 'E ®5l n 1.Jt) is skewadjoint in the sense that
(ßr I s) + (r I ßs) = 0.
Exercise 8.14. If 'Eisa finitely generated projective module over .Jt, show
that there is a bijective linear map between Hom5l ('E, 'E ®5\ 0 1.Jt) and 'E ®5\
0 1.Jt ®.:;t 'E, where 'Eis the left .Jt-module conjugate to 'E. 0
If we identify 'E "" p.Jt m, where p is a projector in Mm (.Jt), we can then
write p.Jtm ®5\ 0 1.Jt ®5\ m.Jtp = pMm(0 1.Jt)p. The involution on .Jt Ieads
to an involution on 0 1 .Jt by setting
(adb)* := d(b*) a* = d(b*a*)- b* da*. (8.28)
Thus a skewadjoint element oc of Hom5l ('E, 'E ®5\ 0 1 .Jt) is the same thing
as a matrix of 1-forms oc E Mm(0 1.Jt) suchthat
oc = poc = ocp = pocp and oc* = -oc. (8.29)
Therefore, any Hermitian connection is of the form \7 := pd + oc where oc
satisfies (8.29).
We remark that the involution as defined in (8.28) is consistent with our
convention that the Fredholm-module differential be da:= i [F, a] so that
(da)* = i [F, a*] = d(a* ), using the operator adjoint.
Definition 8.10. Consider the map \7 2 : 'E ®5l n· .Jt - 'E ®5\0-+ 2 .Jt. Apriori,
it is only complex linear; but a calculation reveals it to be a module map:
\7 2 (sw) = \7(\7s w+s dw) = (\7 2s) w- \7s dw+ \7s dw+s d 2 w = (\7 2s) w.

Therefore \7 2 is a homomorphism of .Jt-modules, entirely determined by


its restriction to 'E, called the curvature K" of the connection.
lt is easy to compute the curvature of a connection d + oc on a free module:
K"s = \7(ds + ocs) + d(ocs) + ocds + oc 2 s
= d 2s

= docs- ocds + ocds + oc 2 s = (doc + oc 2 )s,


as thoroughly expected from ordinary geometry.
8.3 Connections and the Chern homomorphism 339

Lemma 8.4. lf'E 0 and 'E 1 are projective modules over a commutative alge-
bra 5'\. with respective connections V' o, V' 1 and curvatures Ko, K1, there is
an associated connection V' on Hom5'l ('E 0 , 'E 1 ) given by

(V'T) U := V'I(TU)- T(V'ou), (8.30)

whose curvature is

(8.31)

Proof. On rewriting (8.30) as \?1 (Tu)= (V'T) u + T(V'ou), the Leibniz rule
for V' is easily checked. Replacing T by V'T E Hom5'l ('E 0 , 'E 1 ) ®5l 0 15'\., the
Leibniz rule in degree one yields V' I( (V'T)u) = (V' 2 T)u- V'T(V' 0 u), so that

(V' 2 T) U = V'I((V'T)U) + V'T(V'oU)


= V'I(V'I(Tu)- T(V'ou)) + V'I(T(V'ou))- T(V'5u)
= (K1 T - T Ko) u. 0

In particular, any connection V' on 'E over a commutative algebra 5'\. in-
duces a connection on End5'l 'E, also denoted by V', given by V'T := V' o T-
T o V'. The Bianchi identity then is

(8.32)

since K V E End5'l 'E ®5'l n2 5'\..


~ Let M be a smooth compact manifold, without boundary. The classical
Chern character eh isaring homomorphism from K0 (M) to the even part
of the rational Cech cohomology of M. The kernel of eh is a torsion group,
i.e., a group whose elements all have finite order; thus eh is an isomorphism
whenseenasamapfromK 0 (M)®zQ toHeven(M, Q) -andfromK 0 (M)®zC
to Heven (M, ((). The original formulation was in terms of curvature of gauge
fields on the bundles. Herewe plunge directly into an analogous treatment
with a noncommutative flavour. lt follows the lines of Fedosov's algebraic
treatment of the Chern characters of vector bundles [172].
Recapitulating: in the smooth, as in the continuous, category, vector bun-
dles over M are described as finitely generated projective modules, over the
algebra C"" (M) of smooth functions on M. Recall that the module of smooth
sections of a vector bundle E- M is denoted by f"" (M, E) or by f"" (E), for
short. Any vector bundle E can be embedded as a direct summand of a
trivial vector bundle of rank N, say; the results will be independent of the
choice of embedding. As in Chapter 7, we denote the Riemannian curva-
ture by R, rather than K vn. Acting on a smooth section s E f"" (E) suchthat
ps = s, R is given simply by

Rs := (V'B) 2s = (pd)(pd)s = pdpds. (8.33)


340 8. Noncommutative Differential Calculi

The procedure serves to detect nontrivial elements of Ko(C(M)). Before


defining the classical Chern character, we learn to compute using the for-
mula (8.33). Differentiating p 2 = p, we obtain p dp + dp p = dp, so that

p dp = dp (1- p) and dp p = (1- p) dp, thus p dp p = 0. (8.34)

If s = ps, then ds = dp s + p ds, or dp s = (1- p) ds. We arrive at

Rs = p d p ds = d p ( 1 - p) ds = d p d p s = d p d p p s.
Therefore, we can write R = dp dp p = dp ( 1 - p) dp = p dp dp.
Exercise 8.15. On identifying 'E ""' p.J\. m, every Hermitian connection on 'E
is given by
V' ( p s) = p ds + cx s,
with cx E pMm (0 1.Jt) p skewadjoint. Prove that

Kv = pdpdp + pdcxp + cx 2 • <>

Definition 8.11. The Chern character is defined on projectors p E Mm (.Jl.)


as
1
I I
00 00

chp := tr(expR) = ch2k p := k' tr p(dp) 2 k. (8.35)


k=O k=O '

For the algebra .J\. = coo (M), we can regard each dp as a matrix of 1-forms,
i.e., dp E Mm(.J\. 1 (M)), so that ch2k p E .J\. 2k(M) for each k, and the sum
is actually finite.
Proposition 8.5. When .J\. = coo (M), the Chern character defines de Rham
cohomology classes.

Proof. We need to checkthat ch2k p is closed for all k. We compute

d(trp (dp) 2k) = tr(dp) 2k+I = tr(p (dp) 2k+I) + tr((1- p) (dp) 2k+ 1 ).
(8.36)

Both terms on the right vanish; for example,

tr(p (dp)2k+I) = tr(p2 (dp)2k+I) = tr(p (dp)2k+I p)


= tr(p(1- p) (dp) 2k+ 1 ) = 0,
where the trick (8.34) of moving p across each dp is repeatedly used. D

A superficially different proof consists of remarking that, as we can write


V' = d + cx locally (the Leibniz rule shows that a connection is determined by
its restrictions to chart domains), then by (8.30), V'T = dT + [cx, T] locally
8.3 Connections and the Chern homomorphism 341

for T E End5\ CE). The trace operator tr: End5\ (1") ®5\ J'l" (M) - J'l" (M)
vanishes on commutators, giving (globally) the useful identity

tr(V'T) = d(trT) E J'l"(M).

Then d(tr Rk) = tr(V' Rk) = k tr(Rk- 1V' R) = 0, by the Bianchi identity (8.32).
Many authors, e.g. [197], define the Chern character as tr(exp(iR/2rr))
rather than tr(expR) in (8.35). This has the advantage that the associated
cohomology classes are integral, i.e., lie in Heven(M, 7L). However, since we
only need to deal with de Rham cohomology -with complex coefficients-
we may use the simpler normalization (8.35), provided we remernher to
keep track of the various 2rr factors when integrating ordinary differential
forms. This normalization fits better with the algebraic approach [403] to
the Chern character.
lt is known that the curvature of the Levi-Civita connection measures the
noncommutativity of parallel transport in two dimensions. In [451], Tele-
man explains how chk measures the noncommutativity of parallel transport
in 2k dimensions.
Proposition 8.6. The Chern character classes [chzk p] depend only on the
K-theory class [p] E Ko(J'l).
Proof. lt is enough to prove that each p - ch2k p is a homotopy invariant.
Let { Pt : 0 ::5 t ::5 1 } be a smooth one-parameter family of projectors, and
write Pt:= dpt!dt. We need to prove that

:t tr(pr (dpt ) 2k) = tr(fJt (dpt ) 2k) + tr(pr :t (dpt ) 2k)

is an exact form. The first summand vanishes, because, just as in (8.34),


Pt Pt = pt(1- pr) and PtPtPt = 0. Then, like (8.36),

tr(fJr (dpr) 2k) = tr(PtPt (dpt) 2 k) + tr((l- pdfJt (dpt) 2 k),

and both terms on the right are zero; for instance,


tr(PtPt (dpr) 2k) = tr(PtPt (dpd 2kpr) = tr(PtPtPt (dpr) 2k) = 0.
On the other hand,

d (dpt) 2k) 2~1 (


tr ( Pt dt = f;;o d (dpr) (dpt) 2k - J-
tr Pt (dpr )1. dt . 1)

= . L tr ( (dpr)zk-1 Pt :t (dpt)) + .L tr( (dpt)zk-1 ( 1 - pt) :t (dpt))


J even JOdd

i
= k tr( (dpr)zk-1 :t (dpt)) = :t tr(dpt)zk,

and the last expression is an exact 2k-form. D


342 8. Noncommutative Differential Calculi

Proposition 8.7. The Chern character eh is a ring homomorphism from


Ko(C""(M)) = K0 (M) toHd~en(M).

Proof. Recall, from Definition 3.12, that K0 (M) is a commutative ring; the
ring Hd~en (M) of even-degree de Rham cohomology classes (with the mul-
tiplication induced by the wedge product of forms) is also commutative.
The homomorphism properties may be stated as

ch(E E9 F) =ehE+ chF, ch(E ® F) = (chE)(chF),

for (Hermitian) vector bundlesE and F over M; here ehE means chp when-
ever f"" (E) "" p5l m, of course. The first equality is immediate from the def-
inition, since ch2k (p E9 q) = ( 1 I k!) tr(p (dp ) 2k E9 q (dq) 2 k) = ch2k p + ch2k q.
For the second, after writing f"" (F) "" q5l n, we use the representation
[""(E ® F) "" (p ® q)5lmn, discussed in Section 2.6. We also note some-
thing that shall be quite useful later on: given any two connections 'V on
f""(E) and 'V' on f""(F), the operator 'V® 1 + 1 ®'V' is a connection on
[""(E ® F) whose curvature is given by K-v ® 1 + 1 ® K-v·. Thus ch(p ® q) =
tr(exp(pdpdp ® 1 + 1 ® qdqdq)) = (chp)(chq). 0

~ In order to see what is going on, we apply the map eh to our old friend,
the Bott projector on the sphere § 2 = { (x, y, z) : x 2 + y 2 + z 2 = 1 } :

1 ( 1+ Z X- iy) d = _!. ( dz dx- i dy)


PB =Z X + iy 1- Z ' PB 2 dx+idy -dz ·

Then PB dpB dpB = !A, where Ais a matrix with diagonal elements

an= i(1 + z) dxdy- (x- iy) dz (dx + idy),


a22 = - i ( l - z) dx dy + (x + iy) dz (dx- i dy), (8.37)

and therefore

ch(pB) = 1 + i(xdydz- ydxdz + zdxdy) = 1 + ivol(§ 2),

where vol denotes the volume form on the sphere; the nontrivial part of eh
detects the nontriviality of the line bundle.
Exercise 8.16. More generally, for § 2m = {(XI, ... , X2m+d : xf + · · · +
x~m+l = 1}, consider the projector p := !O
+ Y1X1 + · · · + Y2m+IX2m+d·
Prove that
im (2m)! 1 2m)
ch p = 1 + zm+l m! vo (§ . 0

lt is well known [134] that the only nontrivial element of Hd~en(§2m)


is vol( § 2m). Thus, Exercise 8.16 is all we need to tackle some unfinished
business in Chapter 3, strongly reinforcing Proposition 8.7. The following
result, despite the name, is due to Atiyah and Hirzebruch.
8.4 Hochschild homology and cohomology 343

Corollary 8.8 (Chern isomorphism theorem). The Chem character eh is a


ring isomorphism (rom Ko(C"" (M)) ®z Q = K0 (M) ®z Q onto Heven(M, Q)
and from K0 (M) ®z (( onto Hj~en(M).

Proof. First, H•(M,?L) ®z Q "" H•(M,Q) for allfinite CW-eomplexes. In


effeet, let cp be the natural homomorphism H• (M, ?L) ®z Q - H• (M, Q). It
is clear that c/J(M) is an isomorphism when M is a sphere. Then, aeeording
to Lemma 3.30, it is an isomorphism for all finite CW-eomplexes. Now,
putting tagether Proposition 8.7, Exereise 8.16 and Lemma 3.30 again, we
obtain the eonclusion for the ease of finite CW-eomplexes.
The general result follows from eontinuity of both funetors; or one may
invoke Morse theory, to the effeet that every eompaet manifold has the ho-
motopy type of a finite CW-eomplex. Indeed, if f isaMorse funetion on M,
its eritieal points are isolated, by the Morse lemma, henee finite in number.
Let PI, ... , Pm be the eritieal points of index 0. By the main theorems of
Morse theory [41], M is eonstrueted, up to homotopy, from PI, ... , Pm by
attaehing eells of dimension n for eaeh eritieal point of index n. D

Exercise 8.17. Prove that KI(M) ®z Q ""Hoctd(M,Q). 0

There is, then, a hexagon exaet sequenee like (3.37), but with K 0 , KI
replaeed by Heven, Hodd. The tensoring with Q is needed sinee the K -groups
of a compact space may have torsion, so that eh: K• (M) - H" (M, Q) may
not always be injeetive; it is, however, injeetive if M isafinite CW-eomplex
[447, p. 174].
The moral of the story is that, in the commutative ease, K -theory recov-
ers the essentials of ordinary cohomology via the Chern eharaeter. In the
noncommutative ease, we need a replaeement for the de Rham complex.

8.4 Hochschild homology and cohomology


In this section and the next, we shall deal with unital algebras only.
We anticipate that the map that plays the role of the Chern eharacter
will be a homomorphism from the Ko of an algebra into a suitable homol-
ogy dass of the same algebra. In preparation for that, we first introduce
the simplest homology theory for associative algebras, namely Hochschild
homology, and the corresponding cohomology.
An aneestor of the Chern character is the trace map in algebraie K -theory.
If e E Qn (.Jt) is an idempotent, let tr e := L.r=I ekk; this gives a well-defined
map from Qoo (.Jt) to .Jt that is additive under bloek direct sums. If v E
GLoo(.Jt), write w = ev-I; then

trvev-I- tre =I VJkWkJ- I WkJVJk =I [Vjk. Wkj],


j,k j,k j,k
344 8. Noncommutative Differential Calculi

which belongs to the commutator ideal [5\, 5\]. Thus, tr drops to a well-
defined homomorphism from Kg1g(.A.) to .A./[5\,5\]. The latter turnsout
tobe the lowest-level modulein the Hochschild homology of 5\.
On account of the contravariant nature of the Gelfand cofunctor, we
should expect that the cohomology of spaces becomes homology of al-
gebras, and the homology of spaces becomes cohomology of algebras; and
that is indeed what happens. The obvious clash between the degree-raising
coboundary homomorphisms of a cohomology theory and the degree-low-
ering boundary homomorphisms of a homology theory is resolved by find-
ing another natural map in Hochschild cohomology that lowers degree and
matches the de Rham boundary map: see Proposition 8.18.
Later on, in Chapter 10, we construct the noncommutative Chern charac-
ter taking values in the (periodic) cyclic cohomology of the algebra, a variant
of Hochschild cohomology in which both the Hochschild coboundary map
and the degree-lowering map play complementary roles. The domain of this
Chern character, dually to K-theory, will consist of Fredholm modules .
.,.. Before starting the formal proceedings, we recall a few concepts from
homological algebra. We write (c., d) for an arbitrary chain complex whose
homology modules are denoted H n ( C).
Definition 8.12. A chain map f: (C., d) - (C;, d') is a sequence of maps
fn: Cn - C~ making the following diagrams commute:

l
Cn-Cn-1

fn lfn-1
'
Cn d' C'
- n-1

Thus f takes cycles to cycles and boundaries to boundaries, and induces


homomorphisms Hnf: Hn(C)- Hn(C') for each n.
Definition 8.13. A chain homotopy between two chain maps f, g : (C., d) -
(c;, d') is a sequence of morphisms hn: Cn - C~+ 1 satisfying the relations
d'hn + hn-1d = fn- Bn.

usually abbreviated to d'h + hd = f - g. Clearly Hnf = Hng whenever


f and g are chain homotopic, since j(x) = g(x) + d' (h(x)) whenever
dx = 0. If the identity map on C. is chain-homotopic to the zero map, the
chain complex (C., d) is called contractible, and the graded map h satisfying
dh + hd = 1 is called a "contracting homotopy".
A chain complex (C.,d) is called acyclic if Hn(C) = 0 for n > 0. Any
contractible complex is acyclic. Some authors use the word "acyclic" to
mean that Ho (C) = 0 also; we shall not.
Similar definitions apply for cochain complexes.
8.4 Hochschild homology and cohomology 345

Consider now the chain complex of algebras (C.(~). b), where Cn(~) :=
~®(n+ll,with the boundary map b defined on Cn(~) by
n-1
b(ao ® a1 ®···®an):= 2:: (-1)lao ® · · · ® aJaJ+1 ®···®an
j:O

+ (-l)nanao ® a1 ® · · · ® an-1· (8.38)


Also, b = 0 on Co(~) = ~- For example, b(ao ® a!) := aoa1 - a1ao, while
b(ao ® a1 ® az) := aoa1 ® az- ao ® a1a2 + azao ® a1.
1t is easy to checkthat b 2 = 0, by cancellation of terms with opposite signs.
Definition 8.14. The Hochschild homology of ~ is the homology of the
complex (C.(~).b).lt is denoted by H.(~.~) or, more simply, HH.(~).
Lemma 8.9. HHo(O = (, and HHn (() = 0 for n > 0.
Proof. For ~ = (, we get Cn(() = (®n+l == (, where ao ® a1 ® · · · ®an =
aoa1 ... an is the ordinary product. The formula (8.38) reduces to b(1) =
L.}:o (-1 )J = 0 or 1 according as n is even or odd, so the Hochschild chain
complex is just

which has trivial homology except at n = 0. D


There is a more general definition of Hochschild homology, obtained
by replacing the first copy of ~ in ~ ®(n+ll by any ~-bimodule 'E. Put
Cn(~. 'E) := 'E®~®n. Notice that the products anao and aoa 1 that appear
in the definition (8.38) of b(ao ® a1 ® • · ·®an) make sense when ao E 'E.
The homology of this complex is then denoted H. (~. 'E). We use the stan-
dard notations z. (.JI., 'E) for the Hochschild cycles, and B. (.JI., 'E) for the
Hochschild boundaries, so that H. (~. 'E) = Z. (~. 'E) I B. (~. 'E).
Any homomorphism of algebras f: ~ - .JI.' (which need not be unital)
induces a chain map in the obvious way, and then a degree zero homomor-
phism HH.f: HH.(.JI.)- HH.(.JI.'). Thus each HHn is a functor from the
category of (complex) algebras to the category of (complex) vector spaces.
Exercise 8.18. Formulate and prove functoriality of Hochschild homology
in the bimodule case. <>
Exercise 8.19. Prove that HHn(~ EB ~') == HHn(.JI.) EB HHn(~'). <>

The chains of the form a = ao ® · · · ® 1 ® · · · ®an, with ak = 1 for some


k > 0, generate a subcomplex D.~ since a E Dn~ entails ba E Dn- 1 ~.
Introduce the family b' of boundary maps given by
n-1
b'(ao ® a1 ®···®an):= 2:: (-1)lao ® · · · ® aJaJ+l ®···®an, (8.39)
j:O
346 8. Noncommutative Differential Calculi

that is, by dropping the last term in the sum (8.38). When b' is combined
with the degree-one map s: ao ® · · · ® an ..... 1 ® ao ® · · · ® an, the result is

b' s(ao ® ... ®an)


n-1
= ao ®···®an+ L' (-1) 1. + 1 1 ® ao ® · · · ® ajaj+I ®···®an
j;Q

= (1- sb')(ao ®···®an). (8.40)

We conclude in particular that (bs + sb)a = a for a E DnJ\ with an = 1.


By composing s, if necessary, with a cyclic permutation of the factors, we
obtain a chain homotopy (between the identity and the zero map) s' satis-
fying s'ba + bs'a = a for all a E DnJ\, n > 0. Therefore this subcomplex
is acyclic, i.e., Hn(D.Jl, b) = 0 for n > 0. We can form a reduced complex,
the bimodule C.Jl/ D.Jl, which is none other than our old friend n· Jl.
The boundary operator on the quotient n· Jl, still called b, is written as

(8.41)
n-l
+ L (-1)iaodal ... d(ajaj+I) ... dan + (-l)nanaodal ... dan-1·
j;l

Exercise 8.20. Show that (8.41) can be rewritten as b(w da) = ( -1 )k[w, a]
for w E ok 5\, a E Jl, and check again the boundary property of b using
this formula. o
We refer to [53) for further treatment of the Hochschild complex along
this line.
~ We look at the lower-degree homology spaces. First of all, Ho(Jl, '.E) =
'.E 1['.E, Jl], with an obvious notation. If .5\ is commutative and '.Eis a sym-
metric bimodule, then H 0 (Jl, '.E) = '.E. In particular, HHo(Jl) = J\j[J\,.5\];
and HHo(Jl) = .5\ if .5\ is commutative.
Suppose that .5\ is commutative and '.E is a symmetric bimodule. Then
H1 (Jl, '.E) = '.E ®.Jt O!bJt. In particular, HH1 (Jl) = O!bJt when .5\ is com-
mutative! This is easily seen [319): by definition, H 1 (Jl, '.E) is the quotient
of '.E ® .5\ by the relation sa ® b + sb ® a = s ®ab. Therefore the map
cf>: [s ® a] ..... s 4a is well defined, and its inverse is sa 4b ..... [sa ® b].
Exercise 8.21. Check carefully the homomorphism properties of cf>. 0
More generally, there is the following result about the full Hochschild
homology of modules over (complex, unital) commutative algebras.
Proposition 8.10. For any commutative unital algebra .5\ and any symmet-
ric .5\-bimodule '.E, there is a natural map from '.E ® .Jl Q~bJl to Hn (Jl, '.E). In
particular, there isanatural map from O~bJl to HHn (Jl).
8.4 Hochschild homology and cohomology 347

Proof. This canonical map comes from the skewsymmetrization operator


An: 1' ®An .A. - Cn (.A., 1') given by

On applying the Hochschild boundary (8.38) to the right hand side, we ob-
tain in!
terms of the form

( -1) rr ( -l)l s ® · · · ® (arr(j)arr(j+l) - arr(j)arr(j+l)) ® · · · ® arr(n)

for each j = 1, ... , n- 1, which cancel since .A. is commutative; the remain-
ing terms make up the sum

L (-1)rr(sarr(l) ® · · · ® arr(n) + (-l)narr(n)S ® arr(l) ® · · · ® arr(n-1))

= L (-1 )u (sau(l) - au(l)S) ® au(2) ® · · · ® au(n),


O"ESn

which cancels since the bimodule 1' is symmetric. Notice that ( -1) n is the
sign of the cyclic permutation (XI, x2, ... • xn) - (Xn,XI, ... , Xn-d· There-
fore, the image of the map An consists of Hochschild cycles. When n = 1,
A1 is the identity on 1' ® .A.; notice that

(8.43)

so that A1: 1' ®.Jl. - Z1 (.A., 1') drops to the isomorphism (already described)
1' ®5t n!b.A. == H 1 (.Jl., 1') after factoring out terms of the form (8.43) on
each side. In general, n~b.A. = A10!b.A., so there is a natural .A.-module
morphism from 1' ®An .A. to 1' ®5\ n~b.A., and it turnsout that An maps its
kerne! into Hochschild boundaries. That can be verified by a tedious check;
for instance,

A2(sao ® a1 ® a2 + sa1 ® ao ® a2- s ® aoa1 ® a2)


= L (-l)u(-l)u(O) b(s ® au(O) ® au(l) ® au(2)).
uES3

Therefore, An: 1' ®An .A. - Zn (.A., 1') drops to an .Jl.-module map from
1' ®5t O~b.A. to Hn (.A., 1'). D
Proposition 8.11. If 1' is a symmetric bimodule over a commutative unital
algebra .A., then 1' ®5\ n~b.A. is a direct summand of Hn (.A., 1').
Proof. Let rr: C. (.A., 1') - 1' ®5t n;b.A. be the .Jl.-module map of degree zero
defined by

(8.44)
348 8. Noncommutative Differential Calculi

Then rr(b(s ® a1 ® · · ·®an)) equals

n-1
sauia2 ... 4an + L (-l)is4a1···4(aJaJ+1) ... 4an
j=1

( -l)n(ans 4a1 ... 4an-1 - s 4a1 ... 4an-1 an)


= (-l)n(ans-san)4a1···4an-1 =0,

since 'E is symmetric and the identity a 4b = 4b a allows us to pass the


rightmost an to the left. We conclude that rr ob = 0, and so the restriction
of rr to z. (.Jl, 'E) factors through Hn (.Jl, 'E) and yields a module map TTn
from H.(.Jl, 'E) to 'E ®51. n;b.Jl.
Consider now the map TTn 0 An from 'E ®51. n~b.Jl to itself. It is clear that

TTn (An (s 4a1 ... 4an)) = ~1 L (-l)rr rr(s ® arr(l) ® • • · ® arr(n))


. TTESn

= ~1 L (-l)rrs4arr(l) ... 4arr(n) = s4a1···4an,


• TTESn

since n~b.Jl is the exterior product, over .Jl, of n copies of n~b.Jl. Thus An
is an injective module map, with left inverse TTn. 0

Inparticular, n;b.Jl is a direct summand of HH. (.Jl) when .Jl is commuta-


tive; in fact, there is a multiplication onHH. (.Jl), the "shuffle product" [319,
§4.2], making it a differential algebra, suchthat A.: n;b.Jl - HH. (.Jl) is a
morphism of graded differential algebras. We shall not go into that.
Example 8.1. In the case .Jl = C"' (M), we can replace Cn (.Jl) by the .Jl-
module C"' (Mn); this corresponds to passing to continuous Hochschild
homology, as will be explained in the next section. The Hochschild bound-
ary b: C"' (Mn+ 1 ) - C"' (Mn) is given by
n-1
(bF)(xo, ... ,Xn-1) = L (-l)iF(xo, ... ,Xj,Xj, ... ,Xn-Il
j=O
(8.45)

To see that, it suffices to consider the case where F is a simple tensor


F = ao ® · · · ®an, since thesegenerate a dense subalgebra of C"' (Mn+1 ).
In that case, (8.45) is just obtained by evaluating both sides of (8.38) at an
arbitrary point (xo, ... , Xn-Il of Mn.
Example 8.2. Wodzicki's discovery that the noncommutative residue is a
unique trace on the algebra 'P(M) of classical symbols (Theorem 7.6) can
8.4 Hochschild homology and cohomology 349

be understood as an assertion about the Oth Hochschild homology module


of that algebra. To the sequence (7.33) there corresponds a short exact
sequence in Hochschild homology of differential and pseudodifferential
operators, studied in [492,493). (The algebra of smoothing operators on a
compact manifold has the same Hochschild homology as C) The proof of
Theorem 7.6 actually shows a little more: when n > 1,

HHo('P(M)) ""'P(M)/['P(M), 'P(M)] ""H 2 n(T* M \ M) = C,

where T* M \ M is the cotangent bundle with the zero section removed.


In the case n = 1, T* M \ M is disconnected, there are two such residues,
and HHo('P(§l)) "" C2. More generally, Wodzicki was able to prove that
HHk('P(M)) ""H 2n-k(T*M\M).

~ We pass to cohomology by introducing the dual complex to (c. (.JI.), b ).

Definition 8.15. A Hochschild n -cochain on .JI. is an (n + 1) -linear functional


an .JI.. This is the same thing as alinear form on .JI. ®(n+l), or an n-linear form
an .JI. with values in the (algebraic) dual space .JI. *. We mention that .JI. * is
an .JI.-bimodule, where for cp E .JI.* we put (a'cpa")(a 1) := cp(a"a 1 a').
The coboundary operator, also called b, is the transpose of the boundary
operator of homology:

n
bcp(ao, ... ,an+l) := L(-1)lcp(ao, ... ,aJaJ+l•· .. ,an+l>
j=O
(8.46)

The cohomology of this complex is the Hochschild cohomology of .JI., de-


noted by HH• (.JI.).

In particular, a Hochschild 0-cocycle T on the algebra .JI. is a trace, since


TE .JI.* = Hom(.JI.,C) and T(aoad- T(aiao) = bT(ao,ai) = 0.
When the algebra .JI. is 1:2-graded, one can define a graded version of
Hochschild cohomology [280): it is enough to modify the coboundary ope-
rator b by introducing a "Koszul sign" in the formula whenever two argu-
ments are permuted. (The same remark applies to the cyclic permuter A of
(8.48) below, and the other cochain operators that appear in Section 10.1.)
The graded Hochschild (and cyclic) cohomologies of the exterior algebra
N!Rl.n have been computed by Coquereaux and Ragoucy [121).
The Hochschild cohomology of .JI. is often written as H• (.JI., .JI. * ), as it is
clear from the formula (8.46) that in general one can define the Hochschild
cohomology of .JI. with values in an A-bimodule 'E. Let cn (.JI., 'E) denote the
vector space of n-linear maps cp: .JI.n- 'E, regarded as a bimodule under
350 8. Noncommutative Differential Calculi

(a' <Pa") (a1, ... , an) := a' QJ(al, ... , an) a". The coboundary map is
(b<P) (al, ... , an+l) := a1 QJ(az, ... , an+l)
n
+ 2.:(-l)lQJ(al, ... ,aJaJ+l•····an+d
j=l

+ (-l)n+l<P(al, ... ,an)an+l· (8.47)


Exercise 8.22. Verify that H 0 (.A, T) = { s E T : as = sa for all a E .Jt }.
Verify that H 1(.Jt, T) = Der(.Jt, T) I Der' (.Jt, T), i.e., H 1(.Jt, T) is the space
of "outer derivations". 0
The cohomology space H 2 (.Jt, T) classifies the extensions of .Jt by T. We
shall not go into that (but see the discussion at the beginning of Section 8.5).
The interested reader may consult [482].
Definition 8.16. To extend the trace property of 0-cocycles to higher or·
ders, we say that an n·cochain <P on .Jt is cyclic if i\QJ = <P. where
(8.48)
The ( -1) n is again the sign of the cyclic permutation. A cyclic cocycle is
a cyclic cochain <P for which b<P = 0.
For instance, a cyclic 1·cocycle satisfies QJ(ao,al) = -QJ(al,ao) and
QJ(aoal,az)- QJ(ao,alaz) + <P(azao,al) = 0.
Clearly, a cyclic 1-coboundary is a linear function of the commutator:

... There is an important relation between cycles over an algebra and cyclic
cocycles on that algebra.
Definition 8.17. Suppose we are given a n-dimensional cycle over .Jt. The
Chern character ofthat cycle is defined tobe the (n + 1)-linear functional
on .Jt given by
T(ao, ... , an):= I ao da1 daz ... dan.

The whole business of cohomology is to extract information about the


underlying structure -for instance, a Fredholm module (.Jt, Jf, F)- from
the character alone. lt is amazing how much mileage can be obtained from
this approach. To begin with, bT = 0, since

II (-1) 1aodal···d(ajaJ+d ... dan+l + (-l)n+l I an+laodal ... dan

I I
J=Ü

= ( -l)n (ao da1 ... dan) an+l + ( -l)n+l an+lao da1 ... dan = 0.
8.4 Hochschild homology and cohomology 351

The last equality is just the trace property f aw = f wa for a E 0°, w E


nn. Thus T is an n-cocycle. Moreover, T is cyclic: by permuting dan and
ao da 1 ... dan-1, we obtain

T(ao,a1, ... ,an) = (-l)n- 1 I danaoda1 ... dan-1

= (-l)n I andaoda1 ... dan-1 = (-l)nT(an,ao, ... ,an-d.

where we have used that dan ao +an dao = d(anao) and the closedness
of f. Also, T ( 1, a 1, ... , an) = f da 1 ... dan = 0 since f is closed.
In particular, the character of the de Rham complex of a closed manifold
enjoys these formal properties. Theorem 8.2 can be thought of as giving a
cohomological equivalence of two characters, for suitable manifolds.
Example 8.3. An outstanding example of a cyclic 1-cocycle [8] is the Schwin-
ger term oc of Section 6.2. If .Jt is, say, the complex algebra generated by
the elements of o1 (V), it follows from Proposition 6.8 that

i
boc(A, B, C) = B Tr(J[J, AB][], C] - ][], A][], BC] +][],CA][], B])

= 0. (8.49)

In fact, the Schwinger term is the character of a 1-summable cycle on this


algebra, since, by taking F := -i] and using (8.13), we can rewrite the result
of Proposition 6.8 as

oc(A,B) 1 Tr(FdAdB) = -~. AdB.


=-8 I
The index formula (6.44) can now be seen in a new light. Taking F :=
= iF, the operator
E+ - E- in the notation of Chapter 6, so that]

can be regarded as belanging to the summand n 2m- 1 of top degree in a


cycle deterrnined by an odd Fredholm module with operator F. lts integral
gives, according to (8.13),

( -1) m- 1
22 m_ 1
I t [
S[F,S ] ... F,S][F,S]-
t __ __!_ t t
4 m Tr(FdSdS ... dSdS)
= indexS++·

Thus, as already noted in Chapter 6, a finitely summable odd Fredholm


module over .Jt couples integrally with KI(.Jt) via the Chern character.
On the other hand, even Fredholm modules couple integrally with Ko(.Jt)
through the Chern character [91, Prop. IV.l.4].
352 8. Noncommutative Differential Calculi

Proposition 8.12. An (n + 1)-linear functional T: 5\. n+ 1 - C that vanishes


on C EEl 5\. n is a cyclic n-cocycle if and only if it is the character of a cycle
over 5\.

Proof. We have already observed that the character of a cycle is indeed a


cyclic cocycle satisfying T ( 1, a 1 ••.• , an) = 0. Conversely, if T is a cyclic n-
cocycle on 5\. that vanishes on C EEl 5\. n, we may construct a suitable cycle
out of the universal graded differential algebra n· 5\., as follows.
Let n· := EBr=o ok 5\, take d to be the universal differential from ok 5\.
to Qk+ 1.Jl. for k < n and redefine d := 0 on on.J\.. Define f: on.J\.- C by

(8.50)

so that T is automatically the character provided that (0", d, f) is indeed a


cycle. Recall that on 5\. = 5\. ® 5\. ®n so that ao da 1 ... dan is unchanged if,
I

for any j ~ 1, a 1 is replaced by a 1 + A1 with A1 E C. In orderthat (8.50) be


well defined, it is enough that T(ao, ... , an) = 0 whenever any a 1 = 1; but
this is implied by cyclicity of T and the relation T (1, a 1 •... , an) = 0. The
same relation shows that f da 1 .•• dan = 0, so that f is closed.
To see that f has the graded trace property, first consider the case where
Wn E Qn.J\. and a E 5\.. Now Wna- aWn = [Wn,a] = (-l)nb(Wnda),
by Exercise 8.20. Since bT = 0, it follows that f Wna = f awn. Next, if
Wn-1 E on- 15\. and da E 0 15\., then

Wn-1da- (-l)n- 1daWn-1 = [Wn-1,da]


= (-l)n- 1 (d[Wn-1,a]- [dWn-1.a])
= (-l)n- 1d[Wn-1,a] + b(dWn-1 da),

also by Exercise 8.20; since f is closed, bT = 0 implies that f Wn-1 da =


( -1) n- 1 f da Wn-1· Repeated use of these two cases gives

JWn-k ao da1 ... dak = ( -1 )n- 1 Jdak Wn-k ao da1 ... dak-1
= · · · = (-1)k(n-1) Jaoda1 ... dak Wn-k

-- (-1)k(n-k) f ao d a1... d ak Wn-k· 0

The condition T ( 1, a 1 , ••. , an) = 0 is linked to the use of the universal


graded differential algebra in the form Qk 5\. = 5\. ® 5\. ®k. Had we followed
the alternative convention ok 5\. := 5\. + ® 5\. ®k (suitable for nonunital alge-
bras also), as is donein [91], this condition could have been avoided.
Some ambiguity can be detected in the definition of the character asso-
ciated to a Fredholm module, as the precise value of n is only subject to a
lower bound. lf T n is a cyclic cocycle, we get a sequence of cyclic cocycles
8.4 Hochschild homology and cohomology 353

with the same parity T n+2k. for k E ~. where n is the smallest integer com-
patible with summability. We shall remove this ambiguity in Chapter 10,
after introducing a periodicity operator that links this sequence of cocy-
cles.
~ It is often possible to compute the Hochschild homology of an algebra
or module by using the following device.
Definition 8.18. Let 'B be an algebra and :Fa left 'B-module. Aresolution
of :F (over 'B) is an acyclic complex of left 'B-modules of the form

ß ß ß ß E
···- Pn- Pn-1 - · · · - P1 - P o - :f- 0. (8.51)

We say the resolution is projective if each Pi is a projective left 'B-module.


If .Jl is any unital algebra with opposite algebra .Jl o, we can form the
algebra 'B := .Jl ® .Jl o (tensor product over ([), with product (a1 ® a:;)(a3 ®
a4) := a1a3 ® (a4az)o. Any .Jl-bimodule can be regarded as a left 'B-module
by setting (a1 ® a:; )s := a1saz. In particular, the .Jl-bimodule .Jl ®(n+Z) can
be identified with 'B ® .Jl ®n through

Also, .Jl itself is a left 'B-module via (a1 ® a2)c := a1ca 2. The modules
Pn := .Jl ®(n+Z) provide a standard projective resolution of .Jl over 'B, called
the bar resolution [319]:

... _ .Jl ®(n+2) ~ .Jl ®(n+l) ~ ••• ~ .Jl ® .Jl ~ .Jl- O, (8.52)

where m is the multiplication map m(a 1 ® a 2 ) := a 1az, and the maps b'
are given by (8.39). Now, (8.40) says that b' s + sb' = 1, so s is a contracting
homotopy and the complex (8.52) is acyclic; moreover, m is onto because
.Jl is unital, so (8.52) is indeed a resolution. lts modules are all of the form
'B ® Vn for some vector spaces Vn, so they are free 'B-modules.
We can now tensor the bar resolution with any .Jl-bimodule 'E, which
we may regard as a right 'B-module. Indeed, 'E ®:B .Jl ®(n+2) "" 'E ® .Jl ®n =
Cn (.Jl, 'E) via the map <I>: s ® (ao ® • • • ®an+ I) .- an+lsao ® a1 ® • • • ®an.
Then <I> o (1x ® b') = b o <I>, so we get the Hochschild complex; therefore,
H. (.Jl, 'E) is the homology of the tensored complex.
That would be a fairly trivial remark, were it not that the bar resolution
can be replaced by any other projective resolution, so that the Hochschild
homology can be computed as the homology of the complex 'E ®:B P •. This
follows from the next proposition.
Proposition 8.13. Any two projective resolutions of the same 'B-module :F
are chain-homotopy equivalent.
354 8. Noncommutative Differential Calculi

Proof. Suppose :F has two resolutions, the upper one being projective:

ß ß
· · · -Pz -PI-Po
I I I~
I I I ~ (8.53)
h I !I I fo I IJ J.
f b f b f ~
· · · -.Qz -ei-.Qo

Since Po is projective and 17 is onto, there is a map fo: 'Po - .Qo so that
11fo = E (see Definition 2.13). Now 11foß = 0, so foß maps 'PI into ker 17 =
im b ~ .Qo; since 'PI is projective, we get a map !I: 'PI - .QI such that
b!I = foß. lt is clear how this process may be continued inductively, to
obtain a chain map j.: 'P. - .Q. satisfying 11fo = E.
If the lower resolution is also projective, we obtain another chain map
g.: .Q. - 'P. satisfying Ego = 11· lt remains to show that f.g. and g.j. are
chain-homotopic to the identity. For that, we may simplify by supposing
that .Q. = 'P. (after replacing each fn by Bnfn, say), and we must then
show that f. is chain-homotopic to the identity. We need to find maps
Sn: 'Pn - 'Pn+I suchthat ßsn +Sn-Iß = fn -1p", as in the following diagram:

Now E(fo- lp0 } = Efo- E = 0, so fo -lp0 maps Po into kerE = imß ~Po,
and then we can find so: 'Po- 'PI so that ßso = fo -1p0 • Next, ß<!I -1p1 -
soß) = ß(fi -1p1 ) - <fo -1p0 )ß = ß!I- foß = 0, so !I -1p1 - soß maps into
ker(ß: 'PI- Po)= im(ß: 'Pz- 'PI). We can now find a map si: 'PI- 'Pz
suchthat ßsi = !I - 1p1 - soß, and so !I - 1p1 = ßsi + soß. On repeating
this argument inductively, the chain homotopy s. emerges. D
In the next section, we use this machinery to examine the Hochschild
homology of the commutative algebra coo (M).
Hochschild cohornology may also be computed using projective resolu-
tions. Indeed, if 'B = 5\ ® 5\ o, there is an isomorphism of cochain com-
plexes Hom 21 (5l ®(•+Zl, '.E) ::: C (5\, '.E) under the correspondence T ..... <p
determined by <p(ai, ... , an) := T( 1, ai, ... , an, 1); notice that
(ao<pan+I)(ai, ... ,an) = (ao®a~+I)T(1,ai, ... ,an,l)
= T(ao,ai, ... ,an,an+d·

Exercise 8.23. Compute the map Hom21(5\ ®(n+Zl, '.E) - Hom 21 (5l ®(n+ 3 l, '.E)
got by transposing b': 5\ ®(n+ 3 l - 5\ ®(n+Zl, and show that it corresponds
to the Hochschild coboundary b given by (8.47). <>
8.5 The Hochschild-Kostant-Rosenberg-Connes theorem 355

8.5 The Hochschild-Kostant-Rosenberg-Connes


theorem
Wehave already proved that HHI(C""(M)) = 5'1. 1 (M) if M is a compact
manifold and that 5'1." (M) is a direct summand of HH. (C"" (M)) (as conse-
1

quences of Propositions 8.1 and 8.11 and the remarks preceding Proposi-
1

tion 8.10). This raises the possibility that


HH.(C""(M)) = 5'1."(M). (8.54)
That would be tantamount to a homological construction of the de Rham
complex. lt turns out that (8.54) is true! This is a theorem by Connes [86];
we call it the Hochschild-Kostant-Rosenberg-Connes theoreml as it ex-
tends a well-known algebraic result by Hochschild Kostant and Rosen- 1

berg [249] that computes the Hochschild homology of certain finitely ge-
nerated Noetherian rings.
Actuallyl Connes does not prove the direct analogue of the HKR theoreml
but formulates his result in a dual way. Let C denote a de Rham current of
degree k -by definitionl a continuous linear form on 5'1.k(M); its value on
the k-form 11 is denoted fc TJ. The boundary of Cis the current ac defined
by
Iac
w := f dw for w E 5'1.k- 1 (M);
Je
the notation d here means simply the ordinary exterior derivation of formsl
and we hope that the readerwill not confuse it with the universal d which 1

in the commutative case is the finite difference operator. The skewsym-


metrization of a Hochschild k-cocycle <p on 5'1. with respect to all arguments
but the first yields a k-current Ccp:

ICcp
aoda1A···Adak:= ~1
.
L (-l)rr<p(ao~arro)l"'larr(k)).
TTESk
(8.55)

Over a commutative algebral skewsymmetrizing takes cocycles to cocy-


cles and kills coboundaries. Indeedl if f./.1 is a (k- 1)-cochain then 1

Ak{bfjl)(aol"'lak) = ~1 L (-l)rr bf./J(ao~arr(l)l"'larr(k))


' TTESk

= ~1 L (-1) rr (f./.1 (aoarr(l) 1 arr(2) 1 ... 1 arr(k))


• TTESk

+ ( -1)k f./.1 (arr(k)aol arr(l). ... arr(k-1))) 1

1 k .
= k! j~1 (-1) 1 ~(-l)a(f./J([aolajl.aa(l)l ... aa(k)) 1 = 0 1 (8.56)

where u ranges overthe permutations of {1 ~}-1 1 }+1 1 k}. The other


1 ••• ••• I

terms arising from hf./.11 of the form f./J(aol ... arru>~ arr(j+l) arr(k) can-
I 1 ••• I )I

cel on summing over rr. Thereforel (8.55) yields a map from Hochschild
356 8. Noncommutative Differential Calculi

cohomology classes to de Rham currents. The claim is that this map is an


isomorphism of differential algebras (we shalllater exhibit the differential
on H· (.Jt) with which the operator agets intertwined).
We first discuss why this result is reasonable (see [258]). We have al-
ready mentioned that, when .Jt is commutative, H H. (.Jt) is a commutative
superalgebra with the shuffle product. The Hochschild-Kostant-Rosenberg
theorem asserts that, for a suitable dass of commutative algebras, called
smooth algebras, in particular for coordinate rings of affine varieties, the
natural map of Proposition 8.10 is an isomorphism. A commutative algebra
.Jt is "essentially of finite type" if it is a localization of a finitely generated
algebra. We say that an algebra .Jt, essentially of finite type, is smooth [482]
if H} (.Jt, 1:) = 0 for all.Jt-modules 1:, where H} (.Jt, 1:) denotes the submod-
ule of H 2 (.Jt, 1:) consisting of equivalence classes of symmetric factor sets
u: .Jt x .Jt - 1:. This means that there are no commutative algebras that
can be extensions of .Jt by any module. The algebra C" (M) has no commu-
tative extensions; so the Hochschild-Kostant-Rosenberg-Connes theorem
is plausible.
The Hochschild cohomology H•(.Jt,.Jt) plays a central role in the theory
of deformation of algebras. An important tool of this theory is the Ger-
stenhaber bracket on Hochschild cochains (that becomes the Schouten-
Nijenhuis bracket in the commutative differentiable case); consult [64, 195,
288]. The book [64] describes a wealth of "geometric models" for non-
commutative spaces. lt should, however, be clarified that the "differential
Hochschild cohomology" Hciiff(C (M)) alluded to in [64, §19.3] isavariant
00

of H• (51., .Jt) for .Jt = coo (M), as opposed to the much subtler cohomol-
ogy H· (.Jt, .Jt * ), here designated by HH• (.Jt); thus, the algebraic corres-
pondence between that cohomology and multivector fields on M, obtained
in [61], while suggestive, is not a cohomological partner of the HKR theo-
rem.
~ Since coo (M) is far from being finitely generated, its topology must play
a role in going beyond the HKR theorem. We must first review the bidding
to see how Hochschild (co)homology works for algebras that carry natural
locally convex topologies. (This is already indicated by our discussion of the
K-theory of pre-C*-algebras in Section 3.8.) We then speak of continuous
Hochschild (co)homology. We also need to adapt the machinery of projective
resolutions to the case of topological.Jt-bimodules.
Suppose, then, that .Jt is a locally convex algebra whose product opera-
tion (a, a') - aa' is a (jointly) continuous bilinear map from .Jt x .Jt to .Jt.
This is a fairly restrictive assumption in general, but it does hold for Frechet
algebras, as a consequence of the uniform boundedness principle [383]. A
topological right .Jt -module is a right .Jt -module 1: for which the module Op-
eration (s, a) - sa is a jointly continuous bilinear map; again, it is enough
that 1: and .Jt be Frechet spaces. We shall always assume that both .Jt and
1: are complete; an ordinary tensor product like 1: ® .Jt must then be re-
8.5 The Hochschild-Kostant-Rosenberg-Connes theorem 357

placed by the completed tensor product 'E ® 5\.. This is the completion of
the algebraic tensor product of 'E and 5\. in the projective tensor product
topology, determined by all the seminorms of the form s ® a .... p(s)q(a),
where p and q are seminorms on 'E and 5\. respectively. (This extends the
definition given for Banach spaces in Section l.A.) For instance, the Frechet
algebra C"" (M), whose topology was described in Section 3.8, is also what
is called a nuclear space, which implies that there is an isomorphism of
topological vector spaces [224, Thm. II.l3]:

Free right 5\.-modules, of the form V® 5\. with V a complex vector space,
are now replaced by V ® 5\., where V is a complete locally convex space;
examples are the multiple tensor products 5\. ®n := 5\. ® 5\. ® · · · ® 5\.
(n times). We say that a right 5\.-module 'Eis topologically projective if it is
a direct summand of some such V® 5\.. We leave to the reader the task of
stating the analogaus definitions for left modules and bimodules.
In particular, if 5\. = C"" (M), then 'B := 5\. ® 5\. o is just 5\. ® 5\. ""
C""(MxM), and 5\. is a topologicalleft 'B-module via (a1 ®a2)c := a1ca2 =
a 1a 2c, whose value at x E M equals (a 1 ® a2)(x,x)c(x); more generally,
(bc)(x) := b(x, x)c(x) if b E 'B. In other words, the action of 'Bon 5\. is ob-
tained by transposing the diagonal embedding ß: x .... (x, x) : M .... Mx M.
The "bar resolution" of 5\. over 'B has kth component 5\. ®(k+ 2 > "" C"" (M x
Mx Mk), where Mk denotes the cartesian product of k copies of M .

.,.. We now show, following Teleman [452], that the continuous Hochschild
homology of C"" (M) is local in that it depends only on the values near the
diagonal of functions on Cartesian powers of M. Let

ßk(M) := { (x,x, ... ,x) E Mk: x E M}

denote the diagonal submanifold of Mk. Choose a function 1J!: [0, oo)
[0, 1] satisfying IJ!(t) = 1 for 0 ::5 t ::5 !. which decreases smoothly from 1
to 0 for! ::5 t ::51, and suchthat IJ!(t) = 0 fort~ 1. Write IJ!s(t) := lj!(t/s)
for s > 0, so that supp IJ!s = [O,s]. Let p be a distance function on M (in-
duced, say, by some Riemannian metric); then a system of neighbourhoods
of the diagonal ßk+l (M) is given by Vk+l (E) := {X E Mk+l : Yk+l (X) ::5 E},
for E > 0, where

rk+l (xo, ... ,Xk) := p(xo,xd 2 + p(x1,x2) 2 + ···


+ p(Xk-I.Xk) 2 + p(Xk,Xo) 2. (8.57)

Notice that, for k ~ 1,

Yk+dXo, ... ,Xk) = p(Xo,XI) 2 + Yk(XI. ... ,Xk)- p(Xk,XI) 2 + p(Xk.Xo) 2


= rk(XI, ... ,Xk) + 2p(xo,XI)P(Xk.Xo)costh
358 8. Noncommutative Differential Calculi

for a certain angle (h. Also, p(Xk-I.xo) 2 does not exceed 2p(Xk-I,Xk) 2 +
2p (Xk, xo ) 2 • These considerations, tagether with (8.5 7), yield the estimates

Yk(Xlo···oXk) ::5 3Yk+I(Xo, ... ,Xk), (8.58a)


Yk+I(Xo, ... ,Xk-I,Xo) ::5 2Yk+I(Xo, ... ,Xk). (8.58b)

Write also Xk+l,E := t.Jh o Yk+l : Mk+l - [0,1]; then supp(Xk+l,E) s;


Using formula (8.45) for the Hochschild boundary b: C"" (Mk+l) -
Vk+l (E).
C""(Mk), one sees that b(Xk+l.EF) = Xk,E bF in C""(Mk). Therefore, if F E
C"" (Mk+l) satisfies bF = 0, then it can actually be written as the sum of
two Hochschild cycles:

F = Xk+l,EF + (1- Xk+l,E)F,

with the first supported in Vk+l (E) and the second vanishing on Vk+l (E /2).
We wish to show that the second type make only a minor contribution to
the Hochschild homology of C"" (M).
=
Let 'Et := { F E C"" (Mk+l) : F 0 on Vk+dE) }. These ~-modules form a
subcomplex 'E: of C.(C""(M)); each 'Et is a closed subspace of C""(Mk+l)
-so it is complete- and 'Et c 'Ef
for 0 < 8 < E. Let 'EZ denote the inductive
limit limEJo'Et. (As a topological vector space, 'EZ is in generalnot metriza-
ble, but is complete, being a strict inductive limit ofFrechet spaces [412,
Il.6].)

Proposition 8.14. The complex 'E~ is acyclic.

Proof. Define continuous linear maps aE: C"" (Mk+ 1 ) - C"" (Mk+Z) by

If F E 'Et and Yk+Z (xo, . .. , Xk+ I) < EI 3, then Yk+ dx1, ... , Xk+ I) < E from
(8.58a), so that F(x1, ... , Xk+l) = 0 and also aEF(xo, ... , Xk+l) = 0. In other
wor d s, aE maps rrE rrE/3
L.k mto L.k+lo
o

The map aE does not yield a chain homotopy between the identity and
the zero map, but it comes fairly close to doing so:

[ (baE + aEb)F](xo, 0 0 0, Xk) - F(xo, 0 0 0, Xk)


k
~ .
= L.(-1) 1 aEF(Xo, ... ,Xj,Xj, .. o,Xk) + (-1) k+l aEF(Xo,o .. ,Xk,Xo)
j=O
+ I.JJE(p(xo,xd 2 )bF(xi, ... ,Xk) -F(xo, ... ,xk)
8.5 The Hochschild-Kostant-Rosenberg-Connes theorem 359

This equals t/h(p(xo,x!) 2 ) times


k
~ . k
L.(-1)1F(XI, ... ,XJ,XJ, ... ,xk)- (-1) F(xi, ... ,Xk,Xo)
j;l

vEF(xo, ... ,Xk)


:= (-1)kiJIE(p(xo,xd 2 )(F(xi, ... ,Xk,Xo)- F(XI, ... ,Xk.XI)). (8.59)

It follows from (8.58b) that vE maps 'Ef into 'E~ 12 •


On iterating (8.59), we find that v}F(xo, ... ,Xk) equals

«J~t(p(xo, xd 2 )IJIE(p(x~o xz) 2 ) (F(xz, ... , Xk, xo, xd


- F(xz, ... , Xk,Xo,xz)- F(xz, ... ,Xk.XI,xd + F(xz, ... ,Xk. x1. xz) ).
After k iterations, we arrive at

n
k-1
v:F(xo, ... ,Xk) = IJIE(p(Xj,Xj+d 2) G(xo .... ,Xk),
j;Q

where G(xo, .. . ,Xk) is a signed sum ofvalues F(yo, ... ,yk) with eachy; in
{xo, ... ,xk}. This vanishes unless p(x1 ,x1+d <JE for j = O, ... ,k- 1,
and so also p (Xko x 0 ) < kJE, by the triangle inequality. In that region,
Yk+dxo, ... ,xk) < k(k + 1)E; we conclude that v:F = 0 for FE 'E:(k+l)E.

Exercise 8.24. Show that bvEF = vEbF for FE C"" (Mk+l ). 0


To finish the proof, consider the operators sE := aE(1 + vt + · · · + v:- 1).
They satisfy bst + sEb = (baE + aEb) I~:~ vi = 1- v:, so (bsE + Stb)F = F
whenever F E 'Ek(k+l)E
k
and therefore s E'· 'Ek(k+l)E
I k
- 'Ek(k+l)E/ 3 · 2k-l 1·s a
k
chain homotopy between the corresponding inclusion and the zero map.
In the Iimit E I 0, the complex 'E? is acyclic. 0

Corollary 8.15. HH.(C""(M)) depends only on the germs of functions FE


C"" (Mk+l) on the diagonals Llk+l (M). a
This construction is important because it shows the local character of
(continuous) Hochschild homology and, by duality, of Hochschild coho-
mology. The precise formulation of this locality may now be given, namely
that the Hochschild dass of any cycle F E C"" (Mk+ 1) is determined by
the restriction of F to any diagonal neighbourhood Vk+l (E). A Hochschild
k-cocycle is then a continuous k + 1-linear functional on C"" (M), or equiva-
lently a continuous linear functional on C"" (Mk+l ), depending only on the
360 8. Noncommutative Differential Calculi

restriction of its argument to any neighbourhood of ßk+ 1 (M): that is to say,


it is given by a distribution on the compact manifold Mk+l whose support
lies in the diagonal ßk+l (M).
lt is possible to compute the continuous Hochschild homology groups
of coo (M) by continuing the previous line of argument, using not just the
germs of smooth functions along the diagonals but also their higher-order
jets. Thereby one arrives at a differential-geometric proof of the direct ana-
logue of the HKR theorem; we refer to [452] and [47] forthat proof.
~ Connes' extension of the (dual) HKR theorem proceeds by finding a pro-
jective resolution of coo (M) over 'B = coo (M x M) that consists of finitely
generated projective modules. These are, of course, by the Serre-Swan the-
orem, modules of (smooth) sections of vector bundles over M x M. In fact,
one can take 'Ek := [ 00 (M X M,Ek) where Ek comes from pulling back the
kth exterior power of the (complexified) cotangent bundle

via the map pr 2 (x,y) := y. With these modules, we form a complex


tx tx tx tx ß•
0 - 'En- · · · - 'E2- 'E1- C (MX M)- C (M)- 0
00 00
(8.60)

that looks like the de Rham complex over M, pulled back to M xM, but with
maps tx: 'Ek- 'Ek-l "going the wrong way". This can be achieved if these
maps are contractions with a certain vector field X E *(MxM), since t~ = 0
automatically, provided we can construct a chain homotopy consisting of
continuous linear maps sk: 'Ek- 'Ek+ 1, making (8.60) acyclic.
The sequence (8.60) is exact at 'Eo = coo (Mx M) if and only if the vector
field X satisfies X(x,x) = 0 and X(x,y) * 0 for x * y. To achieve that
near the diagonal, we can choose a Riemannian metric on M and use the
corresponding exponential map expy: TyM - M (see [238], for example),
which is one-to-one near the origin of TyM, to define

X(x,yl := exp_;;l (x) E TyM

whenever x is close enough to y. If E > 0 is small enough, V2 ( E) is a tubular


neighbourhood of ß 2 (M) c Mx M, over which Xis a smooth section of the
bundle pr! M, satisfying X(x,y> = 0 for x = y only.
To proceed, we can either restriet everything to V2 ( E), or extend X to all
* *
of Mx M, keeping the property that X(x,y> 0 for x y. By construction,
X is a real vector field, but a complex one would do just as well. Suppose,
then, that M carries a real vector field Y' that never vanishes; then we can
set Y(x,y) := Y~, and replace Xby X2,EX +ip(x,y)Y. Such anonvanishing Y'
8.5 The Hochschild-Kostant-Rosenberg-Connes theorem 361

exists if and only ifthe Euler characteristic x(M) := Ir=o( -1)k dimH~R (M)
is zero; that is automatic if n = dimM is odd, by Poincare duality on the
compact manifold M. If dimM is even, x(M) can be nonzero, but in that
case x(M x lr) = 0. For the time being, then, we assume that X(M) = 0, so
that X can be taken as globally defined and vanishing only on flz(M).
Lemma 8.16. Ifx(M) = 0, then (8.60) is a projective resolution o(C 00 (M)
over coo (M x M).
Proof. The idea of the proof is to establish a sort of "wrong-way Poincare
Lemma" near the diagonal, in order to construct a contracting homotopy.
Choose any 1-form 11 E 'E1 suchthat 17(X) = 1 on Vz(E). Define c/>r
on Vz(E) by cJ>dx,y) := expx(tX(y,x)). Then, for w E 'Ek. we can define
s(w) E 'Ek+1 by

f1 dt
Sk ( w) := X2,2E Jo cl>i (dy (Xz,Ew)) T + (1 - Xz,E) Tl A w.

Here dy denotes the exterior derivative in the second variable. If f E


C (M), then (jw)(x,y) := j(x)w(x,y) makes each 'Ek a left C (M)-
00 00

module, for which s(jw) = f s(w).


In a neighbourhood of x E M, we can use Riemannian normal coordinates
such that the exponential map is given by c/>r (x, y) = t y for 0 :5 t :5 1, and
X(x,y) = -y. Thus, if w E 'Ek vanishes outside Vz(E) and if w(x,x) = 0,
then

1 1
0
dt +LX
cJ>i(dyLXW)-
t
1 1
0
dt =
cJ>i(dyW)-
t
1 1
0
cJ>i((dyLX + Lxdy)W) -dtt
f1 dt
= Jo cJ>i (Lxw) T = w,
where we have used the Cartan formula dtx + txd = Lx. The last equality
comes from the following identity, which we express in the notation of
Section 7.3: if R is the Euler vector field on ~n, generating the dilations
{Pdt>O. and if a E .Jl.k(~n) with ao = 0, then

a=
e
Jo
d 1 d dt f1 dt
dt(pia)dt= Jotdt(pia)T= JoPi(DRa)T.
f

Now we compute
f1 dt
(SLx + Lxs)(w) = X2,2E Jo cJ>i(dy(Xz,ELxW)) T + (1- xz,E) 11/\ LxW
f1 dt
+ XZ,ZELX Jo c/>i(dy(Xz,Ew)) T + (1- Xz,E) tx(l1 A w)

= X2,2E (Xz,Ew) + (1- Xz,E) 17(X) w


= (Xz,E + 1- Xz,E) w = w,
since X2,2E = 1 when XZ,E > 0. Thus, stx + txs = 1 on each 'Ek. 0
362 80 Noncommutative Differential Calculi

Proposition 8013 now assures us that both the bar resolution (8o52) and
the resolution (8o60) compute the Hochschild cohomology of .J\ = C"" (M),
through the isomorphisms of cochain complexes

Hom11CE.,.Jt*) ""Hom11 (.Jt®<•+ 2 l,.Jt*) ""C"(.Jt,.Jt*),

where 'B = C""(M x M)o Indeed, the first isomorphism is obtained from
a chain map of 'B-modules j.: 'E. - .J\ ®<•+ 2 l, which is worth describing
explicitlyo The boundary map b' on .J\ ®<k+ 2 l = C"" (Mx Mx Mk) is given by

b'F(x,y;xl, 000,Xk-d = F(x,y;x,x1, 000,Xk-1)


k-1
·-1
+ L...(-1)1
~
F(x,y;x 1, 000,Xj,Xj, 000,Xk-1), (8061)
j=l

for k ~ 1. The case k = 0 is m(a ® b) = ab E C""(M)o that is, mF(x) =


F(x,x) = ß*F(x); in particular, the rightmost link in the desired chain
map is fo = id11, the identity on C""(M x M)o
The other links are given by the recipe

fkw(x,y;xl. 000,Xk)
:= 2: (-1)TT W(X, y)(X(Xrroi>Yh ooo, X(xrrlk),y) ), (8062)
TTESk

where w E 'Eko To see that, we check directly that it intertwines the bound-
ary maps:

b'fkw(xo,y;xl,ooo,Xk-1) = fkw(xo,y;xo,xl.ooo,Xk-d
= 2: (-l)rrw(xo,y)(X(xrr 1o>,y),ooo,X(xrrlk-li>Yl)

Only the first term on the right of (8061) contributes, since the skewsym-
metry in x1, 000, Xk-l kills the terms with a repeated x i 0The case k = 1 also
fits the pattern:

b'fiw(x,y) = fiw(x,y;x) = w(x,y)(X(x,yJ) = txw(x,y),

so b' !I w = txw = fotxWo


.,.. Denote by 'Dk (M) the space of de Rham k-currents on Mo Given any
CE 'Dk(M), we define a Hochschild k-cocycle cpc by

cpc(ao,al,ooo,ak) := fcaodaiA 0 0
ol\dako (8o63)
8.5 The Hochschild-Kostant-Rosenberg-Connes theorem 363

The cocycle property follows at once from the proof of Proposition 8.11;
moreover, it is clear that Äk<pc = <pc, where Äk here denotes the skewsym-
metrization map on cochains,

Äk<p(ao,a1, ... ,ak) := ~1 L (-1)rrcp(ao,arrob····arr(k)). (8.64)


• TTESk

lt follows from Proposition 8.10 that Äk<p is a Hochschild cocycle.


The equations (8.63) and (8.55) yield two maps ß: 'Dk(M) - HHk(Jt) :
C ...... [cpc] and cx: HHk(Jt,5t*) - 'Dk(M): [Akcp) ..... C<P, which satisfy
cx o ß = id. The map cx is well-defined on account of (8.56). Therefore, ß is
an injective 5t-module map, identifying 'Dk(M) with a direct summand of
HHk(Jt).
We wish to show that ß is onto, i.e., that ß o cx = id on HHk(.Jt); or,
equivalently, that cx is injective. Suppose for a moment that this is true
when x(M) = 0. If M is even-dimensional, let C := C 00 (M x lf) and define
r: C- 5t by r(c)(x) := c(x, 1).
Exercise 8.25. Show that Hkr (the map in cohomology induced by r) em-
beds HHk(.Jt) as a direct summand of HHk(C). <>

Thus it is enough to show that ß o cx = id on H Hk (C). Bett er yet, we may


assume that x(M) = 0, replacing M by Mx lf if necessary. Thus, we may
use the projective resolution (8.60).
Theorem 8.17 (Connes). If M is a compact manifold, there is a canonical
isomorphism, given by (8.55), between the continuous Hochschild cohomol-
ogy module HHk(C 00 (M)) and the space 'Dk(M) of de Rham k-currents
onM.

Proof. Suppose, as we may, that X(M) = 0. Since (8.60) is a projective res-


olution of 5t, it follows that the continuous Hochschild cohomology of 5t
is the cohomology of the complex Hom21Cfk. 5t * ); our task is to identify
the latter concretely. The diagonal embedding ~: M - M x M satisfies
pr 2 o~ = idM, so ~*Ek = ~*pr~ AkT(!M = AkT(!M. Now Proposition 2.12
(or rather, its analogue in the smooth category) shows that

'Ek ®11Jt = po (Mx M, Ek) ®c~wxM) Coo (M) ""roo (M, ~ *Ek) = 5t k (M).
The (topological) dual space is then

(8.65)

The boundarymap tx: 'Ek - 'Ek-1 induces amap 'Dk(M) - 'Dk- 1(M) that is
the transpose of w ...... ~* (txw); butthislast map is zero, since X vanishes
on the diagonal. Therefore, the complex Hom21 ('E., 5t *) has zero maps,
and its cohomology consists of the component modules themselves; that
is, HHk(C 00 (M)) ""'Dk(M).
364 8. Noncommutative Differential Calculi

To any Hochschild k-cocycle cp corresponds Tcp E Hom:B (.Jl ®(k+ 2 l, .Jl *)


given by

(8.66)

Now Tcp o fk E Hom:B CEk. .Jl *); Iet C~ be the associated current, given
by (8.65); then [cp] - C~ is an isomorphism from HHk(.Jl) to 1Jk(M).
1t remains only to compute C~ explicitly.
Take w E .Jl k (M) and Iet w := pr2 w, so that ~ * w = w. The current C~
is given by fc<p w := Tcp(fkw)(1). If, say, w = aoda1 A • · • A dak, then

n
k
fkw(x,y;xl, ... ,Xk) = I (-1)7Tao(y) daj(y)(X(Xrr(j),y))
TTESk j=l

n
k
= I (-1)0"ao(y) dau(j)(y)(X(xj.y)). (8.67)
uESk j=l

The skewsymmetry of this expression shows that Tcp o fk remains un-


changed if cp is replaced by A.kcp; thus we can as weil suppose that A.kcp =
cp. By the construction of X, da(y)(X<x.yJ) = X(x,y)(a(y)) - a(x) as
y - x; combining that with (8.66) and (8.67), we obtain

Tcp(fkw)(l) = I
0"~
(-l)ucp(ao,au(l), ... ,au(k)) = k! J aoda1A· · ·Adak.
~

In other words, oc[cp] = Ccp = (1/k!)C~. Thus, oc is indeed injective. As a


result, ß is an isomorphism when x(M) = 0, and so also in general. D
To finish, we want to express the de Rham boundary map C ..... ac in
terms of a functorial degree-lowering operation in Hochschild cohomology.
If CE 1Jk(M), then Stokes' theorem gives

fac ao dal 1\ .•. 1\ dak-1 = r dao


Je
1\ dal 1\ ••• 1\ dak-1

= cpc(1,ao, ... ,ak-d· (8.68)

Moreover, CfJc(ao, ... , aj-1, 1, aj+l• ... , ak-d = fc ao da1A · · ·Ad( 1) A · · ·A


dak-l = 0 if the 1 occurs in the jth position and j > 0. We can therefore
add or subtract such terms to the right hand side of (8.68). In particular,
fac ao da1 A • • • A dak-l equals

Bocpc(ao, ... ,ak-1) := CTJc(1,ao, ... ,ak-d- (-l)kcpc(ao, ... ,ak-l.l).


(8.69)

In view of the skewsymmetry of cpc (in all variables but the first), we may
also introduce some permutations of these arguments.
805 The Hochschild-Kostant-Rosenberg-Connes theorem 365

Definition 8.19. If .\ denotes the operation (8.48) of cyclic permutation, let


N := 1 + .\ + · + ,\k on k-cochains; since A.k+ 1 = 1 on k-cochains, it is
0 •

clear that N(l- ,\) = ( 1- .\)N = 0 and that any cochain of the form NtJ.! is
cyclic. In particular, the operator B := NBo maps Hochschild k-cochains to
cyclic (k- 1)-cochainso Writing I for the inclusion of cyclic cochains into
all Hochschild cochains, we obtain maps IB: ck (A, A *) - ck- 1 (A, A * ):
k-1
lBtJ.!(ao, .. o,ak-1) := 2:(-l)i<k- 1ltJ.1(1,aJ.····ak-1.ao, ... ,aJ-1) (8.70)
j=O
+ (- 1) (j-1)(k-1),,,(
'Y aJ, ... ,ak-1.ao, ... ,aJ-1•
1) .

Proposition 8.18. The de Rham boundary o:


'Dk(M) - 'Dk-dM) corre-
sponds to the operation (1/k)IB: HHk(A)- HHk- 1(5\).

Proof. Since <Pc(ao, ... , ak-1. 1) = 0 for CE 'Dk(M), we get

k-1
IB<Pc(ao, .. 0, ak-d = 2: (-1 )i<k- 1l<Pc(1, a1, ... , ak-1, ao, ... , a1-d
j=O

2: (-l)i<k- 1) J da1 1\ • • • 1\ dak-1 1\ dao 1\ • • • 1\ da1-1


k-1

j=O C
k-1
= 2: J
j=O C
dao 1\ • · • 1\ dak-1 = kJ
iJC
aoda1 1\ • 0
• 1\ dak-1,

by using (8.68); thus IB<Pc = k<P;;c. Since C - [<Pc] is an isomorphism,


the operations aand (1/k) IB are matched at the level of cohomology. D
Part 111

GEOMETRY

And don't speak too soon,


for the wheel's still in spin

-Bob Dylan
9
Commutative Geometries

In this chapter, our goal is to construct geometries on commutative spaces.


That is to say, given a noncommutative space admitting a differential cal-
culus that is in fact commutative -in other words, an algebra .J\ = C"" (M)
for some differential manifold M- we shall find the extra structures that
give rise to geometries.
We must maneuver within two constraints. First of all, our construction
must yield classical differential geometry on M as its output. Secondly, it
must be framed in algebraic (or rather, operatorial) terms so that it extends
to a general recipe for geometries on noncommutative pre-C* -algebras.
The bridge between classical differential geometry and modern noncom-
mutative geometry is the Dirac operator. This Ieads us to Iimit our study
to those manifolds on which a Dirac operator may exist (the spinc mani-
folds, described below). On the other hand, within that restriction, all the
geometric information we need can be teased out of the Dirac operator: a
farnaus example [88] is how this operator determines the geodesie distance
between points of M.
The main part of the chapter develops the general theory of Dirac opera-
tors. lt is followed by two appendix sections that develop detailed examples
of particular Dirac operators; these are placed at the end only in order not
to interrupt the generalline of argument.
370 9. Commutative Geometries

9.1 Clifford modules


Throughout this chapter, M will denote a connected smooth manifold of
dimension n, without boundary. As in Chapter 7, we take M tobe com-
pact, mainly for convenience of notation. Unless the contrary is explicitly
indicated, 5l will denote the pre-C*-algebra coo (M) of (complex) smooth
functions on M, and A := C(M) its C*-completion.
Definition 9.1. A Euclidean vector bundle of M isareal vector bundle E-
M, of rank r, say, endowed with a positive definite pairing
g : [ 00 (E) X [ 00 (E) - coo (M, ~).
that is coo (M, ~)-bilinear, and extends to a C(M, ~)-bilinear form g:
f(E) x
[(E) - C(M, We also denote by g the obvious complexifications to
~).
bilinear forms on the complexified roo (E) and [(E), with values in coo (M)
and C(M), respectively. By the Serre-Swan theorem, g is the image, under
the r functor, of a bundle map E X E - M; it thereby induces a quadratic
form Bx on each fibre Ex. Therefore, we can form the complex Clifford
algebras CHEx) for each x E M. Since Cl( Ex) "" A •E~ as vector spaces,
these Clifford algebras are the fibres of a complex vector bundle Cl E - M
of rank zr (which is just the exterior bundle A•Ec - M under another
name).
The A-module of continuous sections B := f(GE) is thus an involu-
tive algebraunder the pointwise Clifford product (vi\)(x) := v(x)i\(x),
uniquely determined by the relation v 2 = g(v, v) for v E f(E), where
(v*)(x) := v(x)* comes from the involution in each Cl(Ex). The normal-
ized trace in each fibre endows f(GE) with a Hermitian pairing (5.7):
(v I i\)(x) := T(v(x)*A(x)), for each x E M,

making it a right C* A-module. In fact, B is a C* -algebra in its own right,


under the norm llvll := sup{ llvxll: x E M}, where llvxll is the C*-norm in
the finite-dimensional algebra Cl( Ex).
The prime examples of Euclidean vector bundles are the tangent and
cotangent bundles over a Riemannian manifold: see Definition 7.1. The
corresponding pairings are the metric g on vector fields *(M) and g- 1
on 1-forms 5l 1 (M).
Exercise 9.1. Show that the bundle isomorphism g: TM - T* M induced
by the metric on a Riemannian manifold extends to an isomorphism of the
Clifford bundles Cl TM and Cl T* M. o
Definition 9.2. The Clifford bundle over a Riemannian manifold (M, g) is
the bundle of complex Clifford algebras Cl(M) - M generated by the cotan-
gent bundle T* M- M with g- 1 as its Euclidean structure. In other words,
Cl(M) :=Cl T* M. When convenient, weshall identify Cl(M) with Cl TM via
the isomorphism of the previous Exercise.
9.1 Clifford modules 371

If h isanother metric on M, then T := ft- 1 o g is an automorphism of the


tangent bundle and h(T(X),X) = X'(X) = g(X,X) > 0 for any nonzero
vector field X; therefore, T is positive definite with respect to the metric h
and thus has a positive definite square root, i.e., a bundle automorphism
a of TM satisfying h(a(X), a(X)) = h(T(X), X) = g(X, X). lf G' (M) pro-
visionally denotes the Clifford bundle formed from the Riemannian space
(M,h), then a(X) 2 = g(X,X) in G'(M); on applying Proposition 5.1 to
each fibre, a extends to a bundle isomorphism ä": Cl(M) - G' (M). We
conclude that the isomorphism dass of G(M) depends only on the mani-
fold M and not on the chosen metric.
Definition 9.3. A (left) Clifford module over a compact Riemannian mani-
fold (M,g) is a finitely generated projective right C(M)-module 'E = f(E),
corresponding to a complex vector bundle E - M, tagether with a C(M)-
linear homomorphism c: f(Cl(M))- [(EndE). In other words, 'Eisa B-A-
bimodule for A = C(M) and B = f(Cl(M)).
If E - M is endowed with a Hermitian pairing, so that 'Eis a pre-C*-
module over A, we say that the Clifford action is selfadjoint if c ( K) t = c ( K*)
for K E B, or equivalently, if c(oc)t = c(oc) for any real-valued oc E 5\ 1 (M);
that is, (s I c(K)t) = (c(K*)s I t) for s, t E 'E.
The obvious example of a Clifford module is the de Rham algebra of
differential forms .Jl• (M), with the Clifford action given by

c(oc)w:=OCI\W+L(OC~)W for OCE.Jl 1 (M), WE.Jl.(M), (9.1)

or, more briefly, c(oc) := E(oc) + t(oc~) where E denotes exterior multipli-
cation on the left and L denotes contraction by vector fields. Clearly each
c(oc) interchanges differential forms of even and odd degrees, and

so that c extends to a Clifford action of f(Cl(M)). This isanatural global-


ization of formulae (5.1), (5.2) and (5.3).
On each fibre (A"Ti M)c, the Clifford action of Cl( TiM) is reducible,
by what was remarked in Section 5.3, and one might hope that the global
Clifford module should also decompose into a direct sum of several sub-
modules. Therefore we seek irreducible Clifford modules, where each fibre
be linearly isomorphic to a Fock space of dimension 2m if dimM = 2m
I

or 2m + 1. There is, however, a topological obstruction to the existence of


such an irreducible module, which we shall discuss in detail in the next
section.
There is a naturall2-grading of Clifford modules over an orientable ma-
nifold M, induced by the chirality elements of the fibres Gx(M). Indeed,
each (TiM,g; 1 ) is an oriented Euclidean vector space. The chirality ele-
ment y in f(Cl(M)) is locally given by y(x) := ( -i)m 9I(x) ... 9n (x), when
372 9. Commutative Geometries

dimM = n = Zm or Zm + 1, where {9}, ... ,9n} is any local oriented or-


thonormal basis of 1-forms. (By the way, since we can suppose this local
basis to be smooth, it follows that ;y E f"" ( Cl(M)) .) Recall from Section 5.1
that ;y = ;y* and y 2 = 1.
Definition 9.4. On any selfadjoint Clifford module 'E over an even-dimen-
sional manifold, the operator c(y) is thus a grading operator (i.e., it is
formally selfadjoint and involutive). Thus 'E = 'E+ Eil 'E- where 'E± is the
(±1)-eigenspace of c(y). The algebra A = C(M) acts evenly on 'E, so that
'E± = [(E±) where E = PEilE- (Whitney sum) is a ;l2-graded vector bundle;
on the other hand, since a;y = -;ya for a E .J\ 1 (M), each c(a) anticom-
mutes with c(y) and so it interchanges 'E+ and 'E-. Therefore c(K)'E± c;; 'E±
for K E [((l+(M)), while c(,\)'E± <;; 'E"' for ,\ E [(Cr(M)); that is to say,
the Clifford action is a graded action.
However, if the dimension of M is odd, it is natural to consider modules
under the Clifford action of the even subalgebra [(O+(M)) only, whereas
;y is a section of o- (M). We can extend any such action of even sections
to the whole of [( Cl(M)) by the following device (already hinted at, in the
algebraic setting, in Section 5.3): for ,\ E [(0- (M) ), the section .\.y is even,
so we can define
c(.\.)!JI := c(.\.y)!JI. (9.2)
Since y 2 = 1, this is consistent provided c(y)IJI = c(y 2 )!JI = tJI always;
therefore, c(y) must act trivially on 'E. To sum up: in odd dimensions, the
grading is effectively lost, and we shall always consider ungraded Clifford
modules when dim M is odd.

9.2 Spine structures: the algebraic way


Let us first suppose that M has even dimension n = Zm. A Clifford module
'E for B := [((l(M)) is at the sametime a (right) module for A = C(M),
therefore a B-A-bimodule. lt is natural to ask whether a suitable bimodule
S can be found that can implement a Morita equivalence between the C*-
algebras A and B.
We thus need a full right C* A-module S = [(S) such that End~ (S) ""
[(Cl(M)). This means that there is a vector bundle isomorphism EndS""
Cl(M). Now, by Lemma 5.5, Cl(M) is a bundle of simple matrix algebras
whose rank is zn. If S is a vector bundle of rank N over M, then EndS
has rank N 2, so that N 2 = zn and consequently N = zm. We then re-
quire that the action of Ox(M) = 0 T; Mon the fibre Sx be given by its
zm-dimensional representation, which is irreducible; we say that S is an
irreducible Clifford module.
lt is now useful to recall the concept of a continuous field of Hilbert spaces
over M: this a family of Hilbert spaces g := {Ex : x E M} together with a
9.2 Spine structures: the algebraic way 3 73

given set ß(~) of continuous sections, satisfying conditions (a-c) of Defini-


tion 2.9.
Lemma 9.1. The set ~ := { Sx : x E M} is a locally trivial continuous field of
Hilbert spaces, whose space of continuous sections is S itself.

Proof. Each fibre Sx of the vector bundle S - M carries a scalar product


coming from the Hermitian pairing on the C*-module S, namely, (rx lsx) :=
(r I s) (x). Indeed, if x E M, the set is { Sx : s E S} the vector space Sx. If
r,s ES, thenx .... (rxlsx) isjustthe continuous function (rls) in C(M). The
local triviality of the vector bundle means that each point z E M has a neigh-
bourhood U suchthat the restriction of S to U is trivial, i.e., there is a bundle
isomorphism T: Slu....::... Sz x U; the corresponding map I= { Tx : x E U}
trivializes ~ over U, since it carries Slu = f(U,S) into C(U-Sz). Finally, if
a general section {Yx} xEM lies in the local uniform closure of S, it can be
approximately uniformly on compact subsets of such neighbourhoods U
by elements of C(U-Sz), so that it liesinS also. In summary, ~ fulfils the
three requirements to be a continuous field of Hilbert spaces. D
Definition 9.5. An elementary C* -algebra Ais one that is isomorphic to an
algebra of all compact operators over some Hilbert space that may be either
finite-dimensional or infinite dimensional and separable; thus A == MN ( <[)
for some NE ~ or else A ==X. Ifwe do not wish to discriminate on grounds
of dimension, we may say that an elementary C*-algebra is simple, that
it contains projectors of rank one, and that any two rank-one projectors
p,q E Aare related by q = vpv* for some v E A.
Since Clx(M) == EndSx for each x E M, the fibres of the Clifford bundle
form a (locally trivial) continuous field of elementary C* -algebras over M,
which we denote by Cl(M). The space of continuous sections of this field
is just the C*-algebra f(Cl(M) ).
Next, consider the alternative case that M has odd dimension n = 2m+ 1.
The fibres of Cl(M) are no langer simple algebras, since Ox M == MN ( () Eil
MN ( C) where N = 2m. Thus, to get a field of elementary C* -algebras we
shall take the subbundle of even Clifford subalgebras o+ (M), whose fibres
are OxM := o+ TiM == MN(<[); the corresponding field of elementary
C*-algebras is written o+ (M). In this case, we seek a Morita equivalence
between A := C(M) and B :=[(Cl+ (M)), via a bimodule S on whose fibres
o+ (M) acts irreducibly.
To unify the two cases as far as possible, we shall use the notation
o(+l (M) to mean O(M) or o+ (M) according as dimM is even or odd.
Thus o(+l (M) is always a continuous field of elementary C*-algebras .
..,. Locally trivial continuous fields of elementary C*-algebras over M are
classified, up to isomorphism, by a cohomological invariant called the Dix-
mier-Douady dass [137, Chap. 10]. (An isomorphism of two such fields A
and JI is a family of C* -algebra isomorphisms cf> = {cf> x: Ax ....::... Bx} that
374 9. Commutative Geometries

matches their algebras of continuous sections.) The construction of this


invariant by Cech cohomology follows the procedures setforthin the proof
of Theorem 1.13. Weshall now go through it in some detail.
Proposition 9.2. Let !1. = {Bx : x E M } be a locally trivial continuous field
of elementary C* ·algebras over M. Then !1. determines a cohomology dass
8 (!1.) E H 3 (M, Z), and another such field !1.' is isomorphic to !1. if and only if
8 (!1.' ) = 8 (!1.).
Proof. Let r denote the space of continuous sections of !1.. If x E M, there
is a neighbourhood U of x for which !l.lu is trivial; therefore, if Px is a
projector of rank one in Bx. there is a family of projectors of rank one
{ Py : y E U} in flu that contains Px· Now let Hy := ByPy for y E U.
These left ideals become Hilbert spaces under the scalar product

(aypy I bypy) := Tr(pya;bypy),

which is easily seentobe well defined. The ketbra laypy)(bypyl takes


CyPy to aypyb;cypy. Now since By is elementary, it is generated by its
operatorsofrank one, which are all ofthe form aypyb;; thus X(Hy) ""By.
If we write H for this continuous field of Hilbert spaces and X(H) for
the continuous field of elementary C*-algebras determined by H, we have
shown that !l.lu ""X(H).
Now we can cover the locally compact manifold M with an open cov-
ering {u1 }, where in each u1 we can find a continuous field of Hilbert
spaces H 1 satisfying !l.luJ ""X(H). The question is whether we can patch
them tagether to a global field of Hilbert spaces with the same property.
Suppose we choose a particular isomorphism fl_1 : X (H1 ) - !1.1 uj for each j;
then fl.i 1fl.1 : X(H 1 ) - X(Hi) is an isomorphism over each nonempty
Ui n u1 . By refining the covering if necessary, we can find continuous fam-
ilies of unitary isomorphisms '!!iJ: H 1 - Hi, defined over Ui n UJ. so that
Ad ('!!iJ) = '!!:.iJ ( ·) '!!i/ equals fl.i 1fl.1 in its domain.
Clearly, Ad('!!i) Ad('!!Jk) = Ad('!!ik) over each Ui n UJ n Uk. Therefore

(9.3)

for some continuous maps !liJk: Ui n u1 n Uk - lr. Finally, the relation

' '-1'
ß.Jkl ßikl ßiJL = ('!!:.Jt-1 '!!:.Jk '!!:.kt )( '!!:.kt-1 '!!:.ik
-1 )( -1
'!!:.a '!!:.a '!!:.iJ '!!:.JL
)

-1 -1 -1 ' -1 '
= '!!:.jt '!!:.jk '!!:.ik '!!:.ij '!!:.jl = '!!:.jt '!!:.jk ß.ijk '!!:.ij '!!:.ij '!!:.jl = ß.ijk

holds, over each Ui n u1 n Uk n Uz. This says that !1 = {fliJd is a Cech


2-cocycle with values in lr. Several choices have been made to obtain {l,
namely the covering {u1 }, the isomorphisms {fl.1 } and the unitary maps
{'!!:.iJ }. If we begin again with alternative choices, we reach a new 2-cocycle
!1'; we leave the reader to check (see Exercise 9.2 below) that !1 and !1' differ
9.2 Spine structures: the algebraic way 375

by a Cech 2-coboundary. Thus [~] E H 2 (M I) is well-defined and depends


1

only on the original datum ]i.


Just as in the proof of Theorem 1.13 from the short exact sequence of
1

sheaves:
0 - 7L ..3!!..!:.. 2rri!Rl. ~ l f - 0
- - I

we extract the isomorphism H 2 (M I) ""H 3 (M 7L) -compare (1.9). We de-


1 1

note by 8 (Ji) the image of [~] in H 3 (MI 7L); this is the Dixmier-Douady dass
of fi.
If cp: Ii - Ji' is an isomorphism of continuous fields of C* -algebrasl we
can construct this invariant for Ji' by replacing {~i} with {cp o ~i}; thusl the
family Qj 1 ~i is unchanged and we condude that 8(Ji') = 8(Ji). We leave the
reader to establish the conversel after we investigate the case 8 (Ji) = 0. D

Exercise 9.2. Show that the dass [~] E H 2 (M I) is independent of the


1

choices of {Uj} {~i} and {:!fij} made in the previous proof.


I 0
Theorem 9.3 (Plymen). Let M be a compact Riemannian manifo/d. Then
the C* -algebras C (M) and r (((] <+) (M)) are Morita-equiva/ent if and only if
8(((]<+) (M)) = 0 in H 3 (M 7L).1

Proof. If C(M) ~ nn<+) (M) )I then the equivalence bimodule S determines


a locally trivial continuous field of Hilbert spaces $_ = { Sx : x E M} for
which X($_) = ((] <+) (M); hence the {~i} of the previous proof may be taken
tobe restrictions of a Single isomorphism ~:X($_) - n<+)(M). Thus we
may take each {:!fij} to be the identity map and each ßiik} to be the con-
stant function 1 so that 8(((](+) (M)) = 0.
1

Converselyl if 8 ( ((] <+) (M))= 01 we can find an open cover {Ui} of M and
maps {~i} I {:!fij} and {~ijk} as abovel so that [~] = 0 in H 2 (MI Il. Therefore
~ is a 2-coboundaryl i.e. there are continuous functions Y.ij: Ui n Ui - lf
1

suchthat ~iik = Y.ij Y.jk Y.il on each Ui n Ui n Uk. Now J:!..ij := Y.i/ :!fij defines
unitary maps J:!..ij: Hi - Hi over Ui n Uj1 which satisfy J:!..ij J:!..jk = J:!..ik·
This means that these J:!..ij are transition functions that allow to patch
tagether the local fields of Hilbert spaces over each Ui into a global field
$_ = {Sx: x E M }; in other words by redefining the Sx we may assume that
1

:!fij = id over each Ui n Ui. But thenl the families of isomorphisms ~i and
~i coincide on each Ui n Uj so they are all restrictions of a global isomor-
1

phism ~:.X($_) - n<+) (M). Passing to the spaces of continuous sectionsl


we obtain an isomorphism End~ (S) - r (((] <+) (M)) and Theorem 4.26 nowI

gives the desired Morita equivalence. D

Exercise 9.3. Now show that if 8 (Ji') = 8 (Ji) in H 3 (MI 7L) then the locally 1

trivial continuous fields of elementary C* -algebras Ji and Ji' are isomorphic.


0
376 9. Commutative Geometries

Definition 9.6. Let A and B be two Morita-equivalent C*-algebras. Consider


the set of isomorphism dasses of the C* B-A-bimodules 'E that implement
a Morita equivalence A ~ B; wederrote this set by Mrt(B,A).
When A = B, the isomorphism dasses of Morita selfequivalences of A
form a group, under the operation ['E] · ['E'] := ['E ®A 'E']. Weshall call this
the Picard group Pic(A) := Mrt(A,A).
The identity element of Pic(A) is [A], where A carries the obvious A-
bimodule structure; recall from Section 2.5 that End~ (A) = A. Moreover,
if Ais unital, then any dass in Pic(A) is represented by an A-bimodule 'E
suchthat A !::' End~ ('E) via an isomorphism that takes lA to lx; in view of
Proposition 3.9, 'Eis finitely generated and projective as a right A-module.
In particular, if A = C(M) for a compact space M, then Pic(A) coincides
with the Picard group of line bundle dasses discussed in Section 2.6. We
recall from Proposition 2.19 that Pic(C(M)) "'H 2 (M, 'l.) in that case.
The case A = C(M), B := f(G<+>(M)) possesses an important simplifi-
cation. The equivalence bimodules S constructed from locally trivial fields
of Hilbert spaces are finitely generated and projective A-modules; this can
be seen directly, by patehing finitely many local sections of trivial bun-
dles with a finite partition of unity, or indirectly, by invoking the Serre-
Swan theorem, since the construction also produces a vector bundle S
for which S = f(S). Because the C*-algebra A is unital (by compactness
of M), Proposition 3.9 tells us that the identity operator ls is A-compact,
so that the algebras End~0 (S), End~ (S) and EndA (S) all coincide. More gen-
erally, if 'E and :F are two finitely generated and projective A-modules,
then Hom~0 ('E,:f) = Hom~('E,:f) = HomA('E,:f) since each adjointable
A-module map is A-compact. We therefore may and shall omit the super-
script 0 from all such spaces of A-module maps.
Proposition 9.4. Let M be a compact Riemannian manifold, A := C(M) and
B := f(G<+\M)). Then the Picard group Pic(A) acts (on the right) transi-
tively and freely on Mrt(B, A).

Proo(. As we have seen, each dass in Mrt(B, A) is represented by a bimodule


of the form S = f(S) for some vector bundle S - M of rank 2m, where
dimM = 2m or 2m+ 1. These bimodules may be permuted by twisting by
an A-module .L of rank one, via the recipe of Section 4.5

Such an .L is a finitely generated projective A-module with a partner Z


satisfying .L ®A Z "' A; recall, from Section 2.6, that we may take Z to
be the dual A-module .[# = HomA(L,A). Then .L and .[# may be fitted
with compatible pairings to make them C* A-modules, so they become
selfequivalence A-bimodules, with [.L#] = [.L]- 1 in Pic(A). By Lemma 4.24,
T .... T®lL maps EndA(S) intoEndA(S®A.L) andisinverted by T' .... T' ®lL#,
9.2 Spine structures: the algebraic way 377

so that by Theorem 4.26,

and thus [S ®AL] E Mrt(B,A) also. From any A-module isomorphism


L ....::..z we obtain a B-A-bimodule isomorphism S ®AL....::.. S ®Al. Thus,

[S] · [L] := [S ®AL] (9.4)

is a well-defined right action of Pic(A) on Mrt(B, A).


Suppose that S' isanother equivalenee bimodule yielding A ~ B. Then it
defines a loeally trivial eontinuous field of Hilbert spaees ~, over M, together
with an isomorphism !!.':X(~') - Cl(M), as in the proof of Theorem 9.3.
We ean then ehoose loeal isomorphisms between the fields ~ and ~' that
intertwine the aetions of Cl (+l (M). Sinee the Clifford algebras 01+ l (M) aet
irreducibly on the Hilbert-spaee fibres, these loeal isomorphisms pateh to-
gether to yield a line bundle whose seetions form HomB ( S, S'), the spaee of
adjointable maps of left B-modules T: S- S'. This spaee is an A-bimodule
via (a1 Ta2)(!/J) := T{!fJada2, for a1, a2 E A. Now if R E HomB(S', S), then
RT E EndB(S), and (a' R) (Ta")(!fJ) = R (T(!fJ)a' a"), for a', a" E A. There-
fore the map T®R- RT and the faet that [S] E Mrt(B,A) give A-bimodule
isomorphisms

HomB(S, S') ®A HomB(S', S) ""EndB(S) ""A, (9.5)

whieh shows that HomB(S, S') is a rank-one A-module. Moreover, the map

S ®A HomB(S, S')- S': !fJ ® T- T(!fJ)

is a B-A-bimodule homomorphism, beeause !fJ®aT evaluates to T(!fJa) and


b(T(!fJ)) = T{b!fJ) for bEB. Indeed, it is an isomorphism: by eonsidering
the analogaus homomorphism from S' ®AHomB(S', S) toS and using (9.5),
it is easily seen to be bijeetive.
By taking L = HomB ( S', S) in (9.4), we find that the aetion of Pie (A)
on Mrt(B, A) is transitive. Similarly, if [L] E Pie(A), there is a eanonieal
isomorphism cp: L - HomB(S, S ®AL) given by cp(t): !fJ- !fJ ® t. Thus, if
S®AL"" S®AL' asB-A-bimodules, thenHomB(S, S®AL) ""HomB(S, S®AL')
and eonsequently L "" L' as A-modules. This means that the aetion of
Pie(A) on Mrt(B,A) is free. D

The result of the Proposition ean be summarized by saying that Mrt(B, A)


is a principal homogeneaus space for the group Pie (C (M)) = H 2 (M, ~), that
is to say, it is a eopy of this group but with no distinguished basepoint .
.,. 1t often happens that when a eohomology dass of some strueture is
guaranteed to be zero, a new dass of lower degree appears that is an in-
variant for that strueture. In the present situation, the Morita equivalenee
378 9. Commutative Geometries

of certain algebras is signalled by the vanishing of a third-degree dass, so


we may expect to find a second-degree characteristic dass of this pair of
algebras. Such a dass has been identified by Plymen [377], whose lead we
follow.
Let Ji = <Cl<+>(M) and suppose that c5(Ji) = 0. Given any B-A-bimodule
S with [S] E Mrt(B,A), let 5# := HomA(S,A) be its dual A-module. Since
HomA(S,A) :=Horn~ (S,A), any element of s# is a "bra operator" (1/11: cf> .....
( !fJ I cf>). There is a natural Clifford action of B on S#, given by

b <!Jll := (1/11 o x<b'). (9.6)

where b ..... x(b') is the canonical A-linear antiautomorphism of f(<Cl(M))


that extends the sign reversal cx ..... -cx of .Jt 1 (M); here x = B-1 is the Bo-
goliubov automorphism that supplies the l2-grading of Cl(M) -see Sec-
tion 5.1. [Notice that the presence of x in (9.6) is moot when M is odd-
dimensional, since x = 1 on f( <Cl+ (M)) .]
With this protocol, S# becomes a B-A-bimodule. Therefore it is of the
form S# = f(S'), where S' .... M is again a bundle of rank 2m; consequently,
[S#] E Mrt(B,A). Let T := HomB(S#, S); then T is arank-oneA-module for
which S# ®AT:::::: S. If [L] E Pic(A), we find that
(S ®AL)# ®A (T ®AL ®AL) :::::< (S# ®AL#) ®A (L ®AT ®AL)
:::::: s# ® A T ® A L :::::: s ® A L,

where the commutativity of the group H 2(M ,l) has been used. Thus the
action [S] ..... [S ®AL] induces the transformation [T] ..... [T ®AL ®AL] on
H 2(M,l). Let j*: H 2(M,l) .... H 2(M,l 2) be the homomorphism of "mod-2
reduction" that is obtained from the exact sequence of abelian groups
2 j
0 -+ l -+ l -+ l2 -+ 0

(where '2' derrotes multiplication by two) and the corresponding long exact
sequence in Cech cohomology:

-- H 1 (M,l2) ~ H 2(M,l) ~ H 2(M,l).:!.!.. H 2(M,l2) ~ H 3 (M,l)--

where 2*[L] := [L ®AL]. Therefore, the action [S] ..... [S] · [L] passes to
the sum [T] ..... [T] + 2*[L] in H 2(M,l), so the dass K(!i) := j*(T) E
H 2(M,l 2) is independent of S, and depends only on the C*-algebra fieldJi.
Definition 9.7. Let M be an orientable compact manifold satisfying the con-
dition c5 (<Cl<+> (M)) = 0. A pair (v, S), where v is a specific orientation on M
and S is a Morita equivalence bimodule for A = C (M) and B = f( <Cl<+> (M)),
is called a spinc structure on M. If such a pair is given, we refer to M as a
spinc manifold.
If the orientation v of M is fixed, we refer to S alone as the spinc struc-
ture; two such, S and S', are called equivalent if they are isomorphic as
9.2 Spine structures: the algebraic way 379

equivalence B-A-bimodules; by Proposition 9.4, these equivalence classes


of spine structures are parametrized by H 2 (M.~).
This definition of a spine structure is due to Karrer [273, 274] andin-
dependently to Connes; Plymen [377] was the first to put in print Connes'
ideas in this respect. The idea of supplementing the orientation in this way
goes back to Atiyah, Bott and Shapiro [15], who defined a K -orientation on a
manifold M as an orientation plus a suitable extra ingredient, which turns
out to be a spine structure as usually described in books on differential
geometry (see [314, App. D], for instance). One may also refer to the pair
(v, S) as a K-orientation, as is done by Connes in [91, 11.6.;y].
Negation in the abelian group H 2 (M.~) corresponds to reversing the
orientation: -[ (v, S)] = [ ( -v, S) ]. The effect of reversing the orientation is
to change the chirality element ;y to -;y. Thus, if dimM is even, the grading
ofthe spinor module is changed to its opposite: ( -v, s+ $S-) - (v, s- $S+ );
while if dimM is odd, the Clifford action of ,\ E [ ( ( r (M)) is changed from
c(.\) to -c(.\), on account of (9.2).
Fixaspine structure (v, S) on a compact spine manifold M. The relation
S = [(S) defines the corresponding spinor bundle S - M. By Proposi-
tion 9.4, any other spine structure is of the form (v, S 181A L) where L = [(L)
is a rank-one A-module and L - M is a complex line bundle. By Proposi-
tion 2.6, S 181A L = [(S 181 L), so that the twisting of A-modules S - S 181A L
corresponds to the twisting ofvectorbundles S- S181L. More generally, one
can twist by any vector bundle F; if :f = [(F), one forms S 181A :f = f(S 181 F).
This is still a left B-module under the obvious Clifford action:

c(v)(f./J 181 t) := c(v)(/J 181 t, (9.7)

if v E B, f./J E S, t E ;:. In fact, any Clifford module over M is of this form.


Proposition 9.5. Any selfadjoint Clifford module over a compact spinc ma-
nifold with prescribed spinc structure (v, S) is of the form 'E = S 181 A :f for
some finitely generated projective A-module :f.
Proof Let :f := HomB(S, 'E). Since S and 'E are A-modules, so also is ;:,
under the action (at)((/J) := t((/Ja) fort E ;:, a E A, tJ1 E S. Now tJ1 181 t -
t (f./J) is a well-defined and bijective map from S 181 AJ to 'E, carrying c (v) f./J 181 t
to t(c(v)(/J) = c(v)(t((/J)) if v E B. The action (9.7) on S 181A :f makes this
correspondence a B-A-bimodule map. D
Exercise 9.4. Show that EndB (S 181 A :f) "" EndA (:f), by suitably adapting
Lemma 4.24 to the present case. 0

.,. Our goal is to construct Dirac operators on spine manifolds; we take up


this task in the next section. However, in order to construct a globally well
defined operator, having a spine structure does not suffice. It turns out
that the obstruction to such a global construction is precisely the invariant
380 9. Commutative Geometries

K(.!J.). When it vanishes, we find another ingredient for our toolkit, namely,
a conjugation operator on the spinor module.
Theorem 9.6. Let M be a spinc manifold, A := C(M), B := rco<+l (M) ). The
dass K (JJ.) in H 2 (M, 1l.2) vanishes if and only if at least one Morita equivalence
B-A -bimodule S admits a bijective antilinear map C: S - S satisfying the
following conditions:

C(I.Jia) = (CI.JI) ä {or a E A, (9.8a)


C(bi.JI) = X(b) CI.JI for bEB, (9.8b)
(Ccp I CI.JI) = (I.JII cp) {or cp, 1.J1 ES. (9.8c)

Proof. The condition K(.!J.) = 0 means that if S is any equivalence B-A-


bimodule and 'I':= Hom8 (S#, S), then j* ['I'] = 0, so that ['I']+ 2* [L] = 0
for some [L] E Pic(A). In that case, S ~ s# ®A 'I' ~ S# ®A J:# ®A J:#, so
that S ®AL ~ 5# ®A J:#. Replacing (if necessary) the original module S by
S ®AL, we see that K(JJ.) = 0 if and only if S "" 5# for a suitable equivalence
B-A-bimodule S, where the action (9.6) on 5# is understood.
The "Riesz theorem" of Section 4.5 identifies the conjugate A-module
S with Horn~ (S, A) = S#, where r.jJ E S corresponds to ( I.JII E 5#. As a
C* A-module, S carries the hermitian pairing (r.jJ I cjJ) := (cp II.JI), so any
B-A-bimodule isomorphism T: S# - S can be composed with the maps
1.J1 .... r.jJ .... (I.JII to yield an antilinear bijection C: S- S: 1.J1 .... T(I.JII.
The commutativity of the algebra A yields

(I.Jia I cp) = ä(I.JII cp) = (I.JII cp)ä = (I.JII cpä),

that is to say, (I.Jial = (I.JII ä (no surprise there), and since T is an A-module
map, (9.8a) follows. Moreover, if bEB, then (bi.JII cp) = (I.JII b*cp) -recall
that B "" EndA (S)- and so

(bi.JII = (I.JII 0 b* = X(b) (I.JII,


where we invoke (9.6) for the second equality. Since T is a left B-module
map, this is equivalent to (9.8b).
Conversely, if Cis a given antilinear bijection satisfying (9.8a) and (9.8b),
then T(I.JII := C(I.JI) defines a B-A-bimodule isomorphism from S onto S#,
showing that K(JJ.) = 0.
It remains to show that such a C can be made an isometry by a suit-
able normalization. Let ct be the (antilinear) adjoint of C, defined, as in
Section 5.4, by (cp 1 cti.JI) := (I.JII Ccp). Then, if bEB and cp, 1.J1 ES,

(c/J 1 ctcbc- 1 cc- 1 )ti.JI) = (Cbc- 1 (c- 1 )ti.JII Cc/J) = <x<i1Hc- 1 )ti.JII Cc/J)
= ccc- 1 )ti.JII x<b')CcfJ) = cc- 1x<b')CcfJ II.JI)
= (b*cp II.JI) = (cp I bi.JI),
9.2 Spine structures: the algebraic way 381

so that (ctc)b = b(ctc) for all b E B. Thus ctc E End 8 (S) "" A, and
therefore ct C = a 1s where a is a positive invertible element of A. (This
uses the irreducibility of S as a Clifford module, since EndB(S) "" A is
a version of Schur's lemma). If we now replace C by a-112c, we thereby
normalizesothatctc = 15 .Inthatcase, (C</>IC!JI) = (!JJICtC</>) = (1/JI</>),
as daimed. 0

Lemma 9.7. Any conjugation C satisfying (9.8) has square C2 = ±1.

Proof. Since X(b) = Cbc- 1 for all b E B, and b ..... x(b) is involutive,
it follows that C 2 bc- 2 = b, so that C2 E EndB(S) "" A = C(M). Write
C2 = u 1s with u E C(M); then (Ct) 2 = ü 1s, so that lul 2 = (Ct) 2 C2 =
(ctc) 2 = 1 in C(M).
Moreover, uC = C3 =Cu= üC, so that ü = u and therefore u 2 = 1 in
C(M). Since we are assuming that M is connected, this implies that u =
±1c(M}· 0

Even so, the choice of C is still not canonical, since we can replace it by
,\C for any scalar function ,\ E C(M- T); indeed, by antilinearity, (,\C) 2 =
.\C.\C = AÄC 2 = c 2 and (,\C)t(,\C) = ,\,\ctc = ctc = 1s, for any such.\.
Weshall regard ,\C and C as equivalent conjugations on S.
Definition 9.8. A spin structure on an orientable compact manifold M is a
triple (v, S, C) where (v, S) is a spinc structure on M and Cis an antilinear
conjugation on S satisfying (9.8). If M carries such spin structure, we refer
to M as a spin manifold.
Exercise 9.5. If K(Ji) = 0, show that the various spin structures on Mare
dassified by H 1 (M, 7L. 2 ). o
What we have donein this section, following Plymen [3 77], is to introduce
spinc and spin structures directly from the noncommutative point of view,
based on the notion of Morita equivalence. There is, to be sure, a standard
"commutative" treatment of these structures in books on differential geo-
metry -that of [314] is quite comprehensive- starting from the theory of
principal bundles. The paper [377] also contains a comparison of these ap-
proaches, showing that a manifold M carries a spinc structure, in the sense
described here, if and only if the structure group SO(n) of the tangent
bundle can be lifted to the group Spinc(n) = Spinc(~n). The obstruction
to doing that is a cohomology dass W3(M) in H 3 (M,7l.), which turnsout
tobe equal to 8(1(]<+l(M)). For spin structures, the lifted structure group
is Spin(n) = Spin(~n), and the obstruction to achieving that is the second
Stiefel-Whitney dass w 2 (M) E H 2 (M, 7L. 2 ), which is the mod-2 reduction of
W3 (M) and can be identified with K(G<+l(M)).
We are happy to report that the program of relating integral Cech co-
homology groups of a manifold M to C*-algebraic invariants, begun with
Theorem 1.13 and Proposition 2.19, has arrived so far.
382 9. Commutative Geometries

We remark also that a parallel "bosonic" theory exists for Mpc and Mp
(metaplectic) structures on symplectic vector bundles [400), with the Clif-
ford bundle replaced by the "Weyl bundle" [376].

9.3 Spin Connections and Dirac operators


We return to the smooth category in order to address the subject of con-
nections. Wehaveseen in Section 7.1 that the Levi-Civita connection is the
unique torsion-free metric connection VB on X(M), and it determines by
duality (7.10) a Levi-Civita connection, also denoted by VB, on .J\ 1 (M).
The algebra f"" (l[l(M)) is generated by .J\ 1 (M). The connection VB on
.J\ 1 (M) now extends to all of f"" ( 1[1 (M)) via a recursive application of the
Leibniz rule:

VB(v,\) := (VBv),\ + v(VB,\) for v,,\ E f""(l[l(M)). (9.9)

This yields a connection VB: f"" (l[l(M)) - .J\ 1 (M, l[l(M) ), once again called
a Levi-Civita connection. The Clifford product in f"" (l[l(M)) extends in an
obvious way to allow us to multiply elements of f"" (l[l(M)) and elements
of .Jl 1 (M,O(M)), so that the right hand side of (9.9) makes sense. After
contracting with any vector field X E X(M), this equation implies that the
operator V~:= tx o VB+ VB o tx is a derivation of the algebra f""(l[l(M)).
We may express the skewsymmetry (7.11) of the Riemannian curvature
by saying that R(X, Y), for any pair of vector fields X, Y, is a section of the
bundle so(TM) - M of g-skewsymmetric endomorphisms of TM. We also
write so(T* M) to denote the g- 1 -skewsymmetric elements of End(T* M).
If U is the domain of a chart that trivializes T* M, we can choose a local
orthonormal basis of 1-forms (9 1 , •.• ,9n) over U and express the Levi-
Civita connection on .J\ 1 (M) by the following analogue of (7.13):

VB 9ß = _[ß dxi
[()(
® 9 01 '

where rfot are the corresponding Christoffel symbols. (We use Latin letters
for coordinate basis indices and Greek letters for orthonormal basis in-
dices.) The metric compatibility of VB is expressed by the relations

that is, f{p + rfot = 0. Thus [fot are components of a skewsymmetric matrix.
We summarize this by writing

SO that VB = d - f over the chart domain U.


9.3 Spin Connections and Dirac operators 383

By transposing with respect to the metric, the curvature tensor R also


acts on sections of T*M, with components Rr 5 kz := grigsJRiJkl = g 5 JRjk 1;
with this understanding, we write R E .J\. 2 (M, so(T* M) ).
The local expression VB = d- f implies that V~ = X- f(X) for any
X E X(M); therefore,

R(X, Y) =[V~, V}]- Vfx.YJ =[X- f(X), Y- f(Y)]- [X, Y] + f([X, Y])
= -X(f(Y)) + Y(f(X)) + f([X, Y]) + [f(X),f(Y)]
(9.10)

ThusR = -df+f Afin.Jl 2 (U,so(T*M)).


lt is also worthwhile to recall that the metric compatibility property (7.8)
of VB holds because the Christoffel symbols are real-valued. Therefore, VB
is a Hermitian connection in the sense of Definition 8.9 .
.,... Now let us assume that M is orientable and gifted with a spin struc-
ture (v, S, C), where S is an irreducible Clifford module of the form f(S).
Weshall commit a slight abuse of language and refer to the .Jl-module of
smooth sections S := roo (S) as the carrier of the spin structure, instead of
the module of all continuous sections. The algebra 'B := roo (Cl(M)) acts
on S -bearing (9.2) in mind if dimM is odd- so that S is a 'B-.Jl-bimodule.
As such, it carries a particularly important connection, called its spin con-
nection, whose existence is given by the following theorem.

Theorem 9.8. Let M be a spin manifold with spin structure (v, S, C). There
is a unique Hermitian connection V 5 : S - .J\. 1 (M, S) that commutes with C
and satisfies the following Leibniz rufe:

(9.lla)

that can be written more succinctly as a commutation relation:

[V 5 ,c(v)] = c(VBv) for v E 'B. (9.llb)

Proof. We first establish existence of V5 locally. By restricting to a chart do-


main U, we mayreplace .Jl, 'Band Sby C0 (U), C0 (U, Cl( !Rn)) and CQ'(U,F)
respectively, where F is an irreducible representation space for Cl( !Rn).
Then VB= d-f on U, withf E .J\. 1 (U)®so(!Rn ). Nowf acts by derivations on
CQ' (U, Cl( !Rn)): that is, the obvious map oc - f oc on .J\. 1 (M) extends to Clif-
()(1 ..• OCk) := Ij:1 ()(1 ... r ()( j ... OCk.
. "' k "'
ford products of 1-forms by settmg [(
Now apply the infinitesimal spin representation p: so(!Rn)- EndF to f,
to obtain
V5 := d- p(f) E .Jl 1 (U,EndS),
384 9. Commutative Geometries

that is, V 5 (/J := d(/J- p(f)(/J for 1/J E f""(U,S). lf v E f""(U,I(](M)), then
the commutation property (5.29) entails that

V 5 (c(v)(/J) = d(c(v)(/J)- p(f) c(v)(/J


= c(dv)(/J + c(v) d!fJ- [p(f), c(v)]I/J- c(v)p(f)(/J
= (c(dv)- c(fv))!Jl + c(v)(d!Jl- p(f)!Jl)
= c(VBv)(/J + c(v) V 5 !Jl.

Next, observe that V~ E f0°(U,EndS) commutes with C, for any real


vector field X. Indeed, in view of (5.12), the p(f) part yields a sum of terms
of the form - ~ L.a,ß Xi[f01 c ( 9 01 ) c (9ß), and it is enough to note that C com-
mutes with c(9 01 ) c(9ß) by (9.8b) and that the Christoffel symbols 01 are rf
real.
Suppose that V' isanother C-invariant Hermitian connection onf"" (U, S)
satisfying (9.11). If XE X(M, IR(), each V~- Vx commutes with each c(v), so
it is a multiplication by a scalar. Hermiticity (8.27) implies that this scalar is
purely imaginary, while C-invariance implies that it is purely real; therefore,
it must be zero. This argument establishes the uniqueness of V5 locally.
But it also means that the locally defined connections must coincide on
spinors supported on the overlap of two charts, so that V5 can be globally
defined while maintaining uniqueness. 0

Proposition 9.9. The curvature ofthe spin connection V 5 is {1 (R), where R is


the Riemann curvature tensor, regarded as an element of .Jl 2 (M, so( T* M)).
Proof. The calculation (9.10) showed that R = -d[ + [ 1\ [ on any local chart
where VB = d- f. The curvature of V 5 is given by (X, Y) ..... [Vt Vf] -
Vfx.YJ· The same calculation (9.10), with f replaced by {l(f), shows that this
curvature equals -d{l(f) + {l(f) 1\ {l(f) = {l(R). 0

~ To write the spin connection in local coordinates, we must first express


the Clifford action locally. Fix a local chart with Coordinates x 1 , .•• , x» E
=
C"" (U), and let { }' 01 Ya : oc = 1, ... , n} be a fixed set of gamma matrices
-for instance, those provided by (8.20)- that are selfadjoint unitary ope-
rators on the Fock space for 1(](11((»), satisfying ;y 01 ;yß + ;yß;y 01 = 28 01 ß. The
metric is given by a real matrix-valued function G = [Bii] on U that is pos-
itive definite, so we can find another real matrix-valued function H = [hj]
suchthat HtH = G. Let H- 1 = [hß] be the inverse matrix; then

Then the Clifford action of 1-forms is locally given by

c(dxr) := hß ;yß. (9.12)


9.3 Spin connections and Dirac operators 385

This corresponds to choosing local orthonormal bases of 1-forms {9a} and


vector fields {Ep} by

(9.13)

so that g- 1 (9a, 9ß) = 8aß and g(Ea,Ep) = 8ap; it is clear that (9a)~ = Ea
and (Ep)' = 9ß. The ambiguity of such choices comes from that of the
matrix H; if G1 i 2 is the positive definite square root of G, we may take
H = AG 1 12 where A: U - SO (n) is any smooth function (we need det A = 1
so that the orthonormal bases (9.13) be oriented). This can be thought of as
expressing a smooth choice of complex structure on each TiM, in order to
specify the Fock space at each point of U on which (9.12) yields the Clifford
action. Weillustrate this "gauge freedom" in Section 9.A.
The spin connection is now expressed locally, using the formulas (5.28)
and (5.12) for the infinitesimal spin representation {.1, by

(9.14a)

or alternatively,
..,s
v il; =
:::~
Vi - Wi, wh ere Wi =
1 ~rß
4 ia Y Yß·
a (9.14b)

Exercise 9.6. Show that hj'tfa = -oi(h1) + [i~h~. Deduce the formula
1 k ß . l
Wi = 4(fi1Bkt- oi(hj)8aph 1 ) c(dx1 ) c(dx ),

that expresses the spin connection in a coordinate basis.


Definition 9.9. The Leibniz rule (9.11) can be stated for any selfadjoint Clif-
ford module. We say that a Hermitian connection V''E on a selfadjoint Clif-
ford module 'E is a Clifford connection if it satisfies the Leibniz rule

V''E (c(v)s) = c(V'Bv)s + c(v) V''E s, for v E 'B, s E 'E, (9.15a)

that can also be written as a commutation relation:

[V''E,c(v)] = c(V'Bv) for v E 'B. (9.15b)

Notice that if 'E = S ®51. J', for a given spin structure (v, S, C), and if
v:r is an arbitrary Hermitian connection on J', then the tensor product
connection V' 5 ® 1:r + 1s ® v:r -recall (8.25)- is a Clifford connection. In
fact, any Clifford connection on a spin manifold is of this form.
Proposition 9.10. Let 'E = S ®51. J' be a selfadjoint Clifford module over
a compact spin manifold with a given spin structure (v, S, C). If V'E is a
Clifford connection on 'E, then there is a unique Hermitian connection V' :r
on J' such that vx = \7 5 ® 1:r + 1s ® v:r.
386 9. Commutative Geometries

Proof. Choose any Hermitian connection V' on :J; then vx- (Vs ® 1.r +
1s ®V') lies in EndACE) ®YJ. .JI. 1 (M), and it commutes with the Clifford
action of 'Bon '.E, on account of (9.11) and (9.15). Since T .... 1s ® T is an
isomorphism from EndA(r> onto EndB('E), by Exercise 9.4, we can find
oc(X) E EndA(:f) = f(EndF) suchthat

vf- (V~® 1.r + 1s ®V~)= 1s ® i oc(X), for XE *(M),


and the hermiticity of the left hand side implies that the function oc(X) is
real-valued. Since the left hand side is tensorial in X, we end up with a real
1-form oc E .JI. 1 (M,EndF) suchthat vx = vs ® 1.r + 1s ®(V'+ ioc). Now
v.r := V' + ioc is the desired Hermitian connection on :f. D
We may summarize this result as follows: all Clifford connections on
finitely generated projective .JI.-modules are abtairred by twisting the spin
connection of the given spin structure.
In particular, if (v, S') isanother spinc structure on the spin manifold M,
then S' "" S ®Y1. .L for some line bundle .L, so that any Clifford connection
on S is given by V= vs + ioc where oc E .JI. 1 (M, ~). By writing oc = A 1 dx1,
this Clifford connection on S' is then expressed locally as
o
Vai = 1 - w 1 + iA1,

so that oc is a gauge potential. We remark that S'# "" s# ®yt .[# "" S ®yt .[# ""
S' ®yt .[# ®YJ. .[#, so that .[# is a "square root" of the rank one module
T = HomB(S'#, S').
Clifford connections may also be constructed on S' when it does not nec-
essarily arise by twisting a spin structure. For that, choose any Hermitian
connection v'l' on the associated rank one module '1 = f""(T). Locally,
we can write v'l' = d + ioc with oc E .J\. 1 (U, ~). In such a local chart, we
can find a suitable spin structure S = f"" (U, S) and a rank one module
.L = f""(U,L) so that S' = S ®c"'Wl .Land '1 = .L ®c"'Wl .L. Moreover, the
connection V' := d + !ioc on .L provides a square root for v 7 . Therefore
V:= vs ® 1L + 1s ®V', given by d- p([) + !ioc, is a connection defined
on f"" ( U, S'), and it is not hard to show that these local definitions are con-
sistent. Notice that there is a matehing Clifford connection on S' #, given
locally by d- p(f)- !ioc. In brief, the set of Clifford connections on S' is
an affine space, parametrized by .J1. 1 (M, T).
This local construction of "spinc connections" dependent on scalar gauge
potentials is sometimes described by saying that "while we cannot con-
struct the spinor bundle and we cannot construct [a square root of the
associated line bundle), we can construct their product" [314, p. 397). This
viewpoint has been exploited by Seiberg and Witten [429,488) in their work
on partial differential equations for monopoles .
.,.. We now come to the main object of spin geometry: the Dirac operator.
As we shall see in subsequent chapters, this operator carries the whole
9.3 Spin Connections and Dirac operators 387

weight of the geometrical structure when the theory is exarnined from the
noncommutative point of view.
Definition 9.10. Let 'E be a selfadjoint Clifford module over a compact Rie-
mannian manifold M. We can rewrite the Clifford action c: f(Cl(M)) -
EndA('E) as the operator c: f(([l(M)) ®..?t 'E- 'E given by
c(v ® s) := c(v)s. (9.16)
Let 'V be a Clifford connection on 'E. The generalized Dirac operator as-
sociated to the connection '11 and the Clifford action c is the composed
map
D := -i(c o v). (9.17)
The factor - i is introduced to make D formally selfadjoint rather than
skewadjoint (see Proposition 9.12 below). 1t is a ([-linear endomorphism of
'E; if j: .J\. 1 (M) ..... f( Cl(M)) is the inclusion map, D can be written as the
composition

'E ~ .J\. 1 (M) ®..?l 'E ~ f(Cl(M)) ®..?l 'E _!__. 'E.

Proposition 9.11. If D is a generalized Dirac operator on a selfadjoint Clif-


ford module 'E, and if a E CO" (M) is regarded as a multiplication operator
on 'E, then

[D, a] = -ic(da). (9.18)


Proof. Scalar multiplication by a commutes with the Clifford action on 'E,
so for any s E 'E we obtain
i[D, a]s = c('V (as)) - ac('ll s) = c('V (as) - a'V s) = c(da ® s) = c(da)s.
0

The generalized Dirac operator is easily expressed in local coordinates:


choose a local basis of 1-forms to express the Clifford action, and contract
'11 with the duallocal basis of vector fields; for instance,

Ds = -ic(dxi) 'llais = -iycx 'VEas.


Here ycx = c(91X) for some local orthonormal basis {9 1 , ... , 9n} of 1-forms,
and Ecx = (9cx) # is the duallocal orthonormal basis of vector fields.
When a spin structure is given, the (generalized) Dirac operator on the
spinor module is usually referred to as the Dirac operator.
Definition 9.11. Let M be a compact spin manifold, with prescribed spinor
module structure S, and let vs be the spin connection on S. Then the Dirac
operator on S is the operator I/) defined by
(9.19)
388 9. Commutative Geometries

Locally, we may express it by

l/}f.JJ = -i c(dxi) 'V~j f.JJ = -i y 01 vt, f.JJ. (9.20)

Since any other selfadjoint Clifford module and connection on M is given


by twisting, we see that the generalized Dirac operator on 1: = S ®Jt :J' is
of the form D = QJ ® 1.r - i c ® 'V J' .
.,.. The Dirac operator QJ on a spin manifold is constructed from the met-
ric g, via the Levi-Civita connection underlying 'V 5 and the Clifford action
determined by (T* M,g- 1 ), in the presence of the topological condition
K(Ji) = 0. Connes astutely noticed that, conversely, the Dirac operator de-
termines the metric [90]. Although this is to some extent a tautological
observation, it allows us to change our viewpoint by establishing l/) as the
main object of study.
Definition 9.12. Given two points p,q of a Riemannian manifold (M,g),
there exists a piecewise smooth curve connecting them. The distance be-
tween them is defined as the infimum of the lengths of piecewise smooth
curves running from p to q (the infimum being attained by a shortest curve
if and only if M is complete). The length -t' ( y) of such a curve y: [ 0, 1] - M
is given by
-t'(y) := J: ly(t)l dt := J: ~g(y(t), y(t)) dt.

Then the distance d ( p, q) = d 9 ( p, q) is defined as

d 9 (p,q) := inf{ -t'(y): y piecewise smooth, y(O) = p, y(1) = q }. (9.21)

Among the required properties of a distance, it is perhaps not entirely


clear that p * q implies d 9 (p,q) > 0. However, on a local chart around
p E M of Euclidean radius E, not containing q, the relation

Bii(x)vivi ~ 8iv1 2

holds for some 8 > 0. Therefore the distance from p to q is at least Eß.
The topology on M determined by d9 is the same as the original manifold
topology. The argument is basically the same: in each chart, there is an
Euclidean-distance ball containing a d 9 -distance ball, and conversely.
It is a basic fact of Riemannian geometry that the distance function d 9
determines the metric g. Since a smooth manifold can always be provided
with a metric, a precise way of formulating that statement is the Myers-
Steenrod theorem [353]. This says that if cp: (M, g) - (N, h) is a bijective
distance-preserving map, i.e., dh (cp(p), cp(q)) = d 9 (p, q) for p, q E M, then
cp is smooth and an isometry, that is, cp* h = g. Fora proof, see, for instance,
[372, Thm. 5.9.1].
Thus, it suffices to show that the Dirac operator determines the distance
function. In fact, what we do is to appeal to the Gelfand cofunctor once more
9.3 Spin Connections and Dirac operators 389

in order to redefine the distance d(p, q) in terms of (a dense subalgebra of)


the algebra C(M).
Definition 9.13. The spinor module S = f"" (S) is a prehilbert space under
the positive definite Hermitian form:

(c/> 11./J} := L(cp 11./J) lv9 1, (9.22)

where lv9 1 is the Riemannian density on M -compare (7.22). The com-


pleted Hilbert space L 2 (M, S) is called the space ofL 2 -spinors determined by
the spinor bundle S-M. Elements of C(M) will be regarded as (bounded)
multiplication operators on L2 (M, S).
Given a distance function d, a d-Lipschitz function a satisfies by definition
an inequality
la(x) -a(y)l ~ Cd(x,y)
for x, y E M and some positive constant C; the least possible C is the
d-Upschitz seminorm of a. Any Upschitz function is continuous. Sup-
pose now that a E C"" (M) and consider the Riemannian distance d 9 . If
y: [0, 1]- M is a piecewise smooth curve with y(O) = p, y(l) = q, then

a(q)- a(p) = a(y(1))- a(y(O)) = Jofl d


dta(y(t)) dt

= J: day(t)(y(t))dt = J: B:r(t)(grady(t)a,y(t))dt.
Applying the Schwarz inequality to the integrand yields the estimate

la(q) - a(p) I ~ J: lgy(t) (grady(t) a, y(t)) I dt ~ Jo1 Igrady(t) ally(t) I dt


~ II gradalloo s: ly(t)l dt = II gradalloo f(y). (9.23)

Thus II gradalloo ~ 1 implies la(q)- a(p)l ~ f(y) for any y, and hence
la(q)- a(p)l ~ d 9 (p,q). Therefore

sup{ la(p)- a(q)l: a E C""(M), II gradalloo ~ 1}

is at most d 9 (p,q).
The estimate (9.23) holds not only for smooth functions a but for any
(continuous) function a whose gradient is defined, almost everywhere, as
an essentially bounded measurable vector field. If we write the length of
a vector Zx E (TxM)c as IZxl := Bx(Zx, Zx) 112 , this means that the func-
tion x . . . I gradx a I should be measurable, with finite essential supremum
II gradalloo. To see that the supremum taken over all such functions sat-
isfying llgradalloo ~ 1 actually equals d 9 (p,q), we can use the function
390 9. Commutative Geometries

ap(x) := d 9 (p,x). This is not smooth at p and will become generally sin-
gular on the boundary of the segment domain of p (the "cut locus" of the
point p, or "caustic", in physicists' parlance); but in any case, ap is d 9 -
Upschitz with C = 1, due to the triangle inequality for the distance func-
tion (9.21), that is, ld9 (p,x)- d 9 (p,y)l :s; d 9 (x,y). We conclude that

d 9 (p,q) = sup{ la(p)- a(q)l: a E C(M), II gradalloo :s; 1 }. (9.24)

Now, the point is that the Upschitz subalgebra of C(M), and thus the
distance function itself, can be characterized in terms of the Dirac ope-
rator, by the condition that [D, a] be a bounded operator on L 2 (M, S) -
opening the way for defining Upschitz subalgebras ofnoncommutative C*-
algebras [480], after Definition 9.16.
Proposition 9.12. Let p and q be two points of a compact spin manifold M.
Then·

d(p,q) = sup{ la(p)- a(q)l: a E C(M), II[JP,a]ll :s; 1 }. (9.25)

Proof. The commutation relation (9.18) shows that II [1)>, a] II is the operator
norm of the Clifford action c (da) on the spinor space L 2 (M, S). This is given
by

llc(da)ll 2 = sup llc(da)(x)ll 2 = supg_; 1 (dä(x),da(x))


XEM XEM

= SUPBx(gradxä,gradxa> = llgradall~. (9.26)


XEM

and the result follows from (9.24). D

Exercise 9. 7. Show directly that the formula (9.2 5), regarded as a definition
of d(p, q), has the formal properties of a distance function: it is symmetric,
vanishes only when p = q, and satisfies the triangle inequality. <>

Example 9.1. The unit circle § 1 is a I-dimensional compact Riemannian


manifold. Elements of coo (§ 1 ) can be regarded as periodic functions j(O),
of period 1. The spinor module has rank one and lP = -idjdO, and so
[JP,J] = -if' as multiplication operators on, say, L2 [0, 1]. Now

lf(ß)- f(a) I = IJ: f' ((}) d(} I :5 lß- al if llf' lloo :5 1,

when a, ß E [0, 1]; thus, d(a, ß) = lß - al. The supremum is attained


by h 01 (6) := 16- al for (} E [a- !,a + !J, which is obviously Upschitz
with C = 1. The cut locus for a is the antipodal point of the circle with
coordinate a ± !. and (9.25) gives the arc length on the circle (of course) as
the distance function.
9.4 Analytical aspects of Dirac operators 391

The only property of the operator I/) used to prove Proposition 9.12 was
the commutation relation (9.18) that is valid for generalized Dirac operators
too. Thus (9.25) remains valid when I/) is replaced by any D satisfying (9.18).
In particular, on non-spin manifolds with a suitable D, we can still measure
distance between points in the noncommutative way.
The distance formula (9.25) suggests an important reformulation that
may be generalized to noncommutative algebras. Write la(p)- a(q)l =
IEp(a) - Eq(a)l; the characters Ep and Eq arepure states (i.e., states that
arenot convex combinations of other states) on the C*-algebra A = C(M).
Thus we could try to extend it to a distance function on the full state space.
A generalized Dirac operator may be used, as in the following exercise.
Exercise 9.8. Let D be a generalized Dirac operator on a Clifford module
over a compact Riemannian manifold M. Show that the following recipe
defines a distance function on the state space of C(M):

d(</>,1./J) := sup{ 1</>(a)- tfJ(a)l: a E C(M), II[D,a]ll ::;1 }. 0

The full state space of a unital C*-algebra Ais the closed convex hull of
the set of pure states, and is contained in the unit ball of the dual Banach
space of A. As such, it is compact in the weak* topology (see Section 1.2).
A very important question is whether the topology determined by the dis-
tance function of Exercise 9.8 will coincide with the weak* topology. Rieffel
has found [397] that this is the case for important examples, including the
noncommutative tori discussed in Chapter 12. He has also shown by simple
examples [398] that, even in the commutative case, the distance function
on the state space induced by (further suitably) generalized Dirac operators
is not determined in general by its restriction to the set of pure states.

9.4 Analytical aspects of Dirac operators


We want to show now that the Dirac operator I/), defined initially on S, ex-
tends to a selfadjoint (unbounded) operator on the Hilbert space L2 (M,S).
(As hinted earlier, we could replace the spinors by "half-densitized spinors",
i.e., decide to work on the Hilbert space completion of f"" (S) ® I.J\ 1112 , with-
out using the metric. This makes sense in noncommutative geometry when
one wants toseparate the "kinematical variables" -the algebra C"" (M) and
the spinors- from the "dynamical variables" -the Dirac operator- as in
prequantization of gravity. In this book, even so, we stick to the Hilbert
space of Definition 9.13. A fine study of the dependence of the Dirac ope-
rator on the metric is found in [43].)
The first step is to establish that D is a formally selfadjoint operator on
its original domain.
392 9. Commutative Geometries

Proposition 9.13. The Dirac operator I/) is {ormally selfadjoint, that is,

(QJ cf> I 1/J) = ( cf> I I/) 1/J) for all cf>, 1/J E S. (9.27)

Proof. By using a partition of unity, it suffices to verify (9.27) for the case
that cf>, 1/J E roo(S) are a pair of sections that vanish outside some chart
domain; thus we use the local expression I/)= -ic(dxi) V'f. J
Now

i(cf> I I/)1/J)- i(I/)cf> 11/J) = (cp I c(dxi)V'~J 1/J) + (c(dxi)V'fcf>


J
11/J)
= (cp I V'~J c(dxi)I/J)- (cf> I c(V'~ dxi)I/J) +(V'~ cf> I c(dxi)I/J)
J J

= a1 <cJ> I c(dxi)I/J)- <cf> I c(V'g dxi)I/J),


J

where we have used the selfadjointness of c(dxi), the hermiticity of the


spin connection, and the relation [V' 5 ,c(dxi)] = c(\i'B dxi). LetZ be the
vector field determined by

a(Z) := (cp I c(a)ljJ), for a E .Jl 1 (M).

Then

a1 ( cf> I c (dxi) 1/J) - ( cf> I c('vg dxi) 1/J) =


J
aJ (dxi (Z)) - (V'~ dxi )(Z)
J

= dxi (V'gJ Z) = div Z,

using the duality relation (7.10) for the Levi-Civita connection, as weil as
the divergence formula (7.18). The divergence theorem now yields

(I/Je/> 11/J)- (c/> I I/JI/J) = -i L (div Z) lv9 I = 0. D

Corollary 9.14. Any generalized Dirac operator on a selfadjoint Clifford


module is {ormally selfadjoint.

Proof. On replacing the Leibniz rule (9.11) for \7 5 by the more general ver-
sion (9.15), the proof of Proposition 9.13 applies, mutatis mutandis, to the
generalized case. D

Next, we observe that the Dirac operator I/) is an elliptic differential ope-
rator of order one. For that, we need merely Iook at the local formula

to realize that I/) is indeed a differential operator, of first order, with some
zeroth-order terms depending linearly on the Christoffel symbols rfc~ via
(9.14). The principal symbol a1 (QJ)(x, ~) is invariantly defined on T* M, so
to compute it we may use normal coordinates centred at Xo E M, whereby
9.4 Analytical aspects of Dirac operators 393

c(dxi) = yi at x = x 0 ; thus O"l(1,7))(x 01 ~) = ~iyi = c(~). Therefore the


principal symbol of 1,7) is given by

(9.28)

In particular I 0"1 (1,7)) 2 = c ( ~) 2 = g - 1 (~I ~) is a nonzero scalar and thus


0"1 (.1,7)) is invertiblel off the zero section ofT* M. This establishes ellipticity
of 1,7). The theory of pseudodifferential operators (see Section 7.A) shows
that 1,7) has a parametrix i.e. a pseudodifferential operator Q of order -1 on
1 1

the spinor bundle S suchthat R :=I -l,l)Q and T :=I- Ql,l) are smoothing
operators. Ellipticity is the key to the improvement of the formal selfad-
jointness of 1,7): we can now show that 1,7) has a unique selfadjoint extension.
Definition 9.14. Let A be an unbounded operator on a Hilbert space J-{
whose domain DomAis a dense subspace of J-{.Its adjoint is the operator
At given by (At~ 117) := (~I A17) for 17 E DomA whose domain is the
1

subspace of all ~ for which the functional17 ...... (~ I A17) is continuous (and
can thus be represented by a vector At~).
If Ais formally selfadjoint that isl (A~ 117) = (~ I A17) for ~~17 E DomA
1 1

then DomA ~ DomAt, with At~ = A~ for ~ E DomA; that is, At is an


extension of A. Since Dom At is therefore also dense, the second adjoint
A := (At) t is an extension of A that is a closed operatorl i.e. its graph is a
1

closed subspace of J-{ a> J-{. In fact, G(A) is the closure of the graph G(A),
and so A is called the closure of A. The adjoint of Ais again At, so that At
is also a closed operator.
Definition 9.15. A densely defined unbounded operator A is selfadjoint if
At= A; in other words, Ais formally selfadjoint and DomAt = DomA.
A formally selfadjoint operator A is called essentially selfadjoint if its
closure A is selfadjoint.
In general, if A is formally selfadjoint, it is to expected that the domain
of At is larger; while if A is extended to a larger domain, the domain of At
will shrink. The goal is then to extend A to a domain just large enough so
that it coincides with the domain of its adjoint. Now a formally selfadjoint
operator might have no selfadjoint extensions at alll or it may have infin-
itely many; for examplesl see [383, VIII.2]. On the other hand, an essentially
selfadjoint operator has a unique selfadjoint extension, namely its closure.
Theorem 9.15. The Dirac operator on a cornpact spin rnanifold M with
spinor rnodule S = roo (S) is an essentially selfadjoint operator on the spinor
space J-{ = L 2 (M, S).

Proof. Since 1,7) is formally selfadjoint on the domain S, its closure 1,7) exists
and its graph is the closure of the graph of 1,7). This means that 1./J E Doml,l)
if and only if there is a sequence of spinors 1./Jn ES suchthat 1./Jn - 1./J and
the limit cp := limn DI./Jn also exists in J-{; of coursei we set 1,7}1./J := cp.
394 9. Commutative Geometries

If 11 ES and 1./J E Dom.l,l), then

(1./J IW'7} = n-oo lim(l,l)I.Jln I 17} = (cp I 17},


lim(f./Jn l.l,l)17} = n-oo

so that 11 ...... ( 1./J l.l,l) 11} is continuous for the L 2-norm on S; that is to say,
1./J E Dom.l,l)t. Thus we need only establish that Dom.l,l)t s;; Doml,l).
Now since l,l) is a pseudodifferential operator, the formula (.l,l)!.JJ I 17} :=
(1./J IW'7} may be interpreted as defining l,l)I.Jl, for a generalf./J E J-f, in the
distributional sense; and the condition 1./J E Dom.l,l)t means that this distri-
butional imagelies also in J-f. Notice that there is no problern in defining
T!.JJ if T is a smoothing operator on S; since T has a smooth Schwartz kernel,
T!.JJ will belang to S.
In order to use pseudodifferential calculus, it is useful to suppose that
the L2-section 1./J has support in a chart domain; this can always be arranged
by using a finite partition of unity {ji} (remember that M is compact) and
applying the argument to each Jii.Jl separately. Now let Q be a parametrix
for .l,l), suchthat R :=I -l,l)Q and T :=I- Ql,l) are smoothing operatorsover
this chart domain, and notice that l,l)T = Rl,l). Since it is a pseudodifferential
operator of order -1, Q extends to a bounded operator on J-f.
Choose a sequence {c/>n} c S with cf>n - l,l)I.Jl in J-f, and introduce 1./Jn :=
Qcf>n + T!.JJ. This new sequence also lies in S, and it converges in J{ to
Ql,l)!.JJ + T!.JJ = 1./J, while

lim l,l)Qcf>n + l,l)TI.Jl = n-oo


lim Wcf>n = n-oo
n-oo
lim (c/>n- Rcf>n) + Rl,l)I.Jl = l,l)I.Jl.

This shows that 1./J E Dom.l,l), as claimed. Therefore, l,l) is selfadjoint. D


The compactness of M is not a necessary condition for a (generalized)
Dirac operator tobe essentially selfadjoint; what is needed, rather, is that
the Riemannian manifold be complete. Forthismore general result, we refer
to [494] or [190, §4.1].
From now on, we shall suppress the closure bar and denote also by l,l)
the unique selfadjoint extension of the Dirac operator originally defined
on S. Its domain is given in the proof of Theorem 9.15, and consists of all
1./J E J{ for which l,l)I.Jl E J{ also, where l,l)I.Jl is tobe understood in the
distributional sense .
.,.. Since l,l) locally has the principal symbol c ( ~), its square -a positive self-
adjoint operator- has principal symbol a-2(.l,l) 2) = c(~) 2 = g- 1 (~.~) 12m=
gii~i~J 12m. Now the scalar Laplacian on M has principal symbol giJ~i~J
given by (7.29), and the analogaus connection Laplacian on the spinor bun-
dle has, in view of the remark at the end of Section 7.2, the matrix-valued
principal symbol gii~i~J 12m. Weshall call it the spinor Laplacian l::!.s; it is
given by the direct analogue of (7.24):

(9.29)
9.4 Analytical aspects of Dirac operators 395

The difference [/) 2 - tl5 is thus a differential operator an S of order less


than 2. It turns out to be of order zero; in fact, as Uchnerowicz discov-
ered [315] and the following theorem shows, it is proportional to the scalar
curvature of (M,g).
Theorem 9.16 (Lichnerowicz). On a compact spin manifold, the Dirac ope-
rator [/), the spinor Laplacian t:;"5 and the scalar curvature s are related by
the equality

(9.30)

Proof. It is enough to prove this an any local chart, so we may use local
expressions for [/), tl5 and s = gi 1R1z = gikgJIRiJkl· First of all,

I!J 2 = -c(dxi)V'~j c(dxi)V'~j


= -c(dxi) c(dxi) V'~; V'~j- c(dxi) c(V'gjdxk)V'~k
. . s s k s
= -c(dxt)c(dx 1 )(Y'a;Y'aj -[iJY'ak).

Using the symmetry ri~ = r}i and the relation {c(dxi), c(dxi)} = 2gii, this
simplifies to

(9.31)

The first term is the spinor Laplacian, while the second involves the cur-
vature of V' 5 , which equals iJ(R), by Proposition 9.9. Thus, [V'~k' V'~1 ]
iJ(R)(ak, az) = ~RiJkl c(dxi) c(dxi), and so

[/) 2 - tl5 = -!RiJkl c(dxk) c(dx 1) c(dxi) c(dxi)


= !RJikl c(dxk) c(dx 1) c(dxi) c(dxi).

We can replace the term c(dxk) c(dx 1) c(dxi) by its skewsymmetriza-


tion Q(dxk A dx 1 A dxi) plus extra terms, given by Exercise 5.1. Since RJikl
is skewsymmetric in i, k, l by (7.16), the last expression simplifies to

Since RJiklBkl = 0 by skewsymmetry of RJikl in k, l, the first term vanishes,


while the second and third give equal contributions. Therefore,

[/J 2 - t:;"5 = ~RiJklBik c(dx 1) c(dxi) = ~RJ! c(dx 1) c(dxi) = ~glJRJ! = ~s,
where we have used the symmetry R1t = Rz1 of the Ricci tensor. D

The Uchnerowicz formula (9.30) has an important generalization, appli-


cable to generalized Dirac operators.
396 9. Commutative Geometries

Corollary 9.17 (Bochner-Weitzenböck formula). Let D = -ic o \JE be a


generalized Dirac operator on a Clifford module 'E = S ® .J\ 5 over a compact
spin manifold. Write \JE = 'V 5 ® 1:r ~ 15 ® 'VF and Iet KF be the curvature
of'VF. Then

(9.32)

Q: .Jt • (M) - f(G(M)) being the quantization map given fibrewise by (5.4).

Proof. In the proof of Theorem 9.16, replace I/) by D and !::;,.5 by t:;,.E. Then
(9.31) becomes t:;,.E- ~c(dxi) c(dxi)KE (ai, aJ ), where KE is the curvature of
\JE. Squaring \JE Ieads to the decomposition KE = iJ(R) ® 1:r ~ 1s ® KF. Now
recall that the f(Cl(M)) acts effectively on the first factor only of S ®.J\ J',
so that to the original (9.31) we just add the extra term

-~c(dxi) c(dxi) ®KF (ai, aj) = -~Q(dxi 1\ dxi) ®KF (ai, aj) =: -~Q(KF).

The factor ~ on the right hand side comes from the relation

~(c(a) c(ß)- c(ß) c(a)) = Q(a 1\ ß) = ~Q(a 1\ ß- ß" a)

for any a, ß E .Jt 1 (M). 0

The lichnerowicz formula has several useful consequences. First of all,


if the scalar curvature of M is strictly positive, say s(x) ;:: s0 for x E M,
then \II/Jl/J\1 2 ;:: ~so\\l/1\\ 2 , so that kerQJ = {0}. (Actually, it is enough that
s(x) > 0 at one point of M [247].)
Secondly, the positive selfadjoint operator QJ 2 is a bounded perturbation
of the spinor Laplacian and has the same asymptotic spectral behaviour for
large eigenvalues. Clearly I/) also has compact resolvent.
Example 9.2. Consider the circle § 1 again. Here QJ 2 = -d 2 I dfJ 2 = !:l, since
the metric is flat; indeed, .Jt 1 (§ 1 ) is the free C"' (§1 )-module with the single
generator dfJ, and 'VB = d. Also, S "" C"' ( § 1 ). The eigenspinors for I/) =
-idjdf} are l/Jr(fJ) := e 2rriril for r E 71., since I/JlJlr = -iljJ~ = 2TTrlJlr, so
sp(QJ) = 2rr7l. with multiplicity one .

.,.. When the spin manifold M is even-dimensional, so that S = s+ ~ s- is


a 71. 2 -graded Clifford module, the grading operator c(y) commutes with
the spin connection operator 'Vi, as is clear from (9.14), and therefore
anticommutes with QJ:

c(y) I/)= -I/) c(y). (9.33a)

Recall that c(y) exchanges s+ and s-, c(y) 2 = 1 on S, and (c(y)c1>\c(y)ljJ) =


(cP \l/J) because y*y = 1. Therefore, c(y) extends to a selfadjoint unitary
operator on the spinor space J-( = L 2 (M, S) that weshall denote by x. [This
grading operatorisoften denoted by r, but we have too many uses forthat
9.4 Analytical aspects of Dirac operators 397

letter already.) Thus J{ = J{ + EB J{-, where J{± = L 2 (M, S±) are the com-
pletions of the "half-spinor" modules S±, whose elements are usually called
Weyl spinors, whereas elements of the full space J{ are sometimes called
Dirac spinors. Clearly, (9.33a) extends to

xW=-QJx, (9.33b)

as an anticommutation relation between selfadjoint operators on Jf. In


particular, QJ maps a dense subspace of Jf+ into J{- and conversely. With
respect to the decomposition J{ = J{ + EB J{-, we write

o
w = (w+ w-)
o (9.34)

Here, w+ and w- are mutually adjoint Operators.


The anticommutation (9.33b) means that the spectrum of the selfadjoint
operator QJ is symmetric about 0. Indeed, if QJI./J = Al/J, then QJ(Xl/J) =
-x(QJI./J) = -A(Xl/J), so that x exchanges the eigenspaces for ;\. and -;\.;in
particular, these eigenvalues have the same multiplicity. More pictorially,
if 1./J = 1./J+ EB 1./J- is an eigenspinor for QJ 2 with eigenvalue ;\. 2 , then

(9.35)

This exhibits the (±A)-eigenspaces for QJ in terms of the eigenspaces for QJ 2 .


.,.. The kernel of the Dirac operator acting on a ~z-graded Clifford module
can be written as ker QJ+ EB ker QJ-, bearing in mind that QJ± takes S± into
s+; this is sometimes called the "index space" of QJ. The index of the odd
operator QJ is defined as

indexQJ := dim(ker QJ+)- dim(ker QJ-). (9.36)

In other words, the index of QJ is defined as the Fredholm index of QJ+ -of
course, the Fredholm index of the selfadjoint operator QJ, in the sense of
Chapter 4, would vanish. (Apart from that, the difference with the frame-
work of Chapter 4 is superficial: conventionally, Fredholm operators are
taken tobe bounded, whereas QJ is not. But one can remedy this by redefin-
ing QJ as an operator between different Sobolev spaces. Or instead, one
can use the alternative definition [132, §15.12) of a Fredholm operator as
a closed, possibly unbounded, operator between Hilbert spaces which has
dense domain, finite-dimensional kernel and finite-codimensional, there-
fore closed, range. In this sense, the operator closures of QJ and ß are un-
bounded Fredholm operators.)
398 9. Commutative Geometries

Example 9.3. On the torus lr 2, again with its usual flat metric, 5\. = C"" (lr 2)
and 'B = f""((l(lr 2)) ""M2(5\.); there are two possible spin structures [247]
(apart from change of orientation), and we choose the one whose spinor
module is free of rank two: S "" 5\. 2. We identify the generators of the
Clifford action ;y 1, ;y 2 with the Pauli matrices u 1, u2. If K denotes complex
conjugation on .J1. 2, then the operator C satisfying (9.8) is given by

(9.37)

If a 1, a 2 are angular variables parametrizing lr 2, and if oi = oI oai, then

I/)= -i(;y 1 01 + ;y 2 02) = -i ( ::. o


·::. o1 -
0 io2) (9.38)
v1 + tv2
can be thought of as a "Cauchy-Riemann operator" on the elliptic curve
lr 2 = (/7!. 2. The spin connection is '\7 5 = d, acting on .J1. 2, and the spinor
Laplacian is !l5 = !l ® 12 (two copies of the scalar Laplacian !l = -of - o~).
From (9.38) it is immediate that QJ 2 = !l5 (of course, s = 0). For each
r = (r 1, r 2) E 7!. 2, there are two eigenspinors for I/) 2, namely 1/J~ = 1/J~ EB 0
and 1/J; = 0 EB 1/J~ where IJ1F(e< 1,a 2) := exp(2rri(r1a 1 + r 2a 2)). The cor-
responding eigenvalues of QJ 2 are 4rr 21rl 2 = 4rr 2(rf + rf), and the eigen-
spinors for I/) can be read off from (9.35) .

.,.. Finally, the Dirac operator on a spin manifold can be used to give an easy
but important simplification of the noncommutative integral (7.83). From
(9.28), the principal symbol of QJ 2 is

CT2(I/J2) = ui(I/))2 = c(~)2 = g-1 (~,~)12m = 0"2(/ls),

as is also clear from the Lichnerowicz formula. lt follows that

After multiplying by any a E 51., taking the matrix trace (that contributes
a factor of 2m, the rank of the spinor bundle) and integrating over the
cosphere bundle §* M, we arrive at

Alternatively, Corollary 7.21 allows us to rewrite this in operator form:

Of course, II/JI-n belongs to the Dixmier trace dass .[ 1+(Jf); this may be
seen directly from the Lichnerowicz formula, since
II/JI-n = (!ls + ~s)-n/2 = (!ls)-nf2(l + ~s(!ls)-1)-n/2,
9.5 KR-cycles and the eightfold way 399

since (!l5 )-n!Z E L 1+ and the second factor on the right is a bounded ope-
rator.
To sum up, the usual integral of functions in C"" (M), with respect to the
Riemannian density, may, when M is a spin manifold, be rewritten as

IMa IVg I = n(2rr)n + I 1-n


2mnn Tr (a [/J ),

for n = 2m or 2m+ 1; compare (7.82). lt is worth stating both cases sepa-


rately [465]:

I {
m! (2rr)m f if dimM =2m is even,
f all/JI-zm-
a[/J-zm
alv9 1 =
M (2m+1)!!rrm+ 1 1 ifdimM=2m+1isodd.

The noncommutative integral (7.83) can also be rewritten as

fa 1[/JI -n ·- n(2rr)n T +( II/JI-n)


.- 2ln/2JOn r a ' (9.39)

where the salient property of [/J is that it is a selfadjoint operator for which
II/JI-n lies in the Dixmier trace dass. This is the form of the noncommuta-
tive integral that we shall use from now on; it generalizes directly to the
noncommutative case.

9.5 KR-cycles and the eightfold way


In Chapter 5 we considered the charge conjugation K of the Clifford algebra
b- x(b), implemented by the antilinear operator C -see equation (5.23).
That is globalized by the conjugation operator C of this chapter, defining a
spin structure. This operator C can be regarded as a partner of the Clifford
action of the algebra BonS, subject to the intertwining rule (9.8b), namely,
C b c- 1 = x(b) for bEB. (9.40)
Since Ca c- 1 = ä for a E C"" (M) by (9.8a), C acts by an antilinear operator
on each fibre Sx of the spinor bundle, and (9.8c) says that this operator is in
fact antiunitary. Thus C can be thought of as a smooth assignment x - Cx
of antiunitary operators on a Fock space carrying an irreducible represen-
tation of the algebra l(l(+l (~n ), each satisfying the analogue of (9.40). From
Lemma 9. 7, we know that C2 = ± 1. Recall that, if { e1o ... , en} is an oriented
orthonormal basis for ~n and yi := c(ej) on the Fock space, then (9.40)
reduces to Cyic- 1 = -yi for j = 1, ... , n when n is even; when n is odd,
we can only say that Cyiyic- 1 = yiyi for i,j = 1, ... , n.
One way to talk about this structure is to say that the fibres of the
spinor bundle carry representations of the pair (G(+l (~n ), K) -see Defini-
tion 5.11. We already mentioned in Chapter 5 that this is equivalent to clas-
sifying the representations of the real Clifford algebras Clp,q with p +q = n.
400 9. Commutative Geometries

In any case, this led Atiyah [11] to define a "Real K-theory" for topologi-
cal spaces X with an involution (a homeomorphism whose square is idx ),
leading to a family of abelian groups KRP,q (X) with the same periodicities
as the algebras Clp,q· These can be rewritten as "KR-cohomology groups"
KRp,q(C(X)) and can then be generalized to the case of C*-algebras with
antilinear involutions. (The use of the terms homology or cohomology in
these contexts is mainly determined by convention; also, the ward "Real",
with a capital R, was introduced by Atiyah as shorthand for "complex with
a given involution".)
For the spin structure classification, we actually need a dual theory called
KR-homology, which was introduced -for involutive Banach algebras- by
Kasparov [275]. Earlier, the K-homology of topological spaces had been
developed as a functorial theory whose cycles pair with vector bundles in
the same way that currents pair with differential forms in the de Rham
theory. Such cycles are given, interestingly enough, by spinc structures:
see [23] for a clear exposition. However, the index theorem shows that the
right partners for vector bundles are elliptic pseudodifferential operators
(with the pairing given by the index map), and Atiyah [14] sketched how K-
cycles should be recast in terms of elliptic operators. Kasparov found the
right equivalence relation for such cycles and, more importantly, showed
that the correct abstraction of "elliptic operator" is the notion of a Fredholm
module over an algebra.
A K -cycle over a pre-C* -algebra .Jl is nothing other than a pre-Fredholm
module (.Jl,J.f,F): see Definition 8.4. Homotopic pre-Fredholm modules
are declared equivalent, and degenerate Fredholm modules (those for which
[F, a] = 0 for all a E .Jl, as well as F = Ft and F 2 = 1) are factared out,
i.e., the direct sum of a given pre-Fredholm module and adegenerate Fred-
holm module is declared equivalent to the former. The equivalence dass
[.Jl, J{, F] is then a K-homology dass for the algebra .Jl. With the obvious
notion of direct sum and an application of Grothendieck's trick, they gene-
rate two abelian groups: the dasses of even (pre)-Fredholm modules make
up the group K0 (.Jl), and those of odd (pre)-Fredholm modules constitute
the group K 1 (.Jl).
Unfortunately, as we have already seen in Section 8.2, computations with
commutators [F,a] coming from Fredholm modules can be quite cumber-
some. An important simplification was introduced by Baaj and Julg [18],
who observed that if D is an unbounded selfadjoint operator, with com-
pact resolvent, and if each [D, a] is at least a bounded operator, then
F' := D(l + D 2 )- 1 i 2 determines a pre-Fredholm module (.Jl,J{,F'). They
also showed that all K-homology dasses of .Jl arise in this way. This moti-
vates the following definition.

Definition 9.16. A spectral triple -also called an "unbounded K-cyde"-


for an algebra .Jl is a triple (.Jl, J{, D), where J{ is a Hilbert space carrying
a representation of .Jl by bounded operators (that we shall write simply
9.5 KR-cycles and the eightfold way 401

~ - a~ for the operator representing a E .Jl.), and D is a selfadjoint ope-


rator on .Jl., with compact resolvent, suchthat the commutator [D, a] is a
bounded operator on Jf, for each a E .Jl..
When the algebra .Jl. comes equipped with an involution T (which need
not be the standard involution a - a*), we can combine this notion with
an explicit action by a real Clifford algebra.
Theorem 5.4 teils us that the real Clifford algebra Clp,q is determined,
up to tensoring by a real full matrix algebra, by

j := (p- q) mod 8.

This can be regarded as an element of ~8 . We shall adopt this notational


convention in what follows.
Definition 9.17. An unreduced KR1-cycle for an algebra with involution
(.Jl., T) consists of a package (.Jl., Jf, D, C, x), where:

(a) (.Jt.,Jf,D) is an unbounded K-cycle for .Jl.;

(b) C is an antilinear isometry on J{ that commutes with D, satisfies


C 2 = 1, and implements T, i.e., CaC- 1 = T(a) for a E .Jl.;

(c) x is a grading operator on J{ commuting with C and anticommuting


withD;

(d) tagether with a representation p of Clp,q by bounded operators on Jf,


where p - q = j mod 8, which commutes with .Jl. and C and anticom-
mutes with the operators D and x.
The representation of Clp,q is generated by unitary operators yk, for k =
1, ... , p + q, satisfying the following relations:

(yk) 2 =+1 for k=1, ... ,p,


(yk) 2 =-1 for k=p+1, ... ,p+q,
yJyk = -ykyJ, if j * k. (9.41)

The supercommutation relations with the other operators are

ayk = yka for a E .Jl., Dyk = -ykD,


cyk = ykc, XYk = -ykx. (9.42)

If desired, the grading operator x can be incorporated into the Clifford


algebra representation as an extra generator; now x2 = + 1 and Cx = xc,
Dx = -xD, so that x and p tagether give a representation of Clp+l,q super-
commuting with .Jl., C and D. This was essentially the convention originally
adopted by Kasparov [275].
402 9. Commutative Geometries

We wish to show that Dirac operators on spin manifolds give rise to


examples of KR-cycles; but it should already be clear that the presence
of the supercommuting representation of Clp,q is incompatible with the
irreducibility of the spinor modules. So our first order of business is to
reduce this structure by eliminating the operators yk, passing to a subspace
of the original Hilbert space J-{. Weshallsee that they leave behind a tell-
tale footprint.
The goal of this reduction is to find a set of operators D', C' and (where
possible) x' defined on a closed subspace J{' of J{ that keep the properties
of their namesakes on J{, Namely, D' should be a selfadjoint operator
on J{', C' should be an antilinear isometry on J{' implementing T, and x',
if it exists, should be a grading operator on J{' that anticommutes with D.
Lemma 9.18. Given (.Jt, J{, D, C, X) and a supercommuting representation
o(Clp+l,q+l on J{, there is a subspace J{' of J{ that reduces .Jt, D, C and x
and carries a supercommuting action o(Clp,q·

Proof. Consider the operator P := yP+ 1 yP+ 2 • It is involutive:

and it commutes with each a E .Jt, with D, C, x. and also with y 1, ... , yP
and yP+3, ... , yP+q+ 2 • Therefore, the ( ± 1 )-eigenspaces J{' := { ~ E J{ :
P~ = ~} and J{" := { ~ E J{: P~ = -~} reduce all the operators in ques-
tion except yP+ 1 and yP+ 2 • The result is obtained by restricting to J{' all
operators except these two, since the algebraic properties listed in Defini-
tion 9.17 survive the restriction. D

Notice that yP+ 1 and yP+ 2 generate a real subalgebra of operators iso-
morphic to Ch, 1 = M 2 (1R?.), and that the two complementary minimal pro-
jectors in this algebra reduce all the other operators in question simultane-
ously; we require M 2 (1R?.) rather than M 2 ((()in order to reduce the antilinear
operator C. Thus, we have in effect used the (1, 1 )-periodicity (5.6a) of real
Clifford algebras.
This procedure can be repeated p or q times (whichever is fewer); at
each stage, eliminating one yr with (yr) 2 = +1 and one y 5 with (y5 ) 2 =
-1, by reducing the remaining operators to the (+1)-eigenspace of yrys.
We thereby reduce to the case that p = 0 or q = 0 (and p - q remains
unchanged).
Remark. In the same way, if p ~ 8, the operators y 1 , ... , y 8 generate a
real subalgebra isomorphic to Cls,o = M16(1R?.), whose minimal projectors
reduce .Jt, D, C, x and the remaining yk, on account of the isomorphism
Clp+B,q ""M15(1R?.) ®Clp,q. given by Corollary 5.3. A similar argument applies
if q ~ 8. In summary, the (1,1)-periodicity and the 8-periodicity of real
Clifford algebras shows that there are at most 8 cases left to consider,
labelled by j = (p- q) mod 8.
9.5 KR-cycles and the eightfold way 403

Theorem 9.19. Any unreduced KRi -cycle for an algebra with involution
(.Jl, T) may be reduced to a representation of 5\ on a closed subspace J-{' s;;
J-{, tagether with:

(a) a selfadjoint operator D' on J-{' making (.Jl, J-{', D') an unbounded
K-cycle;
(b) an antilinear isometry C' on J-{' implementing T; and
(c) when j is even, a grading operator x' on J-{', anticommuting with D'.
The conjugation operator C' obeys the following algebraic relations:

C' 2 = ±1, C'D' = ±D'C', c'x' = ±x'c', (9.43)

where the signs depend only on j mod 8.


Proof. Let (.Jl, J-{, D, C, x>. tagether with operators yk on J-{ obeying (9.41)
and (9.42), constitute the given unreduced KRi-cycle. By applying the re-
ductions described in Lemma 9.18 and in the subsequent remark, which
change neither the algebraic relations among the operators nor the quan-
tity j = (p -q) mod 8, we may assume that either q = 0 and p = 0, 1, 2, 3,4,
or p = 0 and q = 0, 1, 2, 3, 4.
We must now make some further reductions to eliminate the remain-
ing yk, on a case-by-case basis. There are several possible ways to do that;
we shall choose a procedure that in every case defines D' as the restriction
of D to J-{', and also defines x' as the restriction of x to J-{' when j is even.
(Only the conjugation operator C needs tobe modified.) We consider the
even cases first.
Case j = 0: Here p = q = 0 and no yk remains; we take J-{' := J-{ and
C' := C. For a reduced KR 0 -cycle, all signs in (9.43) are plus signs.
Cases j = 2, 6: When (p, q) = (2, 0) or (0, 2), there are two anticom-
muting operators ;y 1 , ;y 2 , with (;y 1 ) 2 = (;y 2 ) 2 = ±1, tobe eliminated. Note
that (;y 1 ;y 2 ) 2 = - (;y 1 ) 2 (;y 2 ) 2 = -1; to get an involutive operator on J-{, we
choose P := i;y 1 ;y 2 . Now each a E 5\ and also D and x commute with P, so
the ( + 1 )-eigenspace J-{' := { ~ E J-{ : P~ = ~ } reduces them.
However, P and C anticommute, since Cis antilinear; but

(9.44)

so we can take C' tobe the restriction of x;y 2 C to J-{'. This is still an anti-
linear isometry. Since (x;y 2 C) 2 = (x;y 2 ) 2 = -(;y 2 ) 2 , we obtain the following
signs in (9.43):

Case j = 2: c' 2 = -1, C'D' = +D'C', c'x' = -x'c';


Case j = 6: c' 2 = +1, C'D' = +D'C', c'x' = -x'c'.
404 9. Commutative Geometries

Case j = 4: When (p, q) = (4, 0) or (0,4), there are four anticommut-


ing operators ;yk, with equal squares ± 1. (The sign of the common square
will not matter, because of 8-periodicity.) Their products yield two com-
muting involutive operators P1 := i;y 1 ;y 2 and Pz := i;y 3 ;y4 that also com-
mute with .Jt, D and X· Therefore we reduce to the joint (+1)-eigenspace
J{' := { ~ E J{: P1~ = Pz~ = ~ }. Again, C anticommutes with P1 and P2,
but the operator ;y 2 ;y 3 C commutes with them, similarly to (9.44). The pre-
vious choice x;y 2 C will not do, since it anticommutes with Pz. We therefore
take C' to be the restriction of ;y 2 ;y 3 C to J{'. The signs are now seen to be

Case j = 4: C' 2 = -1, C'D' = +D'C', c'x' = +x'c'.

Cases j = 1, 7: When (p, q) = (1, 0) or (0, 1), only one ;y 1 is available, so


to make a reduction it must be multiplied by the grading operator x. which
is thereby "lost": there will be no grading operator on the reduced space. We
take P := ix;y 1 if j = 1, or P := x;y 1 if j = 7, in orderthat P 2 = +1 in each
case. Once more P commutes with .Jt and D, andin the case j = 7 it also
commutes with C. In the case j = 1, we find that P;y 1 C = -;y 1 PC = ;y 1 CP;
note, however, that ;y 1 C anticommutes with D. Thus we may define C' as
the restriction to J{' of C when j = 7, but of ;y 1 C when j = 1. The signs
are

Case j = 1: c' 2 = +1, C'D' = -D'C';


Case j = 7: C' 2 = +1, C'D' = +D'C'.

Cases j = 3, 5: When (p, q) = (3, 0) or (0, 3), there are three ;yk with
equal squares ± 1. We can make two commuting involutive operators by
again pressing the grading operator x into service. We take P 1 := ix;y 1
if j = 3, or P := x;y 1 if j = 5, andin both cases we set P 2 := i;y 2 ;y 3 ;
then J{' := { ~ E J{: P1 ~ = Pz~ = ~}, as before. An antilinear operator
commuting with P1 and Pz is easily found. We take C tobe the restriction
to J{' of x;y 2 C when j = 3, but of ;y 2 C when j = 5; again it should be noted
that ;y 2 C anticommutes with D. This gives the signs:

Case j = 3: c' 2 = -1 I C'D' = +D'C';


Case j = 5: C' 2 = -1, C'D' = -D'C'.

Finally, we remark that since the original representation of .Jt on J{ com-


mutes with x and each ;yk, the relation Cac- 1 = T(a) is unaffected by
multiplying C by any of these operators; thus C' aC'- 1 = T(a) holds in all
cases for the reduced representation of .Jt on J{'. D

The reduction process is, of course, not unique. For instance, in the case
j = 1 we could have made the simpler choice P := ;y 1; since C;y 1 = ;y 1C,
the conjugation C now restricts to J{' without modification. However, D
9.5 KR-cycles and the eightfold way 405

no Ionger commutes with P, so we "spend" the grading operator anyway


by defining D' as the restriction to J{' of ixD. Notice that C anticommutes
with ixD because of the i factor that is needed to make D' selfadjoint, so
the signs for the case j = 1 are unchanged by this equivalent reduction.
Similarly, in the case j = 6, we could have taken P := xy 1 , which com-
mutes with D, C and y 2 ; so D', C' are the restrictions of D and C. Having
thereby "lost" x. we recover a new grading x' from iy 2 , and this yields the
minus sign in the relation C'x' = -x'C' .
.,.. We summarize the conclusions of Theorem 9.19 by restating what a KRj-
cycle is in the reduced case, where the Clifford algebra action has been
factored out.
Definition 9.18. Let j E ~s; a (reduced) KRj·cycle for an algebra with
involution (.Jl, T), consists of a package (.Jl, Jf, D, C, X) if j is even, or
(.Jl, Jf, D, C) if j is odd, where:

(a) (.Jl, J{, D) is an unbounded K-cycle for .Jl;

(b) Cis an antilinear isometry on J{ that implements T;

(c) if j is even,x is a grading operator on J{ that anticommutes with D;


(d) the operators D, C, x satisfy the commutation rules (9.43), where the
signs are given by the following tables:

jmod8 0 2 4 6 jmod8 1 3 5 7
C2 = ±1 + - - + C2 = ±1 + - - +
CD= ±DC + + + + CD= ±DC - + - +
Cx =±XC + - + -
(9.45)

A moment's thought shows that the process of reduction is reversible.


That is to say, if a reduced KRj -cycle is given, one can enlarge the Hilbert
space J{ by making a direct sum of extra copies of it, intertwined by ope-
=
rators yk forming a representation of some Clp,q with p- q j mod 8, and
one can extend the operators to the new copies of J{ in order to satisfy
(9.42), thereby manufacturing an unreduced K-cycle.
We exemplify this extension with the circle, where .Jl = C"" (§1) acts by
multiplication operators on J{ = L2 (§ 1 ); we take D := -id!dB and C
as complex conjugation on Jf. Here the antiautomorphism T is complex
conjugation on .Jl, since C ac- 1 is the multiplication operator ä. Since C2 =
+1 and CD= -DC and no grading has been mentioned, these ingredients
form a (reduced) KR 1 -cycle. Let !it := J{ Eil Jf, regarding .Jl, D and C as
406 9. Commutative Geometries

operators on the first copy of J{ only. With respect to this decomposition,


set

yl := (~ ~), x:= (~ ~i), so that ixy = (~ ~1 ).


1

Now define

~ ·- (-idjd()
D.- 0 ~
C :=
(o c) a_ (a ao)
C 0 ' := 0 if a E ./\.

The top left corners of ä, D and y 1 C are a, D and C respectively, and it


x
is clear that D and anticommute, that C2 = + 1 and that C commutes
with both D and x. Therefore (./\, Jt, D, C, x>. tagether with y 1 , forms an
umeduced KR 1 -cycle.
Exercise 9.9. Reconstruct umeduced KRJ -cycles from the reduced ones of
Definition 9.18, in the remaining cases j = 2, ... , 7. 0
Theorem 9.20. Let M be a compact spin manifold of dimension n with a
given spin structure (v, S, C) and corresponding Dirac operator I/) on the
spinor space J{ = L 2 (M,S); and Iet j := n mod 8. If ./\ = C 00 (M), and
x = c(y) is the grading operator when n is even, then (./\, Jf, I/), C, X) for
even n, or (./\, Jf, I/), C) for odd n, is a reduced KRJ -cycle over ./\.

Proof. Recall from (9.20) that I/) is locally of the form I/)= -i yavf.., where
ya = c ( 9 a) gives the action of a local orthonormal basis of (complex) 1-
forms. The spin connection operators V~ are C-invariant when we contract
with real vector fields X, so we may suppose also that the 9a are real 1-
forms; at each point of M, {y 1 , ... , yn} generates a representation of the
real Clifford algebra Cln,o. while the operators {-iy 1 , .•• , -iyn} generate
a representation of Clo,n· To find the commutation relations between I/), C
and x. it is enough to examine them at any point of M (using, say, normal
coordinates at that point to suppress the Christoffel symbols).
When n = 2m is even, C satisfies (5.24), from which the commutation
CI/) = QJC follows immediately. The chirality element of Cl(~ zm) is repre-
sented by x = (-i)my 1 •.. y 2 m, so (5.24) also implies that Cx = (-l)mxc,
in accordance with the first table in (9.45).
When n = 2m + 1 is odd, (5.24) is not directly applicable. To compute
C(-iyk)C- 1, we use the extended action (9.2) of o-(~ 2 m+I) that replaces
-iyk by ( -i)m+I ykyl ... y 2 m+I. Since C commutes with even products
yiyJ, its antilinearity shows that C(-iyk) = (-l)m+ 1 (-iyk)C, and there-
fore CI/) = (-l)m+I I/)C, in accordance with the second table in (9.45).
It remains to check the sign of C2 in all cases. This depends on the type
of the spin representation [46, 54, 439] of the Clifford algebra Clo,n· For
n = 2m, that is the unique 2m-dimensional representation of the algebra;
for n = 2m+ 1, it is the direct sum of the two irreducible representations
9.A Spin geometry of the Riemann sphere 407

of dimension 2m. As outlined in Section 5.3, the type of the spin repre-
sentation of Clq,p depends on - j = q - p mod 8: it is real for j = 6, 7, 0;
complex for j = 1 or 5; and quaternionie for j = 2, 3, 4.
The classification into types involves the effect of conjugating the spin
representation by an antiunitary operator on the representation space; in
the present case, this is precisely the operator C (restricted to the selected
fibre of S). Composing with C( · )c- 1 yields the conjugate representation of
Clo,n. equivalent to the original one if n is even, but possibly inequivalent
if n is odd. In the real cases, C2 = + 1 and the spin representation and its
conjugate are equivalent. In the quaternionie cases, both representations
are also equivalent, but C2 = -1. In the complex cases, the two represen-
tations are inequivalent. This forces C2 = + 1 for j = 6, 7, 0, and C2 = -1
for j = 2,3,4.
For j = 1, the algebra Cl0,1 is just C (with -iy 1 being -i in a one-
dimensional representation). Here C can only be complex conjugation, and
so C2 = +1. For j = 5, it happens that Clo,s "" C ®IR Mz(IHI) ""M4(C) by
Lemma 5.2; and an explicit calculation (which we leave to the reader, Exer-
cise 9.10) shows that there is a unique conjugation that anticommutes with
a given set of generators, and it satisfies C2 = -1. D

Exercise 9.10. Find 5 matrices E1 , .•• , E 5 in M 4 (C) generating a faithful rep-


resentation of Clo,s and an antiunitary operator C on C4 suchthat C EkC- 1 =
-Ek for k = 1, ... , 5. Show that such a Cis unique up to multiplication by
a complex numbers of absolute value 1, and that C2 = -1 in all cases. 0
It is possible to give a more constructive proof of Theorem 9.20 by writing
down a set of matrix generators for Clo,n and a charge conjugation opera-
tor C that commutes or anticommutes with them (according to the sign in
CD= ±DC), in the remairring cases n = 2, 3, 4, 6, 7 also. While straightfor-
ward, this involves considerable bookkeeping: for instance, [58] and [439]
give lists of generators for the algebras Cln,o. and the proof of Lemma 5.2
may be used to assemble another one. We leave the diligent reader to amuse
hirnself with this task.
A spinor tfJ for which CtfJ = tfJ is called a Majorana spinor. Notice that,
in view of (9.45), in some dimensions there can be Weyl-Majorana spinors.

9.A Spin geometry of the Riemann sphere


In this section, we explore a simple but fundamental example: the Dirac ope-
rator on the irreducible spinor module over the sphere § 2 • While the sphere
is undoubtedly the simplest possible even-dimensional compact spin ma-
nifold, its Dirac operator exemplifies the complexity of the general case
while remairring directly accessible by elementary computations; the study
408 9. Commutative Geometries

of eigenspinors of Dirac operators on spheres, begun by Schrödinger [419],


has always been useful to understand spinors.
We give an account of the action of the Dirac operator on spinors, show its
equivariance under the Ue group SU(2) of symmetries of the spinor mod-
ule, compute its spectrum and exhibit a full set of eigenspinors. Our treat-
ment is mainly based on the old article of Newman and Peruase [361], who
introduced several families of functions on the 2-sphere that they called
"spinor harmonics", which generalize the ordinary spherical harmonics and
constitute the eigenspinors. See also Section 2 of [465]; however, in that ref-
erence some signs are not the same as those used here, owing to different
Clifford-algebra conventions.
We shall use the notation of Section 2.6 for complex Coordinates on the
Riemann sphere. Recall that § 2 = UN u Us is the union of two chart do-
mains omitting respectively the north and south poles, with local complex
Coordinates (2.16):
z = e-i<P cot ~.
on UN and Us respectively, and?; = 1/z on UN n Us.
Any spinc structure on § 2 has a spinor module S = f"" (§ 2 , S) where
S - § 2 is a vector bundle of rank 2. Vector bundles over the sphere al-
ways split into Whitney sums of line bundles, as we remarked near the
end of Section 2.6. Thus S = s+ e s- where s+, s- are rank-one projec-
tive modules over § 2, and so are of the form 'E<ml for some integer m.
lf S "' 'E<ml e 'E<nl• then s~ "' 'E<-ml e 'E<-nl "' 'E<-nl e 'E<-ml· Therefore,
S belongs to a spin structure only if S "' 'E(m) e 'E(-m), where the con-
jugation C interchanges the two summands. Moreover, there are C""(§ 2)-
module isomorphisms S ®.Jt s~ "'End_i(S) "'f""(<Cl(§ 2)) "'5\"(§ 2 ); the
second one comes from the Morita equivalence C(§ 2) !'. f(<Cl(§ 2)). Now
5\"(§ 2) "''E(o) e 'Em e 'E(-2) e 'E<Ol -see, for instance, [197, §2.3]- which
forces m = ± 1. We fix the spin structure on § 2 by taking the standard ori-
entation and putting s+ := 'E(l), s- := 'E<-0· In vector-bundle terminology,
we are taking s+ := L, s- := H, these being respectively the tautologicalline
bundle and the hyperplane bundle over § 2 = <CP 1. Recall, from Section 2.6,
that the bundle S = L e H is trivial.
We can give a more concrete description of these line bundles by their
transition functions. Writing z := z1 I zo with (zo, z1) E <C 2, the fibre over z
of the tautologicalline bundleis Lz := { (.\z 0 , .\zi) E <C 2 : .\ E <C }. We may
define local sections aN E f(UN,L) and as E f(Us,L) by

where we have abbreviated q := 1 + zz as in (2.17), and q' := 1 +?;(.These


normalizations are arranged so that (aN I aN)= 1 on UN and (as I as) =
1 on Us, using the standard hermitian pairing on f(L), namely, the one
induced by the inclusion L c <C 2. Clearly, as(z- 1) = ,jqjq' z- 1aN(Z) =
9.A Spin geometry of the Riemann sphere 409

~z/z UN(Z) on UN n Us. In a dual fashion, the hyperplane bundle has local
sections u~, ui related by u; (z- 1) = ~z/z u~(z).
Any (global) section 1./J+ E f(L) is determined by a pair of functions
(1./J"N,I./Js) suchthat 1./J"N(z,z)uN(z) = 1./Js{'l;;,~)us(() on UN n Us. Since
this is equivalent to the relation 1./J"N(z,z) = ~z/zl.fJs(z- 1 ,z- 1 ), we may
dispense with the local sections and define spinor components as such
pairs of functions.
Definition 9.19. The spinor module for the 2-sphere isS= s+ E9 s-, where
s+ := 1:o> = [ (L), s- := 1."(-1) = f 00 (H). Any spinor l.fJ has two compo-
00

nents 1./J±; these can be regarded as pairs of functions on C, with values


l.fJ"tl(z, z) and 1./Jg ((, ~), satisfying the transformation rules:

1./J"N(z,z) =(z/z)1121./Js(z-1,z.-1),
1./J"N(z,z) =(z/z)1121.fJs<z-1,z-1). (9.46)

~ The Levi-Civita connection VB is determined by the metric (2.18a), as


follows. Let us use local real coordinates (x 1, x 2) determined by x 1+ ix 2 :=
z on UN but x 1 + ix 2 := ( on Us; write q := 1 + (x 1) 2 + (x 2 ) 2 in both
cases. Then g = 4q- 2 ( (dx 1 ) 2 + (dx 2) 2 ) on either chart, so BiJ = 4q- 2 8iJ
and gkl = ~q 2 8k 1 • Also, alBiJ = -16x 1q- 3 8iJ. The Christoffel symbols are
given by (7.12):
ri~ = -2q- 1 (xi8~ + xJ8~- xk8ij).

To prepare for the spin connection, it is better to use local orthonormal


bases of vector fields and 1-forms. In other words, we may take the op-
portunity to make a "gauge fixing" according to the general prescription
of (9.13). Over the chart domain UN, weshall use the vector fields

that clearly satisfy g(Ecx,Eß) = 8cxß· The corresponding 1-forms are 9cx :=
E~ = 2q- 1 dxcx, for oc = 1, 2. In the notation of Section 9.3, we have chosen
H := 2q- 1 12 as a particular square root of G = 4q- 2 12 -namely, the
positive definite square root- over UN. (We defer fixing the gauge over Us
until the Dirac operator is introduced.) Now

"V 9iJ; E cx -xia


- cx +!2 q "V 9iJ; acx - 1 U1cxX ßaß --2q- 1 (~·
- - X cxa.+~· U1cxX ßEß -X cxE·)
1 ,

and so

~rß -
icx - 2 q -1 <U1cxX
~. ~ .ßX cx >,
ß - U1 (9.47)

whose skewsymmetry is transparent. The spin connection components are


then given by (9.14b) as wi = if'fcx YcxYß·
410 9. Commutative Geometries

The sphere § 2 = CP 1 is a complex manifold. We can take advantage


ofthat by switching to isotropic bases of vector fields and 1-forms [465];
over UN, these are

qaz = E1- iE2,


q az = E1 + iE2,

The abbreviations az := a;az and az := a;az will be frequentlyused.


Exercise 9.11. Show that the relations 'il~j9ß = -rf()(91X, with (9.47), yield
the formulae:

(9.48)

Find the corresponding relations for V'B(q'- 1d"() and V'B(q'- 1d() con-
tracted with the vector fields q'az; and q'az; over Us. <>

Remark. The relations (9.48) may be thought of as defining isotropic Chris-


toffel symbols [462] for the bases E+ = qaz, E_ = qaz, 9+ = q- 1dz
and 9- = q- 1 di. Thus, V'~p 9P = -I'ffv9v, with J.l, v, p E { +,-}, and the
skewsymmetry of the Symbols (9.47) implies that fffv = Ü When V and p are
opposite signs. The equations (9.48) are therefore equivalent to the relations
"'+ - "'- - .....,+ "'-
[++ = -z, [+- = z, [_+ = z, [ __ = -z.

Any 1-form oc E ..:'1.1(§ 2) maybe written locally as oc = fN(z,i)q- 1 dz +


BN(Z, i)q- 1 di or as oc = - fs((, '()q'- 1 d"(- Bs ("(, '()q'- 1 d(. On UN n Us,
the relations q'- 1 d"( = -(zi/q)z- 2 dz = -(i/z)q- 1 dz and q'- 1 d(
-(zi/q)z- 2 di = -(z/i)q- 1 di show that

fN(Z,i) =(i/z)fs(z- ,r 1 1 ),

BN(z,i) = (z/i)gs(z- 1,r 1 ). (9.49)

By comparing these gauge transformation rules with (9.46), one sees that
.Jt1(§2) ""1:(2) EB 1:(-2)· Now _:;t0(§2) = coo(§2) = 1:(0) and .Jt2(§2) ""1:<o>
since the area form 0 = 2iq- 2 dz A di = 2iq'- 2 d"( A d( of (2.18b) pro-
vides a nonvanishing global section. This proves our previous claim that
.Jt"(§ 2) ""1:(0) EB 1:(2) EB 1:<- 2> EB 1:(0) and therefore justifies the identifica-
tions S± ""1:(±1)·
~ The gamma matrices in dimension two may be taken as the Pauli ma-
trices y 1 := 0'1, y 2 := 0'2, and the grading operator is X:= -iy 1 y 2 = 0'3.
We write y 1 = y 1 and Y2 = y 2, to take full advantage of the summation
convention. lt is convenient to introduce y± := !<y 1 ± iy 2), noting that
[y+,y-] = x as weil. Notice also that x()(y()( = x 1 y 1 + x 2 y 2 = zy+ + zy-
9.A Spin geometry of the Riemann sphere 411

on UN (and xaya = ~y+ + ~y- for the other chart). The components of the
spin connection are then given, according to (9.14b), by

Wi = !rfa YaYß = (2q)- 1 (8iaxß- OißXa) YaYß

= (2q)-1[yi,xaya] = (2q)-1[yi,zy+ + zy-].

Since Y'~i = ai - Wi, this gives


.,..,s
vqaz=q az-zq 1 ( W1-tW2
. - + +zy - J =q az+zZX,
) =q az-z1 [ }' _ ,zy 1-
.,..,s -a 1 ( . ) -a 1[ + - + _ J = q -a z- zZX.
1
V
q-az = q z- 2q 001 + lW2 = q z- 2 y ,zy + zy
(9.50)

(Since Cx = -xc in dimension two, these relations illustrate the charge-


conjugation invariance of the spin connection.)
Exercise 9.12. Show that the spinor Laplacian (9.29) has the local form

(9.51)

on the chart UN. 0

Definition 9.20. The Dirac operator I/) = -i(c o \7 5 ) on § 2 acts on spinors


1/J E f""(UN,S) as

l/)IJJ = -i ya Y'Ia 1/J = -i(y+Y'Sqaz + y-Y'S-a


q z
)1/J
= -i(qaz- ~i) y+I/J- i(qaz- ~z) y-1/J. (9.52)

The last equality follows from y±x = +y±, which is seen from the explicit
representation

y+ =
(0 1)
0 0 '

The particular form of I/) on the local chart UN depends, of course, on


the choice of the local orthonormal basis of vector fields Ea := ~q aa. On
the other chart, the basic vector fields cannot be chosen arbitrarily. Con-
sistency requires that we choose Ea := -~q' aa on Us, where q' := 1 + ~~­
This accords with the matehing of the basic 1-forms q- 1 dz and q- 1 di
on UN with -q'- 1 d~ and -q'- 1 d~ on Us, in order to get the transforma-
tion rules (9.49) for the coefficients. The next Iemma shows that the use of
spinors forces us to adopt this particular gauge fixing.
Lemma 9.21. The operator defined on ["" WN, S) by the formula (9.52) ex-
tends to [""Ws, S) as follows:

(9.53)
412 9. Commutative Geometries

Proof. Fora given spinor 1./J E S, write cf> := l/)1./J. The component cf>+ E s+
is obtained from (9.52) and (9.46):

-i((qoz- ~i)f./JN)(z,i) = cf>"tJ(z,i) = (i/z) 112cf>t(z- 1,z- 1).

On the other hand, since S' = z- 1 on UN n Us,

0'''- ( -1 --1)
- .!.( -)-1/2, 1, - ( -1 --1) _ ( /-)1/2 -2_'+'_S
- 2 zz 'Y s z ,z z z z os- z ,z

= (s-()112(-s-ooi./JJ (s,"() + ~1./Js(s,"()),

so the operator qoz- ~i transforms as follows:

---az
(1 + ZZ_)of./JN( Z,Z-) - 2_ZI./JN
1- -(
Z,Z-)

= (s"()-112(1 + S""()( -s-oiJ (s. "() + ~1./Js(s. "())- ("(;2s-{/2 1./Js(s. "()

= (s/-()112(-(1 + S""()oaf./JJ (s,"() + ~"(f./J5(s,"())


= -(s/"() 112 (q'a(- ~"()f./Js(s."()
= -(i/z)I12((q'a(- ~"()f./JsHz-1,z-1).

We conclude that cf>t = i(q' O(- ~"()f./Js. If we now replace 1./JN by 1./J"t. and
1./Js by 1./Jt, and apply complex conjugation, we obtain

cf>N(z, i) = -i( (qaz- ~ Z)I./J"tJ )(z, i) = i(z I i) 112 ( (q'a(- ~ s) 1./Jt )(z- 1' z- 1),

The previous Iemma is an instance of a more general principle of spin


geometry [93). Given a local first-order differential operator that imple-
ments the Dirac operator on one chart of M, the transition functions for
the spinors produce other local first-order operators that implement l/) on
overlaps with neighbouring charts and may be extended to the whole of
each such chart; by repeating this process as often as necessary, we get
local formulas that implement l/) over the whole manifold. In this way, l/)
is determined by its restriction to an arbitrarily small open subset of M.
In the particular case of § 2 , we can write (9.53) too in the form of (9.52),
i.e., as l/)1./J = -i ya Vt,f./1 on the chart Us, where now Ea = -~q' Oa. Thus,
l/)1./J = i(y+"V 5q'i! + y-v 5 ,-0 )1./J. Using y±x = +y±, the expression (9.53)
yields ' q '
9.A Spin geometry of the Riemann sphere 413

In other words, the spin connection over Us may be obtained from (9.50)
just by replacing z ..... (, i ..... 'and q ..... q'.
~ We shall use a convenient shorthand notation, introduced by Newman
and Penrose [361]:

(9.54)

and its complex conjugate az := q ä"z- iz.likewise, we put Ö( := q' 0(- i'
and äz; := q' oz;- 2 (. Then
- - l

[/) =(w+o qr) .(azo


0 = -t
Öz)
0
= l. (~
Ö(
Öz;)
0 . (9.5 5)

Exercise 9.13. Check directly that the operators appearing in (9.54) are for-
mally selfadjoint, e.g., by showing that (cjJ+ I Özi/F) = -(8zc/J+ I r.jr), where
(cjJ± I f./J±) := fc (b±f./J± n, using the area form n of (2.18b). o
The charge conjugation operator C is determined, up to a multiplicative
constant of absolute value 1, by the requirement that cyac- 1 = -ya. If K
is the ordinary complex conjugation on the spinor components, such a C
is given by

It is immediate that C 2 = -1 and Cx = -xC. The commutation relation


CI/)= +[/JC may be exhibited directly:

Lemma 9.22. The Lichnerowicz formula for the sphere §2 with the chosen
spin structure is

(9.56)

Proof. The spinor Laplacian !l.5 , given by (9.29), can be expressed in the
isotropic basis: see Exercise 9.12. From (9.50) it follows that
414 9. Commutative Geometries

On the other hand, the square of the Dirac operator may be computed
directly from (9.55):

-!
so that l/) 2 = tl5 + over the chart UN. On the other chart, we find the
analogaus formulae (replacing z - l; and q - q'), and so l/) 2 = tl5 + -!
over Us also. D

Corollary 9.23. The sphere (§ 2 , g) has constant scalar curvature s = 2. a


.,. The Dirac operator has discrete spectrum, consisting of real eigenvalues
of finite multiplicity; that much we know from the general theory of self-
adjoint operators with compact resolvent. Also, since the I/) is an elliptic
differential operator, we expect that its eigenspinors are smooth, that is,
they belang to the dense subspace S of the spinor space J{ = L2 ( § 2 , S). In
the present case, we shall exhibit a family of eigenspinors, called spinor har-
monics by Newman and Penrose [202, 361] and identified with "monopole
harmonics" by Dray [147].
Exercise 9.14. Prove the following identities for the ö operators, where
r,s E ~:

Öz(q- 1zr<-zr) = (l +-!- r)q- 1zY(-z)S+ 1 +rq- 1zY- 1 (-i) 5,

-az(q- 1zY(-zn = (l+-! -s)q- 1zY+ 1 <-zr +sq- 1zY(-z) 5 - 1. o


To construct the components 1/J± of an eigenspinor for I/), we may take
certain linear combinations of the terms q- 1zY ( -2) 5 , with land (r- s) held
fixed, and checkthat theselinear combinations obey the gauge transforma-
tion rules (9.46).
Lemma 9.24. Let cp be a smooth function on cx of the form

c/J(z,z) := (1 +zz)- 1 L a(r,s)zr(-i) 5


Y,sEN

{or some coefficients a(r, s) E C Then cp represents a section of S± if and


only if l +-!is a positive integer, and a(r, s) = 0 for r > l + -!
or s >
l±-!.Moreover, the coefficients must satisfy the symmetry relations a (r, s) =
(-l)l±~a(l +-!- r, l ±-!-
s).
9.A Spin geometry of the Riemann sphere 415

Proof. Suppose that cf> represents a section in roo (UN, S +). Then
c/>(z,i) = (i/z)Ii2c/>(z-I,z-I)
= (i/z) 1' 2 (zt) 1(1 +zi)- 1 L a(r,s)z-r(-i)- 5
r,sEN
= (-1) 1+iq- 1 L a(r,s)z 1-i-r(-t) 1+i-s,
r,sEN

where the exponents in the sum an the right hand side must also be non-
negative integers. Thus l- ~ E ~. and the nonnegativity of the exponents
an the right guarantees that r E {0, 1, ... , l- ~} while s E {0, 1, ... , l + ~ }.
The argument for sections in roo WN' s-) is similar. D

The structure of the symmetry relations among the coefficients, and the
allowed rang es of the exponents, suggests the introduction of the following
spinor components.
Definition 9.21. Foreach l E { ~. ~. ~ •... } and m E { -l, -l + 1, ... , l-1, l},
define

(9.57a)

(9.57b)

where the normalization constants Czm are defined as

C := (-l)z-m)2l + 1 (l+ m)! (l- m)!. (9.57c)


lm 4TT (l+~)!(l-~)!

With these components, we form the following spinors:

Y' ·- 1
lm .- J2 (Yz~)
iYz~ '
" . J21(-YiYz~1~) .
Yzm .=

1t is clear that the various Y1~ are linearly independent; in fact, as we shall
soon see, the coefficients have been chosen to make them an orthonormal
family. The same is true of the Y1~ functions.
The notation Y1~ quite deliberately brings to mind the everyday spheri-
cal harmonics Yzm that form an orthonormal basis for L2 (§ 2 ). Indeed, the
spinor harmonics, as originally introduced in [361) and further studied
in [202), were denoted s Yzm with s E { -l, -l + 1, ... , l- 1, l}; the case s = 0
are the ordinary spherical harmonics, and the cases s = ± ~ are our spinor
component functions.
416 9. Commutative Geometries

Lemma 9.25. With 1, m as in Definition 9.21, the relation ÖzYi~ = (1+ ~ )Yi~
holds.

Proof. From Exercise 9.14, it follows that Öz Y1~ (z, i) equals

+ (j + 1) G: t) c~ ~) J 2 J (- t) k.

The right hand side equals (1 + ~) Yi~n, as the term in brackets simplifies to

(1 + c u c c
~) ~ ~) ~ t) + (l + ~) ~ ~) ~ ~)
=( 1 +~)c~~)c:~). D

Exercise 9.15. Show likewise that az Y1~ = - (l + ~) Yi;,1•

Corollary 9.26. The spinors Yim and Yi:n are eigenspinors for the Dirac Ope-
rator:

and each nonzero integereigenvalue ±(1+~) has multiplicity at least (21+ 1 ).

Proof. The explicit form (9.55) for QJ shows that

Y' __ ...!:_(~ Öz)(y~~)=_!__((1+~)YI~)= 1 +! Y'


QJ Im- J2 Öz 0 iYi~ J2 i(1 + ~)Y1~ ( 2 ) Im•

and a similar calculation shows that QJY[:n = - (l + ~) Y/:n. For each eigen-
value ± ( 1+ ~), the ( 21 + 1) possible values of m yield linearly independent
eigenspinors. D
To finish the job, we must establish that this family of eigenspinors is
complete. This is best achieved by appealing to the representation theory
of the compact Ue group SU(2); see [54,287,439] for generalities on rep-
resentation theory and [35] for a useful bestiary of formulas about SU(2).
The properties of the spinor harmonics derive in large part from the cir-
cumstance that the sphere § 2 is the homogeneous space SU(2)(U"; but we
shall not bother to develop this viewpoint here, beyond what we need for
the spectrum of the Dirac operator.
9.A Spin geometry of the Riemann sphere 417

The group SU(2) acts on the Riemann sphere §2 = Coo by Möbius trans-
formations:
g . z = ( ab_
-
~) . z := -a;z+a
a
+ b_'

where aä + bb = 1 in orderthat g E SU(2). These are the rotations of


the sphere, since they preserve orientation and take any antipodal pair of
points {z, -1/.i} to another antipodal pair: just checkthat g · (-1/.i) =
-1 I g · z. Note, in passing, that since g · z =
z if and only if g = ± 1, this
provides a quick proof of the double covering SU(2) - S0(3). The group
SU(2) acts on spinors as follows:
b.i+ä)1/2 ---
T(g)tfJJJ(z,.i) := ( bz + a tfJJJ(g-1. z,g-1. z).

Elements of SU(2) are parametrized by three Euler angles (O<, ß, y), with
a = exp( t (O<+y)) cos ~ ß, b = exp( t (O<-y)) sin ~ ß; in this way, any element
of SU(2) is of the form k(O<)h(ß)k(y), where
eiOt./2 0 )
k( 0<) = ( 0 e-iOt./2 '

Proposition 9.2 7. The Dirac operator I/) is equivariant under the action of
SU(2), that is, T(g)l/) = l/)T(g) on S for any g.
Proof. It is enough to check the cases g = k(O<) and g = h(ß). In view
of (9.55), we need only show that the even operators T(g) on S are in-
tertwined by Öz: s- - S+, since T(g)Öz = ÖzT(g) entails ä"zT(g- 1 ) =
T(g- 1 )ä"z. For that, notice that both T(k(O<)) and T(h(ß)) are of the ge-
neral form
tfJf;,(z, .i) ..... j(z,.i) 1' 2!fJj;,(g(z),g(z)),
tfJ"N(z,.i) ..... j(z,.z)- 1' 2tfJr:,(g(z),g(z)).

lt is readily checked that such a transformation is intertwined by Öz if and


only if f and g satisfy the following differential equations:

(1 + z.i) :~ = j(z,.i)(1 + g(z)g(z)),

(1 + z.i) ~~ = j(z, 2) 2g(z) - .i j(z, .i). (9.58)

In the case of T(k(O<)), the functions j(z,.i) = e-iOI. and g(z) = e-i 01 z
satisfy these equations. For the transformation T(h(ß)), we get

_ .i sin ~ ß + cos ~ ß z cos ! ß - sin ! ß


j(z, z) = 1 1 , g(z) = 2 2
z sin 2 ß + cos 2 ß z sin ~ ß + cos ~ ß'
that also provide solutions to (9.58). 0
418 9. Commutative Geometries

The normalized Haar measure on the compact group SU(2) is given by


dg = (16rr 2)- 1 sinßdocdßdy. The irreducible unitary representations of
this group are labelled by nonnegative half-integers j E {0, ~. 1, ~. 2, ... };
for each j, the matrix elements of the corresponding representation are the
functions 'D-!nn, indexed by m, n E {- j,- j + 1, ... , j - 1,j}, defined by
(j + n)! (j- n)! i(na+my)( · !ß)2j
'D-!nn(oc,ß,y) := . )I ( . -
( J+m.J ) Ie sm 2
m.

x'(-1)j+m-r(J+m)( j-m )(cot!ß)2r-m-n.


Lr r r-m-n 2

The Parseval-Plancherel formula [287] shows that the functions 'D-!nn form
an orthogonal basis for L 2(SU(2)):

f
SU(2)
lh(g)l 2 dg =
2j=0
00

2: <2J + 1) m,n=-
2:
j

j
I <'D-!nn I h) 1 2.

Ifwe compare the definition (9.57) of the functions Y1~ with that of 'D-!nn.
and use z = e-iOI cot ~ß. we find the equalities [202]:

+ -
ylm(z,z) !2i+l 'Dl-~.m ( -oc, ß,-oc')
= V4IT
- - ffll+1
Y1m(z,z) = - --'D
4
l
1
7T 2'm
(-oc,ß,-oc).

Proposition 9.28. The spinors { Y[m, Y[:n : l- ~ E ~. m = -l, ... , l} form


an orthonormal basis for the spinor space L2 (§ 2 , S).
Proof. We associate to each spinor «1J E S a pair of functions h± on SU(2)
by
h±(oc, ß, y) := e+i(OI+y)/2 «JJ"fl(z, i) = e±i(OI-y)/2 «JJfC(, z;),
keeping the identifications z = e-i 01 cot ~ ß, ( = ei 01 tan ~ ß. By integration
over the y variable, we see that h + is orthogonal to 'D-!nn unless m = - ~,
and h- is orthogonal to 'D-!nn unless m = +~. Notice that lh±l 2 is indepen-
dent of y, so that fsu(2) lh±(g)l 2 dg = (4rr)- 1 fs2I«/J"fii 2 0. The Parseval-
Plancherel formula then shows that
II«JJII 2 = («<J+ I «JJ+> + <«JJ- I «JJ-> = 4rr JSU(2) (lh+(g)l 2 + lh-(g)l 2) dg
= 2: 4rr(2l + 1)( I('D~tm I h+) 1
2 +I ('D~.m I h-) 1 2)
l,m - -

= 2: Is2 y~~ «/J + o 2 + Is2 y~~ «/J- o


l,m
I 1 I 1
2

= 2: Y[m I «/J) 2 + Y[:n I «/J) 2 ·


I ( 1 I ( 1

l,m
9.A Spin geometry of the Riemann sphere 419

This is a Parseval identity, showing that the family of spinors {Yzm• Y[:n} is
orthonormal and also establishes its completeness. D

Therefore, there are no other eigenspinors for I/) than those we have
already found, so we now know the full spectrum with its multiplicities.
Corollary 9.29. The spectrum ofthe Dirac operator on the sphere § 2 is given
by

sp(QJ) = { ±(l + ~): l E 1"\:1 + ~} = 7L \ {0}. (9.59)

Each eigenvalue ±(l + ~) has multiplicity 2l + 1. E3

The Uchnerowicz formula (9.56) gives at once the spectrum of the spinor
Laplacian:
sp(ß5 ) = {l 2 + l- i:
l E 1"\:1 + ~ },
with respective multiplicities 2(2l + 1).
A good deal is known about the spectra of Dirac operators on other
particular manifolds. For the sphere §n with the usual metric, the spec-
trum [19, 457] is given by spQJ = { ( ~n + k) : k E 1"\:1 }. The respective
multiplicities are 2Ln/ 2 J (n+t- 1 ); these may be found by constructing eigen-
spinors for I/) by induction on the dimension of the sphere [63].
The eigenvalue 0 is missing from the spectrum (9.59); that is to say,
the Dirac operator on § 2 has no "harmonic spinors". This is a particular
case of the general phenomenon remarked after Corollary 9.17. Concretely,
(I/)+) 2 = -öz"8z and (QJ-) 2 = -8zöz both have spectrum { (l + ~ ) 2 : l E
1"\:1 + ~} = {1, 2, 3, ... }, since Proposition 9.28 implies that the Yi~n yield or-
thogonal bases for the Hilbert spaces J{± = L2 (§ 2 ,S±). The index equals
zero.
One can easily obtain generalized Dirac operators on the two-sphere with
nontrivial kernels by twisting S with a rank-one module over C"" ( § 2 ); for in-
stance, this procedure is developed in [338], where elements of the twisted
spinor module are called "Pensov spinors". That is to say, to get a nonzero
kernel, it suffices to replace the spin structure by some other spinc struc-
ture. The following series of exercises explores the details.
Exercise 9.16. Identify elements of 'E(m) with pairs of functions fN(z, i),
fsC(,,f,) satisfying fN(Z,i) =
(i/z)m1 2j 5 (z- 1 ,z- 1 ). Show that, for each
m E 7L, there is a connection v(m) on 'E(m) determined by
n(m)
v qoz =
a 1 -
q z + -zmz,

Compute the curvature of this connection.


Exercise 9.17. If \7 := \7 5 ® l:E(m) + 1s ® v(m) on s ®c~(§2) 'E(m). and if
IJJm := -i(c o \7) is the corresponding generalized Dirac operator, show
420 9. Commutative Geometries

that it can be expressed over UN as

Wm =: ( 0+ QJ~) . (- 01 Öz + ~mi)
Wm 0
= -t
Öz- 2mz 0 .

Show also that QJ-::n = -iq(m+ 3l1 2 · az · q-(m+ll/ 2 and QJ~ = -iq-(m- 3l/2.
az. q(m-l)/2. 0

Exercise 9.18. If m < 0, show that any element of ker w-::n is of the form
a(z) q(m+ll/ 2 where a(z) is a holomorphic polynomial of degree < Im I;
whereas, if m ~ 0, then ker QJ-::n = 0. If m > 0, show that any element
of ker QJ~ is of the form b(z) q-(m-ll/ 2 where b(z) is an antiholomorphic
polynomial of degree < Im I; whereas, if m ::50, then ker QJ~ = 0. Conclude
that the index of Wm equals -m in all cases. 0

~ Wehave derived some properties of the Dirac operator on the sphere § 2


from the representation theory of the group SU(2). Butthisgame can be
turned around. The previous exercises are not a mere divertimento: they
indicate that the bundle actions of SU(2) over the 'E(ml• at the Ievel of
sections, restriet on the index space of Wm to the finite-dimensional, unitary
irreducible representations of the group. This is clearly aspinvariant of the
Borel-Weil construction.
The same principle works for any compact, connected, semisimple Lie
group. (The remainder of this section presupposes some knowledge of the
structure theory of Lie groups, as found in [54], for instance.) Let G be
such a group, with Lie algebra g, on which the group acts by the adjoint
action Ad( · ); the Lie algebra is endowed with an Ad(G)-invariant Killing
form, given by (X, Y) :=- tr(ad(X) ad(Y)) -that can be used to identify g
with g*. Let T denote a maximal torus of dimension l (the rank of G), and
Iet t be its Lie algebra (a Cartan subalgebra of g). There is an orthogonal
decomposition g =: t 6) m, where m is the space of root vectors. The Weyl
group is the finite group W := N IT, where N is the normalizer ofT in G.
The (lag manifold GI T is the typical orbit of maximal dimension of the
adjoint representation; its dimension is always even, and it is always simply
connected. (Its decomposition as a cell complex can be described purely in
terms of the compact group G, i.e., without invoking complex manifold and
complex group theory, at the price of a little Morse theory [385).) On GIT
we can define several vector bundles by means of the associated bundle
construction. Denote by p the restriction of the adjoint action to the maxi-
mal torus. Consider the principal T -bundle G - GI T and then the bundles
G Xp g, G Xp t and G Xp m, which are respectively the spaces of orbits of
G x g, G x t and G x m under the action p(t)(g,X) := (gt,Ad(t)- 1X) ofT.
There are bundle isomorphisms

G x p g - GI T x g - T g Ic 1T and G x p t - GI T x t - N (GI T),


9.A Spin geometry of the Riemann sphere 421

where these trivializations are given respectively by [g, X] ..... (gT, Ad(g)X)
for XE g and [g,H] ..... (gT,H) for HE t. It is not hard to see that G Xpm-
T(G/T). Therefore, T(GfT) - (G/T x g) e (G/T x t) is stably trivial. This
has an important consequence.
Proposition 9.30. Flag manifolds are spin manifolds in a unique way.
Proof. This generalizes a well-known property of spheres. The Whitney
product formula [341, §4) gives, for the total Stiefel-Whitney classes w E
H"(G/T, ~2),
w(G/T x t) w(G/T) = w(G/T x g).
But w(G/T x t) = 1 and w(G/T x g) = 1 since these bundles are triv-
ial. Therefore w(G/T) = 1, that is, w 0 (G/T) = 1 and w 1 (G/T) = 0 in
Hi(G/T,~ 2 ) for j = 1,2, .... (The same argument gives triviality of the
Pontrjagin classes of the flag manifold.) Now w 0 = w 1 = 0 shows that G/T
carries a spin structure; moreover, H 1 (G/T.~ 2 ) = 0 since G/T is simply
connected, so this spin structure is unique (except for reversal of orienta-
~~- 0
From here, one can proceed to the construction of the irreducible uni-
tary representations of G on the index spaces of Dirac operators. There is
in factasimple relationship between the spinor-based and the Borel-Weil
constructions: one passes from one to the other by twisting or untwisting
with a fixed line bundle that can be obtained directly from the spin struc-
ture of the maximal (co)adjoint orbit GI T; in fact, it is a square root of the
so-called canonicalline bundle An,o(G fT) [7). Now, the Borel-Weil theorem
yields the archetypical geometric quantization construction. Let us digress
further to pointout why and how the Kostant-Kirillov-Souriau "geometric
quantization" scheme can be replaced by a Dirac operator mechanism. The
new approach is a natural variant of the old one, wherein attention was
concentrated on spaces of (cohomology classes of) holomorphic sections
of line bundles. But "holomorphic" means only belanging to the kerne! of
a particular differential operator, and one can replace it profitably by the
Dirac operator; the variants of Borel-Weil-Bott theory for compact groups
that use the Dirac operator are actually cleaner than the holomorphic treat-
ment.
In [472), Vergne noted the "formal analogy" between Kirillov's univer-
sal formula for characters (of representations associated to orbits of the
coadjoint representation) and the index formula for twisted Dirac opera-
tors. To quote exactly: "... at least for orbits of maximal dimension with
compact stabilizers and spin structures, all these indications would lead
us to discover (as Christopher Columbus 'discovered' America) the impor-
tance of the twisted Dirac operator on orbits to construct the quantized
representation ... " As a logical outcome, she was led to propose the re-
placement of polarizations in geometric quantization with operators of the
Dirac type [473).
422 9. Commutative Geometries

Let L be a prequantum line bundle over a spinc manifold M and Iet Dr


be a twisted Dirac operator for L. A quantization of M is the virtual Hilbert
space
Jfv,L := ker Dt - ker Di..

The index theorem gives precisely the dimension of such a space. If M is


already symplectic, it can be made spinc by use of a compatible almost
complex structure; the principal symbol of the associated Dirac operator
is the orientation dass in K-theory, and the index does not depend on the
choice of almost complex structure -see [189]. Experience suggests that
the quantization process can be given a richer and more informative for-
mulation for systems with a high degree of symmetry. Let p: G x M - M
be a group action. A vector bundle E ..!!... M is called a G-bundle if there is
a left action T: G x E - E such that each (T (g), p (g)) is a vector bundle
morphism on E- M, i.e., if the diagram

T(g)
E-E

rr! p(g) !rr


M-M

commutes, and each T(B)x is linear on the fibre Ex. In the hermitian case,
each T(B)x is unitary. If E is a superbundle, we suppose that the bun-
dle action of G is even (so both the even and the odd subbundle are G-
homogeneous). From G-bundles, proceeding in the usual way, one can fab-
ricate the ring Kc (M) of G-equivalence classes of vector bundles over M.
If p is trivial, then Kc(M) "" K(M) ® K(G) "" K(M) ® R(G), where R(G)
is the ring of virtual representations of G. Typically, however, Kc (GJT) i=
K(GJT) ® R(G), and as is weil known [424], R(G) ""R(T)w.
The bundle action gives rise to action on the spaces of sections in an
obvious way. A pseudodifferential operator P: f (E) - f (F) is a G-operator
if P(g · j) = g · P f, for all f E f(E) and all g E G. The basic observation is
that, if P is a G-operator, then ker P and coker P are representation spaces
for G. This is precisely the case for the Dirac operator in the equivariant
context. A finite-dimensional representation is completely determined by
its character [286]. Given P and g E G, define

indexp (g) := tr g iker p - tr g lcoker p.

An equivariant Atiyah-Singer formula gives an explicit way for computing


indexp(g), usually in terms of an equivariant Chern character. Vergne's
"universal formula" in [473] is supposed to accomplish precisely that. In
summary, quantization could be construed as a map in K -theory:

Q: Kc(M)- K(G) ""R(G),


9.B The Hodge-Dirac operator 423

and eventually, in a noncommutative context:

Q: Kc(.Jl)- R(G),

for .Jl a pre-C*-algebra, belanging to a suitable spectral triple (.Jl, Jf, D)


on which the Lie group G acts.

9.B The Hodge-Dirac operator


If (M, g) is an orientable Riemannian manifold that is not necessarily spin
or spinc, we can still construct reducible Clifford modules and define gen-
eralized Dirac operators on them. The simplestand best-known example
is the whole de Rham complex .Jl• (M) of differential forms on M, already
mentioned in the first section. Once an orientation is chosen, the Riemann-
ian volume form v9 makes this a prehilbert space, as already mentioned in
Section 7.2; the scalar product of forms is given by

(cxlß):= L(cxlß)v9 if cx,ßE.Jl.(M). (9.60)

Campare (7.22): here we replace the Riemannian density lv9 I by the volume
form, using the orientation. The integrand is a C"" (M)-valued (sesquilinear)
pairing of forms; for 0-forms, i.e., smooth functions, it is just ör.ß, and for
1-forms it is ( cx I ß) := g- 1( ör., ß), as in (7.2b). More generally, we define
(cx1 1\ ••. 1\ cxk I ß1 1\ .•• 1\ ßt) := 8kt det[ (cxi I ßi)] = 8kt det[g-1 (ör.i' ßi)]

for cx 1, ... , cxk, ß 1, ... , ß 1 E .Jl 1(M); this recipe extends sesquilinearly to a
positive definite pairing on .Jl• (M), whereby forms of different degrees
are orthogonal. The completion of .Jl• (M) in the scalar product (9.60) is a
Hilbert space L 2 •• (M), which is a direct sum of subspaces EBr=o L 2·k (M) of
"square-integrable forms" of each degree.
We keep our working assumption that M is compact; in the noncompact
case, it is clear that a Hilbert space can be similarly defined by completing
the space .Jl~ (M) of compactly supported differential forms.
The Clifford action on .Jl• (M) was given in formula (9.1), for 1-forms.
Obviously the Levi-Civita connection is a Clifford connection for this ac-
tion. The Hodge star operator on .Jl•(M) is * := c(y), the representative
of the chirality element of f"" (I(](M) ). If 9 1, ... , 9n is an oriented local or-
thonormal basis of 1-forms with dual vector fields Ej := (9i) ~, then the
star operator may be locally expressed as

Our conventions differ slightly from those of most differential geometry


books, in that we have included some powers of i in the definition of the
424 9. Commutative Geometries

star operator in order that * * = 1 in all cases, instead of the usual degree-
dependent sign ( -1 )k(n-k) for the square of * on k-forms. The reader can
rest assured that this is the only deviation from custom, by doing the fol-
lowing exercise.
Exercise 9.19. If K = {ii < · · · < }k} and K' = {i1 < · · · < in-kl are
complementary subsets of {1, ... , n}, if 11 KK' is the sign of the shuffle per-
mutation and §.K := 9JI 1\ · · · 1\ 9ik, show that
* §.K = (-i)m(-l)nk-k(k+l)/2/']KK' §.K'. 0

Example 9.4. If M = § 2 with the usual metric g = d8 2 + sin 2 8 dc/> 2 and area
form v = sin 8 d8 1\ dcf>, we can use 9 1 := d8, 9 2 := sin 8 dcf>, so that
*1 = -iv = -isin8d81\ dcf>, *d8 = isin8dcf>,
and therefore *V= i and *(sin8dcf>) = -id8. Alternatively, ifwe use the
local basis 9 1 := q- 1 dz, 9 2 := q- 1 di, with v = 2iq- 2 dz 1\ di, then
* 1 = -iv = 2q- 2 dz 1\ di, * (q- 1 dz) = iq- 1 di.
Exercise 9.20. For cx, ß E .Jl k (M), show that
cX 1\ *ß = (-i)m(-l)nk-k(k+l)l2(cx I ß) V (9.61)
by using local bases as in Exercise 9.19. 0
Both sides of (9.61) are n-forms; by integrating over M, we obtain

(cx I ß} = im(-l)nk-k(k+l)/2 IM cX 1\ *ß. (9.62)

The isometric property of the Hodge star operator is easily deduced from
(9.62). Note that, with our conventions, * cx = (-1) m * ä; thus,

(*CX I *ß} = im(-l)(n-k)(n+k-1)/2 IM *CX 1\ ß


= (-i)m(-l)(n-k)(n+k-1)/2 IM *cX 1\ ß

= (-i)m(-l)(n-k)(n+k-1)12(-l)k(n-k) IMß 1\ *cX


= (-l)m(-l)n(n- 1)/ 2 (ß I cX} = (ß I cX} = (cx I ß},

since ~n(n- 1) = m(2m + 1) = m mod 2. Therefore * extends to the


Hilbert space L 2·" (M) as a unitary operator. Since * * = 1, it is also selfad-
joint; that is, the Hodge star is a grading operator on this Hilbert space. It
exchanges the subspaces L 2·k(M) and L 2·n-k(M).
~ The adjoint dt of the exterior derivative, with respect to the scalar pro-
duct (9.60), maps L2·k (M) into L 2·k- 1 (M), and maps L2•0 (M) = L 2 (M, v9 ) to
zero. It is usually called the codifferential on the de Rham complex, since
it also maps .Jl k (M) into .Jl k- 1 (M), in view of the following identity.
9.B The Hodge-Dirac operator 425

Lemma9.31. dt = (-1)n*d* onL 2·"(M).

Proof. lf oc E .Jtk(M) and ß E .Jtk- 1 (M), then d(ß 1\ *OC) = dß 1\ *OC +


( -1)k- 1 ß 1\ d( * oc) is an exact n-form, whose integral vanishes, by Stokes'
theorem. Therefore,

(dß I oc) = im(-l)nk-k(k+1)/2 L dß 1\ *OC

= im(-l)nk-k(k+1)/2(-1)k JMß 1\ d(*OC)

= (-l)nk-k(k+1)/2(-1)k(-l)n(k-1)-k(k-1)/2 IM (ß 1 *d*OC) v

= (-l)n(ß I *d*OC),
on using (9.62) and the case (k- 1) of (9.61). 0

The operator dt d maps C"' (M) into itself; (7.23) shows that dt (dj) =
- div(gradj) = D.rBf, where D.rB denotes the Laplace-Beltrami operator.
We can now extend this operator to forms of any degree. Notice that dt f =
0 for any 0-form f, so that D.rBf = d(dt j) + dt (dj) also.
Definition 9.22. The Hodge-de Rham Laplacian is the operator D.H on the
Hilbert space L2·"(M), with domain .Jt "(M), defined as

(9.63)

Notice that D.H takes .Jt k (M) to .Jt k (M), since d raises and dt lowers de-
grees by one. A k-form 11 is called harmonic if D.HI1 = 0. lt is obvious from
the definition that D.H is formally selfadjoint. It is in fact essentially selfad-
joint [197]; we denote also by D.H its selfadjoint closure. The kernel of D.H
is just ker d n ker dt; indeed,

(w I D.Hw) = (w I ddtw) + (w I dtdw) = (dtw I dtw) + (dw I dw) ~ 0,


(9.64)

so that D.Hw = 0 implies dw = dt w = 0.


The inequality (9.64) also shows that D.H isapositive selfadjoint operator.
1t commutes with *• d and dt. Indeed, *ddt = (-l)n*d*d* = dtd*
and *dtd = (-1)nd*d = ddt*, from which *D.H = D.H* follows. Also,
dD.H = ddt d = D.Hd since d 2 = 0, and likewise dt D.H = dt ddt = D.Hdt.
Weshallsee soon that D.H is an elliptic second-order differential operator,
so that the space of harmonic forms is finite-dimensional and contained in
.Jt • (M), and that for each k = 0, 1, ... , n, there is an orthogonal direct-sum
decomposition
426 9. Commutative Geometries

In particular, any closed k-form w can be written as w = doc + dt ß + 17 with


17 harmonic, and then ddt ß = dw = 0, so that (dt ß I dt ß) = (ß I ddt ß) =
0 and thus dtß = 0. Therefore [w] = [17] in H~R(M), so the (complex)
de Rham dass of w is represented by its harmonic component 17; indeed,
[ w] - 17 is a bijective correspondence. The finite dimensionality (over <C) of
the de Rham cohomology groups H~R (M) "" ker D.H n .Jl. k (M) follows from
the ellipticity of D.H.
Definition 9.23. The isomorphism [17] ,_ [ * 17] from H~R (M) to H:lik (M)
is the Poincare duality for compact oriented manifolds. (lf one prefers
to match cohomology groups with real coefficients, one can use instead
[17] ,_ [im*l1] onHciR(M.~).)
Definition 9.24. The Hodge-Dirac operator on the compact oriented ma-
nifold M is the operator D := -i(d- dt), whose square is D.H.
In fact,
co\i'B=d-dt;
that is to say, the Hodge-Dirac operator is simply the Dirac operator cor-
responding to the Clifford action defined by (9.1). To prove that, we need
a couple of lemmata. By analogy with (9.16), we introduce the operators
i: .Jt 1 (M) ®.Jt .Jl."(M) - .Jl."(M) and i: *(M) ®.Jt .Jl."(M) - .Jl."(M) by
E(OC ® 11) := E(OC)I1 = OC 1\ 11 and t(X ® 11) := LXI1·

Lemma 9.32. Let \7 be a torsionless connection on the cotangent bundle


T*M- M. Then foranyw E .Jl."(M),

i('Vw) = dw. (9.65)


Proof. In view of the Leibniz rule, it is enough to establish this for 0-forms
(where it is evident) and for 1-forms. Write \7 w = ßk ®l7k E .J\. 1 (M)®.Jl. • (M);
then 'Vxw = ßk(X) 11k by contraction. Therefore, if X, Y E *(M) and w E
.J\. 1 (M), then

i(\lw)(X, Y) = (ßk "11k)(X, Y) = ßk(X)17k(Y)- ßk(Y)17k(X)


= (\i'xw)(Y)- (\lyw)(X)

= X(w(Y))- Y(w(X))- w('ViY- 'V~X)

= dw(X, Y)- w('ViY- \l~X- [X, Y]) = dw(X, Y),


where v~ is the dual connection to \7, defined like in (7.10), and in the
last line we have used the torsionless condition. This establishes (9.65) for
1-forms. D
The next lemma needs more structure.
Lemma 9.33. I( 'VB denotes the Levi-Civita connection on the cotangent bun-
dle of an oriented Riemannian manifold M, and dt: .Jl. • (M) - .Jl. • (M) is the
codifferential, then i o \i'B = -dt.
9.B The Hodge-Dirac operator 427

Proof. By the previous Lemma,

d =f o \7B = \lB o f - [\lB, f].

On a chart domain U with local coordinates x 1 , .•. , xn,

[\lgj,f(dxk)](a) = \lgj(dxk A a)- dxk A \lgi(a) = \lgj(dxk) A a


= -r~.dxJ
y A a = -r~.f(dxJ)(a)
y .

On the other hand, by Proposition 7.2 and (7.18),

Thus,

d = f(dxi) 0 \lgj = \lgj 0 f(dxi) + rfi f(dxi) = -(\lgi)t 0 i(dxi).

Since i:((\) = f(dxi)t, it follows, using a partition of unity, that 'i o \lB =
-dt. 0

Using the notation ajw := dxJ A w, the previous two lemmata can
be written in the language of creation and annihilation operators, as in
Chapters 5 and 6. They translate, respectively, to d = LJ a j \lg1 and dt =
- LJ aJ"ilg1 .
The operators d and dt are not elliptic, although they can be regarded
as operators in elliptic complexes [197]. However, it is clear from the last
two results that the Hodge-Dirac operator has principal symbol c(~) and
so is elliptic, and the Hodge Laplacian ßH, with principal symbol c(~) 2 =
g-1 (~. ~). is also elliptic.
If dimM is even, we can change the rules of the game by redefining the
grading of differential forms: one splits .J\. • (M) into the ( ± 1 )-eigenspaces
for the Hodge star operator: thus .J\. + (M) := { w : * w = w} is the space of
selfdual (orms, and .J\.- (M) := { 11 : * 11 = -17} is the space of antisel(dual
forms. Now dt * = - * d implies that D * = - * D, so the same operator D =
-i(d - dt) interchanges selfdual and antiselfdual forms. In this context,
the operator Dis called the signature operator.
If M is a spin manifold, the operator D is obtained from flJ as a twisted
Dirac Operator. The twisting isomorphism iss ®Jl s~ ""' .Jt•(M) (one can
use S itself as twisting bundle, since S and s~ are antilinearly isomorphic);
twisting the spin connection on S yields precisely the pure Levi-Civita con-
nection \lB on .Jt•(M). lf we treat s~ as a superbundle, we recover the
de Rham complex, whereas if we treat it as an ungraded bundle we obtain
the signature complex: see in this connection Lemma 3.3.5 of [197], which
in turn globalizes a classical construction by Brauerand Weyl [50].
The Hodge-Dirac operator has been much studied from the differential
geometry viewpoint. As for examples, for the n-sphere §n, Folland [185]
428 9. Commutative Geometries

constructed a complete set of eigenforms, using the rotational invariance


of the Hodge-de Rham Laplacian. For the 2-sphere, the spectrum is given
in [337]:
sp(-i(d- dt)) = { ±~l(l + 1): l E ~}
with respective multiplicities 2(2l + 1) for l = 1, 2, ... and multiplicity 2
for the zero eigenvalue. The harmonic forms are, of course, the constant
functions and the area forms; there are no harmonic 1-forms on § 2 • No-
tice that index(- i(d - dt)) = 2 - 0 = 2, which gives the Euler character-
istic x(§ 2 ) = 2. As a signature operator, however, it is weil known that
index(-i(d- dt)) = 0 on §2.
10
Spectral Tripies

The central role of the Dirac operator in the geometry of spin manifolds,
illustrated in the previous chapter, reveals the central thesis of noncom-
mutative geometry: that the structures we call geometrical are at the same
time, and perhaps more fundamentally, operator-theoretic in nature. The
transition to the noncommutative world entails putting the metric-genera-
ting operatorfront and centre. This modern approach to geometry is played
out an a stage which is a Hilbert space :Jf, on which act both an algebra 5\.
and an operator D; together, they form a spectral triple (5\., J-(, D). Guided
in part by index theory, we develop in this chapter the cohomological struc-
ture of spectral triples; from that structure there emerge several operatorial
properties that allow us to assemble the necessary data for noncommuta-
tive geometries.

10.1 Cyclic cohomology


In this section, 5\. is a unital algebra over IC. Our immediate concerns are
purely algebraic, so we shall not mention any topology an 5\. until the need
arises.
There are two main avenues to cyclic cohomology. The direct cohomo-
logical raute was pioneered by Connes [86] and is explained at length in
Chapter 3 of his book [91]. The homological approach, introduced by Tsy-
gan [461] and by Loday and Quillen [320], which is setforthin [53,272,319],
first defines cyclic homology and passes by duality to cohomology; it is
430 10. Spectral Tripies

better adapted to studying cyclic theories over general commutative rings.


Since we shall deal only with complex algebras and modules, we shallleave
the homology theory in the background.
We have seen already, in Section 8.4, the Hochschild cochain complex
(C(.Jt,.Jt*),b), where cn := cn(.Jt,.Jt*) = Hom(.Jt®(n+ll,C) is the set of
(n + 1 )-linear functionals on .Jt. We shall work with several operators be-
tween these sets of cochains. Wehave already met the Hochschild cobound-
ary Operator b: cn - cn+ 1 , given by (8.46), and the cyclic permuter t\: cn -
cn, given by (8.48). Let us abbreviate zn := ker(b: cn - cn+ 1 ) and Bn :=
im(b: cn- 1 - cn), so that HHn(.Jt) := zn;Bn is the nth Hochschild co-
homology module of .Jt. An n-cochain cp E cn is cyclic if Acp = cp. Defini-
tion 8.19 introduced the cyclic skewsymmetrizer N := 1 + A+ A2 + · · · +An :
cn - cn; notice that N(1- A) = (1- A)N = 0, since An+ 1 = 1 on cn.
Dropping the last term in (8.46) gives the truncated Hochschild cobound-
ary b' : cn - cn+ 1 ' namely
n
b'cp(ao, ... ,an+I> :=I (-l)lcp(ao, ... ,aJaJ+1·····an+1),
j=O

that satisfies b' 2 = 0. The difference r := b- b' is also useful [461]:

We recall the definition (8.48) of A on n-cochains:

One can then observe [53, 258] that

for j = 0, ... , n, and so


n n+1
b' =I "-J-1rAi, b = b' + r = I A-J-1rAi, (10.1)
j=O j=O

since An+ 1 = 1 on cn and A-n- 2 = 1 on cn+ 1 .


There are two "degeneracy Operators" s' s': cn+ 1 - cn' given by

scp(ao, ... , an):= cp(1, ao, ... , an),


s'cp(ao, ... ,an) := (-l)ncp(ao, ... ,an, 1).

(Here is where we make essential use of the unit of the algebra .Jt.) Of
course, s' = -sA - 1 , but it is handy to have both available. Their sum Bo :=
s+s' = s' (1-A) is an auxiliaryoperator (8.69) much used by Connes [86,91].
10.1 Cyclic cohomology 431

Definition 10.1. The Connes boundary map B: cn+ 1 - cn is given by


B :=NEo= Ns'(1- .\). (10.2)

This was introduced already in Definition 8.19, in the context of Hochschild


cohomology.
Lemma 10.1. The following identities hold on each cn:
b' (1 - ,\) = (1 - .\)b, bN =Nb'. (10.3a)
b' s + s b' = b' s' + s' b' = 1, (10.3b)
b' Bo + Bob =1- .\, bB + Bb = 0. (10.3c)

Proof. From (10.1) we obtain


n n
(1- .\)b = 2:: ,\-i-1y,\i + r _ 2:: ,\-iy,\i _ ,\-n-1y
j=O j=O
n n+1
= 2:: ,\-i-1y,\i- 2:: ,\-iy,\i = b'(1- .\).
j=O j=1

Next,
n+1
bN = 2:: "A-i- 1 r.\iN = (.\- 1 + · · · + .\-n-l + 1)rN = NrN
j=O
n
= Nr(1 + .\ + · · · + .\n) = N 2:: ,\-i- r,\i =Nb',
1
j=O

since .\N = N"A = N.


The chain homotopy identity b' s + sb' = 1 comes from dualizing the
calculation (8.40); let us check the second equality of (10.3b) explicitly:

(b's' +s'b')<p(ao, ... ,an) = (-l)nb'<p(ao, ... ,an,1) +b's'<p(ao, ... ,an)
n-1
= 2:: (-l)n+i<p(ao, ... ,ajaj+1·····an, 1) + <p(ao, ... ,an)
j=O
n-1
+ 2:: (-l)n- 1+i<p(ao, ... ,aiai+1•···•an,l),
j=O

so b' s' + s' b' = 1 by cancellation.


The identities (10.3a) and (10.3b) conspire to yield (10.3c):

b'Bo +Bob= b's'(1- .\) + s'(1- .\)b = (b's' + s'b')(1- .\) = (1- .\),
bB + Bb = bNs' (1- .\) + Ns' (1- .\)b = N(b' s' + s'b')(1- .\)
= N (1 - .\) = 0. D
432 10. Spectral Tripies

Definition 10.2. The sets of cyclic n-cochains er := er (.Jl, .Jl *) are pre-
served by the Hochschild boundary operator, since (1 - A)<P = 0 implies
(1- A)b<P = b'(1- A)<P = 0. Therefore, (e_\(.Jl,.Jl *), b) is a subcomplex of
(C(.Jl,.Jl*),b). The cyclic cohomology HC(.Jl) of the algebra .Jl is the
cohomology of this subcomplex. In other words, Hen(.Jl) is the quotient
of the cydic n-cocydes zr := zn n er by Br
:= b(er- 1).
In Hochschild cohomology, by contrast, one quotients all cocydes zn
by all coboundaries Bn := b(en- 1 ). So the modules Hen(.Jl) arenot the
same as the Hochschild modules H Hn (.Jl). lf <P E zr, we denote by [<P] E
HHn(.Jl) its Hochschild dass and by [<Ph E Hen(.Jl) its cydic dass. We
shall soon see the precise relationship between the two theories.
The letters He stand for "homologie cydique" [272].
Definition 10.3. The identities (10.3a) show that there is a bicomplex ec•,
introduced by Tsygan [461, Prop. 1] and developed in detail by Loday and
Quillen [321]:

b I I I I
1-.\
-b'
N
b -b'
1-.\ N
e2-e2-e2-e2-···

b I I I I
-b'
1-.\
e1-e1-e1-e1-···
N
b -b'
1-.\ N
(10.4)

b I I I I
-b'
1-.\
eo-eo-eo-eO-···
N
b -b'
1-.\ N

The row and column numbers startat 0, so eepq := eP (.Jl, .Jl *) for p, q E
~-In a bicomplex, all rows and all columns are complexes, and each square
anticommutes: (1- A)b + (-b')(1- A) = 0, bN + N(-b') = 0. Therefore,
we can form a total complex Tot• ee by lumping together the diagonals of
the bicomplex:
Totn ee := EB eepq = EB
eP.
p+q=n O,;;p,;;n

All arrows from one diagonal to the next constitute the boundary operator
of the total complex.
Actually, the bicomplex defined in [461] is the homological analogue
of (10.4), with .Jl ®(P+ 1l at each entry of the pth row, and with the arrows
running the other way. The bicomplex (10.4) has two special properties.
Firstly, every row is an acydic complex. Indeed, on n-cochains, ker( 1- A) =
er= imN and also im(l- A) = ker N. In order to see that NlfJ = 0 implies
10.1 Cyclic cohomology 433

1/J E im(l - i\), we solve the equation 1/J = cp- i\cp by putting

This allows us to define a contracting homotopy [349] with the pair of ope-
rators h, h': en - en defined by

h!JJ := _ _!_1 (i\ + 2i\2 + ... + ni\n)I/J, h ' cp := --1cp,


1 (10.5)
n+ n+
because h(l - i\) + Nh' = 1, which can be rewritten as h' N + ( 1- i\)h = 1.
Secondly, the odd-numbered columns ec(Zr+l), with the coboundary
maps - b', are contractible complexes, in the terrninology of Definition 8.13.
Indeed, (10.3b) says that the maps -s provide one contracting homotopy,
and the -s' maps yield another. (By defining s and s' on e 0 as the zero
map, (10.3b) becomes sb' = s'b' = 1 on e 0 .) The homological dual of any
such odd column is just the bar resolution (8.52) for Hochschild homology.
Given those properties, a standard procedure of homological algebra
shows that HC (.Jt) equals the cohomology of the bicomplex; the latter
is, by definition, the cohomology of the associated total complex. We aug-
ment the rows with the sets er = ker(1- i\), like this:

(10.6)

where i: er - en is the inclusion map. This yields an inclusion that we


also denote by i, of e_\ into the total complex Tot• ee, by insertion of cyclic
cochains into the top component of each diagonal:

i( cp) := (cp, 0, ... , 0) E en $ en -l $ · · · $ e0 ,


which intertwines the coboundary maps (since (1 - i\)cp = 0 by cyclicity),
and thereby we get homomorphisms i*: Hn(Tot ee) - Hen(.Jt).
Exercise 10.1. Show that the maps i*: Hn(Totee)- Hen(.Jt) are in fact
isomorphisms. o
434 10. Spectral Tripies

.,. In Connes' treatment -see [86] and [91, Ill.l]- cyclic cohomology is
computed from another bicomplex, whose differentials are the Hochschild
coboundary operatorband Connes' operator B. We regard the latteras the
composite B = Ns'(1- A), operating between even columns (only) of the
Tsygan bicomplex:
1-.\
cn+l ~cn+l

-b' l!s'
cn~cn.
The operator B thus appears as a bridge between the even columns, that
remains after the contractible odd columns are removed. The role of the
chain homotopy s' is to provide an isomorphism, at the Ievel of cohomol-
ogy, between the old bicomplex and the new one that remains.
The identity B2 = 0 is immediately seen, since (1-A)N = 0. On account of
(10.3c), Bandbare horizontal and vertical boundary maps for a bicomplex.
However, since B takes CCP· 2 r to CCP-l,Zr+Z, it is better to adjust the even
columns of the old bicomplex upward to make the map B horizontal.
Definition 10.4. Define BCPq := Ccp-q,Zq = CP-q (.JI., .JI. *) for p ~ q ~ 0,
BCPq := 0 otherwise. With the arrows B: BCPq - BCp,q+I and b: BCPq -
BCP+l,q, this yields the Connes bicomplex:

(10.7)

The cochains of its total complex are direct sums of Hochschild cochains
of the same parity:
Totn BC := cn Eil cn-z Eil cn- 4 Eil • • • Eil c#n,

where #n = 0 or 1 according as n is even or odd.


We indicate briefly the argument that the complexes Tot• CC and Tot• BC
yield the same cohomology; for a more complete argument in homological
10.1 Cyclic cohomology 435

notation, see [319, Lemma 2.1.6]. Take an element rjJ = (1/Jn, 1/Jn-2 •••• , 1/J#n)
in Totn CC (supported in even columns only), and consider

regarded as an element of Totn BC $ Totn- 1 BC. Since BN = 0, we get

(b + B)p = (b + B)(1 $ Ns') = (b + B) $ bNs' = (b + B) $Nb' s'

=(b+Ns'(1-.\))E9(N-Ns'b')=p( 1 ~.\ ~~)


The matrix on the right is recognizable as the coboundary operator for
two terms of the total Tsygan complex, Totn CC $ Totn- 1 CC. Thus, p ex-
tends additively to an operator from Tot• CC to Tot• BC that intertwines
the coboundary operators and therefore induces a homomorphism in co-
homology.
Exercise 10.2. Show that there is an exact sequences of complexes

0 -(C, -b') ..L Tot• CC _E.. Tot• BC- 0,


j being the inclusion of the first (odd) column in the Tsygan bicomplex. o
Since the complex (C, - b') is contractible, via the chain homotopy (-s'),
its cohomology is trivial, and p*: H• (Tot BC) - H• (Tot CC) is an isomor-
phism.
Corollary 10.2. H• (Tot BC) ""HC (.J\.). B

Example 10.1. We compute HC(C) using the Connes bicomplex. Multi-


linearity gives the relation <p(ao, ... , an) = ao ... an<p(1, 1, ... , 1), so that
cn"" c, with basis element <pn determined by taking <pn(l, 1, ... ' 1) := 1.
Clearly, b<pn = :Lj:J(-1)i<pn+ 1 = 0 or <pn+ 1, according as n is even or
odd. Also, (1- .\)<pn = 0 [n even] or 2<pn [n odd]; s'<pn = (-l)n-1<pn-1;
andN<pn-1 = :L~;::J(-l)r(n- 1 )<pn- 1 = O[neven]orn<pn- 1 [nodd].lnsum-
mary, B<pn = 0 or 2n<pn- 1, according as n is even or odd. A glance at the
bicomplex (10.7), where all entries are C, shows that the total complex is
of the form

where each d i is injective with one-dimensional cokernel; for instance,


d 2(<p3,<p1) = (<p 4 , ?<p2,2<p 0 ). We conclude that

HCn(C) = {C if n is even,
0 if n is odd.

Exercise 10.3. Compute HC(C) from the Tsygan bicomplex (10.4). 0


436 10. Spectral Tripies

.,. One very obvious feature of the bicomplex (10.4) is that its columns
repeat in pairs, and the map that shifts everything two columns to the right
is an isomorphism of complexes. The cokernel of such a map is identified to
the subcomplex consisting of the first two columns alone; since the first odd
column is contractible, this two-column complex has the same cohomology
as the zeroth column alone, that is to say, the Hochschild cohomology of .Jt.
We then getan exact sequence of bicomplexes:

o- cc· ~ cc· _!_ cc 0 e cc 1 - o,


and a corresponding exact sequence of their total complexes, with the maps

S('Pn-2. 'Pn-3 •... , 'Po):= (0, 0, 'Pn-2• 'Pn-3, ... , <po) E Totn CC,
H!J!n. !J!n-1. ... ' lJ!o) := (!J!n, !Jin-I> E ccnO $ ccn-1,1.
A standard procedure of homological algebra then yields the following lang
exact sequence in cohomology, originally described by Connes in [86]:

- HCn(.Jt) _!_ HHn(.Jt)..!!.. Hcn- 1(.Jt) ~ Hcn+ 1(.Jt) _!_ HHn+ 1(.Jt)-
(10.8)
where I is the canonical homomorphism from cyclic to Hochschild coho-
mology defined by J([<pJ.\) := [<p] for <p E Zf, S is the "periodicity ho-
momorphism" given by the two-step shift, and B is the connecting homo-
morphism. Since the complex (CC 1, -b') is contractible, the cohomology
of the first two columns isthat of (CC 0 , b), namely, the Hochschild coho-
mology of .Jt.
To describe the connecting homomorphism B more explicitly, and to
justify its name, we only have to apply the "snake Iemma" [310, III.9] to the
following diagram with exact rows:
s
0 - Totn- 1 C C - Totn+ 1 C C - ccn+ 1·0 e CCn 1

ldn-2 ldn
I
ld~
Totn- 2 CC ~ Totn CC ~ ccno e ccn- 1·1 - 0,

where s- 1dni- 1: kerd~- cokerdn-2 induces B: HHn(.Jt)- Hcn- 1(5\.).


Explicitly, we pull back (lJ!n,lJ!n-I> to its preimage (lJ!n,lJ!n-1,0, ... ,0) in
Totn CC, whose coboundary is (b!J!n, ( 1- A>!J!n- b' !J!n-1. N!J!n-1. 0, ... , 0).
Forthis to lie in the image of S, we need that

b!J!n = 0, (1- A)!J!n = b'!J!n-1. (10.9)

which is just the condition that d~ (!Jin, !J!n-1) = (0, 0). We end up with the
element (N!J!n- 1, 0, ... , 0) E Totn- 1 CC. Tothis result we are free to add a
coboundary term without changing the induced map in cohomology.
10.1 Cyclic cohomology 437

We can use that freedom to obtain the connecting homomorphism from


a cochain map B: cn- cr- 1. Suppose we Start afresh with t/Jn E zn, so
that btJJn = 0. Now b'(1- i\)tJJn = (1- i\)btJJn = 0, and it follows that
(1- i\)tJJn = (b's' + s'b')(1- i\)tJJn = b's'(l- i\)tJJn, so we mayintroduce
t/Jn-1 := s' ( 1 - i\) t/Jn in order to satisfy (10.9). Now NtJJn-1 = N s' ( 1- i\) t/Jn
meets our needs; this is precisely BtJJn as defined in (10.2). On the other
hand, if we had chosen t/Jn- 1 arbitrarily, subject only to (10.9), we would
find instead that N!JJn-1 = Ns'b'tJJn-1 + Nb's'tJJn-1 = BtJJn + bNs'tJJn-1•
so that N!JJn-1 = BtJJn mod BÄ- 1 in any case. In summary: the Connes map
B of (10.2) implements the connecting homomorphism in (10.8), which is
why we call it B, too.
Similar manipulations of chain homotopies enable us to determine a map
S: zr- 1 - zr+ 1 which implements S: Hcn- 1(.JI.) - Hcn+ 1(.JI.). Given <P E
zr- 1 it suffices to add a coboundary of the form dn ( tJJ 17 0,
1 I0) to its
I ",I

image (0, 0, <P. 0, ... , 0) in Totn+ 1 CC, so that in the sum

(btJJ, (1 - i\)tJJ- b' 17. <P + N17, 0, ... 0)I

only the leading term is nonzero. Using the homotopies (10.5), we obtain
<P = N h' <P + h ( 1 - i\) <P = N h' <P, since <P is cyclic; we may then take
17 := -h'<P = -(1/n)<P. Next, we must select a cochain tJJ E cn so that
(1- i\) 1/J = b' 17 = -b' h' <P· Since 1/J = Nh' 1/J + h( 1- i\) 1/J, it is enough to take
1/J := -hb'h'<P, because thenN(/J = -hNb'h'<P = -hbNh'<P = -hb<P = 0,
for <Pisa cocycle. We finally arrive at b(jJ = -bhb'h'<P E zn+ 1. Applying
1 - i\, we get

(1- i\)b!IJ = -b'(1- i\)hb'h'<P = b'h'Nb'h'<P = b'h'bNh'<P


= b'h'(b<P- bh(1- i\)<P) = 0,

SO that -bhb' h' <P lies in zr+l as desired,


I

We remainfree to add to it acoboundaryterminBÄ+ 1 , in order toidentify


a preferred cochain S<P. Before doing so, however, we pause to examine one
more way to set up cyclic cohomology.
~ One can introduce Hochschild and cyclic cohomology in a more "cate-
gorical" way, to allow for extensions of these theories beyond the realm of
unital algebras. The general idea is to characterize abstractly the b and B
Operators of the Connes bicomplex, as follows. The Operator b: cn- 1 - cn
may be written as b = If=o( -1)iöi, where

Öi<P(ao, ... ,an) := <P(ao, ... ,aiai+l·····an), i = 0, 1, ... , n- 1,


Ön<P(ao, ... ,an) := <P(anao, ... ,an-1).

We also introduce maps aj: cn+ 1 - cn I for j = 0, 1 I ... In, given by


438 10. Spectral Tripies

For convenience, we redefine the cyclic permuter on cn as

Tcp(ao, ... ,an) := cp(an,ao, ... ,an-I),

or, more simply, T := (-l)n,\. Notice that Tn+I = ,\n+ 1 = 1 on cn. Also,
N = Ir=o"k = Ir=o(-l)nkTk, while s' = (-l)n<Tn and s = T<ToT- 1 as
Operators on cn, SO that

where we have used the relation NT = ( -l)n N.


We can now give a "generators and relations" presentation of cyclic co-
homology by identifying the necessary relations satisfied by the various
Oi, <TJ and T maps. Leaving T aside for a moment, we first check that
the relations among the oi and Gj arise from the so-called simplicial cate-
gory. This is a small category, called ~. whose objects are the ordered sets
[n] := {0, 1, ... , n}, one for each n E ~.andin which the morphisms from
[n] to [m] are the nondecreasing maps f: [n] - [m]. For instance, let
oi: [n -1]- [n] be the injective map that misses i, and let <TJ: [n + 1]-
[ n] be the surjective map that identifies j and j + 1, for i, j = 0, 1, ... , n.
Any injective increasing map is a product of such Oi, and any surjective
increasing map is a product of such Gj; thus, these maps generate all mor-
phisms of ~. They are subject to the following relations (only):

OJOi = oiOJ-1 if i < j, (10.10a)


O'jO'i = O'iO'j+1 if i ::5 j, (10.10b)
Oi<TJ-I if i < j,
{
O'JOi= id ~f~=!orj+1, (10.10c)
Oi-10'j 1ft> J + 1.

Exercise 10.4. Verify the relations (10010) in the category ~0 Check that
they also hold for the homonymous maps between Hochschild cochainso 0
It is customary to call a cofunctor ~ - C a "simplicial object" in a given
category C; a (covariant) functor ~ - Cis called a "cosimplicial object" in C.
The latter amounts to finding a family of objects { cn : n E ~ } together
with morphisms oi: cn- 1 - cn and <TJ: cn+ 1 - cn for i, j = 0, 1, ooo, n,
satisfying (10olO)o The relations (lOolOa) show that b := I7= 0 (-1)ioi satis-
fies b 2 = 0, so ( C", b) is a cochain complex, whose cohomology generalizes
the Hochschild cohomology of algebras; indeed, for any unital algebra .J'I.,
the complex C" (.J'I., .J'I. *) constitutes a particular cosimplicial objecto
Another example is given by the family of standard n-simplices

~n := { t E ~n : 0 ::5 t1 ::5 o o o ::5 tn ::5 1}


::: { S E ~n+l : Sj ~ 0, So + 0 0 0+Sn = 1}, (10oll)
10.1 Cyclic cohomology 439

where the "face map" Öi: ßn-1 - ßn is the injective affine map whose
image is the ith face of ßn, and the "degeneracy map" ai: ßn+l - ßn
is the surjective affine map which collapses the edge joining the jth and
(j + 1 )st vertices to a point. These structures are discussed in many books
on homological algebra, for instance [482, Chap. 8].
To incorporate cyclic cohomology in this framework, Connes [85] intro-
duces a "cyclic category" i\ with the same objects as ß, but which also allows
as morphisms the cyclic permutations T: [n] - [ n] given by T (0) := n
and T(k) := k - 1 if k = 1, ... , n. This category is described in detail
in [91, liLA], [319, §6.1] and [482, §9.6]. Sufftee it here to say, following [319,
Thm. 6.1.3], that the morphisms of i\ are generated by the various c5i, ai
and T, subject only to (10.10) and the following extra relations:

Tc5i = c5i-IT: [n-1]- [n] for i = 1, ... ,n, and TÖo = 8n,
Taj=aj-IT:[n+1]-[n]forj=1, ... ,n, and Tao=anT 2 ,
(10.12)

Exercise 10.5. Verify the relations (10.12) in the category i\. Check that
they also hold for the homonymaus maps between Hochschild cochains. <>
The map a-1 := aoT- 1 : [n + 1] - [n], satisfying a_ 1(k) = k for k =
0, ... , n and a_ 1(n + 1) = 0, is sometimes called the "extra degeneracy"
(it is a morphism in i\ but not in ß); it satisfies Ta_ 1 := anT. A "cocyclic
object" is a category C is a functor i\ - C; that is to say, a cosimplicial
object {Cn} tagether with morphisms T: cn - cn satisfying (10.12). We
may define N := Lk=o(-l)nkTk on cn, and then B: cn+l- cn by

(10.13)

With bandBin hand, we proceed as before: write BCP'l := cp-q for p ~ q,


BCP'l := 0 otherwise. Then (Tot• BC, b + B) is a cochain complex, whose
cohomology is, by definition, the cyclic cohomology of the object {cn}. For
the particular case cn = cn(.J\, .J\ * ), this repeats the previous construction
of HC(.J\). Later on, in Chapter 14, we shall show how this categorical
viewpoint enables to define the cyclic cohomology of Hopf algebras [114,
116] with minimal extra effort.
~ We now return to the identification of the cochain Scp E zr+l corres-
ponding to cp E zr-l under the periodicity map. We define an auxiliary
cochain map y: cn-l - cn by
n
;y := :L (-1)kk8k.
k=l

Lemma 10.3. lfcp E zr- 1, then ;ycp = -(n + 1)hb'cp.


440 10. Spectral Tripies

Proof. From (10.12), we see that ?1.8i = -8i_ 1 ?1. for i = 1, ... , n. Therefore,
since b' cp = (-1) n- 18n cp because bcp = 0, we conclude that
nn
-(n + 1)hb'cp = (-l)n-1 =
2: (-l)n-k-1k8n-k?l.kcp
2: k?l.k 8ncp
k=1
k=1
n n
= 2:<-1)n-k-1k8n-k'P= 2:<-l)r-1(n-r)8rcp
k=1 r=O

= -nbcp + ycp = ycp,


using both ?l.cp = cp and bcp = 0. D

ltfollowsthatn(n+1)Scp = -n(n+1)bhb'h'cp = bycp,sinceh' = 1/n


on cn- 1 •
The following recipe [86) gives the periodicity operator S as a map of
degree +2 between cyclic cocycles.
Lemma 10.4. The periodicity homomorphism S: Hcn-1 (.Jl) - Hcn+1 (.Jl)
is implemented by the OperatorS: 1 - z;:+ 1 given by z;:-
1 n
Scp(ao, ... ,an+d :=- ( 1 ) 2: cp(ao, ... ,aJ-1aJaJ+1•· .. ,an+1)
n n+ J= 1
1
( 1) 2: (-1)t+Jcp(ao, ... ,ai-1ai, ... ,aJaJ+1•· .. ,an+d·
n n + 1:o>i<J:m
(10.14)

Proof. By definition, by = "Ii,J ( -1 )i+ J j 8i81 ; on cn- 1 , the sum ranges over
j = 0, ... , n and i = 0, ... , n + 1. By splitting that sum into 0 :5 i :5 j :5 n
and 0 :5 j :5 i- 1 :5 n and applying (10.10a), we obtain
by= 2: (-1)i+j(j-i)8i8j.
O:o>i:o>j:o>n

and the analogous computation for yb leads to


yb = 2: (-1)i+J(i- j - 1) 8i8j.
O:o>i:o>j:o>n

The terms with j = n in both sums add up to


n n
2:(-1)i+n- 18i8n = 2:<-1)i+n- 18n+18i = (-l)n- 18n+1b.
i=O i=O

Replacing bycp by (by + yb + (-l)n8n+1b)cp since bcp = 0, we arrive at

-n(n + 1) Scp =
O:o>i:o>j:o>n-1
(10.15)
10.1 Cyclic cohomology 441

The formula (10.14) is then obtained by separating out the terms with i = j.
Since the terms with i = 0 in the sums for by + yb add up to -8ob~ we
can also take the first sum in (10.15) over the range 1 5 i 5 j 5 n. Thusl
-n(n + 1) ?I.Scp = 2: (-1)i+JA8i8Jcp
lsisjsn
2: (-l)i+Jöi-IÖj-IA'P = -n(n + 1)Scpl
lsisjsn
0

Example 10.2. Consider the cyclic cocycle of Definition 8.17 1 namely1 the
character of an n-dimensional cycle (0" 1d 1J) over 5\:

T(aol ... Ian):= I ao da1 daz ... dan. (10.16)

Then -(n + 1)(n + 2)ST(aol .. ·1 an+z) equals


n+l
2:
j=l
I ao da1 ... d(aj-lajaJ+d ... dan+2

+ 2:
lsi<jsn+l
(-1)i+j I ao da1 ... d(ai-lai) .. . d(ajaJ+d ... dan+2

=
n+l
2:
j=l
I (ao da1 ... daj-1 (ajaj+l) + Xj d(ajaj+l)) daj+2 ... dan+2 1

where X1 = 0 andl for j = 21 .. 'In+ 11


Xj = (-1) 1.- l aoa1 da2 ... daj-1
j-1
+ 2: (-l)i+Jao da1 ... d(ai-Iad ... daj-1 + ao da1 ... daj-2 a j-11
i=2
which telescopes to zero. This yields the very handy formula [86]:

ST(aol' .. I an+2) (10.17)

(n
1 n+l
+ 1 )(n + 2 ) j~
I ao da1 ... daj-1 (ajaj+l) daj+Z ... dan+ 2.

We can rephrase the exactness of the long sequence (10.8) as the State-
ment that there is an exact triangle of complexes:
s
HC"(J\) HC"(.A.)

~/r
HH"(.A.)
(10.18)
442 10. Spectral Tripies

where S, I and B have respective degrees +2, 0 and -1. There is a standard
machinery in homological algebra for dealing with such exact couples: see,
for instance, [258, Chap. 1] or [245]. The diagram leads to a "spectral se-
quence" that enables one to compute effectively the cyclic cohomology of
many algebras. The first step is to define a "derived" exact couple, whose
lower complex is the cohomology of the originallower complex -in our
case, HH" (.Jt.)- under the differential given by the composition IB; notice
that BI = 0 entails (IB) 2 = 0. In other words, in the derived triangle, the
lower vertex is ker(IB)/im(IB). Now, we have already met the homomor-
phisms IB: HHn(.Jt.) - HHn-l(.Jt.) in Chapter 8, in connection with the
HKRC theorem: Proposition 8.18 says that when .Jt. = coo (M), IB = n a
equals (a multiple of) the de Rham boundary operator on currents. There-
fore, the derived exact couple for coo (M) has as lower vertex the homolog-
ical de Rham complex of M.
The exact couple (10.18) allows us to compute the cyclic cohomology of
the algebra C (M), for a compact manifold M. The following result [86]
00

moves the HKRC theorem (Hochschild classes correspond to de Rham cur-


rents) into the terrain of (co)homology proper.
Theorem 10.5. Foreach k E ~. there is a canonical isomorphism

HCk(C 00 (M)) == zeR(M) Ef!Ht~ 2 (M) Ef!He~4 (M) Ef! · · · Ef!Hgf(M), (10.19)

where zeR(M) is the set of closed k-currents an the compact manifold M,


H~R(M) is the de Rham homology group of degree r, and #k = 0 or 1
according as k is even or odd.

Proof. Recall from Theorem 8.17 that the isomorphism HHk(C 00 (M)) ==
1h(M) is implemented by [<p] ..... C<P, where

JCep aodaii\"·1\dak:= ~.1 2::


rrESk
(-l)rr<p(ao,arr(I), ... ,arr(k)). (10.20)

Westart with [<p h E HCk (C (M) ), represented by a cyclic cocycle <p E Zf;
00

then (1 1k)B<p = 0, so ac<P = 0; in other words, C<r is a closed current.


Let Äk<p denote the skewsymmetrization of <p given by (8.64) -so that
Ak<p(ao, a 1 , •.. , ak) equals the right hand side of (10.20)- which is also a
Hochschild k-cocycle. lt is in fact a cyclic cocycle, since
AAk<p(ao,aJ, ... ,ak) = (-l)kAk<p(ak,ao, ... ,ak-d
= (-1)k r
Jeep
daol\ .. ·1\dak-lak

= ( ao da 1 1\ · · · 1\ dak
Jeep
k-l
- 2:: (-1)1 J da 0 1\ · · · 1\ d(a1a 1+I) 1\ · · · 1\ dak.
j=O eep
10.1 Cyclic cohomology 443

so that AA.kcp = ÄkCfJ since Ccp is dosed. lt follows from the proof of Theo-
rem 8.17 that cp and ÄkCfJ are Hochschild-cohomologous, and their common
Hochschild dass corresponds to Ccp. We can now express this by saying that
I[cp- ÄkCfJh = 0.
However, cp and ÄkCfJ need not be cydic-cohomologous. Connes period-
icity tells us that [cp- ÄkCfJh E imS, so we can find «J1 E Zf- 2 suchthat
[S«Jlh = [cp- ÄkCfJh. Such a «J1 is not unique, but is determined modulo
ker S = imB, so the corresponding current CI/J E 'Dk- 2 (M) is determined
modulo the de Rham boundaries Bt~ 2 (M); that is to say, [CI/'1 E Hf~ 2 (M)

Next, Äk-2«/J E zr
depends uniquely on [cpl\.
2 and l[«jJ- Äk-2«/Jh = 0, and we may repeat this
process to find a cydic cocyde X E zr
4 with [Sxh = [«Jl- Äk-2«/Jh; and
so on. Let us write CfJk := ÄkCfJ, CfJk-2 := Äk-2«/J, and CfJk-2i for the skewsym-
metrized cocyde produced by the algorithm at degree k- 2j; also, let fk- 2j
be the (dosed) current corresponding to the Hochschild dass [CfJk-2j1· The
algorithm terminates when k - 2j = 0 or 1, according to the parity of k.
Therefore,

cp = L Si CfJk-2i mod Bf. (10.21)


0$2j$k

The skewsymmetric cocydes CfJk- 2j are recovered from the currents fk- 2i
by (10.20) in reverse:

CfJk-2j(ao,ai, ... ,ak-2j) := f aoda1 1\ • • · Adak-2i·


. J~-~
Therefore, the map taking [cph = [cpkl.\ + [Scpk-2h + [S 2CfJk-4h + · · · +
[Slk1 2Jcp#kh to fk + [fk-21 + [fk-41 + · · · + [f#Ü where fk = Ccp, is a well-
defined bijective homomorphism. D
Under the isomorphism (10.19), the periodicity map S: HCk(C""(M)) -
Hck+ 2(C""(M)) corresponds to

In other words, S takes the dosed current Ccp = rk to its de Rham dass, and
reproduces the lower-degree dasses. The Zf~ 2 (M) component of S[cph is
zero, since im S = ker I, so the Hochschild dass [S cp 1 is zero.
For k > dimM, the right hand side of (10.19) is just the full (even or
odd) de Rham homology complex of M, with S: Hck - Hck+ 2 yielding
the identity map in homology. This stabilization property motivates the
following definition.
Definition 10.5. The periodicity maps S: Hcn - Hcn+ 2 define two di-
rected systems of abelian groups; their inductive limits
444 10. Spectral Tripies

form a :;l 2 -graded group HP" (.Jt) := HP 0 (.Jt) EDHP 1(.Jt), called the periodic
cyclic cohomology of the algebra .Jt. (We follow the notation of [53) for
these groups).
Exercise 10.6. Show directly that HP 0 (C) = C and that HP 1(C) = 0. 0

An alternative definitionis to say that HP" (.Jt) is the quotient of HC (.Jt)


under the equivalence relation [<p],\ - [S<p],\. Although HC"(.Jt) is :;l-
graded, the quotient has only a :;l 2 -grading. (However, it conserves a natural
filtration that comes from dropping the leftmost columns of the Connes bi-
complex one by one.) We can now finish the rewriting of de Rham homology
in noncommutative language.
Theorem 10.6. The periodic cyclic cohomology of the algebra coo (M) is
canonically isomorphic to the :;l 2 -graded de Rham homology of M:

E3

Another important example of periodicity [86) concerns the Chern char-


acter of a finitely summable Fredholm module.
Proposition 10.7. Let (.Jt, J-{, F) be a Fredholm module of parity #n, such
that [F, a) E Ln+ 1(J{) for each a E .Jt, with character

Tn (ao, ... , an):= I ao da1 ... dan = !in Tr(XF[F, ao) ... [F, an]),

in the notation of Section 8.2, and Iet Tn+Z be the analogaus cyclic (n + 2)-
cocycle. Then S[Tnh = -(2n + 2)- 1 [Tn+ 2 h in Hcn+ 2 (.Jt).

Proof Consider the following (n + 1)-cochain [91, IV.l.ß):

!J!(ao, ... , an+d := in+ 1 Tr' <xFao [F, ad ... [F, an+l ]),

!
where Tr' T := Tr(T + FTF), as in (8.12). Then N!J! E Cf+ 1, so bN!J! E
Bf+ 2 • Weshallshow that STn+ (2n+2)- 1Tn+Z isamultiple of bN!J!, thereby
establishing that STn and -(2n + 2)- 1Tn+ 2 are cohomologous.
Using da:= i[F, a], we may write N!J! = :Lj:J (-1)(n+ 1U!J.11 , where

If j = 1, ... , n + 1, the coboundary b!J.I1 (ao, .. . , an+2) is given by

2:: (-1)k IFaJ+1 daJ+2 ... dan+2 dao ... d(akak+d ... daJ
j-1

k=O

2:: (-1)k I Fa J daj+1 ... d(akak+d ... dan+2 dao ... daJ-1•
n+2
+
k=j
10.1 Cyclic cohomology 445

which telescopes to

I FaJ+1 daJ+2 ... dan+2 (ao) da1 ... daJ

I
+ ( -1) 1.- 1 FaJ+1 daJ+2 ... dan+2 dao ... daJ-1 aJ

+ ( -l)n I FaJ daJ+1 ... dan+2 (ao) da1 ... daJ-1· (10.22)

When j = 0, the sum over k < j is void, so the telescoping gives only

blJ.Io(ao, ... ,an+2) = (-l)n I[F,an+2]aoda1···dan+1

=i I ao da1 ... dan+1 dan+2·

For j > 0, we abbreviate

so that dRJ = ( -l)n-J+ 1daJ+2 ... dan+2 dao ... daJ-1; then (10.22) becomes

I Faj+1RjdaJ + (-l)nFajdaJ+1RJ + (-l)nFaJ+ 1dRJaJ

= (-l)n I FdaJ aJ+1RJ + FaJ daJ+1 RJ + i(aJFaJ+1F + FaJFaJ+1)RJ

= i(-l)n I (2aJaJ+1- daJ daJ+1)RJ

= -2i(-1)(n+ 1)J I aoda1 ... daj-1 (aJaJ+1) daJ+2 ... dan+2

+ i(-1)(n+l)J I aoda1 ... dan+2·

Therefore, bNlJ.I = ~j:J(-l)(n+llJblJ.IJ = i(n + 2)((2n + 2)STn + Tn+2),


by invoking (10.17) (to which blJ.Io does not contribute) and summing over
all j. D

Definition 10.6. To banish the factor - (2n + 2)- 1 in the periodicity rela-
tion, it is enough to normalize the character by setting

TFn( ao, ... ,an ) .=


. -(-i)nf(n
-1- 2 + 1) Tn( ao, ... ,an )
n.
ni+1)
= 2 1 Tr(xF[F,ao] . .. [F,an]). (10.23)
n.
Then the result of Proposition 10.7 simplifies to S[Tpth = [Tpt+ 2h. There-
fore, {Tpt+ 2k}ho determines a well-defined dass [TF] in HP#n(.Jt). This
dass is the Chern character of (5\, Jf,F) in periodic cyclic cohomology.
446 10. Spectral Tripies

~ The cyclic cohomology of C* -algebras-as opposed to pre-C* -algebras-


is regrettably trivial in many cases, because cyclic (or Hochschild) cocycles
on a pre-C*-algebra often do not extend continuously to its C*-algebraic
completion. As evidence for that, we mention that an everywhere defined
derivation on a unital C* -algebra A is bounded [366, §8.6] and in many
cases is inner (for instance, if A is abelian or simple or a von Neumann
algebra), and so H 1 (A, A) = 0 by Exercise 8.22. Several variants of periodic
cyclic cohomology have appeared in the literature, which seek to avoid that
Iimitation. To get a nontrivial theory for C*-algebras, Puschnigg [381] has
constructed a theory of "asymptotic cyclic cohomology". In another direc-
tion, the "analytic cyclic cohomology" of Meyer [335] is developed in the
setting of bornological spaces. Both these theories satisfy the excision ax-
iom and can then be used to show that excision also holds in entire cyclic
cohomology, which we consider next.

10.2 Chern characters and entire cyclic cocycles


The initial development of cyclic cohomology by Connes was geared to
solve a problern in index theory. Fredholm modules, or more general K-
cycles, provide the ingredients for a theory dual to K-theory (that has come
to be called K -homology), and there is an index pairing that gives the du-
ality. In the even case, the pairing between Ko(.Jt) and the K-homology
group K0 (.Jt) is given as follows. One matches [p] E K 0 (.Jt), represented
by a projector p in Mr(.Jt), with the homotopy dass in K 0 (.Jt) of a Fred-
holm module (.Jt, J-{, F), by assigning them the integer index(p(F ® 1k)p).
If F := DIDI- 1 is the phase operator for a spectral triple (.Jt, J-(, D), we
may write the result as index(p(D ® h)p); these indices are equal since
D and F are connected by a homotopy t - Dt := DIDI-t for 0 ::5 t ::5 1,
and each p (Dt ® 1k) p is a Fredholm operator on the Hilbert space pJ-(k. In
the context of commutative geometry, when .Jt = C"" (M), this pairing gives
the index of a generalized Dirac operator on the vector bundle represented
by p.
The notationF := DIDI- 1 is apt when ker D = 0. Otherwise, we can work
instead with the pre-Fredholm module determined by F' := D(1 + D 2 )-112;
or we can redefine F := 1 on the (finite-dimensional) kernel of D and, to
conserve the index, we supplement J-( with a second copy of ker D and Iet
F act as a partial isometry that exchanges the two copies [91, IV.2.y]. We
shall adopt the latter convention in all that follows, and continue to write
IDI- 1 in the general case, by declaring IDI- 1 := 0 on ker D; then Fis a
canonical symmetry satisfying F := DIDI- 1 .
The practical problern of how to compute this index is solved, in the
case .Jt = C""(M), by passing from K-theory to de Rham cohomology [16]
by the Chern isomorphism of Corollary 8.8, and the index pairing is given
10.2 Chern characters and entire cyclic cocycles 447

by (D, p) ,.._ IM eh p l(D), where l(D) is a characteristic dass depending


only on D. The de Rham homology dass of the current w ,.._ IM w l(D)
is then given by a periodic cydic cohomology dass on C"" (M), in view of
Theorem 10.6, which is none other than the Chern character chF := [TF].
This pairing of K0 (.J'l) with HP 0 (.J'l) -and a similar pairing of K1 (.J'l) with
HP 1 (.J'l)- is the main achievement of the seminal paper [86] of Connes.
~ One can reformulate this result by defining a Chern homomorphism di-
rectly from the K-theory of .J'l to the dual theory of periodic cydic homo-
logy, and write the index in the form of a pairing (eh p, chF). Weshall not
go into that, except to say that there are obvious duals to the Tsygan and
Connes bicomplexes, with reversed arrows, whose columns are complexes
for Hochschild homology; these are fully treated in [319] and [321].
Recall that the Hochschild homology of .J'l may be computed from the re-
-®n
duced complex with modules .J'l 181 .J'l = 0n .J'l and the boundary Operator
b given by (8.41). Define B: nn.J'l- nn+l.Jl by
n
B(ao da1 ... dan) := L (-l)nidai ... dan dao ... daj-l·
j=O

Then B2 = 0 and one checks easily that bB + Bb = 0. If p E .J'l is a projector,


then for n even,
b(p (dp)n) = p (dp)n-l, b((dp)n) = (2p -1) (dp)n-l,
B(p (dp)n) = (n + 1) (dp)n+l, B((dp)n) = 0, (10.24)
andfornodd,b(p(dp)n) = b((dp)n) = OandB(p(dp)n) = B((dp)n) = 0.
More generally, if p E Mr(.J'l) is a projector, we form tr p := Z::k Pkk E .J'l
and likewise define tr p (dp)n and tr(dp)n in nn .J'l. It is easily checked that
the relations (10.24) continue to hold, with p (dp)n and (dp)n replaced by
tr p (dp)n and tr(dp)n respectively. Partly guided by Definition 8.11, we
introduce

(10.25a)

The formal sum of chains chp := Z::k'= 0 ch2k p satisfies (b + B)(chp) = 0.


Apart from the normalization -which can be altered by adjusting the maps
band B by convenient factors [91, p. 204]- this differs from the ordinary
Chern character component (8.3 5) by the factor p-! replacing p in (1 0.2 5a).
This nuancewas introduced by Getzler and Szenes [196), for compatibility
with (10.24). It can be shown that p ,.._ eh p leads to a well-defined map
from K0 (.J'l) to periodic cydic homology of .J'l: see [319, §8.3).
In the odd case, where a dass in K1 (.J'l) is represented by a unitary u E
Mr(.J'l), we may construct
ch2k+l U := (-1)kk! tr(u- 1 dud(u- 1 )du ... d(u- 1 )du) E 0 2k+l.J'l.
(10.25b)
448 10. Spectral Tripies

Notice that ch2k+ 1 u = k! tr(u- 1 du) 2k+ 1. Since

the formal series eh u := Lk=O ch2k+1 u satisfies (b + B)(ch u) = 0. By


pairing eh u with a suitable cyclic cocycle, Perrot [369] has shown that lo-
cal anomaly formulas can be computed in the noncommutative geometry
framework.
There is an obvious bilinear pairing between cn (.J\, .J1. *)I the space of
Hochschild n-cochains, and the space .J1. ®(n+ 1l of Hochschild n-chains, de-
fined by

(10.26a)

With respect to this pairing, the Hochschild coboundary is the transpose


of the Hochschild boundary: (bcp, c) = (cp, bc). Consequently, there is an
induced pairing between cohomology and homology classes. To pair the
elements (1 0.2 5) with cocycles that vanish on degenerate Hochschild chains
(as do all the examples we consider), we may replace them by corresponding
elements of .J1. ®(n+ll. This leads to
. k(2k)! 1
(cp,ch2kP) .= (-1) ---;(!cp®tr(p- 2,p, ... ,p), (10.26b)
(lJI,ch2k+1 u) := (-1)kk!lJI®tr(u- 1,u, ... ,u- 1,u), (10.26c)

when cp E C2k (.Jl.,.Jl. *) and lJI E C2k+ 1(.Jl., .J1. * }.


.- In several applications, there appear Fredholm modules which are not
finitely summable, and the question arises as to how the basic setup of pe-
riodic cyclic cohomology should be modified to accommodate their char-
acters.
One possibility is to extend the Connes bicomplex downward to the
left, that is, BCPq := cp-q (.Jl., .J1. *) for all p ~ q, without the restriction
q ~ 0. Cochains of the modified total complex are now infinite sequences
of Hochschild cochains of a definite parity. There are two cases: even se-
quences cjJ := (cpo, 'P2. cp4, ... ) and odd sequences rjJ := (lJI1,lJI3,lJI>, ... );
the coboundary operator b + B maps each kind to the other. Unfortunately,
unless some growth conditions are placed on these sequences, the result-
ing cohomology is trivial [91, IV.7]. In [87], Connes introduced the following
growth condition in order to accommodate pairings with series of chains
such as (10.25), and he showed that it yields a nontrivial cohomology when-
ever .J1. is a unital Banach algebra.
Definition 10.7. Let .J1. be a Unital Banach algebra. Any 'Pn E cn (.Jl., .J1. *)
is norrned, as a multilinear function on .Jl., by

llcpnll := sup{ lcpn(ao, .. . , an)l: each lla}ll =:;; 1 }.


10.2 Chern characters and entire cyclic cocycles 449

An even sequence of cochains cp, and an odd sequence of cochains ljJ, are
called entire cochains if each of the series

in a complex variable z, has infinite radius of convergence. Denote the set of


all even, respectively odd, entire cochains by CE0 (.Jl.), respectively CE 1 (.Jl.).
With the coboundary operator b + B in both directions, these form a two-
term complex whose cohomology HE 0 (.Jl.) e HE 1 (.Jl.) is called the entire
cyclic cohomology of .Jl..
Exercise 10.7. If cp and ljJ are respectively even and odd entire cochains
on .Jl., show that the functions

jcp(a)

.= I<-1) k -(2k)!
00
k <P2k(a, ... ,a),
1
k=O .

·- ~ k (2k + 1)!
f.p(a) .- L (-1) k' f./J2k+da, ... ,a)
k=O .

are entire analytic functions on the Banach space .Jl..


Example 1 0.3. We compute H E• (([),in order to compare it with the periodic
cyclic cohomology of I( (see Exercise 10.6). Denoting once more by <Pn the
basis element of cn "" I( normalized by <Pn(l, ... , 1) = 1, it is clear that
II <Pn II = 1. Thus an even entire cochain is given by a series <P = l:k'=o a2k<P 2k
with coefficients a2k E I( suchthat k- 1 1a2k 11 /k - 0 as k - oo, and an odd
entire cochain is a series tfJ = l:k'=o c2k+1 <P 2k+ 1 suchthat k- 1 1c2k+11 1ik - 0.
We know that for each k E ~. b<P 2k = 0 and B<P 2k = 0, whereas b<P 2k+ 1 =
<P 2k+ 2 and B<P 2k+ 1 = ( 4k + 2) <P 2k. Thus, (b + B) tfJ = 2c1 <P 0 + l:k'= 1 ( c2k-1 +
(4k + 2)c 2k+d<P 2k, so that (b + B)f./J = 0 only if tfJ = 0. On the other band,
every even entire cochain <P is a cocycle, and it is a coboundary if and only
if and we can solve the equations 2c1 = ao and c2k-1 + (4k + 2)c2k+1 =
a2k for k ~ 1, subject to the growth conditions. lt is easy to check, using
Exercise 10.7, that jcp(l) = 0 if and only if <P = (b + B)f./J for some f./J, so
that <P- fcp(l) provides a isomorphism from HE 0 (1C) to C Therefore,

HE 1 (1() = 0.

Exercise 10.8. Make this check by showing z 2 f<b+BJ.p(Z) = 2(1- z 2 )j.p(z)


for z E C <>

~ We now consider a spectral triple (.Jl., Jf, D), as introduced in Defini-


tion 9.16: .Jl. is an algebra represented by bounded operators on the Hilbert
space J-f, D is a selfadjoint operator on Jf, with compact inverse, such
that [D, a] is bounded for every a E .Jl.. Weshall assume also that there is
450 10. Spectral Tripies

a grading operator x on Jf, commuting with .JI. and anticommuting with D


(i.e., the spectral triple is even). Odd spectral triples can be treated in par-
allel by taking x = 1 and dropping the anticommutation condition. We may
use the following norm on .JI.:

llallv := llall + II[D,a]l!. (10.27)

Since llabl!v :s; llabll + II[D,a]bll + lla[D,b]ll :s; llallvllbllv. this makes
.JI. a normed algebra. Now, entire cyclic cohomology is a theory for Ba-
nach algebras, so we shall assume in this section that .JI. is complete in
the norm II · llv. It should be noted, however, that the theory can also be
developed for Frechet algebras [91, 285].
Definition 10.8. A spectral triple (.JI., Jf, D) is called:

(a) p-summable if (1 + D 2)-1i2 E LP(Jf);

(b) p+-summable if (1 + D2)-1i2 E LP+(Jf);

(c) 8-summable if e-tD 2 is traceclass for all t > 0.

Since Dis invertible, (.JI.,Jf,D) is p-summable if and only if D- 1 E LP,


or p+-summable if and only if D- 1 E LP+.
Lemma 10.8. If (.JI., Jf, D) is p-summable for any finite p, then it is also
8-summable, andTrrtD 2 = O(t-Pi 2) as t I 0.

Proof. We can write e-tD 2 = (1 +D 2)Pi 2e-tD 2 (1 +D 2 )-P/ 2 where (1 +D 2 )-p/2


is traceclass by hypothesis, and (1 + D 2)Pi 2e-tD 2 is bounded. This follows
from functional calculus, since the function .\ - ( 1 + ,V)Pi 2 e-t,V has supre-
mum (p/2e)Pi 2 t-PI 2 et. 0

There are genuinely infinite-dimensional spectral triples where the sum-


mability condition must be replaced by "e-tD 2 is traceclass for some (not all)
t > 0": see the discussion on p. 395 of [91]. The definition of 8-summability
is modified accordingly. Here we adopt the original definition [90] .
.,. The Chern character for 0-summable spectral triples was introduced
in [87]; the original version, also discussed in [91, IV.8.y], was technically
rather involved. A simpler variant was then proposed by Jaffe, Lesniewski
and Osterwalder [262]-now called the JLO cocycle- and later strearnlined
byGetzler and Szenes [196], and is based onideas from the quantum theory
of systems with infinitely many degrees of freedom. Think of D as a super-
eharge and of D 2 as a Hamiltonian (the Uchnerowicz formula is also sug-
gestive in this regard); for any operator A E L(Jf), let A(t) := e-tD 2 AetD 2 •
Then an operator product A 0 AI(t 1 ) •• • An(tn) can be averaged over the
simplex 0 :s; t1 :s; • • • :s; tn :s; 1 to yield the following correlations.
10.2 Chern characters and entire cyclic cocycles 451

Definition 10.9. For Ao,AI, ... ,An E L(Jf), write

(Ao,AI, ... ,An}D := J Tr(xAoe-t 1 D2 A1e-<trt!lD 2 ••• Ane-0-tnlD 2 )dnt

J
Lln

= Tr(xAoe-soD 2 A1 e-s 1D2 ••• Ane-s"D 2 )dn s (10.28)


Lln

where the integral extends over the n-simplex (10.11). The even JLO cocy-
cle is the even entire cochain Ch" (D), whose components Ch 2k(D) are

Chn(D)(ao, ... , an) := (ao, [D, ad, ... , [D, an]}v


= J Tr(xaoe-soD [D,ade-s D
Lln
2 1 2
••• [D,an]e-snD 2 ) dns. (10.29)

The odd JLO cocycle has components Ch2 k+ 1(D) given by the same formula,
with the understanding that x = 1.
To show that Ch" (D) is indeed an entire cocycle, we need a preliminary
lemma.
Lemma 10.9. If jE {1, ... , n- 1} and A 1 E Dom(adD 2 ), then

(10.30)

Proof. There is a standard commutator identity

(10.31)

since the integrand is the derivative of -e-sD 2 Ae-O-slD 2 • This is easilymod-


ified to give

Substituting this in (Ao, ... ,A1_1 , [D 2 ,A1 ],AJ+l• ... ,An}v gives a differ-
ence of two integrals over (n- 1)-simplices that coincide with the right
hand side of (10.30). 0

Proposition 10.10. The cochain Ch" (D) satisfies (b + B) Ch" (D) = 0.

Proof. According to (8.69), Bo Chn+l (D) is given by

Bo Chn+l (D) (ao, ... , an)


:= Chn+l (D) (1, ao, ... , an) - ( -l)n+l Chn+l (D)(ao, ... , an, 1)
= (1,[D,ao], ... ,[D,an]}v. (l0.32a)
452 10. Spectral Tripies

Since B = NBo, cyclic skewsymmetrization of the right hand side yields


n
BChn+l(D)(ao, ... ,an) = ~([D,ao], ... ,1,[D,aj], ... ,[D,an])n
}=0
= ([D, ao], ... , [D, an])n. (10.32b)
The second equality follows from the general identity
n
(Ao,Al, ... ,An)D = ~(Ao, ... ,Aj-I.1,Aj, ... ,An)v, (10.33)
}=0

JJ
where (Ao, A1, ... , An) D = (Ao, A1, ... , An) D du by introducing a trivial
extra integration; the polyhedron ll.n x [0, 1] can then be subdivided by the
inequalities t 1 ::5 u ::5 tJ+l into n + 1 simplices, each of which is a copy of
tl.n+l; integration over these simplices yield the terms an the right hand
side of (10.33).
Next, when n ~ 1, b Chn-l (D)(ao, ... , an) equals
n-1
(aoa1, ... ,[D,anl>D + ~ (-1)l(ao, ... ,[D,ajaJ+l], ... ,[D,an])n
j=l
+ (-l)n(anao, [D,ai], ... , [D,an-Il)n

n-1
+ ~(-1)1(ao, ... ,aj[D,aJ+Il •... ,[D,an])v
j=l
+ (-l)n ([D, ai], ... , [D, an-d. anao)D
n
= ~ (-1)1-l (ao, [D, ai], ... , [D 2 , aj], ... , [D, anl>v. (10.34)
j=l
The last equality follows from the previous lemma.
The supercommutator [D, aoe-soD 2 [D, a1 ]e-s1D2 ••• [D, an]e-s"D 2 ] is an-
nulled by the supertrace Tr(x· ). Integration over ll.n gives
n
'
L. ( -1)J-
. 1 (ao, ... , [D, aJ-11. [D, [D, aJ ]], [D, aJ+l], ... , [D, anl>D = 0.
}=0
(10.35)
Since [D, [D, a1 ]] = [D 2 , a1 ], (1 0.34) simplifies to
b Chn-l (D) (ao, ... , an) = -([D, ao], [D, a1], ... , [D, anl>D
= -B Chn+l (D)(ao, ... ,an).

Also, for n = 0, we get B Ch1(D)(ao) = ([D, ao])n = Tr(x[De-D 2 , ao]) = 0


since D is odd. 0
10.2 Chern characters and entire cyclic cocycles 453

The JLO cochains thus have the right algebraic properties to be cocycles.
The growth condition which ensures that they are indeed entire cocycles is
provided by the following estimate of Getzler and Szenes [196, Lemma 2.1].
Lemma 10.11. Jf Aj, Bj. j = 0,1, ... ,n, are bounded operators and ifat
most k ofthe A 1 are nonzero, then forO < E < 1/2e,

I(Ao D + Bo, ... ,An D + Bn)v I ~ (eE/2)-k/2


( _ k)' Tre
-(l-E)D2 nn (IIAJ II + IIBJII ) .
n . J=O
(10.36)

Proof. The generalized Hölder inequality (Exercise 7.22) shows that

I Tr(xAo . .. An)l ~ IIAoiiitso ... IIAniiitsn

if so + · · · + Sn = 1. Therefore,

The factors in this product can be estimated thus:

IlA jDe-siD 2 IIItsi ~ IlAJ II IIDe-ESiD 2 IIIIe-O-E)sjD 2 IIItsi,


IIBJe-siD 2 IIItsi ~ IIBJIIIIe-siD 2 IIItsi = IIBJII (Tre-D i. 2
)
5

The function ,\ ..... Ae-E5i'v, for i\ ~ 0, attains its maximum (2eES1 )- 112 at
i\ = (2Es1)- 112 , and thus

Plugging this into the previous integral, we obtain

I(AoD + Bo, ... ,AnD + Bn)D I


~ Tre-O-EW 2
fi (IIAJII + IIBJII)(2a)-k J (sb ... s~)- 112 dns',
j=O
12
~n

where sj = 1 or s1 according as A 1 = 0 or not. Since

the estimate (10.36) follows. 0


454 10. Spectral Tripies

Taking Bo = ao, Bi= [D, ai] for j = 1, ... , n gives

IChn(D)(ao, ... ,an) I!'> ~Tre-(l-E)D 2 IIaoll


n.
Il II[D,aj]ll
i=l

!'> ~Tre-(l-E)D 2
n.
Il llaillv.
i=O
(10.37)

so that II Chn(D)II !'> Tre-(l-E)D 2 /n! and thus Ch"(D) is a cocycle for the
entire cyclic cohomology of the Banachalgebra (.Jl,ll · llv ).
.,. The essential property of the JLO cocycle is that its entire cohomology
dass is unchanged on applying a differentiable homotopy to the opera-
tor D. Just as with the invariance of the K-theory Chern character in Sec-
tion 8.3, this property is established by means of a transgression formula
which expresses the difference of two JLO cocycles as a coboundary.
Topave the way for this formula, we introduce an auxiliary entire cochain

(10.38)
n
:= L (-1)k#V (ao, [D,a!], ... , [D,ak], V, [D,ak+d .... , [D,an])v,
k=O
where V is an operator on J{ that is either even or odd, with parity #V = 0
or 1. More generally, if each Ai is either even or odd, we may write

t(V) (Ao, ... , An) v


n
·- L'(-1)(#Ao+···+#Ak)#V(A Or" · • A k, V • A k+lr .. · • A n ) Dr
.- (10.39)
k=O
so that ~n(D, V):= t(V) Chn(D).
If Visa bounded operator on Jf, or eise if V= D, the sequence ~· (D, V)
whose components are ~ 2 k(D, V), is an entire even cochain, and similarly
in the odd case. This can be checked directly using the estimates (10.36).
We can now determine the coboundary of ~· (D, V). The argument lead-
ing to (10.35) shows more generally that
n
'(-1)#Ao+···+#AJ- 1 (A o, .. ·, A J-l•
L · [D , A 1·] , A J+lr - 0·
· .. ·, A n )D-
j=O

Applying this to the right hand side of (10.38) yields a sum of several terms,
adding up to zero. The terms containing [D, ao] are
n
L (-1)k#V ([D, ao], ... , [D, ak], V, [D, ak+d •... , [D, an])v
k=O
= (-1)#VL(V)([D,ao], ... , [D,an])v = (-1)#VB~n+l(D, V)(ao, ... ,an),
10.2 Chern characters and entire cyclic cocycles 455

where the last equality is found by adapting (10.32). The terms containing
[D, V] add up to thn(D, [D, V])(ao, ...• an). The remaining terms are
I (-1)k#V+j- 1(ao, ... , [D 2 , aj], ... , [D, ak], V, ... , [D, anl>D
+ I (-l)<k+ 1l#V+j- 1(ao, ... , [D, ak], V, ... , [D 2 , aj], .. . , [D, anl>D
j>k
= (-1)#Vbthn- 1(D,V)(ao, ... ,an)
n
+ I (-1)k(#V+ 1l- 1(ao, [D, ai], ... , [V, akl, ... , [D, anl>D.
k=1
on applying Lemma 10.9. Denoting the last sum by an(D, V)(a 0 , •. • , an),
we end up with
( -1)#V (bthn-1(D,V) + Bthn+1 (D,V)) + thn(D, [D, V])+ ()(n(D, V) = 0.
(10.40)
We come now to the key transgression formula.
Proposition 10.12. Ift ..... Dr is a continuously differentiable one-parameter
family of odd selfadjoint operators on J{ such that either all Dr are bounded,
or Dt = tD with (.A., J{, D) 0-summable, then
d Chn (Dr) = -bthn-1 Wr,Dr)
dt · -Bth n+1 Wr,Dr).
· (10.41)

Proof. We can write Chn(Dr) = (Ao, ... ,An)v1 where Ao = ao and Aj =


[Dr,aj] for j = 1, ... ,n. We replace D by Dr on the right hand side of
(10.29), and differentiate with respect tot. That gives a sum of terms com-
ing from the derivatives of the [Dr, a j], plus other terms arising from the
derivatives of the exponentials. The first sum is
n
I (ao, [Dr, ai], ... , Wr, ak], ... , [Dr, an])D 1 = -an Wr, Dr )(ao, ... , an).
k=1
For the second sum, we need the Duhamel equation
d
-e-D;
dt
?
+ 11
0
e-sD; [D
?

to
D ]e-O-slD; ds = 0

t
?

I
(10.42)

that Ieads to

Replacing e-<tJ+I-tJW~ by this expression in (10.29) gives an integral over


an (n + 1)-simplex that simplifies to
n
-I (Ao, · · ·, Aj, [Dr, Dr], Aj+1o · · ·, An)D1 = - thn (Dr.[Dt, Dr]).
j=O
456 10. Spectral Triples

To see why equation (10.42) holds, consider the operator function

Clearly A(t, 0) = 0 for all t; to get A(t, 1) = 0, it is enough to checkthat


A(t, u) satisfies the heat equation (o I ou + Df )A(t, u) = 0. Since o /ou + Df
vanishes on e-uD~, it follows that

(a~ +Df}A(t,u) = [a~ +Df, :tJe-uD~ + [Dt,Dde-uv;


= (- :tDf + [Dt, Dd )e-uDl = 0.

To justify these operator derivatives, it is enough that the function t ....


Tre-Dl be uniformly bounded on compact intervals. In the case Dt = tD,
this is guaranteed by the 8-summability assumption. The other case, where
t .... Dt is a continuous path in L(J-f), requires some finer estimates, for
which we refer to [196, Thm. C]. Collecting now the derivative terms and
using (10.40) with V = Dt. we end up with

·
d Ch n (Dr) = -a n Wt,Dd-
dt ·
thn Wt.!Dt,Dd)
= -bthn- 1 Wt,Dd -Bthn+ 1 Wt,Dd. D

Corollary 10.13. I{ (5\, J-f, D) is a 8-surnrnable spectral triple, the entire


cyclic cohornology dass o{Ch"(tD) is independentoft > 0.
Proof. If 0 < s < t, then

Chn(sD)- Chn(tD) = b S: thn- (uD,D) du+ B S: thn+ (uD,D) du,


1 1

and the right hand side is the nth component of a coboundary. D

... We now restriet our attention to the p-surnrnable case, with p finite.
Replacing D by tD in (10.37) gives the estimate

(10.43a)

Since Tre-0-E)t 2 D2 = O(t-P) by Lemma 10.8, it follows that

limChn(tD) = 0, for n > p. (10.43b)


t!O

Lemma 10.11 shows likewise that II thn(tD,D)II = O(tn-P) as t l 0, so that


thn(tD,D) is integrable on any interval [0, t 0 ], if n > p.
This integrability and the limit (10.43b) also hold when (Y.., J-f, D) is a
p+ -surnrnable spectral triple, whereby IDI- 1 E LP+ c Lr for any r > p. For
10.2 Chem characters and entire cyclic cocycles 457

any integer n > p, of the same parity as the spectral triple, and any t > 0,
there is then a Hochschild n-cochain [112] given by

T,Ch~(D)(ao, ... , an) := J~ thn(uD, D)(ao, ... , an) du. (10.44)

We can now step back from entire to periodic cyclic cohomology, by replac-
ing the tail of the sequence Ch • (D) by this cochain.
Proposition 10.14. If (.J\., Jf, D) is a p+ -summable spectral triple, then for
any integer n ~ p that is even or odd according as the spectral triple is
graded or not, and any t > 0, there is a cocycle ch~(D) in Totn BC(.J\.)
given by

ch~(D) := I Chn-Zk(tD) + BT.Ch~+l (D). (10.45)


Os2ksn

If 0 < s < t, then eh~ (D) and eh~ (D) are cohomologous, and determine the
same class [chn(D)].\ in HCn(.J\.). Moreover, S(ch~(D)) and ch~+Z (D) are
cohomologous, so that there is a well-defined class ch(D) E HP#n(.Jl).
Proof. First of all, the top term in the sum (10.45) is a Hochschild cocycle.
Indeed, the transgression formula of Corollary 10.13 and Proposition 10.10
show that

and the right hand side vanishes as s l 0; taking this limit, we arrive at

Applying Proposition 10.10 again, we get the cocycle property (in the total
Connes complex):

2s2k,;n Os2ksn-2
I b Chn-Zj (tD) + B Chn-Zj+Z (tD) = 0.
2s2jsn
lf 0 < s < t, Corollary 10.13 again gives

=I Chn-Zk(sD) -Chn-Zk(tD) -BJ thn+I(uD,D)du


Ln/2J t
ch~(D) -ch~(D)
k=O 5

=(b+B) I rthn-Zk-I(uD,D)du, (10.46)


Os2k<n 5

so that ch~(D) and ch~(D) differ by a coboundary in Totn BC(.Jl).


458 10. Spectral Tripies

Lastly, S (chf (D)) is represented by the same Hochschild cochains as


chf(D), in Totn+ 2 BC(.Jl). Since

chf+ 2 (D)- chf(D) = Chn+ 2 (tD) + BT.Chf+ 3 (D)- BT,Chf+ 1 (D)

=- J: b~n+ 1 (uD,D) du- BT.Chf+ (D) 1

= -(b + B)T,Chf+ 1 (D),

which is a coboundary, S(chf(D)) and chf+ 2 (D) are cohomologous. D

Fix an integer n > p, even or odd according as (.Jl, J-{, D) is a graded


or ungraded spectral triple. We claim that chf (D) converges as t - oo
and that the limit defines a cyclic n-cocycle over .Jl. However, when the
(finite-dimensional) kernel of D is nontrivial, this limit is rather involved;
the full computation is laid out in [112]. To avoid this complication, weshall
adopt the simplifying assumption, in the rest of this section and whenever
convenient from now on, that D is invertible.
We may use Hölder's inequality, in the form Tr(H 2 ) ::s; /IH 11 Tr H for a
positive traceclass operator H, to refine the estimate (10.43a) to

(10.47)

lt follows that Chn(tD) - 0 in norm as t - oo; indeed, the right hand


side of (10.47) is dominated by tne-0-Elt 2 t. 2 12 , where ,\ 2 is the firstpositive
eigenvalue of D 2 • Notice that no restriction is placed on n for this limit to
hold; therefore, the terms Chn- 2 k(tD) on the right hand side of (10.45) all
vanish as t- oo, and there remains only

Tß := limchf(D) =BlimT.t'hf+ 1 (D) =B


t-oo t-oo Jof"" ~n(uD,D)du. (10.48)

The integral on the right hand side exists, provided n > p to ensure conver-
gence at the lower boundary; convergence at the upper boundary is guar-
anteed by the exponential falloff of the integrand that follows from the
estimate (10.36), since D has no zero eigenvalues by assumption. The right
hand side of (10.48) is now a Hochschild n-cocycle, by the proof of Propo-
sition 10.14, and it is cyclic since it lies in B( c;:+I (.Jl) ).
By letting t - oo in (10.46), we find that

chf(D) -Tß = (b+B) L J"" ~n- 2 k- 1 (sD,D)ds, (10.49)


Os2k<n t
10.2 Chern characters and entire cyclic cocycles 459

so that the periodic dass of T Jj is still cohomologous to eh~ (D) for any
0 < s < oo. Explicitly, T}j is given by the formula

T}j(ao, ... ,an) = B fo"" t(D)Chn(tD)(ao, ... ,an)dt

= N Jo"" t(D) (1, [tD, ao], ... , [tD, an])w dt

=N i tn+1
oo
--
o n+ 1
t(D)([D,ao], ... ,[D,an])wdt,

where we have used (10.32); here N is again the cydic skewsymmetrizer in


the arguments ai. The right hand side can be written more explicitly as

N ioo tn+ 1
-- 1
I ,, ,,
Tr(xDe-sot·D· [D, ao]e-s 1 t·D· ... [D, an]e-snt·D·) dns dt.
,,
0 n+ 6n
(10.50)
We can simplify this expression, up to cohomology at any rate, by apply·
ing another differentiable homotopy [llO]. This time, weshall use
Du:= DIDI-u, for 0 ::5 u ::51.

This homotopy starts at D and ends at F = DIDI- 1 . Even though Du is


unbounded for 0 < u < 1, we may apply the method of Proposition 10.12
to compute the derivative (d/du)(T}j). Replace D by Du in (10.50); to
ensure that the integrand is differentiable in u, we can estimate it by the
generalized Hölder inequality (7.104). The term Due-sot 2 D~ is uniformly
bounded by (2es 0 t 2 )- 1 ' 2 and the other e-sJt 2 D~ are tracedass, and weshall
soon verify (Exercise 10.9 below) that each [Du,aj] belongs to, say, the
Schatten dass Lrlu for any r > p.
One can then show that (d/du)T}ju is a coboundary in the total Connes
complex. In view of (10.49), it is enough that (d/ du) chrWu) be a cobound·
ary. Makinguse ofidentities similar to (10.40) -see Proposition 3 of [ll2]-
this proceeds as in the proof of Proposition 10.12, with the result that

ddu ehr (Du)= -(b + B) L


Os2k<n
f° ddu thn-zk- 1 (sDu,Du) ds,

Adding d/du of (10.49) or, equivalently, letting t - oo, wegetat once

ddu T}ju = -(b +B) L


Os2k<n
i""° ddu thn-2k-1(sDu,Du)ds.
This establishes the coboundary property in the total Connes complex. lt
can in fact be expressed as a Hochschild coboundary; some further manip-
ulation along the lines of [112] eliminates the lower-degree terms on the
right hand side, leading to

d Tvn =- b B
-d
u u
i"" L(SDu)
0
. ,.... (sDu,Du) d s,
..,ll.,n
460 10. Spectral Tripies

which lies in Bf(.JI.). Therefore, T}j and T]J are cohomologous in Zf(.JI.).
This last transgression formula allows us to replace the character (10.50)
by a much simpler cocycle. Namely, since P = 1, all the exponential terms
in (10.50) are scalars, and factor out to give

l oo
o n +1
e
tn+l -t2 J dsdt-(
n
Cl.n
_ 1 00
1.! n/2 -u _ [(I+ 1 )
n + . o 2 u e du- 2 ( n + 1 ),..
1 )1

Therefore, taking into account that N = (n + 1) on n-cocycles that are


already cyclic, we arrive at

n f(I+1)
Tp (ao, ... , an) = 2 n.1 Tr(xF[F, ao] ... [F, an]).

This is precisely the cyclic cocycle (10.23) representing the Chern character
of (.JI., 3f, F) in periodic cyclic cohomology, obtained in Section 10.1!

10.3 Tameness and regularity of spectral triples


Before tackling the main result of this chapter (computing the image in
Hochschild cohomology of the Chern character), we must sharpen our tools
by making a closer inspection of the operator-theoretic properties of spec-
tral triples. We collect in this section several such tools: estimates for com-
mutators between elements of .JI. and nice functions of D, the permutability
of operators under Dixmier traces, and the stability of some bounded ope-
rators on 3{ und er certain unbounded derivations. The end product will be
a concept of regularity for spectral triples which provides an operatorial
formulation of the calculus of smooth functions.
The crucial algebraic feature of the (generalized) Dirac operators, intro-
duced in Section 9.3, is the formula (9.18), whereby the commutators [D, a],
for a any smooth function, are given by bounded operators on the spinor
space. It forms part of the general definition of a spectral triple (Defini-
tion 9.16). To proceed further, we require several estimates on such com-
mutators, which may be derived by a Fourier-transform technique, origi-
nally due to Helton and Howe [239].
Lemma 10.15. Let a be a bounded operator and D a selfadjoint operator
such that [D, a] is also bounded; and Iet g E V ( ~). Then [g (D), a] satisfies
the estimate

ll[g(D),a]ll5 217TII[D,a]ll Joo-oo ltg(t)ldt. (10.51)

Proof. Using functional calculus, we can write

[g(D), a] = -21 Joo g(t) [e 1.tD, a] dt.


7T -oo
10.3 Tameness and regularity of spectral triples 461

Now, it is clear that

[eitD,a] = s: :s (eistDaei(l-s)tD) ds = it s: eistD[D,a]ei<l-s)tD ds.


(10.52)

The space of vectors in J{ that have compact support with respect to the
spectral measure of D is a core for D and also for g(D). If ~. 11 E J{ lie in
this common core, then

(g(D)ryl a~}- (171 ag(D)~) (10.53)


· Joo tg(t) dt
= _t_
2rr -oo
il
o
(De-istD'11 aei<l-s)tD~)- (171 eistDaDei<l-s)tD~) ds,

Therefore, the form (g(D)ryl a~} - (171 ag(D)~) is bounded on this core;
that is to say, a lies in the domain of the derivation T .... [g(D), T], and the
advertised bound applies. 0

Corollary 10.16. Let D be a selfadjoint operator, and suppose that a, [D, a]


and [D, [D, a]] are bounded. If g E V(~). then

ll[g(D),a]- g'(D) [D,a]ll ::5 41rr II[D, [D,a]]ll Joo-oo t 2 1g(t)l dt. (10.54)

Proof. From (10.52), it follows that ll[eitD,a]ll ::5 t II[D,a]ll for any t > 0,
and thus

ll[eitD,a]- iteitD[D,a]ll = IIJ: :s (eistDaem-s>tD)ds- iteitD[D,a]ll


= llit s: eistD [D,a] ei<l-s)tD) ds- iteitD[D,a]ll

= t IIJ: eistD [ei<l-s)tD' [D, a]] dsll


::5 t s: (1 - s)t ds II [D, [D, a]]ll = ~t 2 II [D, [D, a]] 11.
Since g' (D) = (2rr)- 1 Coo itg(t) [eitD, a] dt, the argument ofLemma 10.15
now yields (10.54). 0

Weshall need the following variant [81] of Lemma 10.15.


Lemma 10.17. Let a be a bounded operator and Da selfadjoint invertible
operator such that ID 1- 1 and [D, a] are bounded. Then, for any 0 < r < 1,
the operator [I Dl r, a] is bounded; moreover,

II[IDir,a]ll ::5 Crii[D,a]ll. (10.55)

with Cr independent of a.
462 10. Spectral Tripies

Proof. By continuity, the equality (10.53) holds when g is smooth enough


that tg(t) is integrable. Choose 8 > 0 so that sp IDI ~ [8, oo), let g(x) :=
lxlr for lxl ;e: 8, and extend g smoothly to [ -8, 8]. Then g(D) = IDir.
Since g' is smooth and t g(t) oc 9'
(t), this function decays rapidly at
infinity, so the only issue is whether it is locally integrable (at the origin). As
was noted in Section 7.B, the Fourier transform of a homogeneaus function
of degree r- 1 is homogeneaus of degree -r; therefore, as g (x) - Ix Ir has
compact support, 9' is the sum of a homogeneaus function of degree -r
and an analytic function. The integrability of tg(t) now follows from the
integrability of It 1-r on any compact interval, for 0 < r < 1.
The proof of Lemma 10.15 then shows that a lies in the domain ofT .....
[IDir, T], and (10.55) follows, with Cr = (2rr)- 1 f~~t lt g(t)l dt. D

.,.. Another dass of commutator estimates allows us to compare the Sum-


mability properties of a spectral triple (.Jl, J{, D) and of its associated Fred-
holm module (.JI.,Jf,F), where F := DIDI- 1.
Lemma 10.18. Jf(.Jl,Jf,D) isap-summabl espectraltriple , letF := DIDI- 1.
Then the relation [F, a] E LP (Jf) holds for all a E .Jl.

Proof. Using the spectral formula for the inverse square root of a positive
selfadjoint operator:

we compute

[F,a] = [DIDI- 1,a] = [D,a] IDI- 1 + D [IDI- 1,a]


= !fo""([D,a](i\ +Dz)-1+D[(i \+D2)-1,a]) ~
= ! fo"" ([D, a] (i\ + Dz)-1 - D(i\ + Dz)-1 [Dz' a] (i\ + Dz)-1) ~
= _!_ f"" (i\(i\ + Dz)-1 [D,a](i\ + Dz)-1
rr Jo
- D(i\ + D 2)- 1 [D,a] (i\ + D 2)- 1D) ~· (10.56)

since [D 2 , a] = D [D, a] + [D, a] D. The compactness of [F, a] follows from


that of both terms of the last integrand: [D, a] is bounded by hypothesis,
(i\ + D 2 ) - 1 is compact for each i\ -see the discussion of resolvents in
Section 7.4- and D(i\ + D 2)- 1 = T(i\ + D 2)- 112 where T := D(i\ + D 2)-112
is bounded.
Todetermine the Schatten dass of [F, a], we may assume that a * = -a,
and consequently that [D, a] and [F, a] are selfadjoint. (More generally,
10.3 Tameness and regularity of spectral triples 463

[D, a] ± [D, a]t = [D, a] + [D, a*] = [D, a + a* ].) Replacing [D, a] by its
norm in the last integral of (10.56) gives

!II[D,a]11fooo(,\+D 2 )- 1 ~ = II[D,a]IIIDI- 1 ,
so the obvious estimate -II[D,a]ll s [D,a] s II[D,a]ll yields
-II[D,a]IIIDI- 1 s [F,a] s II[D,a]IIIDI- 1 . (10.57)
Since IDI- 1 E f.P and [D, a] is bounded, each [F, a] lies in f.P too. D

The previous proof is taken from [421], which deals with the converse
problern of reconstructing a spectral triple from a given Fredholm mod-
ule. There are examples [90] of Fredholm modules with every [F, a] E LP
that do not arise from any finitely summable spectral triple. However, in
the case where .J\ is the group algebra of a discrete group of polynomial
growth r and all [F, a] lie in f.P, a construction by Schrohe, Walze and
Warzecha [421], based on [91, Thrn. IV.8.4], produces a spectral triple sat-
isfying DIDI- 1 = F that is q-summable for all q > p + r + 1.
The development (10.56) may be generalized by replacing F by DIDI-u
for any 0 < u s 1: since IDI-u = Cu J;"(.\ + D 2 )- 1 .\-u/ 2 d.\ with C~ 1 :=
fooo (t+ 1)- 1 t-u/ 2 dt, we change rr- 1 d.\1 -IX to Cu.\ -u/ 2 d.\ throughout. When
a* = -a, (10.57) generalizes to
-II[D,a]IIIDI-u s [DIDI-u,a] s II[D,a]IIIDI-u. (10.58)
Exercise 10.9. If v- 1 E .[.P+ I 0 < u < 1, and r > p, show that
II[DIDI-u,alllr;u sii[D,alll (1 + IIIDI- 1 IIr).
and conclude that [DIDI-u,a] E Lrfu for each r > p.

~ When Dis a selfadjoint operator with compact resolvent and g E 1J(IR)


is a smooth function with compact support, then the operator g(D) has
finite rank, and so also does any commutator [g(D),a]. Indeed, if g(D)
has rank m, then since
[1<1>)(1/JI,a] = 1<1>)(a*!/JI- la<j>)(!/JI
for any rank-one operator I<1>) (1/J I, the rank of [g (D), a] is at most 2m. In
particular, the size of [g(D), a] can be measured with any of the symmet-
ric norms discussed in Section 7.C. A particularly useful estimate is the
following one, due to Connes [89] -see also [91, IV.2.8]- which uses the
norm (7.108) of the separable operatorideal f.P-.
Lemma 10.19. Let D be a selfadjoint invertible operatorsuch that v- 1 E
f.P+ with 1 < p < oo, and Iet g E 1J (IR). Then there is a constant Cp (g) such
that
ll[g(tD),a]llp- s Cp(g) II[D,a]IIIID- 1 IIp+•
whenever t > 0 and a is a bounded operator with [D, a] bounded too.
464 10. Spectral Tripies

Proof. Suppose that suppg s;; [ -R,R]. The operator g(tD) has a finite rank
that cannot exceed the number of indices k for which the corresponding
singular value satisfies sk(D- 1 ) ~ t/R. The definition (7.108) of the norm
in LP+ gives the estimate
m-1
am(D- 1) = 2:: Sk(D- 1) ::5 m(p-1)/p IID- 1llp+
k=O

form = 1, 2, ... ; it follows that

(10.59)

where ap := IID- 1IIp+· If we choose m to be the rank of g(tD), so that


t/sm ::5 R, then m ::5 (ap/Sm)P ::5 (apR/t)P. Since the rank of [g(tD),a] is
at most 2m, we get an estimate for the trace norm:

ll[g(tD),a]il1 ::5 2m ll[g(tD),a]ll ::5 2(apRJt)P ll[g(tD),a]ll.

The interpolation inequality (7.109) now shows that the norm of this
commutator in LP- may be estimated by

II [g (tD), a] llp- ::5 ßp II [g(tD), a] II ~IP II [g (tD), a] 11 1- 11 P


::5 2 11 PapßpRt- 1 ll[g(tD),a]ll.

Combining this inequality with (10.51) -with D replaced by tD- yields

ll[g(tD),a]llp- ::5 (27T)- 12 11 PapßpRII[D,a]ll J~oo lug(u)ldu. (10.60)

The right hand side is no Ionger t-dependent. Since ap = IID- 1IIp+. this
gives the desired bound, with Cp(g) := (2rr)- 12 11PßpR f lug(u)l du. D
In the previous proof, the bound (10.59) shows that

Sm(IDI-P) = Sm(D- 1 )P ::5 a~jm

for all m, and so the singular value sums obey an estimate of the form
aN(IDI-P) ::5 (ap + 8)PlogN for any 8 > 0, if N is large enough. In the
estimate (10.60) for ll[g(tD),a]llp-. we can thus replace the term ap =
IID- 1IIp+ by the pth root of any generalized limit of the bounded sequence
(log N) - 1aN (I D 1- P). Each such generalized limit is given by a Dixmier trace
Trw IDI-P, as explained in Section 7.5. Letting t I 0 also, we arrive at the
estimate

liminf!l[g(tD),a]llp- ::5 C~(g) II[D,a]ll (Trw ID1-P) 11 P, (10.61)


t!O

where the constant C~ (g) does not depend on D.


10.3 Tameness and regularity of spectral triples 465

~ At one stage in the development of the theory of spectral triples, the


authors introduced the word "tameness" in order to develop Hermitian
structures from such triples [469]. A p+-summable spectral tripleis called
tarne if each functional
a ..... Trw aiDI-p
is a trace on the involutive algebra Xn generated by .J\ and the commuta-
tors { [D, a] : a E .J\ } . lt was then known that the Standard commutative
spectral tripleis tarne, but there was some contention on whether every p+-
summable spectral tripleis tarne (it is not). The issue was resolved in [81],
wherein it was shown that a weak form of regularity is enough to establish
tameness.
Theorem 10.20. Let (.J\,J{,D) be a p+-summable spectral triple and let
Trw be any Dixmier trace. Then the functional
a ..... Trw aiDI-p

is a hypertrace on .J\; that is,


Trw(aT IDI-P) = Trw(Ta IDI-P),

for any bounded operator T on J{.


Proof. The Hölder inequality (7.74) for the Dixrnier traces shows that
ITrw([a,T)IDI-P)I = ITrw(T[IDI-P,a])l::;; IITII TrwlliDI-P,aJI.
Therefore, we simply need to check that
TrwlliDI-P,aJI =0 forall aE.J\.
Choose r with 0 < r < 1 suchthat k := p/r is an integer. The identity
k
[IDI-rk,a] = L IDI-r(l- 1l[IDI-r,a] IDI-r(k-1)
1;1
k
=- L IDI-ri[IDir,a] IDI-r(k-1+1)
[;1

shows that
k
Trw I [I DI-p, a] I ::;; L Trw IID1-rl [I Dir, a]IDI-r(k-l+ll 1. (10.62)
1;1
We apply the Hölder inequality of Proposition 7.16 to each term of this
sum. Choose PI > 1 and put qz := ptf(PI- 1); then the right hand side
of (10.62) is bounded by
k
II[IDir,a]ll L(Trw IDI-rlp 1 ) 11 P1 (Trw IDI-r(k-l+ 1lql) 11q1 • (10.63)
[;1
466 10. Spectral Tripies

Notice finally that we can choose pz so that both exponents -rlpz and
-r(k - l + 1)qz are less than -p. For instance,
2p 2p
pz := r(2l- 1)' qz = r(2k- 2l + 1)'
will do; thus, the expression (10.63) vanishes. D

Corollary 10.21. Let (.Jt, Jf, D) be a p+ -summable spectral triple. If both


.Jt and [D,.Jt] := { [D, a] : a E .Jt} lie within the domain of the derivation
8 : T ..... [IDI, T), then T ..... Trw TIDI-P is a hypertrace on the algebra
generared by .Jt and [D,.Jt].
Proof. Let 3t'v denote the algebra generated by .Jt u [D,.Jt]. The result is
an immediate consequence of Theorem 10.20, provided we can show that
(3\v,Jf, IDI) is also a p+-summable spectral triple. Since 8 is a derivation,
it is enough that every [IDI,a] and every [IDI, [D,a]] be bounded, which
is what the extra hypothesis guarantees. D
Remark. We stated that not every p+-summable spectral tripleis tarne. An
instructive counterexample is given in [81], showing that some smoothness
condition, to prevent .Jt from becoming too big, is indispensable .

.,. The boundedness of [I D 1. a] is not assured by the hypotheses of Lemma


10.17. Indeed, by writing IDI = FD, one sees that [IDI,a] = F[D,a] +
[F,a]D, so that boundedness of [IDI,a] isanontrivial requirement, even
when [D, a] is already bounded. From now on, the following regularity
property of spectral triples will be crucial.
Definition 10.10. A spectral triple (.Jt,Jf,D) is called regular if, for each
a E .Jt and k E ~. both a and [D, a] belong to Dom 8k, where

8(T) := [IDI, T).

We again denote by 3t'v the algebra (of bounded operators) generated by


.Jt and [D,.Jt] := { [D, a] : a E .Jt }. The regularity condition means that
3t'v ~ Domoo 8, the smooth domain of the derivation 8.
The prime example of a regular spectral triple is the case where D is
a (generalized) Dirac operator and .Jt = coo (M) is the space of smooth
functions on a manifold M. To understand why that should be so, first
recall that in pseudodifferential calculus, the smooth functions are those
which belong to Sobolev spaces of arbitrarily high order. Weshall now Iift
the curtain on an abstract pseudodifferential calculus developed by Connes
and Moscovici in [92,113].
Definition 10.11. Let (.Jt, Jf, D) be a regular spectral triple. A scale of Hil-
bert spaces { J{s : s E ~ } is defined by setting

Jf 5 :=Dom IDI 5 for each s E ~-


10.3 Tameness and regularity of spectral triples 467

Under the norm 11~11~ := 11~11 2 + IIIDI 5 ~11 2 , each J-f5 is a Hilbert space, and
for s > t, the inclusion J-f 5 ..... J-{t is continuous. Notice that J-{ 0 = J-f. The
intersection

J-{00 := n
sEIR
J-{S = n00

k=O
J-[k = Domoo IDI (10.64)

is a Frechet space under the norms II · llk. for k E ~.


Foreach r E 7l, we introduce the vector space Opö of operators T: J-[oo -
J-[oo for which there are constants C5 satisfying

IIT~IIs-r 5 Csll~lls, for all ~ E J-f 00 • (10.65)

so that T extends to a bounded operator from J-f 5 to J-[s-r for every s E IR.
Remark. 1t is enough to know that (10.65) holds for integer values of s,
since the interpolation theory of Banach spaces yields corresponding in-
equalities for the intermediate values. This works as follows: if T is an ope-
rator satisfying (10.65) for s = k and s = k + 1, then since J-[k+ 1 c J-[k
and J-[k+ 1-r c J-[k-r, for each s E [k, k + 1] we can find a constant
Cs 5 max(Ck, Ck+1) thatis anoperatorboundforthenorms ll.lls-r and ll·lls.
Fora useful summary of these matters, see [91, IV.B].
Lemma 10.22. lf (.Jt,J-f,D) is a regular spectral triple, then Av s;; Opß.
Also, b- IDI b IDI- 1 E Opi)1 for each b E Xv.

Proof. Regularity says that any b E Av lies in Dom8k for each k E ~.


Clearly IDI b IDI- 1 = b + 8(b) IDI- 1 is weH defined since 8(b) and IDI- 1
are bounded operators. Also,

On the other hand, IDI- 1 b IDI = b- IDI- 1 8(b), and by pulling IDI to the
left twice we get IDI- 2 b IDI 2 = b- 2IDI- 1 8(b) + IDI- 2 8 2 (b). By induction,
we arrive at

IDik b IDI-k = I (~)81 (b) IDI-J,


j=O J

IDI-kbiDik = I(-1)J(~)IDI-J8J(b),
j=O J

for k = 0,1,2, .... lt follows that IDik b IDI-k is a bounded operator, for
each k E 7l. If ~ E J-[oo, then

llb~ll~ = llb~ll 2 + IIIDikb~ll 2


5 llbll 2 11~11 2 + IIIDik b IDI-k 11 2 111Dik~ll 2 5 Cf 11~11~.
468 10. Spectral Tripies

where Ck =s; max(llbll~ 11/Dik b /DI-k II). This shows that b E Op~.
Regularity also implies that 8(b) E Dom"" 8 whenever b E 3t'v~ so we
may conclude that 8(b) E Op~~ too. lt is clear from Definition 10.11 that
T/D/k E Opf/k whenever T E Opf>; it then follows that b- IDI b IDI- 1 =
-8(b) IDI- 1 E Opi)1. D

We shall also need to consider the unbounded derivation adD 2 : T -


[D 2 T]. Following [113] weshall abbreviate
1 1

y(k) := (adD2)k(T).

This derivation generally does not preserve boundednessl but it may weil
happenl for a given operator T that y(r) /DI-r remains bounded for each
1

r =1 2 3 1 1 we claim that the operators [D a] have this property when


1 ••• ; 1

the spectral triple is regular. For that it is enough to show that each T E
1

Dom"" 8 satisfies y(r) E Opf> for every r. Notice that b(r) /DI-r = Rr(b)
and /DI-rb(r) = F(b) where the transformations L R are defined by
1 1

(10.66)

Indeedl R 2 (b) = [D 2 [D 2 b]IDI- 1J IDI- 1 = [D 2 [D 2 b]] /DI- 2 = b( 2l /DI- 2


1 1 1 1

since D 2 and /DI- 1 commute and L2(b) = /DI- 2b( 2l similarly; induction
1

gives their higher powers.


Exercise 10.10. Using the identity [D 2 b] I = /D I 8 (b) + 8 (b) /D I show that
I

Rr(b) = ±(r)
k=O k
/D/k 8r(b) /D/-k 1

and conclude that Rr(b) is a bounded operator whenever b E Dom"" 8. 0

In fact the proof of Lemma 10.22 shows that if b E Dom"" 8 then each
1 1

summand of Rr(b) lies in Op~; thereforel b(r) E Opf> for any r E ~.


The transformationsLand R commute and LkR 1(b) = /DI-k b(k+l) IDI- 1 1

for kl l E ~. The common domain of all L k R 1 certainly includes Dom"" 8 I

andin fact coincides with this smooth domainl by the following calculation.
Lemma 10.23. I( L1 R are defined by (10.66) and 8 = ad ID/, the common
smooth domain nk,l=O Dom(L kR 1) equals Dom"" 8.
Proof. Assurne that b E DomR n DomL 2. We can express commutators
[IDI~ b] in terms of [D 2 b] as in (10.56): 1 1

[ID/~bJ = [D 2 1DI- 1 ~bJ = [D 2 ~bl IDI- 1 +D 2 [/DI- 1 ~bJ


= ! J: [D2, b] (,\ + D2)-1 _ D2(,\ + D2)-1 [D2 b] (,\ + D2)-1 1 ~
= _!_ f""(A+D 2)- 1 [D 2 b](A+D 2)- 1 v'AdA. (10.67)
rr Jo 1
10.3 Tameness and regularity of spectral triples 469

If we move [D 2, b] to the left of the last integrand, this becomes

Since rr- 1 Jo"" t (.\ + t 2)- 2 JXd.\ = rr- 1 J0"" 2u 2(1 + u 2)- 2 du = ~. this ex-
pression reduces to ~R(b), which is bounded. Torecover [IDI, b], we must
deal with the commutator

[ (.\ + D2) -1, [D2 b ]] = - (.\ + D2 )-1 [D2 [D2 b ]] (.\ + D2) -1
I I I

= -(.\+D2)-1D2L2(b) (.\+Dz)-1.

Since 11(.\ + D 2)- 1D 2 11 ~ 1 for all.\ ~ 0 and since L 2 (b) is bounded, we get
the estimate

The integral is majorized by MIID- 4 11 JX d.\ + ft' .\- 3 12 d.\ = ~ IID- 4 11 + ~;


we conclude that [I D I. b] is also bounded.
Wehave shown thatDomRnDomL 2 ~ Domö. Ifwereplace bbyö(b), the
same calculation yields DomR 2 n DomL 2 R n DomL 4 ~ Domö 2; repeating
this argument shows that nkl=oDom(LkR 1) ~ Domör for any r. D

~ Another property of regular spectral triples that we briefly consider is


whether or not .J'l is a pre-C*-algebra. Since .J'l ~ Dom"" ö, we may confer
on .J'l the locally convex topology generated bythe seminorms a - llök (a) II.
for k E ~- .J'l need not be complete in this topology, but we may complete it
if necessary and represent the completion Jlby bounded operators on Jf,
since the new topology is stronger than the norm topology of .J'l. However,
it is not obvious that the commutators [D, a] will remain bounded for el-
ements of the completion. Provided that this last condition is fulfilled, we
obtain a spectral triple (Jl, Jf, D) where 3l is a Frechet algebra.
Suppose, then, that .J'l is complete in the aforementioned topology; then
.J'l = n .J'ln where .J'I.n is the Banach algebra obtained by completing .J'l
I

in the norm a - Lk=O llök(a)ll. It can then be shown, by suitable norm


estimates [327], that the invertible elements of the C*-completion A = 5\o
that lie in .J'I.n are already invertible in .J'I.n; in other words, .J'I.n n A x = .J'l~.
Taking the intersection over all n gives .J'l n A x = .J'lx, so that .J'l is a good
locally convex algebra.
The upshat is that completeness of .J'l guarantees that .J'l is a Frechet
pre-C*-algebra. We therefore do not hesitate to include this as a desirable
property of spectral triples.
470 10. Spectral Tripies

10.4 Connes' character formula


The Chern character of a finitely summable Fredholm module (5\, Jf, F)
has been identified, in Section 10.1, as the periodic cydic cohomology dass
of the cydic n-cocyde (10.23):
Tf(ao, ... , an):= An Tr(xF[F, ao] ... [F, an]).

where n is any integer of the same parity as (.5\,Jf,F), large enough that
the trace converges, and An:=[(~+ 1)/2n!. For instance, if (5\,Jf,D)
is an n+-summable spectral triple and F = DIDI- 1 , then Lemma 10.18
shows that [F, a] E LP for all p > n, in particular for p = n + 1, so that
[F, ao] ... [F, an] is tracedass.
However, as the examples in Section 8.2 should indicate, the direct calcu-
lation of the Chern character from this formula may be fraught with diffi-
culties. Even in the commutative case, when 5\ = C"" (M), its evaluation in-
volves integrals like fM" Jdx1ox2)f2(x2, X3) .. . fn (Xn,X1) dnx, where the
fi may be singular integral kernels. Contrast this with the Dixmier trace,
which in the commutative case leads to ordinary integrals in one variable.
1t is therefore very desirable tobe find a way to compute the Chern charac-
ter by a "local formula", in which ordinary traces are replaced by Dixmier
traces or suitable generalizations thereof.
Such a local formula was indeed found by Connes and Moscovici in [113),
after considerable effort and under extra technical assumptions on the
spectral triple. We shall comment further on that at the end of this sec-
tion. For our purposes in this book, however, we need only to determine
the Hochschild dass of Tf; that is to say, we must find a Hochschild -not
necessarily cydic- n-cocyde which agrees with Tf on Hochschild n-cydes,
and which is "local" in the sense that it is given by a Dixmier trace. Such
a Hochschild cocyde was constructed by Connes in 1987, and announced
in [89); for further discussion, see [91, Thm. IV.2.8). The detailed construc-
tion, however, has not appeared in print before now. In this section we
develop that construction, based on the original notes [102) which Alain
Connes kindly made available to us.
Westart with an n+-summable spectral triple (.5\,Jf,D), graded or un-
graded according as n is even or odd, and let F := Dl Dl- 1 . We further
assume a weak regularity property of the spectral triple, namely that the
algebra 3\0 lie in the domain of 8 2, where ö(T) := [IDI, T]. (This is enough
to ensure the tameness property of Corollary 10.21).
Now let Trw be any Dixmier trace, associated to a state w of Boo as in
Section 7. 5.
Lemma 10.24. The (n + 1)-linear functional cp[j on 5\ defined by

cp[j(ao, ... ,an) :=An Trw(Xao [D,ai] ... [D,an] IDI-n) (10.68)
is a Hochschild n-cocycle on 5\.
10.4 Connes' character formula 471

Proof. Since ID\- 1 E Ln+ and each [D,aj] is bounded, then ID\-n E L 1+
and the operator ao [D, ad ... [D, an] ID\ -n lies in the Dixmier trace dass,
so cpj!f is finite-valued on 5\. ®(n+ 1>. To see that bcpj!f = 0, notice that since
a ..... [D, a] is a derivation, the expression (8.46) for bcpj!f (a 0 , ... , an+ I)
telescopes to

(-l)ni\n (Trw(Xao [D,ar] ... [D,anlan+11D\-n)


- Trw(Xan+1ao [D,ad ... [D,an] ID\-n)),

which is of the form Trw([T,an+IliD\-n) with TE L(3f), and therefore


vanishes by Theorem 10.20. 0

Recall that the pairing (10.26a) between Hochschild cochains and chains
satisfies (b<p, c) = (<p, bc). lt follows that, if cp 1 and <p2 are cohomologous
n-cocydes, with <p 1 - <p2 = btJ.I, say, their pairings with any Hochschild
n-cyde yield the same values, since

(<p1,c)- (<p2,c) = (btJ.I,c) = (tJ.I,bc) = 0 whenever bc = 0.

The converse -equality of pairings with cydes implying that the cocydes
are cohomologous- holds whenever HH• (5\.) and HH. (5\.) aredual spaces
via the induced pairing between cohomology and homology dasses; but we
shall not need this converse result explicitly. Our goal in this section is to
establish that (Tf,c) = (<pjJ,c) for any Hochschild n-cyde c; this is the
precise meaning of our daim that "cpj!f represents the Hochschild dass of
the Chern character Tf" .
.,. The first step is to approximate the expression (10.23) for Tf by the trace
of a finite-rank operator. For that, we use a cutoff of the type described in
Section 7.B: choose a function g: [ 0, oo) - IR!. with g ( t) = 1 for 0 s; t s; ~,
g decreases smoothly from 1 to 0 for ~ s; t s; 1, and g(t) = 0 fort :2::. 1;
and Iet g(t) := g(-t) fort< 0. Theng E V(IR!.) with suppg ~ [ -1, 1]. Let
At := g(tiDI) fort > 0; this is a positive, finite-rank operator satisfying
P112t s; At s; P1;t. where PN denotes the spectral projector of IDI on the
interval [O,N].
Lemma 10.25. If ao, ... , an E 5\., then

Tr(xF[F, ao] ... [F, an]) = lim Tr(XAtao[F, ar] ... [F, an]).
tlO

Proof. The left hand side may be rewritten as Tr' (X ao [F, a 1] ... [F, an]),
where Tr': T ..... ~(T + FTF) is the conditional trace (8.12). The spectral
projectors P 11 t are finite-rank operators, and P 11 t l 1 weakly as t I 0. If
S E L(3f), then

Tr' S = lim Tr' (Pr;tS) = lim Tr(P1;tS),


t lO t !0
4 72 10. Spectral Tripies

since Tr' and Tr coincide on traceclass operators. The inequality P1;2t 5


At 5 P1 ;t shows that limt~o Tr(Pl/tS) = limt~o Tr(AtS) when S is positive;
the same is true for any S by polarization. For S = xao [F, a 1] ... [F, an],

Tr' (xao [F, ad ... [F, an]) = lim Tr(Atxao[F, ad ... [F, an]).
t10

Since IDI commutes with x. so does At, and the result follows. D
Nowlet c = LJ ao1 ®a 1J®· · · ®anJ be aHochschild n-cycle. The condition
bc = 0 may be reformulated, using (8.41), as LJ b(a 01 dal} ... danJ) = 0 in
on- 1.Jl. By Exercise 8.20, this simplifies to

Lj[aojda1J···dan-1,J•anJ] = 0.
The universal property of the graded differential algebra (0" .Jl, d) shows
that the analogaus relation holds whenever d is replaced by any other
derivation from .Jl to an .Jl-bimodule, andin particular by a ..... [F, a]. We
conclude that

Lj aoJ [F, a1j] ... [F, an-1,}] an} = LJ anJaOJ [F, a1j] ... [F, an-1,} ].
Ifwe abbreviate s1 := a 01 [F,alj] ... [F,an-I,J], we may rewrite this equal-
ity as L, 1 [s1,an1 ] = O.lt follows that A;; 1(Tp-,c) is approximated by

LjTr(xAtSj[F,anj]) = LjTr(xSjFanjAt- xSjanjFAt)


= Lj Tr(xSjFanjAt - xanjSjF At)
= - LJ Tr(XSJF [At, an}]),

which Ieads to the following approximation formula.


Lemma 10.26. Define«!Jt E cn(.Jl,.Jl*) by

«jJt(ao, ... , an) :=-An Tr(xao[F, ad ... [F, an-dF[At, an]).

Then (T?, c) = limtlo(«/Jt, c) whenever c is a Hochschild n-cycle. a


Lemma 10.27. The operator ao[F, aJ] ... [F, an-dFIDin-I is bounded, for
any ao, ... , an-1 E .Jl.

Proof. Since D = FIDI = IDIF, we may write, for any a E .Jl,

[D,a] = [F,a]IDI +F[IDI,a] = [F,a]IDI +F8a,

and similarly [D,a] = IDI[F,a] + 8aF. Therefore, [IDI, [F,a]] = [F,8a]


is bounded and in fact lies in LP for p > n. Indeed, on replacing a by 8a
in (10.57) and using [D,8a] = [D, [IDI,a]] = [IDI, [D,a]] = 8([D,a]), we
see that [F,8a] is dominated by 118([D, a]) IIIDI- 1 when a* = ±a. Thus, it
is enough to show that ao[F,adiDI ... [F,an-diDIF is bounded; and this
follows at once from the relation [F, a] IDI = [D, a]- F 8a. D
10.4 Connes' character formula 473

For convenience, let R := -.\nxao[F, ad ... [F, an-dFIDin- 1 E L(J-f),


so that

(10.69)

We nowuse the normestimate ofLemma 10.19 on [At. an] = [g(tiDI ), an],


replacing D by IDI in the Statement of the lemma. This yields

II[At,anllln-::;; Cn(g) II[IDI,anliiiiD- 1 IIn+ forall t > 0.


Let q = nl(n- 1), so that v-<n- 1l E Lq+. Since Lq+ is the Banach-space
dual of the operatorideal Ln-, as explained in Section 7.C, (10.69) may be
estimated by

ltJ.Idao, ... ,an)l::;; IIRIIIID-(n- 1 )llq+II[At,anllln-


::;; Cn(g) IIRIIII8aniiiiD-<n- 1lllq+IID- 1 IIn+• (10.70)

so the continuous function t - tJ.Idao, ... , an) is bounded for 11t > 3,
say. Evaluation of the state w on the corresponding element of Boo gives a
number which we shall denote by

.,.. Computation of this generalized limit is achieved by the following propo-


sition of Connes [102]. For any continuous function f: [0, oo) - ~ and any
integer k, write

mk(j) := sup{f(u): k::;; logu::;; k + 1 }.

Proposition 10.28. If f: [0, oo) - [0, oo) is continuous, p > 1, and if the
series .L:k mk(j)ePk converges, then Mp := p f0"" j(u)uP- 1 du is finite, and

lim tPTr(j(tiDI)S) = Mp Trw(SIDI-P) (10.71)


t- 1 -w

wheneverS E L(J-f) andD- 1 E LP+(Jf).

Proof. The hypothesis on f and the estimate

establish the finiteness of Mp. lf we subdivide each [k, k + 1] into r equal


subintervals, and write mkJ (j) := sup{ j(e 5 ) : k + j Ir ::;; s ::;; k + (j + 1) Ir}
for j = 0, ... , r - 1, we can likewise estimate

Mp ::;; I mkJ(f) eP<k+J/rl (ePir- 1) ~ I mk(j) ePk(eP- 1).


kJ k
474 10. Spectral Tripies

By choosing r !arge enough, we can sandwich f between two functions h 1 ,


h2 of the general form

k k+1
h(u) := Aku-P for - ::5logu ::5 - - , (10.72)
r r
where each Ak ~ 0 and .l:kAk < oo; suchthat f0"" lhi(u)- hz(u)l uP- 1 du
is as small as we please. lt is enough to replace f by a function of the form
(10.72) and to verify (10.71) for such functions. In that case,

Now Iet Pk be the spectral projector of ID I on the interval [eklr, e(k+ 1llr]
and let NJDI be the counting function of ID 1. By Exercise 7.26, the hypothesis
IDI- 1 E LP+ is equivalent to NJDI (u) ::5 CuP for some C > 0. Thus,

Tr(PkiDI-P) = fe<k+l)/r

eklr
u-P dNIDI (u)

Since this bound is independent of k, the sequence

is also bounded, by (1 + p/r)CIISII. We may therefore consider the genera-


lized Iimit

as defined in Section 7.5, applied to {adho·


If we replace {ak }k;;, 0 by the sequence of Cesära means 01.1 : = z- 1 2.:i":1 ak.
this generalized Iimit yields the desired Dixmier trace. For that, notice that
Po+ · · · + P1- 1 is the spectral projector for ID I on [1, elfr], which contains
N v (e 11r) = O(eP 11r) eigenvalues; in the notation of Lemma 7.17, Po+· · · +
1 1

P1- 1 = EN where logN - pl/r. By invoking that Iemma for the operator
IDI-P E L 1+, we conclude that

On the other hand, if m := -rlog t so that t- 1 = emfr, then tP j(tu) =


Aku-P when k!r ::5log(tu) ::5 (k + 1)/r, that is, when (k + m)/r ::5logu ::5
(k + m + l)Jr. Choosing t suchthat m E ~. we obtain

tPTr(j(tiDI)S) = LAkTr(Pk+miDI-PS) = LAkak+m·


k k
10.4 Connes' character formula 475

Define a new bounded sequence by

(10.73)

Then
!:_MpLw({ßm}) = lim tPTr(j(tiDI)S).
p t-1-w

To finish the proof, we need only ensure that

For that, it is enough to check that ak ,._ oq and ak ,._ ßm are regu-
lar sequence transformations. A linear transformation of sequences bJ :=
L.:k';o c1kak is called regular ifwhenever {ak} converges, then {bJ} also con-
verges to the same Iimit. The Toeplitz-Schur theorem [231, §3.2] states that
such a transformationpreserves convergence if and onlyif (a) YJ := Lk lcJk I
is bounded independently of j; (b) for each k, 8k := limJ-oo CJk exists; and (c)
if c1 := Lk CJk. then 8 := limJ-oo c1 exists; furthermore, the transformation
is regular if and only if 8k = 0 for each k and 8 = 1.
All those conditions are easily checked for the Cesaro case ak ,._ Olz,
where CJk := 1/(j + 1) for k = 0, 1, ... ,j and CJk := 0 otherwise. The other
transformation may be obtained by first setting cJk := Ak-J for k ~ j and
CJk := 0 for k < j. In that case c1 = YJ = Lk .\k for each j, and 8k = 0. The
normalization (10.73) then implies that ak ,._ ßm is regular, too. D

.,. We return now to the functionali/Jt of (10.69). The next step is to replace
[At, an]= [g(tiDI),an] by a more tractable substitute.

Lemma 10.29. If g(t) = h(t) 2 where h E :D(~) is also a cutoff, and if


a E .Jl, then

ll[g(tiDI),a]- ~{g'(tiDI),t8a}lln- = o(t) as t I 0. (10.74)

Proof. Write Rt := h(tiDI) = A~ 12 and St := h'(tiDI); since g = 2hh', the


argument of the norm II · lln- in (10.74) is

[Rf,a]- tRtSt [IDI,a]- t[IDI,a]StRt


= Rt([Rt,a]- tSt [IDI,a])+([Rt,a]- t[IDI,a]St)Rt.
Now Corollary 10.16, with D replaced by IDI, shows that

where C' = (4rr)- 1 Coos 2 ih(s)l ds. The same estimate holds for [Rt,a]-
t[IDI,a]St, too.
476 10. Spectral Tripies

On the other hand,


IIRtlln- ~ IIPittlln- ~ L (k + 1)-l/n- N~n-1)/n,
ksNr

where Nr := NIDI(l/t) = o(t- 1fn) as t I 0 since Ln- c Ln+, using Exer-


cise 7.26. We conclude that IIRrlln- = O(t-<n-ll/n 2 ), and the left hand side
of (10.74) is then of order O(t 2 -<n-ll/n 2 ) as t I 0. 0

Let f(t) := g' (t) /tn- 1 for t * 0 and j(O) := 0; then f E 1J(~). Now,
1./.lt (ao, ... , an) = Tr(R ID 1-(n- 1) [g(t ID I), an])
= !tnTr(Rj(t/DI) 8an)
+ !tnTr(RIDI-<n-ll8anf(tiDI)IDin-I) + o(t)
= !tnTr(j(tiDI) {R,8an})
+ !tnTr(R[IDI-<n-ll,8an]f(tiDI)IDin-l) + o(t).

Since [IDI-<n-ll,8an] =- 'Li:::fiDI-k8 2 (an)IDI-n+k, the second term on


the right is tn Tr(j(t/DI )T) with T = -! 'Li:::fiDik- 1R IDI-k8 2 (an) E Ln+;
therefore, IDI-nT is traceclass, so this term vanishes in the generalized
Iimit as t- 1 - w.
Lemma 10.29 shows that [At, an] maybe replaced by! {g' (tiDI), t 8an}
while keeping the estimate (10.70) that justifies taking the Iimit t- 1 - w,
and therefore
lim f.Jit<ao, ... ,an)=! lim tnTr(j(tiDI){R,8an})
r- 1 -w r- 1 -w
=!Mn Trw( {R,8an} IDI-n) =Mn Trw(R 8an IDI-n).

The last equality follows from the tameness property of Corollary 10.21.
Since Mn= n f0"" j(u)un- 1 du= n f0"" g' (u) du= -n, we conclude that

lim f.Jit<ao, ... , an)


r-1-w
= ni\n Trw(Xao[F,ad ... [F,an-dFIDin- 1 [1DI,an]IDI-n)
= ni\n Trw(Xao[F,ad ... [F,an-dD- 1 [1DI,an]). (10.75)
The second equality comes from exchanging [ IDI, an] and IDI-n under the
Dixmier trace. This does not alter the result, because [I Dl-n, [I D 1. an]] =
- Lk=O IDI-k- 18 2 (an)IDI-n+k is of the order of IDI-n- 1, so that the two
arguments of Trw in (10.75) differ by a traceclass operator.

to show that
Since
"n
11> Denote the right hand side of (10.75) by "n(ao, ...• an). The final step is
is a Hochschild n-cocycle which is cohomologous to cp'fj.
10.4 Connes' character formula 477

because (D- 1, [ IDI, an]] = -D- 18([D, an])D- 1 is of the order of n- 2 , the
computation of bl;;n (ao, ... , an+1) gives

(-l)n- 1n.\n Trw(Xao[F,ad ... [F,an-Ilan[IDI,an+dD- 1)


+ ( -l)nnt\n Trw(Xao[F, ad ... [F, an-d [IDI, anan+dD- 1)
+ (-l)n+ 1n.\n Trw(Xan+1ao[F,ad ... [F,an-d [IDI,an]D- 1)
= ( -l)n+ 1n.\n Trw(Xao[F, ad ... [F, an-d [IDI, an] (D- 1, an+Il),

which vanishes because the last argument of Trw is traceclass.


We introduce more cochains '(;1 , ... , l;;n- 1 by defining '(;k(ao, ... , an) as

Lemma 10.30. The cochains '1;;1 •••• , l;;n are Hochschildn-cocycles which are
rnutually cohornologous.

Proof. We already know that bl;;n = 0, so it is enough to produce (n - 1 )-


cochains 111. ... , 1Jn-l suchthat lJlk - lJlk+1 = bTJk for k = 1, ... , n- 1.
In the formula which defines 'i;;k. we may move n- 1 to the right under
the Dixrnier trace. Indeed, (D- 1, [ ID I, ak]] = -D- 1 8([D, ak])D- 1 lies in
L<n1 2l+ and each [F,a 1 ], for j * k, lies in LP for p > n, so their product
is traceclass; thus, we can replace n- 1 [1DI,ak] by [IDI,ak]D- 1 . To move
n- 1 = FID!- 1 past [F,a 1 ] for j > k, we likewise note that

which yields a traceclass expression when it replaces [F, a 1 ]; and finally,


F[F,a1 ] = -[F,a1 ]F. In summary, moving n- 1 to the extreme right only
changes '(;k by a sign factor ( -1) n-k:

'(;k(ao, ... , an) = ( -l)n-knt\n Trw <xao[F, ad ... [ IDI, ak] ... [F, anlD- 1 ).

Now (l/Jk -l/Jk-d(ao, ... ,an) maybe writtenas

Since

[F, 8(akak+1)] = [F, ak 8ak+1 + 8ak ak+Il


= ak [F, 8ak+Il + Rk + [F, 8akl ak+1.

it follows easily that bTJk = l/Jk - l/Jk+1· 0


478 10. Spectral Tripies

To establish that [ (n] = [ cpyj] in H Hn (51.), it is therefore enough to show


that [(Il + · · · + [(nl = n[cpyj ]. This is the content of the next proposition.
Proposition 10.31. The cochain cpyj- n- 1 ((1 + · · · + (n) is a Hochschild
n -coboundary.
Proof. In the expression (10.68) for cpyj (ao, ... , an), we may exchange the
operators [D, a1 ] and IDI- 1 under the Dixmier trace, by the same argument
we have used several times already. Therefore,

Moreover,

[D,aJ]IDI- 1 = [FIDI,aJ]IDI- 1 = [F,aJ] +F[IDI,aJ]IDI- 1


= [F,aJ] + 8a1D- 1 + [F,8aJ]IDI- 1.
Invoking (10.57) with a replaced by 8a1, and using [D,8a 1 ] = 8([D,a1 ]),
we see that [F, 8aJ] IDI- 1 is ofthe order of 118 ( [D, a J]) II D- 2 , so those terms
vanish under the Dixmier trace, and we may replace each [D,a1 ]1DI- 1
in (10.76) by [F, a 1 ] + 8a1 D-1.
Under this replacement, (10.76) becomes a sum of 2n terms, the first
ofwhich is An Trw(Xao [F,ad . .. [F,an]). Since ao = F[F,ao] + FaoF and
xF[F, ao] ... [F, an] is traceclass, this term becomes

An Trw(xFaoF[F,ad ... [F,an]) = (-l)nAn Trw(xFao[F,ad . .. [F,an]F)


= -An Trw (Fxao[F, ad ... [F, an]F)
=-An Trw(Xao[F,ad ... [F,an]),

and therefore it vanishes.


The terms having a single 8a1 D- 1 factor add up to n- 1 ( (1 + · · · + (n).
lt only remains to checkthat terms with two or more 8a1 D- 1 factors are
coboundaries. For instance, since

it follows that

An Trw(Xao[F, ad ... [F, aJ-Il 8aJ 8aJ+1 [F, aJ+2] ... [F, anJD- 2)

equals biJIJ(ao, ... ,an), where

1J1J(ao, ... , an-d = -! (-1)1 An Trw (xao[F, ad ... 8 2(aJ) ... [F, an-IlD- 2 ),
so two consecutive 8a1 D- 1 factors yield a coboundary. 0

Exercise 10.11. Checkthat the remaining terms in the expansion of cpyj


are Hochschild coboundaries, too. 0
10.4 Connes' character formula 479

Putting it all together, we arrive at the main result of this chapter.


Theorem 10.32 (Connes' character formula). If (5\.,Jf,D) is an n+-sum-
mable spectral triple, and ifthe algebra Av generated by .Jt and [D, .Jt] lies
within the domain of 8 2 = (ad ID I) 2 , then, for any Dixmier trace Trw, the
Hochschild n-cocycle cp'Jj and the cyclic n-cocycle Tf yield the same value
on every Hochschild n-cycle c:

(cp'Jj,c} = (Tf,c}, forall c E Zn(.Jt,.Jt).

More explicitly: with An = [(I + 1) /2 n!,


L.J An Trw(xaoj [D,aiJ] ... [D,anJ] IDI-n)
= L.J An Tr(xF[F, aoJ] [F, a1J] ... [F, anJ]) (10.77)

wheneverc = L.J aoJ ® a1J ® •· • ® anJ satisfies bc = 0. D


Theorem 8.2, which gives a local expression of the Chern character for
tori, is a direct consequence of this formula. First replace the constant An
on both sides of (10.77) by iin, for compatibility with (8.21b). Then, for the
Dirac K-cycle over any compact spin manifold M, Connes' trace theorem
allows us to compute the left hand side of (10.77) as
in-m
n( 2 rr)n Wres(x L.J aoJ c(daiJ) ... c(danJ) 11])1-n)

=
(2i)mnn
n( 2rr)n
JM L.J aoJ da1J 1\ · • • 1\ danJ.

where m = Ln/2J, as before. In calculating this Wresidue, tr(x · ), with the


trace coming from the spin representation, acts as a Berezin integral and
only the top exterior form survives. The constant before the integral is
the one given by (8.21), and Theorem 8.2 is thereby established for any
compact spin manifold. Indeed, arguing as in Corollary 7.22, Langmann's
noncompact flat case [312] can be proved, too.
~ One can go much further than the character formula (10.77).lt is indeed
possible, by imposing some extra conditions on the spectral triple, to write
down a local formula for the full Chern character, rather than merely its
Hochschild dass. This is the celebrated local index formula of Connes and
Moscovici [113]. Before stating it, we indicate briefly how it comes about.
The road to the local index formula for finitely summable spectral triples
begins with the observation, made at the end of Section 10.2, that the char-
acter [Tf] may be obtained by performing the Iimit t - oo on the family
{ ch7(D) : 0 < t < oo} of entire cyclic cocycles, defined by (10.45), which
are all mutually cohomologous. It is natural to enquire also about the Iimit
as t I 0. Indeed, the heat-kernel proof of the Atiyah-Singer index theorem
starts from the observation by McKean and Singer that a supertrace of the
form Tr(xe-tD 2 ) is independentoft for 0 < t < oo; the Iimit t - oo, if it
480 10. Spectral Tripies

exists, gives the Fredholm index of D, while the limit t I 0 gives a local for-
mula for the index. However, that requires some work, because the latter
limit is singular: when Dis a pseudodifferential operator, the local formula
is extracted from an asymptotic development of the heat kernel.
After a lengthy analysis, Connes and Moscovici have shown [113) that,
in favourable circumstances, one can compute a finite part of the limit of
eh~ (D) as t I 0, and that this finite part is cohomologous to any eh~ (D)
with t > 0. The circumstances that allow to control the divergences and
extract the finite part as a certain zeta residue are: (a) finite summability of
the spectral triple (.Jt, Jf, D); (b) regularity, in the sense ofDefinition 10.10;
and (c) the property that, as P runs over the algebra generated by allök(a)
and all8k ( [D, a)), for a E .Jt and k E N, the functions '(p(z) := Tr(PIDI-z)
extend holomorphically to C \ Sd, where Sd is a discrete subset of the
complex plane, depending only on D, where the functions '(p may have at
mostsimple poles. The set Sd is then called the "dimension spectrum" of
(.Jt, Jf, D); for the Dirac spectral triple with .Jt = coo (M), Sd consists of
the integers {0, 1, ... , dimM}.
With these provisos, the local index theorem reads as follows [97, 98), in
the odd case.
Theorem 10.33. Let (.Jt, Jf, D) be a finitely summable, ungraded, regular
spectral triple with discrete and simple dimension spectrum. Then the equal-
ity
f P := ~;8Tr(P IDI- 5)

defines a trace on the algebra generated by .Jt, [D, .Jt] and { ID Iz : z E C } .


The following formula contains only a finite number of nonzero terms and
defines the odd components { CfJn : n = 1, 3, 5, ... } of a cocycle cp in the
complex Tot" BC(.Jt):

CfJn (ao, ···,an) = 2:


ke7ln
Cn,k f ao [D, ad(kJl ... [D, an](kn) IDI-n- 2 1kl,

(10.78a)

denoting y(r) := (adD 2 )Y(T) and lkl := k1 + · · · + kn. The coefficients are
.J2i (-1)1kl f(~ + lkl)
(10.78b)
Cn,k = -k-!- (k1 + 1)(kl + k2 + 2) ... (lkl + n) ·
The pairing ofthe cyclic cohomology dass ofcp in HC" (.Jt) with K1(.Jt) gives
the Fredholm index of D with coefficients in K1 (.Jt).
The last assertion of the theorem deserves some further explanation.
When u E Mr(.Jt) is a unitary representing a dass [u] E KI(.Jt) and 1./J is a
cyclic cocycle of odd degree 2k + 1, the pairing

(r.Jl,Ch2k+Iu) = (-1)kk!r.Jl®tr(u- 1,u, ... ,u- 1,u)


10.5 Termsand conditions for spin geometries 481

of (10.26c) depends only on [u] and on [!JI] E HC 2k+ 1 (5t) and yields a
pairing between K1 (5t) and Hcodd(.Jt), which in fact induces a pairing be-
tween K1 (5t) and HP 1 (5t): see [91, III.3]. When tJ1 = v'2l Tjk+ 1 , the result
is an integer, namely the index of the Fredholm operator QruQr. where
Qr is the projector !o + F) ® 1r on J{ ® e; and [u] ...... index(QruQr)
is a well-defined homomorphism from K1 (5t) to 71.. The theorem therefore
yields a local formula that computes the Chern character [Tp] in the odd
case.
In the even case, there is a similar result, where one replaces ao by xao
in (10.78a) and omits the factor v'2l in (10.78b). However, the lowest-degree
term is given instead by <Po(ao) = Ress~os- 1 Tr(xao IDI- 5 ). We refer to
[113, Thm. 4] for the details. Indeed, the development in [113] deals with
the more general case wherein the dimension spectrum need not be simple
but may have multiplicities; some extra terms then appear in (10.78).

10.5 Terms and conditions for spin geometries


We now bring together several strands and lay out a structure deserving
the name of noncommutative geometry. Briefly, this is a finitely summable,
regular spectral triple, with an "orientation dass" and a "real structure"
that generaUze the orientation and spin structure of a spin manifold. The
formal definitions follow.
Definition 10.12. Areal spectral tripleis a spectral triple (5t, J{, D), to-
gether with an antiunitary operator C on J{, suchthat b ...... Cb*c- 1 deter-
mines an action of the opposite algebra 5t o on J{ that commutes with the
action of .Jl, that is,
[a, Cb* c- 1 ] = 0, for all a, b E 5t. (10.79)
To avoid ambiguity, weshall occasionally denote by rr the given repre-
sentation of 5t by bounded operators; in other words, rr(a)~ = a~. for
a E 5t and ~ E J{. Then
(10.80)
is a representation of 5t o on J{, and (1 0. 79) may be interpreted as the
statement that the operator algebras rr(5t) and TT (5t commute.
0 0
)

We now list the conditions that weshall impose on a real spectral triple
in order to constitute a noncommutative geometry.
Condition 1 (Classical dimension). There is a nonnegative integer n that
we shall call the classical dimension of the ensuing geometry, for which
n-1 E Ln+ (J{) but n- 1 ft. L~+ (J{). This n is even if and only if the spec-
tral triple is even. If the algebra 5t and the Hilbert space J{ are finite-
dimensional, the classical dimension of the geometry is zero.
482 10. Spectral Tripies

These conditions imply that the operator IDI-n lies in the Dixmier trace
dass L 1 + and that Trw IDI-n > 0 for any Dixmier trace Trw. In particular,
the spectral triple is n + -summable.
It is important to note that n is uniquely specified by Condition 1. Indeed,
if T is an operator in the Schatten dass LP(J{) for p 5 n, then ITin is
tracedass, and so Tr+ ITin = 0: thus LP !:;: ; L~+ for p 5 n. In particular,
if v- 1 E Lr+ (J{) for r (not necessarily an integer) less than n, we may
choose p := ~(r + n), so that v- 1 E LP, and a (ortiori v- 1 E L~+(Jf),
contrary to hypothesis.
Here we part company with the 0-summable spectral triples that are not
finitely summable. Indeed, the main point of the dimension condition is the
finite summability; the integrality is needed only for consistency with the
orientation and reality conditions below.
Condition 2 (Regularity). The spectral triple (5'., Jf, D) is regular; as we
have already seen, this means that all operators [D, a] are bounded, and
5\ u [D,.J'.]!:;::; Dom (8) where 8(T) := [IDI, T].
00

The regularity implies that 5\ c Op~, in view of Lemma 10.22. In partic-


ular, J{oo is a left 5'.-module. We shall demand a much stricter finiteness
condition on this module.
Condition 3 (Finiteness). The algebra 5\ is a pre-C*-algebra, and the space
of smooth vectors J{oo := nk Dom(Dk) is a finitely generated projective left
5'.-module.
The hypothesis on J{oo says that one can find m E ~. an idempotent
ein Mm(Y.) and a left 5\-module isomorphism from J{oo onto m.J\e. One
would like to replace this idempotent by a projector in the involutive alge-
bra Mm(Y.); such a p is indeed given, using Theorem 3.8, by Kaplansky's
formula for e* (since we are dealing with left modules), namely, p := e* er- 1
where r = e*e + (1- e)(1-e*) = 1 + (e* -e) (e- e*). The only issue here is
the invertibility of r in Mm (5'.), since Theorem 3.8 has been proved for C*-
modules by using functional calculus to invert r (or equivalently, by using
the block-matrix formulas (3.4) and (3.5) after showing complementability
of the module). Thus we can certainly say that r is invertible in Mm (A),
where Ais the C*-completion of 5\. This is where the condition that 5\
be a pre-C*-algebra comes in: for, by Proposition 3.39, Mm(A) is also a
pre-C*-algebra, and we condude that r- 1 E Mm (5'.), too.
That said, we can give a simpler presentation of the status of the alge-
bra 5\. Write J{oo = m.J\p, and consider the dual right 5'.-module p.J\m.
Then we can identify 'B := pMm(Y.) p = p.J\m ®Jl m.J\p with the endo-
morphism algebra EndJ\ J{oo, acting on J{oo on the right. This (algebraic)
Morita equivalence shows also that

(10.81)
10.5 Termsand conditions for spin geometries 483

(See the example at the end of Section 2.A). Now suppose that T E L(J{)
is an operator in the weak closure of the algebra rr(.JI.), such that T E
Domoo 8. Then the proof of Lemma 10.22 shows that T maps J{k to J.[k for
each k, andin particular T maps J{oo into J{oo. Moreover, by von Neumann's
bicommutant theorem [366, Thm. 2.2.2), the weak closure of rr(.JI.) is just
its bicommutant rr(.JI.)" =: .JI.". Recall that the commutant .JI.' = rr(.JI.)'
means the involutive algebra of all S E L (.1{) suchthat [S, a] = 0 whenever
a E .JI., and that .JI." := (.JI.' )'. The identification J{oo = m.JI.p matches
elements of 'B =End.:<\ J{oo to members of .JI.', and therefore TE .JI." is an
operator that commutes with the right action of 'B. From (10.81) we now
conclude that

.JI. = {TE .JI.": TE Dom"" 8 }. (10.82)

We can now conclude that .JI. is a Frechet pre-C*-algebra. Indeed, the


von Neumann algebra .JI." is in particular a C*-algebra, and (10.82) says
that .JI. is the algebra of its smooth elements under the action of the one-
parameter subgroup of automorphisms generated by the derivation 8; the
result follows from Proposition 3.45. From Theorem 3.44, we therefore
know that the K-theory of .JI. is the sameasthat of its C*-completion.
~ The previous three items may be called the analytic conditions for a spin
geometry. Before stating the algebraic conditions, we recall that x denotes
the grading operator for an even spectral triple (.JI., .1{, D), and that we
write x := 1 when dealing with an odd spectral triple; and that if the conju-
gation operator C satisfies suitable commutation relations with .JI., D and x,
then the whole package (.JI., .1{, D, C, X) forms a cycle in KR-homology of a
certain involutive algebra.
Since .1{ carries two commuting representations rr and 7T of the respec-
0

tive algebras .JI. and .JI. o, we may regard a ® bo - aCb* c- 1 as a represen-


tation of the (algebraic) tensor product .JI. ® .JI. o. The natural involution on
.JI. ® .JI. o is given by

(10.83)

The right hand side is represented on .1{ by the operator b*Cac- 1 =


Cac- 1 b* = C(ac- 1 b*C)C- 1 . lf C2 = ±1 also holds, so that c- 1 b*C =
Cb*c- 1 , then T is implemented on .1{ by C(. )c- 1 •
Condition 4 (Reality). The conjugation C satisfies C2 = ±1, CD = ±DC,
and Cx = ±XC, where the signs are so chosen that (.JI., .1{, D, C, X) is a
(reduced) KRi-cycle over the algebra with involution (.JI. ® .JI. ", T), where
j = nmod 8.

We call the operator Ca real structure on (.JI.,J{,D); the commutation


relations of C with .JI., D and x are set out in the table (9.45).
484 10. Spectral Tripies

Condition 5 (First order). The representation of .Jl o commutes not only


with that of .Jl but also with [D,.Jl]; in other words,

[[D,a],Cb*c- 1 ]=0 forall a,bE.Jl. (10.84)

This definition is symmetric in .Jl and .Jl o. Indeed, the Jacobi identity and
(10.79) show that

[[D,a], [Cb*c- 1 ] + [a, [D,Cb*c- 1 ]] = [D, [a,Cb*c- 1 ]] = o,


andthus
[a, [D, Cb*c- 1 ]] = o.
Note also that [D, Cb*c- 1 ] = ±C [D, b*] c- 1 = +C [D, b]* c- 1 according
as CD= ±DC.
Weshallsee in the next chapter that (10.84) expresses the defining prop-
erty of first order differential operators. However, its present formulation
imposes no obligation that the algebra .Jl be commutative; it is enough that
the Hilbert space be large enough to support commuting representations
of .Jl and .Jl o.
The property (10.84) has an important consequence, namely, it allows us
to construct a representation of Hochschild chains over .Jl with values in
.Jl ® .Jl o. The latter is an .Jl-bimodule in the obvious way:

a' (a ® bo )a" := a' aa" ® bo.

Definition 10.13. Given a spectral triple (.Jl, J{, D, C, X) satisfying the re-
ality and first-order conditions, any Hochschild k-chain in Ck(.Jl, .Jl ® .Jl o)
is represented on J{ by

rrv((a ® bo) ® a1 ® • • · ® ak) := aCb*C- 1 [D,ar] ... [D,ak]. (10.85)

We can now formulate the third algebraic condition. lt is an abstract


version of a volume form on a manifold.
Condition 6 (Orientation). There is a Hochschild cycle c E Zn (.Jl, .Jl ® .Jl 0 )
suchthat

rrv(c) =X· (10.86)

The degree n of this cycle is, of course, the classical dimension of Condi-
tion 1. We elucidate this condition in the commutative case in Chapter 11,
but we may already connect it with the HKRC theorem of Section 8.5. ln-
deed, that theorem teils us that the Hochschild cohomology classes of
C"" (M) are given by de Rham currents on M; the dual Statement is that
Hochschild homology classes of C"" (M) come from differential forms on M,
graded by degree. An n-cycle corresponds to a form of top degree; the claim
is that the nondegeneracy of the volume form is reflected in the algebraic
10.5 Termsand conditions for spin geometries 485

condition rrv(c) = x. We repeat, for emphasis, that in odd-dimensional


cases this means rrv(c) = 1.
~ The last condition, of a topological nature, is the K-theoretic version of
Poincare duality [94]. As one may anticipate from the case of (compact) ma-
nifolds, this involves a homomorphism from the K-theory group Kn-rCJ\)
to the K-homology group Kr(.J\. := K#r(.J\. introduced in Section 9.5.
0
)
0
),

This may be defined directly as the "cap product" of elements of Kn-r (.Jl)
by a fixed dass in Kn (.J\. ® .J\. o), which is none other than the dass of the
unbounded K -cyde (.Jl ® .J\. o, J{, D). In the even case, one can by dual-
ity consider the homomorphism from Ko (.J\.) x Ko (.J\. o) to Ko (.J\. ® .J\. o)
given by ([p], [qo]) ,__ [p ® qo], and contract it with the index map de-
fined by D. Thus, if p E Mk (.Jl) and q E Mz (.Jl) are projectors, then
P := p ® (C ® 1L)q(c- 1 ® 11) is a projector acting on J{ ® ckl, and the
compression P(D ® 1k!)P is an odd Fredholm operator on P(Jf ® Ckl),
whose index is defined as in (9.36). Replacing [q by [q], we end up with
0
]

an additive pairing on Ko(.Jl), given by


([p], [q])- index(P(D ® 1k!)P) E 71.. (10.87a)
In the odd case, given unitaries u E Uk{.Jl), v E Uz (.Jl) defining dasses in
KI(.Jl), the operator U := u ® (C ® 1L)v(c- 1 ® 1d is unitary on J{ ® ck 1;
if Q := ~(1 + DIDI- 1 ) ® hz. QUQ is a Fredholm operator on Q(Jf ® ckz),
(o ut)
U 0 is the symmetry of an even Fredholm module over .J\. ® .J\. o, and
compression with Q E!l Q yields an additive pairing on K1 (.Jl) by

.
([u], [v]) ,__ mdex ( QUQ
0 QUtQ) , (10.87b)
0

that is, the Fredholm index of QUQ.


The duality condition can now be stated as follows.
Condition 7 (Poincare duality). The additive pairing (10.87) on K. (.Jl), de-
termined by the index map of D, is nondegenerate.
This condition is satisfied by ordinary (compact) manifolds, since it is a
restatement, via the Chern isomorphism, of the ordinary Poincare duality
between de Rham homology and cohomology: see Section 11.1.
To finish, we summarize these terms and conditions under the name of
geometry.
Definition 10.14. A noncommutative spin geometry isareal spectral tri-
ple ~ := (.Jl, Jf, D, C, xl fulfilling Conditions 1-7 of the above catalogue.
The namewill be justified in the next chapter, wherein spin geometries
over the algebra C"" (M) are dassified.
There is a natural notion of unitary equivalence of geometries over a
given algebra .J\.. A geometry is unitarily equivalent to ~ if it is of the form
486 10. Spectral Tripies

(5\, J{', UDU- 1, ucu- 1, Uxu- 1) for some unitary isomorphism U: J{ -


Jf'; the corresponding action of 5\ on J{' is a- Urr(a)U- 1 •
The 0-dimensional spin geometries, in which 5\ is a finite-dimensional
semisimple matrix algebra, have been studied in detail by Paschke and
Sitarz [365) and by Krajewski [292, 293), to which we refer. The operator D
is then a Hermitian matrix satisfying the geometrical conditions.
It is not hard to describe the product of two given spin geometries. If §" 1
and §"2 are spin geometries of respective classical dimensions n 1 and n 2 ,
the data of the product geometry §" = YI x §"z come from the product of
the underlying KRj-cycles. In other words, 5\ := J\1 181 5\z, J{ := Jf1 181 Jfz,
and the other terms may be defined as follows. If n 1 is even, we may take

Since D 1 and x 1 anticommute, it follows that D 2 := Di 181 1 + 1 181 D~. The


Hölder inequality for Dixmier traces then shows that the classical dimen-
sion of §" is n1 + nz. lf nz is even, we may take instead D := D1 181 X2 + 1 181Dz;
when both geometries are even, these two versions of the product geometry
are equivalent [207). We also take x :=XI 181 xz. unless both geometries are
odd; in that case, the product geometry is even, and §" is defined by dou-
bling the Hilbert space Jf1 181 Jf2 [91, IV.A) and setting

D ·= ( 0 D1 181 1 - i 181 D 2)
. D1 181 1 + i 181 Dz 0 .

The conjugation operator on §" may be determined systematically by am-


plifying each Yi to an unreduced KRn;_cycle, forming the product of these
tagether with a graded tensor product of their supercommuting Clifford
actions (which yields a supercommuting action of Cln 1 +n 2 ,o on the product
cycle), and reducing according to the procedures ofTheorem 9.19. We leave
this task to the reader, pointing out that the resulting C may be simply ex-
pressed in terms of C1 and Cz. If n1 or nz is even, then in most cases
C = C1 ® C2 [94); but, as noted by Vanhecke [464), this rule has some ex-
=
ceptions. One should take C := C1 181 Czxz if n1 6 and nz 2 mod 8; or =
C := C1X1 ® Cz if n1 is even and n1 + nz =
1 or 5 mod 8.
11
Connes' Spin Manifold Theorem

Let M be a compact oriented n-dimensional manifold without boundary.


The theorem asserts that an n + -summable real geometry over the algebra
C"" (M) determines a unique spin structure on M; and that, among all ab-
stract spin geometries in the sense of Section 10.5, compatible with that
structure, the one determined by the Dirac operator is singled out by a
variational principle.

11.1 Commutative spin geometries revisited


Tobegin with, we show that Dirac operators do indeed determine abstract
spin geometries. Suppose that a spin structure is already given on a com-
pact boundaryless manifold. We now recall the ingredients for a spin geo-
metry, already assembled in Chapter 9.
Definition 11.1. Let (v, S, C) be a spin structure on a compact Riemannian
manifold M without boundary. The Dirac geometry determined by this
spin structure is the real spectral triple (j = (..A,J-f,QJ,C,x), where
(1) 5l. = C""(M),

(2) J-{ = L2 (M,S) is the spinor space obtained by completing the spinor
module S =: f"' (M, S) with respect to the scalar product (9.22),
(3) lj) = -i(c o \7 5 ) is the Dirac operator given by (9.19),
(4) Cis the conjugation operator for the spin structure, and
488 11. Connes' Spin Manifold Theorem

(5) X = c(y), where y is the chirality element of roo (Cl(M) ), is either the
identity operator or the standard grading operator on :J{, according
as dim M is odd or even.

When the algebra .Jl is commutative, the conditions for a spin geometry
admit a few simplifications. First of all, the opposite algebra .Jl o equals .Jl
itself, and the mapping TT of (10.80) can be regarded as a representation
0

of the original algebra .Jl on :}{, Although, in general, one may use two
different representations of this algebra, in the case of the Dirac geometry,
TT coincides with the given action rr of .Jl on :}{, Indeed, Cb*c- 1 = 8_ 1 (b)
0

from (9.8b), when b E roo (Cl(M) )-in this chapter we reserve the letter x
for the grading operator and use the notation ß_ 1 of Section 5.1 for the
Bogoliubov automorphism-and B-1 (b) = b for b E C (M), so that the
00

commutation relation (1 0. 79) just expresses the commutativity of the alge-


bra .Jl = C (M). Therefore, TT (b) = rr(b) denotes the action of b E .Jl as
00 0

a multiplication operator on the spinor space.


A second simplification is that when TT = rr, we may regard the ori-
0

entation cycle as an element c E Zn (.Jl). For in that case, the represen-


tation rrv: Ck{.Jl, .Jl ® .Jl) - .[(:}{) of (10.85) factors through the map
Ck(.JI.,.JI. ® .Jl) - Ck(.Jl) induced by the multiplication .Jl ® .Jl - .Jl; we
denote the other factor also by rrv, from Ck (.Jl) to .[(:}{).In other words:
since aCb*C- 1 = ab for a, b E .Jl, nothing is lost by replacing a ® bo on
the left hand side of (10.85) by the product ab in .Jl. (This factorization
cannot work unless .Jl is commutative, since then a ® bo ..... ab is no langer
a homomorphism.) To summarize, when .Jl is commutative and TT = rr, 0

the module Ck{.Jl) is represented on :J{ by

(11.1a)

and the orientation cycle is an element c E Zn (.Jl).


Notice that the kernel of rrv contains the subcomplex Dk(.Jl) generated
by chains of the form a = ao ® · · · ® 1 ® · · · ® ak with ai = 1 for some
i = 1, ... , k, already introduced in Section 8.4. We may then pass to the
quotient Ok.Jl = Ck.JI.IDk.JI., and regard rrv as an .Jl-module map from
nk.Jt to .[(:}{), given by

(11.1b)

Theorem 11.1. The Dirac geometry is a noncommutative spin geometry.

Proof. We must show that (j camplies with the seven conditions laid out in
Section 10.5, subject to the aforementioned modifications.
(1) The classical dimension of (j equals the dimension n = dimM of the
manifold M. This happens because [/) is an elliptic differential operator,
so that II/JI and its powers are pseudodifferential operators. As already
11.1 Commutative spin geometries revisited 489

observed in Section 9.4, IJ,lJI-n is a bounded measurable operator in the


Dixmier trace dass, and by (9.39),

f I.l,lJ 1-n .-
·- n(2rr)n T +I 1-n-
2ln/2JOn r .l,lJ - 1. (11.2)

(Once more, we assume in the notation that .l,lJ is invertible, by dropping


its finite-dimensional kernel from J{ if necessary.) Now IJ,lJ 1- 1 is an elliptic
pseudodifferential operator of order -1, belonging to Ln+ but not to L8+.
(2) Since, by Proposition 9.11,

[.l,lJ,a] = -ic(da), (11.3)

it is immediate that II [J,lJ, a] II = llc(da) II = II gradall ""'so [J,lJ, a] is certainly


bounded for each a E 5t.
To establish regularity of the spectral triple (5t, Jf, .l,lJ), it is enough,
in view of Lemma 10.23, to show that both a and [J,lJ, a] lie in the com-
mon smooth domain of the commuting operatorsLand R of (10.66). Since
the principal symbol of .1,7J 2 is a scalar matrix, it follows that [J,7J 2 , a] is a
pseudodifferential operator of order one at most; repeated commutation
with .1,7J 2 shows that a<k+l) is of order k + l at most, and thus LkR 1(a) =
IJ,lJI-k a<k+O l.l,lJI- 1 is a bounded pseudodifferential operator, for all a E
C""(M). On replacing a by -ic(da), every LkR 1([.1,7J,a]) is bounded, too.
(3) We know that C""(M) is a pre-C*-algebra, as remarked at the end of
Section 3.8. Also, the smooth domain of .l,lJ = -ic(dxi)V'~ is the mod-
ule of smooth spinors: J{"" = f"" (M, S) = S, which is a finitely ' generated
projective module over C"" (M).
(4) The reality condition for the Dirac geometry is precisely Theorem 9.20,
which shows that this geometry forms a reduced KRi-cycle over C"" (M),
with j := dimM mod 8. (Notice that the required involution on C"" (M) is
just complex conjugation, since cac- 1 = ä for functions.)
(5) The first order condition holds for the Dirac geometry, since

[[D,a], Cb*c- 1 ] = -i [c(da), b] = 0


for functions a, b E C"" (M), acting by multiplication Operators on the
spinor space.
(6) The required Hochschild n-cycle c is given by the volume form v9
defining the given orientation on M. (This is not surprising, since there is a
natural pairing of Hochschild n-cycles with Hochschild n-cocycles, and the
latterare given, in view of Theorem 8.17, by de Rham currents of degree n.)
To compute TTJ)J(C), we need a local expression for c. First of all, choose
a (finite) open covering of M by local charts {(U1, c1)} and take a smooth
partition of unity {f1} subordinate to it. Each c1 : u1 - ~n has components
cJ, ... ,cj that maybe regarded as elements of C""(M) supported in u1. If
490 11. Connes' Spin Manifold Theorem

{8}, ... , Bj} is a local orthonormal basis of 1-forms, where Bj = aj5 dcj,
the Riemannian volume form may be written over Uj as

Vg = e} 1\ ••. 1\ Bj = hjdc) 1\ ••• 1\ dcj,

where hj = det[aj5 ]. Now put cJ := in-m fihj, which is another smooth


function supported in Uj (with m = Ln/ZJ, as usual); this notation allows
us to write in-mv9 globally as a finite sum

(11.4a)

The corresponding Hochschild n-chain is defined as

(11.4b)

Alternatively, c = 2.i An (cJ ® c} 1\ • • • 1\ cj ), which lies in the image of the


skewsymmetrization map An: 5\ ® An 5\ - Cn (.Jl) introduced in Proposi-
tion 8.10; by the proof ofthat proposition, c is a Hochschild n-cycle.
Using (11.1), the commutation relation (11.3), and the fact that the Clif-
ford product of an orthonormal basis of vectors is unchanged by an even
permutation of that basis, we then find that

7TJP(C) = ~ 2.: (-l)rr2.:cJ[JP,cf 0 )] ... [1)),cf(n)]


n. TTESn j

= (-i~n 2.: (-l)rr2.:cJc(dcf 0 )) ... c(dcf(n))


n. TTESn j

= (-i~m 2.:Ji 2.: (-l)rrc(Bf(l)) ... c(By(n))


n. j TTESn

= (- i) m 2.: fi c ( B}) ... c (Bj) = c (Y) 2.: fi = X.


j j

(7) As mentioned already in Section 10.5, the Poincare duality condition


given there is a version of the usual Poincare duality (see [134], for instance)
between the de Rham homology H~R(M) and the de Rham cohomology
Hd_R (M) of the compact oriented manifold M. The latter duality can be re-
stated as the existence of a nondegenerate pairing on Hd_R (M), given by

where, say, oc E 5\ k (M) and 11 E 5\ n-k (M) are closed forms; the integral
depends only on the cohomology classes [oc] and [1]], since it vanishes if
either oc or 11 is exact. The nondegeneracy is due to the orientation dass
11.2 The construction of the volume form 491

[v9 ]. Indeed, the formula (9.61) can be rewritten as EkÖ< 1\ *ß = (oc I ß)v
with Ek := im(-l)nk-k(k+ll/ 2, and then

for any form 17 representing a nontrivial dass in de Rham cohomology. 0

Poincare duality can be developed in a purely KK-theoretic setting; this


is outlined in [91, Vl.4.ß], and a version of it was sketched in Section 10.5.
On specializing to the case of manifolds, Kasparov's theory allows one to
establish an isomorphism between the K-homology K"(C(M)) and the K-
theory K.(f(Cl(+l(M))) for any compact manifold M. The spinc structure
(v, S), which establishes a Morita equivalence between the two C* -algebras,
permits us to replace the latter module by K.(C(M)). One can then transfer
the duality to the de Rham setting via the Chern isomorphism theorem.
(In fact, since taking Chern characters eliminates torsion, this K-theoretic
formulation of Poincare duality is somewhat sharper than the differential
geometric version.)
.,. If pisanontrivial projector in the centre of .J\ and if D, C and x commute
with rr ( p), then (j can be written as the direct sum of two geometries over
the algebras p.J\ and (1- p).J\. This happens, for example, in the case of
the Dirac geometry over a compact spin manifold that is not connected, by
taking p to be the indicator function of one component.
Definition 11.2. We shall call a spin geometry irreducible if no nontrivial
projector on J{ commutes with rr(.J\), D, C and x.
Suppose that (j1 = (.Jl.,J{I,DI,CI,Xd and (j2 = (.J\,Jf2,D2,C2,X2) are
two spin geometries over the same algebra .J\, having the same classical
dimension n. Their Whitneysum (j := (5\, Jf1 ~J{2,D1 ~D2, C1 ~C2, XI ~X2)
then makes sense; the representations of .J\ and .J\ o on 3{1 ~ 3{2 are direct
sums of their given representations on 3{1 and 3{2. The classical dimension
of (j is also n, since n- 1 = D1 1 ~ D2 1 E .[n+ (J{I ~ Jf2 ).
In the case of a Whitney sum, the projector P1 with range 3{1 ~ 0 reduces
the geometry. Any Dirac geometry on a spin manifold is not a Whitney sum
of two subgeometries, because of the irreducibility of the spin representa-
tions of Cl (~ 2 m) and Cl+ (~ 2 m+ 1) on the fibres of the spinor bundles. If M
is connected, then the Dirac geometry is irreducible.

11.2 The construction of the volume form


We come, finally, to the classification of abstract noncommutative spin geo-
metries on manifolds. We suppose that we are given a connected compact
manifold, and a geometry which is not a Whitney sum of subgeometries.
492 11. Connes' Spin Manifold Theorem

Theorem 11.2. Let G = (.Jt, Jf, D, C, X) be an irreducible noncommutative


spin geometry, of classical dimension n = dimM, over the algebra .Jt =
coo (M) of smooth functions on a compact orientable connected manifold M
without boundary. Then:

(a) There is a unique Riemannian metric g = g(D) on M, whose distance


function is given by

d 9 (x,y) = sup{ la(x)- a(y)l: a E C(M), II[D,a]ll :51}.

(b) M is a spin manifold, and the possible operators D' for which g (D') =
g (D) form a finite union of affine spaces Iabelied by the spin structures
onM.

(c) The action functional S(D) := f IDI-n+Z yields a quadratic form on


each of these affine spaces, attaining an absolute minimum at D = I/),
the Dirac operator for the corresponding spin structure; this minimum
is proportional to the Einstein-Hilbert action, namely, the integral of
the scalar curvature:

Before entering on the proof, the expression f ID 1-n+Z requires a ward


of explanation. Up to now, the notation f has been used for a certain multi-
ple (7.83) of (any) Dixmier trace of a measurable operator. Yet the operator
IDI-n+Z does not lie in the Dixmier trace dass; indeed, IDI-n+Z E LP+ for
p = n I (n - 2) > 1. However, in the course of the proof, we shall show
that it is actually a pseudodifferential operator. Following Connes [96], we
define the action functional using Wodzicki residues, as follows:

S(D) =f IDI-n+z .·= 2Ln/2JOn


1 Wres IDI-n+z . (11.5)

Our normalization, chosen for compatibility with (11.2), differs from that
of [96], where f is implicitly taken tobe synonymaus with n- 1 (2rr)-n times
the Wodzicki residue.
Our approach to the proof consists of three steps:
(1) establish the nondegeneracy of the volume form on M,

(2) build the spinor bundle over M, tagether with the corresponding Clif-
ford action and Riemannian metric, and

(3) compute S (D) by pseudodifferential calculus, showing that D = I/)+ p


with S (I/)+ p) being a positive definite quadratic functional of a certain
remainder term p.
11.2 The construction of the volume form 493

Let Trw be any Dixmier trace; then the geometric data for (j provide the
Hochschild n-cochain (10.68):

The Operator xao [D,ai] ... [D,an] IDI-n lies in the Dixmier trace dass
because x. ao and each [D, ak] are bounded, and IDI-n E L 1+. Recall, from
Lemma 10.24, that cpYJ is a Hochschild n-cocycle.
We would like now to use these cocycles to define the noncommutative
integral of a function in .Jl, along the lines of the formulae (7.83) and (9.39).
Since we do not have the Wodzicki residue at our disposal-at this stage, we
do not yet know whether D is a pseudodifferential operator-we must use
the Dixmier traces directly to integrate functions. The first issue that must
be settled is measurability, that is, the result of the integration must be in-
dependent of which of the several Dixmier traces we use to compute it. It is
here that Connes' character formula, Theorem 10.32, decisively intervenes.
Proposition 11.3. If (j is a spin geometry of dimension n over .Jl = C"" (M)
and if a E .Jl, then a ID 1-n is a measurable operator.

Proof. Let c E Zn (.Jl) be the orientation cycle for (j. Then ac is also a
Hochschild n-cycle. Indeed, if c = 2:1 cJ ® cJ ® • · · ® cj, then

b(ac) = L.J
"'·(ac)0c )1 ® c )2 ® · · · ® c»-
) ·· · + (-l)»ac)0 ® c)1 ® • · · ® c»-
)
1c':l
)
+ (-l)n+1c»aco
) )
® c1 ® ..• ® c»-1)
) )
= a(bc) + ( -l)n "'
L.1
·(ac»-
)
c':la)c
)
0 ® cl
J J
® · • · ® c»-
J
1 = 0'
since .Jl is commutative.
Consider now the Fredholm module (.Jl, Jf, F), where F := DIDI- 1 is the
phase of the Dirac operator, and let

denote its Chern character. Then Theorem 10.32 assures us that TJf and
cpYJ take the same values on any Hochschild n-cycle, andin particular,
:\.~ 1 TJ!(ac) = :\.~ 1 cp~(ac) = 2:1 Trw(xacJ [D,cJ] ... [D,cj] IDI-n)
= Trw(xarrv(c) IDI-n) = Trw(xax IDI-n)
= Trw(a ID!-n).

Therefore a IDI-n is measurable, and the noncommutative integral

is well defined and independent of w. 0


494 11. Connes' Spin Manifold Theorem

.,. Our immediate goal is to tease out of this noncommutative integral on


coo (M) a volume form v on M. The difficulty is to show that the inte-
gral (11.6) is nondegenerate, that is, f a IDI-n > 0 whenever a is a nonzero
positive element of coo (M). This condition is clearly necessary is order to
be able to rewrite this integral as fM a v for a suitable volume form v, yet
to be constructed.
Any local chart on M is given by n coordinate functions c 1, ... , cn E
C (M), supported on an open set U c M. These coordinate functions de-
00

termine a diffeomorphism from U onto a bounded open subset V c ~n.


so that any a E coo (M) with support in U can be written in the form
a = j(c 1 , ..• ,cn) for some smooth function f on V. The usual volume
form on V, namely, the restriction of the Lebesgue measure on ~n to V,
may then be pulled back via the chart c = (c 1 , .•• , cn) to give an integral
on functions supported in U:

Using a partition of unity subordinate to an atlas of local charts, we can


add several such local integrals to define an integral over all of M. Now,
although this elementary construction of an n-density is highly coordinate-
dependent, it does serve to define uniquely the Lebesgue measure class
on M, because two integrals built in this mannerare absolutely continuous
with respect to each other [133, 16.22.2]. Furthermore, if the charts are
compatibly oriented, this n-density is in fact a volume form.
From the operatorial point of view, the local Coordinates c 1 , ... , cn on U
are represented on J{ by n commuting (bounded) selfadjoint operators,
whose joint spectrum is a subset of V. Before proceeding further, it is
worth recalling what the spectral theorem tells us about such a system
of commuting selfadjoint operators, say S1 , ... , Sn. on a separable Hilbert
space J{. Theseoperators generate a unital C*-algebra A, whose character
space M(A) is compact. Evaluation ofpolynomials p ..... p(SI, ... , Sn) yields
a surjective morphism from C(0/:= 1 sp(Sk)) onto A, which corresponds,
under the Gelfand cofunctor, to an injective continuous map j: M(A) -
0/:= 1 sp(Sk) c ~n; we call the compact set V := j(M(A)) the joint spec-
trum of S1, ... , Sn.
If J{ contains a cyclic vector ~ for the algebra A, the positive linear
functional f ..... (~I j(SI. ... , Sn)~) =: fv f dJ1~ defines a positive measure
on V, or equivalently, a positive measure on ~n with support in V. More-
over, g(S1 , •.. ,Sn)~ ..... g defines a unitary isomorphism between J{ and
L 2(V, dJ1~ ); under this isomorphism, each Sk corresponds to the fCi\) .....
i\kf(ll.) on V. In general, just as in [383, Thm.VII.3], J{ is a direct sum of
cyclic subspaces, yielding measures J1 1 ,J1 2 , •• • , on ~n with support in V,
such that Sk acts as multiplication by Ak on the direct sum of the spaces
L2(V, J1J)· We may lump tagether these measures to get J1 := J1I E9 J12 E9 • • •
11.2 The construction of the volume form 495

with support in V, called the spectral measure of the cornmuting family


SI, ... ,Sn.
This spectral measure can be decomposed into three parts: an atomic
or "pure point" part and two parts that are respectively singular and ab-
solutely continuous (AC) with respect to Lebesgue measure on ~ n, and the
Hilbert space J-{ splits correspondingly into a direct sum of subspaces:
see [284, X.1] or [383, VII.2] for the case of a single operator. We are mainly
concerned here with the AC part, since the operators arising from coordi-
nate functions an a manifold will only have absolutely continuous spec-
trum. The AC part of the spectral measure is of the form m(x) dnx for a
certain multiplicity (unction m: V - [0, oo)-the Radon-NikodYm. deriva-
tive-so that

(11.7)

gives the AC part of the spectral measure of (SI, ... , Sn).


Coming back now to the noncornmutative spin geometry (j, consider the
operators c], ... , cj E .Jl, obtained from one surnmand of the orientation
cyde c; if V = v1 c ~n denotes the AC part of their joint spectrum, we
can form the integral (11.7), which cannot vanish if f is a positive-valued
continuous function. Taget the desired nondegeneracy, weshall argue that,
in this case, the noncornmutative integral

f ,_ f f<c], ... ,cj) IDI-n (11.8)

cannot vanish either .


.,. Already for a single selfadjoint operator S, the absolutely continuous
part of the spectrum is a delicate flower that can wither until small per-
turbations of S. We refer to [284, Chap. X] for a full discussion of these
matters; here we recall briefly what we need. The basic result is the Weyl-
von Neumanntheorem [360], which says that any seifadjoint operator S on
a separable Hilbert space can be perturbed by a seifadjoint Hilbert-Schmidt
operator A, with IIAII2 assmallas we please, in such a way that S + A has
pure point spectrum. In fact, an improvement by Kuroda [284, Thm. X.2.3]
shows that one can find such an A making sp(S + A) discrete, even if the
Hilbert-Schmidt dass is replaced by any syrnmetric operator ideal other
than the trace dass (for instance, LI+ would da).
On the other hand, the Kato-Rosenblum theorem [284, Thm. X.4.4] states
that if S is selfadjoint and if A is selfadjoint and traceclass, then the abso-
lutely continuous spectra of S and of S + A coincide.
The comparison of integrals like (11.7) and (11.8) depends crucially an a
result that widens the scope of the Kato-Rosenblum theorem, in the case
where S is replaced by n > 1 cornmuting Seifadjoint operators (SI •... , Sn).
Suppose that these operators are perturbed to another set of cornmuting
496 11. Connes' Spin Manifold Theorem

selfadjoint operators (SI + A1 , ... , Sn +An); we need to know the precise


conditions on (AI, ... , An) in orderthat the AC part of the spectral measure
of (S 1 , ••• , Sn) be unchanged.
The answer to that questionwas provided by Voiculescu in a splendid pa-
per [474, Thm. 4.9]. He found that, for n > 1, the AC spectrum is preserved
for certain perturbations that need not be traceclass. The exact result is
that (A 1 , ... , An) should be commuting selfadjoint compact operators in
the symmetrically normed operator ideal .cn-, introduced in Section 7.C,
which includes the Schatten ideal LP whenever 1 ::5 p < n. Rather than
prove Voiculescu's theorem here, we refer the interested reader to the orig-
inal paper; the proof is based on the computation of the following useful
quantity, which measures the strength of the perturbation.
Definition 11.3. Let 1 be a symmetrically normed ideal of compact Ope-
rators on a Hilbert space Jf. Call Rt the partially ordered set of positive
finite-rank operators on J{ of norm at most 1. Let (SI, ... , Sn) be an n-tuple
of commuting selfadjoint operators on Jf. Then Voiculescu's modulus for
(S1, ... , Sn) with respect to 1 is defined as
(11.9)

In the case 1 = .cn-, we write k;; instead of k1 .


To see the point of this definition, consider first the easy case in which
all Si have pure point spectra. Since they commute, they can be simulta-
neously diagonalized. Then, by taking A to be a projector on some finite-
dimensional subspace generated by the largest eigenvalues of the Si, we
see that II [A, Si 1111 will depend only on the trailing eigenvalues and can be
made as small as desired, no matter which symmetric norm is used. Thus,
as A t 1 through a sequence of such projectors, the limit inferior vanishes
and thus kJ(SI, ... ,Sn) = 0. Therefore, the nonvanishing of kJ(SI, ... ,Sn)
signals that the spectral measure of (S1 , ... , Sn) has a nonzero continuous
part.
It turnsout [474, Prop. 4.1] that k;;(S 1 , ..• ,Sn) also vanishes for opera-
tors that have purely singular continuous spectral measure; thus, if the k;;
modulus is nonzero, the absolutely continuous part of the spectral measure
is nontriviaL
On the other hand, if we use the uniform norm by taking 1 = X, we
can choose an increasingly ordered family A(X E Rt that form a quasicen-
tral approximate unit for L(Jf), that is 1 an approximate unit as in Defi-
nition 1.19 that satisfies lim(X IIA(XT- T A(XII = 0 for any T E .L(Jf). lt is
known [9, Thm. 1] that any approximate unit for an ideal of a C*-algebra
can be modified to yield another that is quasicentral. Taking the limit in-
ferior in (11.9) through such a family shows that kx(S 1 , ... ,Sn) = 0 in all
cases. This indicates that k1 (S1 Sn) can only be nontrivial if the ideal
I ••• I

1 issmall enough. In factl if n > 11 it happens that k1(S 1 , ... Sn) = 0 for
I
11.2 The construction of the volume form 497

1 =Ln [474, Thm. 4.21, so that 1 =Ln- is the borderline case for nonvan-
ishing of the modulus (11.9).
The key result of Voiculescu is the following relation [474, Thm. 4.51.
Suppose that the absolutely continuous part of the spectral measure of
(Slo····Sn) is supported an V!;;; ~n. with multiplicity function m. Then
there is a positive constant Cn, independent of (SI, ... , Sn), for which

fv m(x) dnx = Cn(k;;(SI, ... ,Sn))n. (11.10)

In particular, notice that when J-{ = L 2 (M, S) and c 1 , ... , c n E C"" (M) are
local coordinates an a chart forM, then k;; (c 1 , •.. , cn) > 0.
Ta link Voiculescu's modulus for Ln- to Dixmier traces, we must relate
it to continuous functionals an the larger operator ideal Ln+. (The dual
space of Ln- is Lq+, where q = n/(n- 1), rather than Ln+, according to
Section 7.C, so this is not simply a matter of duality.) The link is achieved by
the commutator estimate of Lemma 10.19 and the inequality (10.61) which
follows from that estimate.
Proposition 11.4. Let (SI, ... , Sn) be an n-tuple of commuting bounded self-
adjoint operators on J-{ and Iet D be a selfadjoint operator on J-{ such that
IDI- 1 E Ln+ (J-f) and each [D, Si 1 is bounded. Then there is a constant c~.
independent of(S1, ... , Sn) and D, suchthat the following estimate holds for
any Dixmier trace Trw:
k;;(SI, ... ,Sn):::; C~ max II[D,Sdll (Trw IDI-n) 11 n. (11.11)
I:si:sn
Proof. We apply the estimate (10.61) with p = n, a = Si and a suitably
chosen function g E D(~). By construction, g(tD) has finite rank for any
t > 0; to ensure that it lies in nt, it is enough to require that 0 :::; g :::; 1.
If we demand that g(O) = 1 too, then g(tD) = (2rr)-I/2 JIRg(u)eiutD du
converges weakly to (2rr)- 1 ' 2 f~W.g(u) du = 1J{ as t I 0. Finally, we may
ask that g be an even function, smoothly decreasing from 1 to 0 an the
interval [0, R 1; in fine, g should be a cutoff like those of Section 7.B. Then,
as t I 0, the positive operators g(tD) increase monotonically to 1J{.
These properties show that, to majorize the Voiculescu modulus (11.9),
we may replace the limitinferior over all A E ~i by using only A = g(tD)
with t I 0. Thus,
k;;(SI, ... ,Sn):::; liminf max ll[g(tD),Sdlln-.
t tO 1:si:sn

and the Proposition follows from (10.61) by taking C~ := C~(g), since g


was constructed without regard to D. D

~ We now return to the system of n commuting selfadjoint operators Si,


whose joint spectral measure has an absolutely continuous part J..lac sup-
ported an V, given by (11.7). The relation (11.10) ofthis measurewith Voicu-
lescu's modulus allows us to replace the estimate (11.11) \N:ith the following
498 11. Connes' Spin Manifold Theorem

inequality:

fV
m(x)dnx :5 Cn(C~)n max II[D,SdllnTrw IDI-no
1:s;i:s;n
(11.12)

Now let Aw(f) := Trw (f(S1, 000, Sn) IDI-n), for f E 1J(V), be the measure
determined by an operator D satisfying the hypotheses of the previous
propositiono If f ~ 0 andif /lacU) = fv j(x)m(x) dnx > 0, then?l.w(j) > 0
also, so that llac is absolutely continuous with respect to Awo
Remarko In consequence, if S1, 000, Sn have only absolutely continuous joint
spectrum and if a = j(S1,ooo,Sn) ~ 0 in C""(M), then TrwaiDI-n = 0
implies that llacU) = 0 and hence that f = 0 and then a = Oo We may, for
example, take the Si tobe the operators of multiplication by the coordinates
on some local chart of M, and a tobe a positive function supported on
the chart domain; then Trw a IDI-n > 00 By writing a generalnonnegative
a E C"" (M) as a sum a = LJ !Ja1 with a suitable partition of unity {f1}, we
condude that a .... Trw a IDI-n isapositive measure on M vanishing only
when a = 0, and any measure in the Lebesgue dass is absolutely continuous
with respect to this oneo

.,.. At this stage, we jump ahead a little to mention an important property


of the geometry that is a consequence of the first-order and finiteness con-
ditions (we shall deal with those later), namely that the Hochschild cocyde
cp'ß is cyclico From Connes' character formula, we already that know that
the dass [cp'ß h in cydic cohomology is the Chern character, and therefore
that cpYJ differs from a cydic n-cocyde by a Hochschild n-coboundaryo (The
lower-degree terms in the Chern character, which are matched by Theo-
rem 1005 with de Rham dasses of degree at most n - 2, contribute no
Hochschild n-cochains hereo) However, we daim a little more, namely that
Bcp'ß = 0, where Bis the Connes boundary map (l0o2)o Recall that B = NB 0 ;
the cochain B0 cp'ß is given by
Bocp'/J(ao,ooo,an-d := cp'/J(1,ao,ooo,an-d- (-l)ncp'/J(ao,ooo,an-1. 1)
= An Trw (X [D, ao] o00[D, an-IliDI-n)o (11.13)
Now this is already a cydic (n- 1)-cochain, since ?I.Bocp'ß = Bocp'/Jo In-
deed, the term [D, an- 1 ] may be brought to the front and reinserted before
[D,ao] since, by Theorem 10020, Trw(oiDI-n) is a hypertrace; moreover,
[D, an-1] X = (-1)n- 1x [D, an-1] since [D, an-1] is an odd operator when
n is eveno lt follows that Bcp'ß = nBocp'IJ, so our daim is that B0 cp'IJ = 0,
namely that the right hand side of (11.13) vanishes, for all a 0 , 000, an- 1 E J\o
Finally, B0 cp'ß = 0 if and only if the cocyde cpYJ is cydic, since, from (1Oo3c),

( 1 - ?l.)cp'/J = b' Bocp'ß + Bobcp'IJ = Oo


For the Dirac geometry, this condition amounts to Stokes' theorem; it
should already be dear that the noncommutative integral (llo13) will re-
11.2 The construction of the volume form 499

duce to an ordinaryintegral of closed n-forms like dao/\ · · · 1\dan-1 over M,


which will vanish since M has no boundary.
The HKRC Theorem 8.17 allows us to make the link to Stokes' theorem
without going into the particulars of the Dirac geometry. Indeed, even if q:/[j
is only cyclic modulo a Hochschild coboundary, it is still true that Bcp~ is a
coboundary, since B(bf./J) = -b(Bf./J) for f./J E cn- 1 (5\), and so IB[cp~] = 0
in HHn(Jl). If C~p is the de Rham n-current determined by [cp~]. Proposi-
tion 8.18 shows that IB[cp~] - n oC~p = 0, so that C~p is a closed n-current.
The space z~R (M) of closed currents is one-dimensional, the closed cur-
rents being just multiples of the fundamental dass [M]. (Here we use the
assumption that M is orientable.) In other words: since M is a connected
orientable manifold without boundary, the only available n-dimensional cy-
cle for integration is the whole manifold, although we may give the integral
an arbitrary normalization.
Let us write C~p =: tM, where t is the appropriate normalization constant.
This current is canonically obtained from cp~ by skewsymmetrization, ac-
cording to (8.55):

Ancp~(ao,alo····an) := 1., L. (-l)rrcp~(ao,arr(l), ... ,arr(n))


n. TTESn

= .\~ L. (-l)rrTrw(xao [D,arr(l)] ... [D,arr(nd IDI-n)


n. TTESn

(11.14)

We now evaluate this identity on the Hochschild n-cycle ac that is provided


by Proposition 11.3, for any a E 5\. The proof of Theorem 8.17 shows that
the Hochschild cocycles cp~ and Ancp~ are cohomologous, and so give the
same value when paired with the cycle ac. Therefore,

Tr+ a IDI-n = Trw a IDI-n = .\;:;_ 1 cp~(ac) = .\;:;_ 1 Ancp~(ac)


,_1"'"'
= "n w(
L""n'Pv 0 1
acj,cj, ... ,cjn)
j

= t JM L. acJ dcJ 1\ · · · 1\ dcj.


j
(11.15)

Now the remark after the inequality (11.12) shows that when a > 0 in
coo (M), the left hand side of (11.15) is nonzero, and thus the possibility
that t = 0 is ruled out.
Thus fM ao da1 1\ · · · 1\dan = t- 1 .\;:;_ 1 Ancp~ (ao, a1, ... , an) in general. For
example, if 0::;; ao ::;; 1, with ao = 1 on a neighbourhood of some point y,
and if x 1 , ... , xn are local coordinates near y, then by applying the Hölder
500 11. Connes' Spin Manifold Theorem

inequality of Proposition 7.14 for the case p = 1, we obtain

IJMaodxll\ ... Adxnj

::5--1 -,
1
t 1 n.
2::s Trw lxao [D,xrrOl] ... [D,xrr<nl] IDI-n j
TTE n

::5 - 1
1 1
t k=l
Il II[D,xk]ll Trw(ao IDI-n). (11.16)

Consider now the n-form

v :== t 2:: cJ dcJ " ··· 1\ dcj.


j

The map a - Tr+ ( a ID 1-n) = fM a v is a positive linear functional on


C"" (M), which kills no nonnegative function other than zero. The estimate
(11.16) shows that v does not vanish at any point y E M, and therefore is
a volume form on M.

11.3 The spin structure and the metric


We now examine the role of the finiteness and first-order conditions on
the geometry: these show that the operator D is not too far from being a
Dirac operator on M. To support that claim, we must first produce a spin
structure on M.
A candidate for the spinor module is already available, namely the space
J{"" of smooth vectors for the action of D on J{, which by the finiteness
hypothesis is a finitely generated projective .Jl-module; the algebra .Jl acts
on the left by (the restriction to J{"" of) multiplication operators on J{. The
Serre-Swan theorem allows us to write J{"" as a module of smooth sections
f"" (M, S) for a certain vector bundle S-M.
Proposition 11.5. There is a unique .JI.-valued pairing { · I ·} on J{"" such
that

f{IJ.II <J>} IDI-n = (<J> I IJ.I) forall <J>,tJI E J{"". (11.17)

Proof. As alreadydiscussedinSection 10.5, we mayidentify J{"" with m.JI.p,


for some positive integer m and some projector p E Mm (.Jl); the standard
positive definite Hermitianpairing on m.JI.p is given by {ap lcp }' := apc* =
Ljk a iP jkct for a, c E m .Jl. The notation indicates that the pairing will be
linear in the first variable, as discussed in Section 4.5. We may provisionally
define a new scalar product on J{"" by setting

(<J> I IJ.I)' :== f {IJ.II <J>}' IDI-n.


11.3 The spin structure and the metric 501

The calculation of this scalar product depends on the identification of Jf""


with m.JI.p, so it will be equivalent, thoughperhaps not equal, to the original
scalar product {· I · ). Let T denote the positive invertible operator on J{
suchthat
{</> 11/J)' ={</>I T!JJ) for all </>, 1/J E Jf"".
The tameness of the noncommutative integral shows that, for each a E .Jl,

<<I> I Ta!JJ) =<<I> I a!JJ)' = f {a!JJ I <J>}' IDI-n = f a {1/J I <J>}' IDI-n

= f {1/J I <J>}' a IDI-n = f {1/J I a*<J>}' IDI-n = {a*<l> 11/J)'


= {a*<J> I T!JJ) ={</>I aTIJJ),
for </>, 1/J E Jf"". We conclude that T commutes with the action of .Jl, so
that {1/J I <1>} := { T- 1 1/J I <1>}' defines another positive definite pairing on the
(left) .Jl-module Jf"". The new pairing clearly satisfies (11.17).
To see that the pairing that verifies (11.17) is unique, it is enough to notice
that the difference of two such pairings would be an .Jl-valued bilinear map
killed by all functionals of the form c ,_ f ac IDl-n, for a E .Jl; then, since
f aa* IDI-n = 0 only if aa* = 0 in .Jl, this difference must be zero. D

.,. It is now time to bring in the first-order condition. The identification


Jf"" "" f""(M,S), whereby elements of .Jl act as multiplication operators
on the left and equivalently on the right, shows that Jf"" can be regarded
as a symmetric .Jl-bimodule (or more simply, an .Jl-module); that is, TT = rr. 0

The first-order condition (10.84) simplifies to

[[D,a],b]=O forall a,bEC""(M).

We may extend this formula by continuity to all b E C(M); then it states


that for each a E C""(M) the operator [D, a] belongs to Endc(MJ(f(M,S)) =
f(M, End S). In fact, the regularity condition shows, through the agency of
Lemma 10.22, that [D, a] preserves Jf"" and so it belongs to f"" (M, EndS).
In other words, when a E C"" (M), [D, a] is a matrix-valued multiplication
operator on Jf"".
Furthermore, if a, b E .Jl and 1/J E Jf"", then

[D,ab]IJJ = a[D,b]IJJ + [D,a]b!JJ = a[D,b]IJJ + b[D,a]IJJ,


so that [D, ab] = a[D, b] + b[D, a] in f""(M,EndS). Thus,

Dab!JJ - a Dbi/J - b Da!JJ = ab DI/J

for each 1/); that is to say, D itself is a (matrix-valued) first-order differential


operator acting on smooth sections of the bundle S. (This can also be seen
502 11. Connes' Spin Manifold Theorem

bynoting that eachX(a) := {[D, a]IJ!i </>} satisfies the Leibniz rule X(ab) =
a(Xb) + b(Xa) and so defines a vector field on M.)

.,. The principal symbol ofthisdifferential operator is a function <TI (D) on


the cotangent bundle T* M. We refer again to Section 7.A for the symbol
calculus; we need mainly the formula (7.90):
<TI (D)(x, ~) := lim t-Ie-ita(x) D(eita(x). ),
t-oo
for any a E coo (M) suchthat da(x) = ~-Note that <TI (D) (x, ~) E End(Sx );
furthermore, if oc E .JI.I (M), the assignment x ..- a!(D)(x, OCx) is a smooth
section in roo (M, End S) which weshall denote by c(oc), namely,
c(oc)(x) := a!(D)(x, OCx). (11.18)
In particular, when oc =da and ~ = da(x), this symbol can be computed
with l'Höpital's rule:

c(da)(x) =<TI (D)(x, ~) = lim .!:.e-ita(x) D eita(x) = lim !!..e-ita(x) D eita(x)


t-oo t t-oo dt
= i lim e-ita(x) [D, a] eita<x> = i [D, a](x),
t-oo
where the last equality holds because [D, a] is a multiplication operator.
We conclude that
[D,a] = -ic(da).
The reader may have wondered at the absence here of "quantum differ-
entials" (images of universal forms in n· 5'1. under the representation rrv)
that played so prominent a role in earlier treatments: see [91, VI.1] or [469],
for instance. In that scheme, the de Rham algebra of differential forms was
recovered from the Clifford action of rrv (0" 5'1.) by factaring out the differ-
ential ideal ] = ker rrv + d ker rrv, lovingly called "junk" by the practition-
ers [2 59, 422]. A lucid exposition of how this works, when D is the Dirac
operator, has been given by Frank [186]. But the cantrast with our present
situation should now be clear: we possess the differential forms already
and it is the Clifford action that had to be painstakingly built .
.,. To obtain the promised spin structure, we now construct an algebrathat
is Morita equivalent to 5'1.; the equivalence bimodule will be J{oo, of course.
The first step is to note that the action of 1-forms given by (11.18) satisfies
a crucial commutation relation.
Lemma 11.6. Ifoc,ß E .JI.I(M), then [c(ß) 2,c(oc)] = 0.

Proof. Since we can always write oc = da locally, it is enough to show that


[c(ß) 2,c(da)] = 0 for each a. Moreover, c(ß) 2(x) = (a!(D)(x,ßx)) 2 =
a2(D 2)(x, ßx). using the multiplicative property (7.89) of principal sym-
bols. By the same multiplicative property,
[c(ß) 2,c(da)] = [a2(D 2),<To(i [D,a])] = ia2([D 2, [D,a]]),
11.3 The spin structure and the metric 503

so we must show that the expected principal symbol of [D 2, [D, a]] -of
second order-vanishes. (One may except such a commutator to yield a
first-order operator, with principal symbol given by the Poisson bracket as
explained in Section 7.A, only if the commutands are scalar pseudodifferen-
tial operators; for matrix-valued symbols, as in the present case, the order
of the commutator could be the sum of the orders of the commutands.)
Wehave already met, in Section 10.3, the relation
[D 2,[D,a]] = IDI8([D,a]) +8([D,a]) IDI,
where 8(T) := [IDI, T] as before. Lemma 10.22 shows that 8([D, a]) E Op~
by regularity, and it follows that [D 2, [D, a]] = [D, a]Ol E Opb. In the
notation of the norm estimate (10.65), this means that II [D, a]Oleitb!/111 ::::;
CIIIeitb!/1111 = O(t) for b E 5\., since
lleitb!/llli = (!/11 !/J} + (!/11 e-itbD2eitb!/1} = O(tz).

Therefore, if ~ = db(x),
O'z([D,a]O>)(x,~) = limt-2e-itb(xl[D,a]<Oeitb(x) = 0. 0
t-oo

Fix a point x E M and consider the effect at this point of the operators
c(a) obtained from real1-forms a E 5\. 1(M, ~). From (11.18), the matrix
c(a)(x) E End(Sx) depends only an ~ = ax E TiM; we denote it by
c(~). The finite-dimensional (complex) vector space Sx carries a positive
definite scalar product, namely, the restriction to this fibre of the pairing
(11.17) an roo (M, S). The matrix c(~) is Hermitian, since it is the value at a
real cotangent vector of the principal symbol of the selfadjoint operator D.
Therefore, the algebra generated by all such c ( ~) is an involutive subalgebra
of the matrixalgebraEnd Sx. As such, it is a finite direct sum of full matrix
algebras:
(11.19)

It follows from Lemma 11.6 that each c(~) 2 lies in the centre of this
algebra. Let p 1 be the orthogonal projector in EndSx whose range is the
summand Mk 1 (<[); it isaminimal projector in the centre. Then
(11.20)

where Bx is a scalar-valued quadratic form an TiM.


More generally, ~ - c(~) 2 may yield a direct sum of several quadratic
forms. If all of these are nondegenerate at each point x, then by the continu-
ity of the principal symbol 0'1(D), the number and size of the component
blocks in (11.19) is independent of x, the module J{oo breaks up into a
direct sum of submodules, and the geometry y decomposes as a corres-
ponding Whitney sum of subgeometries. By the irreducibility assumption,
there is only one such submodule.
504 11. Connes' Spin Manifold Theorem

Therefore, it can happen that r > 1 and PI * 1sx in (11.19) only if the
quadratic form (11.20) is degenerate at the point x; if that is so, we can
choose a basis ~I, •.. , ~n for TxM suchthat c ( ~n) 2 = 0 and hence c (~n) = 0.
Consequently,

c(~rr(l)) ... c(~rr(n)) = 0 (11.21)

for each permutation rr E Sn, since one factor of the product vanishes.
In fact, (11.21) holds for any basis of TiM, since a change of basis only
multiplies the left hand side by a determinant. Therefore,

:2: (-1) rr trsx (XxC ( ~rr(l)) ... C ( ~rr(n))) = 0.


TTESn

However, on account of the equality (11.15), this entails that the volume
form v vanishes at the point x, contrary to what we have established in the
previous section. Therefore, degeneracy ofg x is ruled out, and consequently
PI = 1. Note the immediate consequence that the differential operator D
is elliptic.
Wehave shown that, at each x E M, the operators c(~) generate an ir-
reducible representation of an n-dimensional Clifford algebra (TxM,gx)
with a nondegenerate quadratic form. Furthermore, the irreducibility as-
sumption shows that the bundle S-M has rank 2Ln/ 2J.
Taking stock, we have manufactured a Clifford action of .Jt 1 (M) on J{"",
whose square yields a nondegenerate metric on M:

This metric is Riemannian since c(oc) 2 (x) = u2(D 2 )(x, OCx) is the principal
symbol of a positive differential operator. Along the way, we have shown
that the fibres of the vector bundle S form a continuous field of Hilbert
spaces Sx, whose endomorphism algebras EndSx carry an irreducible ac-
tion of a Clifford bundle. When n = dimM is odd, the central elements of
these Clifford algebras act trivially on each fibre, by Schur's lemma, so that
we need only consider the action of their even subalgebras. The Clifford
action generates the following C*-subalgebra of L(J{):

alg(c(oc): oc E .Jt 1 (M)) if n is even,


{ (11.22)
B := alg(c(oc)c(ß): oc,ß E .Jti(M)) if n is odd,

and we have shown that the C*-completion f(M,S) of J{"" implements a


Morita equivalence between A := C(M) and B. Thus (v, J{"") constitutes a
spinc structure on the Riemannian manifold (M,g).
The distance formula follows at once from the relation [D, a] = -i c(da)
and the formula (9.26) which says that llc(da)ll = II gradalloo, where the
gradient is taken with respect to the metric g just constructed. Indeed,
11.3 The spin structure and the metric 505

starting from (9.24), we see that

d 9 (x,y) = sup{ la(x)- a(y)l: a E C(M), II gradalloo :51}


= sup{ la(x)- a(y)l: a E C(M), llc(da)ll :s; 1}
= sup{ la(x)- a(y) I : a E C(M), II [D, a] II :5 1 }.

The first part of the proof of Theorem 11.2 is now complete .


.,.. In order to get the spin structure, we now bring in the conjugation opera-
tor C that specifies the reality condition on the geometry t;i. We may write
Ca*c- 1 = a in view of the identification of 7T (a) with rr(a); in partic-
0

ular, C determines an antilinear bundle isomorphism of S, that is, it acts


fibrewise on each Sx as an antilinear isomorphism.
If oc E .Jl 1 (M.~) is a real1-form, the operator C intertwines c(oc) with
its negative. Indeed, locally we can write oc = da, where a is a real-valued
smooth function, so that
c(oc)(x) = lim t-1e-ita(x) D eita(x).
t-oo
If n is even, so that CDC- 1 = D by the reality condition, then

Cc(oc)c- 1 = limt- 1 euave-ita = c(-oc) = -c(oc) =: 8_1(c(oc)).


t-oo
When n is odd, cvc- 1 = ±D only, but this is enough to establish that
Cc(oc)c(ß)C- 1 = c(oc)c(ß) for oc,b E .Jl 1 (M.~). In either case, we may
conclude that
Cb*C- 1 = B-1 (b),
where b lies in the algebra B of (11.22).
Thus C satisfies (9.8a) and (9.8b). It remains to show that Cis antiunitary
for the pairing {· I ·} on :H"" (which corresponds to (9.8c) in our case, where
the roles of A and B are reversed). Since C is by hypothesis an antilinear
isometry on :H, then, for each a E .Jl,

f a {Ce/> I Ctf/} IDI-n = (Ctf/ I aCcp} = (Ctf/ I Ca*cp}

= (a*cf> I tf/} = <cf> I atf/} = f a {tf! I cf>l IDI-n.

The uniqueness argument of Proposition 11.5 allows us to conclude that


{ C cf> I Ctf/} = {tfJ I cp} for cp, tfJ E :H"". Therefore, C fulfils the require-
ments (9.8) for a conjugation operator on :H"", and so (v, :H"", C) defines
a spin structure on the manifold M.
Observe that the data of the given geometry t;i now form a reduced KRi-
cycle over the involutive algebra .Jl, in accordance with Definition 9.18. The
reality condition supplies, by hypothesis, the correct signs for j = n mod 8.
506 11. Connes' Spin Manifold Theorem

11.4 The Dirac operator and the action functional


Now that we have built a spin structure on M from the data of the given
noncommutative spin geometry § over C"" (M), we can proceed to define
the spin connection and the Dirac operator QJ uniquely associated to the
spin structure (v, J-{"", C); this construction has already been explored in
Section 9.3. However, the Dirac operator will generally not coincide with
the original operator D.
The operators D and QJ have two properties in common. Firstly, they are
both first-order differential operators on M, acting on the spinor bundle S
(or equivalently, on the module J-{""). Secondly, they share the same prin-
cipal symbol, namely, the Clifford action (11.18):
CTI(/J))(x, OCx) = c(oc)(x) := CTI(D)(x, OCx).
Therefore, D and QJ differ only by a zeroth-order term, that is, a matrix-
valued multiplication operator on J-{"":
D =: QJ + p, where p E f""(M,EndS). (11.23)

Three important algebraic properties verified by D and QJ are also shared


by p; these are selfadjointness, (anti)commutation with x and (anti)commu-
tation with C:
cpc- 1 = ±p, (11.24)

where the sign of the third equality is negative if and only if n = 1 or


5 mod 8.
The affine space referred to in part (b) ofTheorem 11.2 is the translate by
QJ of the real vector subspace of f"" (M, End S) satisfying (11.24). By fixing
a metric g in advance, one can regard the problern of classifying geomet-
ries over C"" (M) from the viewpoint of the topology of the Riemannian
manifold (M,g). Now M has only finitely many spin structures, since these
are classified by the finite set H 2 (M, 7L 2 ); fixing the spin structure specifies
the Dirac operator, and the remaining freedom in the geometry is just the
choice of p, subject to (11.24). The second part of Theorem 11.2 is now
established .
.,.. Since the differential operators D and I/) are elliptic, their powers are
elliptic pseudodifferential operators, and noncommutative integrals such
as f a ID 1-n can be computed by means ofWodzicki residues. Connes' trace
theorem yields

f aiDI-n = ~~7;;-b: Tr+(a IDI-n) = 2 Ln/!Jnn Wres(a IDI-n),


which generalizes (9.39). Fora E C""(M), the operator a IDI-n is pseudo-
differential of order -n, with principal symbol
a(x)<Ln(IDI-n) = a(x)CT-n(ß-n1 2 ) 1,
11.4 The Dirac operator and the action functional 507

where the factor 1 is an identity matrix of rank 2Ln/ 2J. The proof of Corol-
lary 7.21 then shows that

wresx(a ID\-n) = 2Ln/ 2JO.n a(x)~detgx ldnxl,

where the trace has been taken over the fibres of the spinor bundle, and
lv9 1(x) = .Jdetgx ldnxl is the Riemannian density of the metric that we
have constructed. Thus the integral f a ID 1-n = fM a v9 is independent of
the zeroth-order term p in (11.23).
Other powers of IDI may also have nontrivial Wresidues. Kastler, andin-
dependently Kalau and Walze, computed Wres 11,7) 1-n+ 2 and showed its rela-
tion to the Einstein-Hilbert action [268, 281]. lt turnsout that Wres IDI-n+ 2
indeed depends on p. Recall that the action functional S (D) of (11. 5) is
defined to be a multiple of this Wresidue; our next task is to compute it
explicitly.
We shall work with local coordinates (x, ~) on T* M, and write c(~) =
yk ~k for the Clifford action of a local1-form. 1t will be handy to abbreviate
11~11 2 := g- 1(~,~),so that, for instance, the principal symbol of any power
IDik is ak(IDik) = ll~llk. Recall from Section 9.3 that the Dirac operator 1,7)
has complete symbol
a(l,lJ) = yJ(~J + iwJ),

where the w J are the coefficients of the spin connection. The complete
symbol of Dis therefore a(D) = a 1(D) + a 0 (D) = c(~) + ao(D), where

ao(D) = ao(l,lJ) + p = iyJwJ + p.


The second-order differential operator D 2 has symbol a (D 2 ) ä2 +
ä 1 + ä 0 , easily found from (1,7) + p ) 2 = 1,7) 2 + {1,7), p} + p 2 and the composition
formula (7.88):

ü2 := a2(D 2) = 11~11 2 ,
ä1 := ai(D 2) = ai(J,7J 2) + {c(~),p},
äo := ao(D 2) = ao(1,7J 2) + i{yJwJ.P}- iyJ OJP + p 2.

We abbreviate a 0 (D 2) = ao(1,7J 2) +L2 (p) +p 2, where L2 (p) is linear in p or its


derivatives OJP· In what follows weshall generallyfind that ak-2(1Dik) =
ak-2(11,7Jik) + Lk(P) + Qk(p), where Qk is quadratic in p and Lk(P) is a
linear term. We do not write Lk explicitly, since it is always traceless and
does not contribute to the Wresidue density. We shall, however, compute
the quadratic part in full.
We now elaborate the complete symbol of v- 2. Write a(D- 2) =: a_ 2 +
a_ 3 +a- 4 + · · ·, andapplythe compositionformula (7.88) toD 2D- 2 = 1; the
principalsymbolsobeyä2a-2 = 1,andsoa-2 = äi 1 = II~II- 2 .Inparticular,
oa_ 2 JoxJ = 0 for eachj, which greatly simplifies the composition formula.
508 11. Connes' Spin Manifold Theorem

For the next two orders we find:

(11.25)

Solving order by order, we get

0"-3W- 2) = -0"2 26"1 = a_3(I/r 2)- 11~11- 4 {c(~). p},


a_4(D- 2) = -6"2 1 (6"10"-3 + <Toa-2- 2igJk~J Ok0"-3)
= 0"-4U/r 2) + L2(p) + 11~11- 6 {c(~),p} 2 -11~11- 4 p 2.

The same technique gives the leading terms in the symbol of IDl- 1 , by
analyzing the product IDI- 11DI- 1 = v- 2. lf, say, a(IDI- 1) = 6"-1 + 6"-2 +
&_ 3 + · · ·, then 0'~ 1 = a_2(D- 2), so 0'-1 = 11~11- 1 . The next two orders
satisfy

which entails

0"-2<1DI- 1) = 0"-2(if/JI- 1)- ill~ll- 3 {c(~),p},


0"-3(!D!-1) = a_4(jf/Jj-1) + LI(p) + ~~~~~~-5 {c(~),p}2- !11~11-3 p2.

Next, we consider theproduct v- 2m+ 2 = (D- 2)m- 1, whose principal sym-


bol is a~- 1 = ll~ll- 2 m+ 2 . The two succeeding terms are

0"-2m+1 (D -2m+2) m-2 0"-3


= ( m- 1) 0"_2
= 0"-2m+1(f/J- 2m+ 2) - (m -l)ll~ll- 2 m {c(~),p}

and

(]"-2m (D -2m+2) = (m _ 1 )0"m-20"


-2 -4
+ (m -1)am-30"2
2 -2 -3

- i(m- 1)a~-2 ~ a~~2 aa:-k3


= 0"-2m (f/J- 2m+ 2 ) + L-2m+2 (p)
+ m(~ -1) ll~ll-2m-2 {c(~),p}2 _ (m -l)ll~ll-2m p2.
11.4 The Dirac operator and the action functional 509

Finally, using IDI- 2m+ 1 = v- 2m+ 21DI- 1, we find that CT-2m+diDI- 2m+ 1) =
ll~ll- 2 m+ 1 and

0"-2m(IDI- 2m+ 1) = 0"-2mWl>l- 2m+ 1)- (m- !)ll~ll- 2 m- 1 {c(~),p},


0"-2m-diDI- 2m+ 1) = 0"-2m-1Wl>l- 2m+ 1) +L2m+1(P)

+ 4m~- 1ll~ll-2m-3{c(~),p}2
_ (m _ !)ll~ll-2m-1p2.

When n =2m or 2m+ 1, we can combine these two cases in one:

O"-n<IDI-n+ 2) = O"-nW.ZWn+ 2 ) +Ln+2(P) + n(nB- 2 ) 11~11-n- 2 {c(~),p} 2


- n;211~11-np2. (11.26)

To compute its wresidue density, we take the trace of this (matrix-valued)


symbol over each EndSx and integrate over 1~1 = 1-where 1~1 2 = Lk ~~.
in cantrast to 11~11 2 = g- 1(~.~).First of all,

tr{c(~). p} 2 = 211~11 2 tr p 2 + 2 tr(c(~)pc(~)p)


= 211~11 2 tr p 2 + 2~k~l tr(ykpy 1p).

The integral J1 ~ 1 =1 ~k ~~ Iu~ I vanishes if k * l and gives equal results, by


symmetry, when k = l; it follows that J1 ~ 1 = 1 ~k~zlu~l = Dkz/n. Therefore,

I1~1=1
11~11-n- 2 tr{c(~),p} 2 1u~l

= ( 2 tr p 2 + ~ ~ tr(ykpykp)) On ~detg ldnxl,

where the integration is performed just as in the proof of Proposition 7.7.


When n =2m is even, the terms in parentheses equal (4/n) tr(p N(p)),
where we introduce
n
N(p) :=! L yk {p, yk}.
k=1
We observe that, if r is odd:

N(yh ... yir) = ~yh ... yir + ~ Lk yk yJJ ... yJr yk


= r yh ... yir, (11.27)

since the kth summand equals yh ... yJr if k E U1 •... , ir} and - yJ 1 ••• yJr
otherwise. Since XP = -px by(ll.24), p(x) is anodd element ofEndSx and
so (11.27) determines N(p). Notice that N is none other than the "number
Operator" on n-(Ti'M) ""A-(Ti'M).
510 11. Connes' Spin Manifold Theorem

When n = 2m+ 1 is odd, N (p) can be computed by replacing each yk by


yky, for k = 1, ... , 2m, since the Clifford action is specified by (9.2). Since
the chirality element y is central, (11.27) holds also for r odd, whereas
N(yh ... yJr) = (n- r)yh ... yJr if r is even.
In either case,

q(p) := tr(p(N(p)- p)) (11.28)

defines a positive (semidefinite) quadratic form.


Lemma 11.7. The quadratic form (11.28) is positive definite.

Proof. Wemaywriteit as q(p) = tr(p(N -1)p). The domainof q is specified


by the conditions (11.24) on p: if n is even, it is the selfadjoint part of
End- S, whereas if n is odd, it is the selfadjoint part of EndS. Now, for
n even, N has eigenvalues 1, 3, 5, ... , n- 1 on each End- Sx; and for n odd,
N has eigenvalues 1, 3, 5, ... , non each EndSx. Thus N ~ 1 on the domain
of q in both cases.
It is clear that q(p) = 0 if and only if (N- 1)p = 0, i.e., if and only if
N(p) = p. Now suppose that n is even, and restriet to a local chart. Then,
since N is the number operator, the condition N(p) = p means that p(x) =
ykAk (x) for some complex functions A 1 , ... , An. and the selfadjointness
of p implies that each Ak is real. We also know that Cpc- 1 = p and that
Cc(cx)c- 1 = -c(cx) for anyreal1-form cx; this implies that Cykc- 1 = -yk.
The conclusion isthat CAkc-I = -Ak; in other words, the function Ak is
purely imaginary. This forces Ak = 0 for each k.
There is a parallel argument when n is odd. Again we work locally; now
N(p) = p and pt = p means that p(x) = ykyAk(x) for some real func-
tions Ak. Since y = (-i)Ln/ 2 ly 1 ••• yn, we find that Cykyc- 1 = ±yky, with
the positive sign if and only if n = 1 or 5 mod 8. By (11.24), Cpc- 1 = +p
holds with the opposite sign, and again CAkc-I = -Ak and Ak is purely
imaginary, so it vanishes.
Therefore, on any local chart, and thus globally too, we find that p = 0
whenever N(p) = p, which shows that the form q is positive definite. 0

.- We are ready to compute the Wresidue of ID 1-n+ 2 • The quadratic terms


in p yield

( n(n-2)
8
4
n n-2 ) n-2
tr(pN(p))- - 2 - tr(p 2 ) = - 2 - tr(p(N -1)p)

times On .y'(Ietg ldnxl = On v 9 in wresx IDI-n+ 2 • On integration over M,


the linear term in (11.26) vanishes, as already noted, and we end up with

Wres IDI-n+ 2 = Wres IWI-n+ 2 +(I- 1)0n L tr(p(N- 1)p) v 9 .


ll.4 The Dirac operator and the action functional Sll

Therefore,

S(D) = f IDI-n+ 2 = f IJ,lJI-n+ 2 +(I -1)2-Ln/ 21 L tr(p(N -1)p) v9 .

Lemma 11.7 now shows that S(D) has a unique minimum when p(x) = 0.
Its value is given by S (l,lJ) = f IJ,lJ 1-n+ 2, and it only remains to compute this
quantity.
A closely related quantity is Wres(!:l5 )-I+ 1 , where !:!.5 is the spinor Lapla-
cian. In Chapter 7, we computed the analogaus Wresidue for the scalar
Laplacian (associated to the Levi-Civita connection). The spinor Laplacian
is obtained from the scalar one by replacing \lB = d- f by \7 5 = d- Ji(f) in
a local orthonormal basis, as explained in Section 9.3. The Kastler-Kalau-
Walze formula expresses Wres!:l-I+ 1 in terms of the coefficient a 2 (x)
of the spectral density kernel of !:!.-I+ 1 , that is computed from the sym-
bol (7.29) of !:!.; on replacing f by Ji(f) and taking the trace over EndSx to
compute the wresidue density, a2 (x) is simply multiplied by the numerical
factor 2Ln/ 2l. Therefore, in view of (7.38), the KKW formula for the spinor
Laplacian is
Wres(!:l5 )-I+ 1 = n- 2 2Ln/ 2Jo
12 n M
sv
B•
J
where s is the scalar curvature of M.
Recall now the Lichnerowicz formula (9.30): J,lJ 2 = !:!.5 + ~ s (where 5 acts
as a multiplication operator on Jf We can use it to express the symbol
00 ).

u (J,lJ-n+ 2) in terms of u ( (!:!.5 )- I+ 1 ) and 5; we just replace the relation D =


1,lJ + p by Lichnerowicz' formula and run the algorithm again, to compute the
first three terms in the symbol of u(IJ,lJik), for k = -2, -1, -2m+2, -2m+ 1
in turn.
For instance, setting Ö"i := Uj(J,lJ 2) and UJ := UJ(J,lJ- 2), we again find that
u-2 = ä"i 1 = 11~11- 2 and that (11.25) holds. In the present case, ä-1 = u 1(!:!.5 )
and ä-0 = u 0 (!:!.5 ) + ~s -where each UJ(!:l5 ) can be read off from (7.29) on
replacing the terms A J by the components w J of the spinor connection.
Therefore,
u_3(J,lJ-2) = -ä"i2ä"1 = u_3((!:ls)-1),
U-4(J,lJ- 2) = -Ö"z 1(Ö"1U-3 + Ö"oU-2- 2igjk~j akU-3)
= u_ 4((!:ls)-1) _ ~~~~~~-4 5 _

Notice that the only contribution of 5 comes from the term -ä"i 1ä"ou-2 =
-11~11- 4 ä"o. In all cases, on comparing u(I.J,lJik) with u((!:l5 )kl 2 ), both the
principal and subprincipal symbols coincide, while Uk-2<IJ,lJik) contains a
term linear in 5. As a shortcut, we may determine it from the previous cal-
culation by setting {c(~). p} and alllinear terms in p to zero, and replacing
p 2 by ~s. We end up with

U-n(IJ,lJI-n+ 2) = U-n((!:l5 )-I+ 1)- n; 2 11~11-n 5.


512 11. Connes' Spin Manifold Theorem

Integration over I~ I = 1 and M then gives the required Wresidue:

Wres(IJJ)I-n+Z) = Wres((Ll5 )-~+ 1 )- n; f2Ln12JII~II-n


2 s(x) la?;dnxl

=Wres((Ll5 )-~+ 1 )- n- 2 2Ln/ZJnnJ sv9


8 M
= (n-2- n-2)2Ln/2J"
12 8 un
JM SVg

=- n- 22Ln/2Jn
24 n
J sv
M B·

Finally, the rninimum of the action functional is

S(JJ)) = - n 2~ 2 JM s Vg,

the required multiple of the Einstein-Hilbert action. The proof of Theo-


rem 11.2 is finished. D
~ Suppose now that the reality condition is omitted from the definition of
a noncommutative spin geometry, so that only the data (C"" (M), J{, D, x),
satisfying Conditions 1-3 and 5-7 of Section 10.5, are given. Then the con-
struction of volume form, Clifford action and metric go through without
change, and they produce a spinc structure (v,J{""). However, since the
conjugation C is not now available, this need not underlie a spin structure.
That leaves open the possibility that wz (M) * 0.
Moreover, a Dirac operator may not be available either, since there may
be no global spin connection. Even so, the spinc manifold M carries gen-
eralized Dirac operators, obtained from Clifford connections on J{"" as
explained in Section 9.3: each of these is of the form V( 0 l + iA, where
v(O> is an arbitrary but fixed Clifford connection and the gauge potential
A lies in .Jt 1 (M, T) = .Jt 1 (M) ®.5t 1:', where 1:' = f"" (M, T) is the rank-one
module associated to the spinc structure. Let JJ)(Al := -iC o (V'(Ol + iA) be
the corresponding generalized Dirac operator. The principal symbols of D
and JJ)(Ol coincide, so we may write D = JJ)(Ol + p, as before. However, the
proof ofLemma 11.7 fails in the absence of the conjugation operator C, and
now q(p) is merely positive semidefinite. When q(p) = 0, all that we can
now conclude is that p (x) is locally of the form yk Ak (x) [ or yk y Ak (x) in
the odd-dimensional case] where the functions Ak are real; the local forms
Ak dxk combine to yield a U (1) gauge potential in .Jt 1 (M, T). Thus the
minimum of the action functional S(D) is no Ionger unique, and instead, is
attained precisely on the set of generalized Dirac operators JJ)(Al.
In other words, by suppressing torsion terms, the minimization of the
action picks out exactly those operators that determine the commutative
geometry of the spinc manifold.
~ Theorem 11.2, which we have named Connes' "spin manifold theorem",
provides a purely noncommutative foundation for spin geometry, in that
ll.A The Riemann sphere as a spectral manifold 513

commutativity only enters through the assumption that the underlying al-
gebra is C"" (M). With its proof, we have achieved the goal of showing pre-
cisely how the realm of dassical spin geometry foreshadows the noncom-
mutative world. It therefore shows what are the essential features of a spin
geometry over a noncommutative algebra: we shall see some examples of
that in the next chapter. Each of the first six of the conditions laid out in
Chapter 10 plays an important role in the recovery of the spin structure
and the Dirac operator from the noncommutative geometric data.
The theorem was announced in the paper [96]; the proof in this chapter
follows the one outlined by Alain Connes in his 1996 winter course at the
College de France [95]. In keeping with the general approach of this book,
we have limited ourselves to that version of the theorem that takes the
algebra coo (M) as given.
One can make the stronger statement [96, 97] that a data set (j over a
commutative algebra 5\, complying with all terms and conditions laid out
in Section 10.5, will yield the same spin geometry, by identifying along the
way a manifold M whose algebra of smooth functions is isomorphic to 5\.
This requires a K-theoretic formulation of the nature of a differentiahte
manifold, beyond being the Gelfand spectrum of a commutative pre-C*-
algebra. Key issues are Poincare duality in its K-theoretic form, and how the
manifold coordinates may be extracted from the geometric terms, provided
that duality holds. This stronger version of the spin theorem has been taken
up by Rennie [386]. An important feature of his paper is the argument
that the natural map from o;b.Jl to HHn(.Jl)-see Proposition 8.10-is an
isomorphism, which then allows the reconstruction of a de Rham complex
from the Hochschild homology, in order to build up the desired manifold.
Connes' spin manifold theorem gives, then, a spectral reformulation of
the dassical geometry of spin manifolds. lt is by now well known that in or-
dinary Riemannian geometry, the spectrum of the Laplacian does not fully
determine the metric, even if its conformal dass is given [55], so the shape
of a drum cannot be heard. For spin manifolds, the situation is better: the
spectrum of the Dirac operatortagether with the volume form, or more pre-
cisely the unitary equivalence dass of its noncommutative spin geometry,
fully determines the metric and the spin structure. In that sense, one can
indeed hear the shape of a spinorial drum.

ll.A The Riemann sphere as a spectral manifold


In the course of this chapter, we have seen that the orientation condition
rrv (c) = x is a defining equation for spin geometries. If D is fixed, one may
regard it as a (highly nonlinear) equation for the coefficients c j of c; we shall
now see, by an example, that it can determine the algebra 5\ which these
coefficients generate. On the other hand, if this algebra and consequently
514 11. Connes' Spin Manifold Theorem

the cycle c are known beforehand, one may ask which operators D obey
the relation rrv(c) = X· In the commutative case, this amounts to finding
all Riemannian metrics having a prescribed volume form; the orientation
condition then becomes a fixed-volume constraint while allowing variations
of the metric, and thus is important for a noncommutative treatment of
gravity [96].
It is pointed out in [101] that if one replaces c by components of the
cyclic-homology Chern character (10.25), these same equations hold a great
deal of geometrical information, provided one replaces the commutative
algebra .Jl by the mildly noncommutative matrix algebras Mr(.Jl). We take
the opportunity here to explain how the 2-sphere § 2 , with its standard
volume form, emerges from such equations.
lf a (not necessarily commutative) involutive algebra .Jl is given, we can
consider a projector p E M2 (.Jl), say, which leads to the chain ch2k p :=
tr( (p- ~) 181 p®2k) E .Jl 2k+l, where in this section we define the partial trace
tr: Mz(.Jl)- .Jl as

trp = tr (Pu Pl2) 1


Pzz := 2<Pu + Pzz), (11.29)
P21
so that tr(1 2 ) = 1. Wehave omitted the normalization factor of (10.25a).
Recall, from Section 10.2, that the various chzk p appear as components of
a cycle eh p in the total complex, with boundary operator b + B, defining
cyclic homology. lt follows that ch2k p is a Hochschild 2k-cycle whenever
the lower components vanish: ch 2j p = 0 for j = 0, 1, ... , k- 1.
For simplicity, we shall consider only the case k = 1. Writing [D, p] as
a shorthand for [D 181 1z, p ], and plugging c = ch 2 p into the orientation
condition, we arrive at a pair of equations
tr(p- ~) = 0, (11.30a)
tr((p- i) [D,p] [D,p]) =X· (11.30b)
Now, however, we may regard this pair of equations form a different
standpoint, by taking them to be simply a pair of equations for unknowns
p and D: we then look for K-cycles which solve this equation. That is to say,
suppose we are given (only) a if 2 -graded Hilbert space J-{, and tr denotes
the normalized partial trace (11.29), taking L(Jf e Jf) to .L(Jf). Then we
seek a family of projectors p on Jf e Jf and a selfadjoint operator D with
compact resolvent satisfying (11.30), where it is implicit that the operator
[D, p] be bounded. If .Jl denotes the algebra of bounded operators on Jf
generated by the components Pii of such projectors p, then (.Jl, Jf, D) is
a spectral triple for which (11.30b) provides an orientation condition.
The defining conditions p* = p and tr(p- ~) = 0 allow us to write the
projector p in the form

1 ( 1+Z X- iy)
p = 2 X + iy 1- Z '
ll.A The Riemann sphere as a spectral manifold 515

where x, y, z are bounded selfadjoint operators on J-(, with -1 ::5 z ::5 1 by


positivity of p. The idempotence p 2 = p implies
(1 ± z) 2 + x 2 + y 2 ± i[x,y] = 2(1 ± z),
(1 + z)(x ± iy) + (x ± iy)(1 ± z) = 2(x ± iy),
which simplify to [x,y] = [y,z] = [z,x] = 0 and x 2 + y 2 + z 2 = 1. Thus,
x, y, z generate a commutative algebra 5\, and their joint spectrum in ~ 3
is a closed subset V of the sphere § 2 , on account of x 2 + y 2 + z 2 = 1.
Simply put, the equation (11.30a) alone, applied to a projector p, shows
already that p is the restriction to V ~ § 2 of the Bott projector! Our hap-
piness will be complete if we can show that V is all of § 2. For that, the
second equation (11.30b) is needed. If we abbreviate da := [D, a] here, it
is enough to compute tr( (p - t) dp dp); in fact, we have essentially done
so already in Section 8.3. The result is t<an + a22) where an and a22 are
similar to (8.37), except that 1 ± z must be replaced by ±z, and dx by dx
and so on in that formula, without forgetting that here dx and dy need
not anticommute. We find that
X= tr((p- ~)dpdp)
= ~(x(dydz- dzdy) + y(dzdx- dxdz) + z(dxdy- dydx)).
This is just the equation rrv (c) = x. where c is the Hochschild 2-cycle
c := ~(x ® (y ® z- z ® y) +y ® (z ® x- x ® z) + z ® (x ® y- y ® x)).
The corresponding volume form is given by (11.4), with n = 2 and in-m = i
in the present case. lt is none other than
CT = x dy A dz - y dx " dz + z dx A d y,
namely, the standard volume form on § 2 ! Since this vanishes nowhere on
the sphere, we conclude that V = § 2. Therefore, the pre-C* -algebra gene-
rated by x,y,z is isomorphic to C""(§ 2 ). Wehave recovered the Riemann
sphere from the equations (11.30).
On the other hand, Connes' spin manifold theorem shows that, if g is any
metric on § 2 -not just the round metric- whose volume form y'(l:"efg d 2 x
equals CT, then the corresponding Dirac operator D = Jj) 9 satisfies (11.30b).
Therefore, (11.30) provides a geometrical framework within which one may
vary the metric while keeping the volume form fixed.
At the next stage, we may consider the system of equations
tr(p- t) = 0, tr((p- t) [D,p] 2 ) = 0, tr((p- t) [D,p] 4 ) = x.
where tr is now the normalized partial trace on 4 x 4 matrices. We re-
fer to [101] for the detailed calculation. One finds solutions p E M4 (5\)
where 5\ = C"" (§ 4 ), the sphere being obtained as the quaternionie projec-
tive space; and D is any Dirac operator whose metric gives the standard
volume form on the 4-sphere.
Part IV

TRENDS

I think this is the beginning


of a beautiful friendship

-from Casablanca
12
Tori

This last part of the book sets out to explore several interfaces between
noncommutative geometry (NCG) and physics. We have chosen to report
on the avenues that lookmore promising, as the century draws to a close,
rather than on the more established "applications".
The noncommutative geometry reconstruction of the Standard Model of
particle physics, originated in [109], has been exhaustively explored. This
is a purely classical model, whose quantumcounterpartstill eludes us [5].
This book is not the appropriate place to review it, and we must refer to
our own survey [329] and to the pathbreaking papers by Chamseddine and
Connes [75, 76].
However, the pivotal role played by the noncommutative view of the
Higgs particle as a vector boson in the rapprochement between noncom-
mutative geometry and physics cannot be ignored altogether. We there-
fore quickly review the relevant facts. The elementary particles discovered
so far in Nature fall into two categories only: spin-1 bosons called gauge
particles, and (three generations oO spin-~ fermions. These two kinds are
very different from both the physical and geometrical standpoints. Phys-
ically, spin-1 bosons are the carriers of fundamental interactions that go
hand in hand with a gauge symmetry: local SU(3) for the strong force, and
SU(2)r x U(1)y for the electroweak interactions. The physics of spin-~
fermians is obtained by applying Dirac's minimal coupling recipe to the
fermionic free action: fermians interact by exchanging the gauge particles
that their quantum numbers allow.
The dynarnics of the spin-1 gauge bosons are governed by the Yang-
Mills action, which is the square of the curvature of the gauge connection.
520 12. Tori

Geometrically, fermions are sections of the spinor bundle. The minimally


coupled Dirac operator, that is, the Dirac operator twisted with a gauge
connection, yields the gauge-covariant derivative of these sections.
Fermionic matter splits into left-handed and right-handed sectors. Chi-
rality or handedness, and the fact that SU(2)r x U(l)y quantum numbers
of the left-handed fermions differ from the corresponding quantum num-
bers of their right-handed partners, must be incorporated as ingredients of
any noncommutative geometry treatment.
In nature, the SU(2) x U(l)y symmetry has been broken down to U(l)em·
The standard way of inducing such a symmetry breaking implies includ-
ing in the spectrum of the theory a scalar particle, called the Higgs bo-
son, by adding to the sum of Yang-Mills and fermionic Lagrangians the
so-called Higgs Lagrangian. The latter has three parts: the gauge-covariant
kinetic term for the Higgs field, the symmetry-breaking Higgs potential,
and the Yukawa terms. Three of the original SU(2) x U(l)y gauge bosons
become massive via the Higgs mechanism. The Yukawa terms, in turn, ren-
der massive the appropriate fermions. Although the Higgs particle has not
yet been detected -so there is no definitive evidence that nature behaves
in accordance with the symmetry-breaking pattern provided by the Higgs
Lagrangian- it Iooks increasingly likely that it will be detected in the com-
ing years.
The Higgs sector of the Standard Model Lagrangian is undeniably an ad
hoc addition to the original Lagrangian, its main virtue being that it pro-
vides a perturbatively renormalizable and unitary way to implement the
breaking of the SU(2)r x U(l)y symmetry. lt Iacks the geometrical inter-
pretations of the Yang-Mills and Dirac actions. Moreover, the Higgs particle
is undeniably a very odd particle in the Standard Model: it belongs to none
of the particle categories discovered so far, being the only scalar elementary
particle in the minimal Standard Model.
If we examine the structure of interaction terms in the Standard Model,
we may notice something peculiar: the point-like couplings between fer-
mions and the Higgs particle Iook very like those between fermions and
gauge bosons. Noncommutative geometry has the capacity to substanti-
ate this remark, by reinterpreting the Higgs field as a gauge potential: one
would then expect Dirac's minimal coupling recipe to yield at once the
Yukawa terms and the Dirac terms. This is possible because in noncom-
mutative geometry external and internal degrees of freedom of elementary
particles are, of course, on the same footing.
A particular spectral triple, able to reproduce the whole gamut of the
Standard Model particle interactions, is defined as follows. Consider the
"Eigenschaften algebra",
.J\p :=I( EB D-!1 EB M3(1(),

with D-!1 denoting the quaternions. Let J{F be a finite-dimensional Hilbert


space with a basis Iabelied by the Iist of elementary particles. A Dirac ope-
12. Tori 521

rator DF, defining a 0-dimensional spin geometry, is appropriately chosen


according to the Kobayashi-Maskawa generation-mixing matrices. Then the
product of this geometry, in the sense of Section 10.5, with the ordinary
spin geometry over (Euclideanized) spacetime gives (a Euclidean version of)
the Standard Model.
The Higgs particle (a quaternionie field) now lies at the outputend of the
model, and the properties of the symmetry-breaking sector are determined.
By means of an extra condition, the so-called unimodularity condition (equi-
valent to cancellation of anomalies in the noncommutative framework [6]),
the model gives also the correct hypercharge assignments. lt has been ar-
gued that "most" models of the Yang-Mills-Higgs type cannot be abtairred
from noncommutative geometry [259, 317,422]; it is all the more remark-
able that the Standard Model can.
The Connes-Lott approach which is reviewed in [329] and the Chamsed-
dine-Connes approach differ mainly in that, in order to obtain the "world
Lagrangian", the former makes heavy use of the "quantum differentials"
briefly mentioned in Section 11.3, whereas the latter relies on the gener-
alization of the action functional calculation of Section 11.4 to the pro-
duct spectral triple (see also [180] in that regard). The calculations are
rather formidable in both cases. Fora recent painstaking summary of those,
see [160]. The first method seems to suggest that the mass of the Higgs
particle (after allowing for some quantum corrections) should be of the
order of 1.3 top masses, at least. The second points to lower values, but
still higher than present estimates. Other vexing issues remain, such as the
(in)compatibility of Poincare duality and massive neutrinos [417], and there
seems to be no clear way forward at the time of writing.
lt is fair to say that some attempts to interpret the Higgs field as a gen-
eralized Yang-Mills model predate noncommutative geometry proper. A
pioneer paper in that respect was [149]. A particularly appealing effort is
the su ( 211) model of electroweak interactions, due to Coquereaux, Esposito-
Farese, and Vaillant [119] and to Häußling, Papadopoulos and Scheck [235],
building on an idea that goes back at least to [356]. Generation mixing is
very naturally described in this model. A good review of it is [234] .
.,.. Chapter 12 belongs to the category of examples; but these are models
that crop up in different areas and enjoy a tremendous popularity. Besides,
maybe (two-dimensional) noncommutative tori are only the first of a dass
of examples describing noncommutative Riemann surfaces; in this sense
the paper [355]looks promising. The most impressively beautiful applica-
tion of the noncommutative tori is undoubtedly that they provide a geomet-
ric model for the integerquantumHall effect: see [29,334,497], or [91, IV.6]
itself, for lucid surveys of the physical theory of this effect and the role of
noncommutative 2-tori in explaining how and why the Hall conductance is
quantized. We are unable to improve on those.
522 12. Tori

Noncommutative tori have been used, for instance, as compactifications


in superstring theories [103] and as first-quantized spaces underlying fer-
mion field theories [470]; see [430] for a comprehensive discussion of these
issues. The full geometrical structure is needed to obtain the physical inter-
pretation of the models. For other approaches which use tori to deal with
spacetime geometry, consult the review [318].
At a deep and perhaps fundamentallevel, quantum field theory (QFT) and
noncommutative geometry are made of the same stuff. To examine their
interconnection, it helps to have an algebraic version of QFT. This is one
reason why we unearth the Shale-Stinespring second quantization appara-
tus (which, in the charged case, is tantamount to the simplest Connes' Fred-
holm module): the transition from the commutative to the noncommutative
manifold case is almost unnoticeable in that framework. And indeed, quan-
tum theory on noncommutative backgrounds, starting with [182] and [470],
has known an explosive development. Doubtless, not everything in that pro-
duction is quite correct, and a lot of work needs to be done to clarify the
issues of principle. Beyond that, Chapter 13 is an unabashed excursion in
physics; the matter of the phase of the quantum scattering matrix is solved
within the Shale-Stinespring formalism, thereby brushing with the diver-
gence troubles and renormalization procedures of quantum field theory.
In Chapter 14, we encounter them again; some melodies insinuated in
the very first chapter in relation with the understanding of noncommuta-
tive symmetries, are woven with new themes in renormalization theory,
substantiating the claim that QFT and noncommutative geometry are cut
from the same cloth. Wehave tried to dissipate the aroma of mystery that
still surrounds the simultaneaus appearance of the algebra of rooted trees
in both theories; but, perhaps fortunately, some of it stilllingers.

12.1 Crossed products


The simplest examples of highly noncommutative spaces that can carry in-
teresting spin geometries are the noncommutative tori. They arose from
problems in ergodie theory, namely, the study of the rotation of the cir-
cle by an irrational angle and the related issue of the Kronecker flow on
the 2-torus with an irrational slope. These dynamical systems generate cer-
tain "transformation group C*-algebras" [153] that are highly noncommu-
tative. On considering rotations of the circle by rational angles too, one ob-
tains a family of C*-algebras that include the commutative algebra C("U" 2 )
as "case zero"; these are the noncommutative 2-tori. About 1980, their K-
theory groups and Morita equivalence classes were computed by Pimsner,
Voiculescu and Rieffel [374, 390], while Connes [83] analyzed their differ-
ential structure. The 2-tori were, indeed, forerunners of the noncommuta-
tive differential geometries introduced by Connes in his 1985 paper [86].
12.1 Crossed products 523

Higher-dimensional torialso exist and provide interesting examples of non-


commutative spaces in any dimension.
Let it be said that, although this chapter is not very short, we barely
scratch the surface of what is known on noncommutative tori. The most
glaring omission is the lack of direct consideration and use of the Chern
character. A treatment rendering full justice to the subject, sorely needed,
would already be of book size .
.,. The term "noncommutative torus" refers to two species of algebras: the
C*-algebras originally called "irrational rotation algebras", as well as the
pre-C*-algebras that consist of the smooth elements of these C*-algebras
under the action of certain Ue groups of automorphisms. Webegin with the
C*-algebras, and turn later to the pre-C*-algebras when we need to focus
on the differential structure.
Definition 12.1. A continuous action of a locally compact group Gon a C*-
algebra Ais a group homomorphism a: G - Aut(A) suchthat t ..... ar(a)
is a continuous map from G to A, for any a E A. The triple (A, G, a) is
often referred to as a C* -dynamical system [366, 456]. The systems of most
interest to us are the cases G = ~. involving one-parameter groups of auto-
morphisms of A, and G = ?L, obtained by iteration of a single automorphism
a 1. We limit our scope-and lighten the notation-by dealing with abelian
groups only.
Several involutive algebras can be associated to a C* -dynamical system.
The first is Cc(G-A), consisting of continuous functions a: G - A with
compact support. (For a discrete group like ?L, this means finite support.)
The algebra operations are the convolution and involution given by

(a * b)(t) := Jc a(s) a 5 (b(t- s)) ds, a*(t) := ar(a(-t)*). (12.1)

This product is continuous for the L1-norm

llaiii := Jc lla(t)ll dt,


and the completion of Cc(G-A) in this norm is an involutive Banach al-
gebra L1(G-A). This is a C*-algebra only in trivial cases; even if A = C
and G = ?L, this algebra is -t 1(7L), not a C*-algebra. However, we can also
associate a C*-algebra to the dynamical system, as follows.
Definition 12.2. If Bis an involutive Banachalgebra with norm II • 11 1, any
involutive representation p of B-that is, a *·homomorphism from B to
L(:Hp) for some Hilbert space :Hp-satisfies llp(b)ll =::; llbll1- Indeed, p is
norm-decreasing (since it shrinks spectra and L(:Hp) is a C*-algebra). The
supremum over all such involutive representations p is bounded:
llbll := sup llp(b)ll =::; llbll1. (12.2)
p
524 12. Tori

and thus defines a serninorm on B. (In many cases, it is already a norm;


otherwise, we quotient B by its kernel to get a normed algebra.) Since
/lp(b*b)/1 = llp(b)/1 2 for each p, this is a C*-norm, and the completion
of B in this norm is a C* -algebra, called the enveloping C* -algebra of B.
Given a C*-dynarnical system (A, G, oc), the enveloping C*-algebra of
L1 (G-A) is denoted by A ~a G, and is called the crossed product of A
by the action oc of G.
Even if the C*-algebra A is itself commutative, say A = Co(Y), and
G is abelian, the crossed product C* -algebra is in general not commuta-
tive. Theseare the so-called transformation-group C*-algebras C* (G, Y) :=
C0 (Y) ~a G [153]. Many of the examples we consider are of this type.
When Gis a discrete group, Ais embedded in L1 (G-A) and thus also
in A ~a G as the functions supported by the identity element of G. The
group Gis also represented inside A ~a G by the delta-functions 8t (t) := 1,
8t(s) := 0 for s * t; it follows that 8ta8i = oct(a) for a E A, t E G. By
construction, each involutive representation p of L 1 (G-A) extends to a
representation of A ~a G (also called p). Given such a p, Iet rr := PIA and
set Ut := p(8t). We thereby obtain an instance of the following definition.
Definition 12.3. A covariant representation of a system (A, G, oc) isapair
(rr, u), where rr is a representation of A and u is a unitary representation
of the group G on the same Hilbert space, satisfying

UtTT(a)ui = rr(oct(a)), for all a E A, t E G.

Given a covariant representation (rr, u) of (A, G, oc), the recipe

p(a) :=Je rr(a(t)) ur dt (12.3)

deterrnines a nondegenerate *-representation of U(G-A). The converse


is also true [366, Prop. 7.6.4]: using approximate units for A and L1 (G), one
can construct a covariant representation of (A, G, oc) from any nondegen-
erate representation of L1 (G-A) and recover the latter from (12.3). Thus,
there is a bijective correspondence between covariant representations of
(A, G, oc) and nondegenerate *-representations of A ~a G .
.,.. The following alternative construction avoids the need to take the supre-
mum (12.2) over all possible involutive representations. If A acts faithfully
on a Hilbert space J{, let !it := L2 (G-Jf) = L2 (G) ® J{ be the Hilbert
space of functions ~: G- J{ for which 11~11 2 := fc ll~(t) 11 2 dt is finite, and
define a covariant representation (rr, ,\) of (A, G, oc) on !it by

[rr(a)~](s) := oc_ 5 (a)~(s), [Ar~](s) := ~(s- t).

The closure of the range of the corresponding "regular" representation


p: L 1 (G-A) - L(L 2 (G-Jf)) is a C*-subalgebra of L(L 2 (G-Jf)) called
12.1 Crossed products 525

the reduced crossed product. Indeed, this C*-algebra is generated by the


operators fi"(a) and At.
The representation {J extends to a morphism of C*-algebras from A ~a G
onto the reduced crossed product. lt turns out that, when G is an amenable
group-a dass that includes all abelian groups and compact groups, but not
semisimple noncompact groups like SL(2, ~)-this is anisomorphism [366,
Thm. 7.7.7]. Since we shall deal with abelian groups only, all our crossed
products will be reduced.
Consider the particular case where A = C(X), G = l; the automorphism
a 1 of C(X) is determined, using the Gelfand cofunctor, by a homeomor-
phism cp: X -X via ai(j(x)) =: j(cp- 1 (x) ). Suppose that Xis endowed
with a regular Borel measure J1 which is invariant under cp. In this case, the
crossed product C(X) ~a l acts on the Hilbert space .f!2(l) ® L 2 (X, df.l) by

[,\m~Jn(X) := ~n-m(X). (12.4)

Exercise 12.1. Show that, if J1 is a cp-invariant measure on X with full sup-


port, then the C* -algebra of operators on L 2 (X, df.l) generated by the mul-
tiplication operators ~(x) .... j(x) ~(x) and the shift operator V~(x) :=
~(cp(x)) is a faithful representation of C(X) ~a l. 0

Weshalltake the representation of this exercise as a working definition


of C(X) ~a l.
~ Another special case is A = C with trivial action a. Then (12.1) is the
ordinary group convolution, and the crossed product, now written C* ( G),
is just the enveloping C*-algebra of U (G) [137]; it is usually called the
group C* -algebra of G.
A useful variant of the group C* -algebra construction arises in the the-
ory of projective group representations, where one needs another datum,
namely a 2-cocycle or "multiplier" determining a central extension of the
group (a concept that appeared already in Chapter 6, in multiplicative no-
tation).
Definition 12.4. If Gis a locally compact abelian group, a 2-cocycle on Gis
a function a: G x G - lr suchthat a(O, 0) = 1 and

a(r,s)a(r + s, t) = a(r,s + t)a(s, t) for all r,s, t E G.

lt follows easily that a(t,O) = a(O,t) = 1 and a(t,-t) = a(-t,t) for all
t E G. Two cocycles 0" and T are called cohomologous if O"(S, t)T(S, n- 1 =
1J(S)1J(t}1J(S + t)- 1 for some 17: G- lr. The Banachalgebra L1 (G,a) is
defined by introducing a twisted convolution and involution

(a * b)(t) :=Je a(s) b(t- s) a(s, t - s) ds,


a*(t) := a(t,-t)a(-t). (12.5)
526 12. Tori

The cocycle property ensures that this convolution is associative. The en-
veloping C*-algebra of L 1 (G, <T) is denoted C* (G, <T); it is called the twisted
group C* -algebra of G by the cocycle <T. Weshall usually denote products
in the completed algebra C* (G, <T) simply by ab rather than a b. *
Exercise 12.2. Show that C*(G,<T) ""C*(G,T) if and only if the cocycles
<T and T of G are cohomologous. 0

12.2 Structure of NC tori and the Moyal approach


The two-dimensional noncommutative tori are transformation group C*-
algebras obtained from the action of 7L on the circle X = lr by iterates of the
rigid rotation by an angle 2rrB. An obvious invariant measure J.1 is then, of
course, the Lebesgue measure on the circle.
Definition 12.5. Let (} be any real number. For convenience, we identify
functions F E C(lr) to continuous periodic functions on f: ~ - I( with
period 1 by j(t) = F(e 2rrit ). The rigid rotation of the circle by the angle 2rrB
yields the automorphism cx(j)(t) := j(t + B) of C(lr). The torus algebra
A~ is defined as
A~ := C(lr) ~a 7L.
This can be regarded as the algebra of operators on L 2 (lr) generated by the
multiplication operators and the unitary shift operator V~(t) := ~(t + 0).
We can expand elements of C(lr) as Fourier series:

j(t) = I Cn e2rrint.
nEZ

The unitary multiplication operator

generates the representation of C(lr) on L 2 (lr). Therefore, the algebra A~


is generated by the two unitaries U and V. Since (VU~)(t) = (U~)(t + B) =
e2rri(t+Ol~(t + (}) = e2rri9 (UV~)(t), these unitaries obey the commutation
relation VU = e 2rrie UV.
Proposition 12.1. Let A be the universal C* -algebra generated by two uni-
taries u, v such that

vu = e 2rri9 uv. (12.6)

The map u - U, v - V determines an isomorphism of A onto A~.


Proof. The linear map that takes finite sums Lm,n amnumvn to the re-
spective sums Im,n amnumvn is multiplicative, on account of (12.6), and
12.2 Structure of NC tori and the Moyal approach 527

extends to a unital morphism cf>: A- A~. Todefine an inverse morphism,


it suffices to obtain a covariant representation of (C('U'), lL, oc) whose image
is isomorphic to A. We can suppose that A is faithfully represented on a
Hilbert space J{ and define (rr,A) on J{ as follows: set Am:= vm, and if
j(t) = Ln CneZrrint is a finite Fourier series, let rr(j) := Ln CnUn. Clearly
Amrr(j),\~ =Ln CneZrrimnOun = rr(ocm(j)), therefore 1T extends to C("U')
to give the desired covariant representation. The associated representation
of A~ is an isomorphism p: A~ - A satisfying p(U) = u and p(V) = v,
which is inverse to cf>. 0

In consequence, any C*-algebra generated by two unitaries satisfying


(12.6) is a quotient of A~. In fact, since A~ is simple for irrational 8, as we
shall see, any such algebra is isomorphic to A~.
~ It is clear from (12.6) that A~ is abelian if and only if (} is an integer. We
may identify Ao with the C*-algebra C("U' 2 ), i.e., functions on the familiar
2-torus with angular coordinates (c/> 1 , c/> 2 ), by taking u := eZrric/> 1 and v :=
e2rric/>2.
The "generators and relations" presentation of the C*-algebra afforded
by Proposition 12.1 immediately gives certain isomorphisms between the
various A~. Firstly, there is the isomorphism A~ "' A~+O for any n E 7L, since
(12.6) is unchanged by (} ...... n + 8. Therefore we can, whenever convenient,
restriet the range of the parameter (} to the interval 0 ~ (} < 1. On the
other hand, since uv = e 2rriO-tll vu, the correspondence u ...... v, v ...... u
extends to an isomorphism A~ "' AL 0 • We shall see shortly that these
are the only isomorphisms between the torus algebras; so that the interval
[0, ~] parametrizes a family of nonisomorphic C*-algebras.
For nonintegral values of (}, the rational and irrational cases are very
different. We are chiefly interested in the algebras A~ for irrational 8, but,
to gain some perspective, we first examine the structure of the rational
rotation algebras, following [251].
Proposition 12.2. Suppose that (} = p I q where p, q are relatively prime
integers, with q > 0. The torus algebra A~ 1 q is the algebra of continuous
sections of a vector bundle over 'U' 2 , whose fibres are full q x q matrix alge-
bras.

Proof. Let R, S be the unitary q x q matrices

1 0 0 1
1 0
R:= s := 1 (12.7)
0
0 0 1 0
528 12. Tori

where we abbreviate i\ := e 2rrip/q. It is clear that Rq = 1q, Sq = 1q and that


RS = i\SR.
Define an action of the finite abelian group H = 7Lq ES 7Lq on the algebra
= C("ß"" 2 ) ® Mq(<C) by
Mq(C("IT" 2 ))

ocw(j(z, w) ® A) := j(z, i\ - 1 w) ® RAR- 1 ,


oco1 (j(z, w) ® A) := JCA.z, w) ® SAs- 1 •

The fixed-point subalgebra of Mq (C CU" 2 )) und er this action is generated by


the elements u := w ®Sand v := z ® R. Since

(z®R)(w®S) =i\(w®S)(z®R),

this subalgebra is isomorphic to A~tw Any element of A~fq is of the form


2.~ s=l frs(Zq, wq) urvs.
Let E - -u- 2 be the trivial bundle with fibre Mq ( <C), and let To 1 , T 10 be
the bundle isomorphisms defined by T 10 (z, w;A) := (z, i\- 1 w;RAR- 1 ),
T 0 I(z,w;A) := (i\z,w;SAS- 1 ). These maps commute and generate com-
patible free actions of the group H on E and its base space -u- 2 • Now -u- 2 1H ""
T2 via the wrapping map (z, w) - ( zq, wq), so the quotient E 1H is the to-
tal space of a vector bundle F - T2 , whose typical fibre is Mq ( <C), with
twisting indexed by p. It is then easy to see that f(F) is isomorphic, as a
C("U" 2 )-module, to the fixed-point subalgebra of [(E) under the action of H,
which is just A~lw D

Corollary 12.3. I{B is rational, the centre of A~ is isornorphic to C("U" 2 ).

Proof. Central elements are those of the form g(zq, wq) ® 1q. D

For irrational 8, the algebra A~ is in fact an inductive limit of finite direct


sums of matrix algebras over C("U") [158) .
.,. When e = 0, any element a E AÖ = C("U" 2 ) can be developed as a Fourier
series:
a(cf>I, cf>z) ~ I ars e2rrirc/>Je2rrisc/>2,
Y,SE?l

where the series converges uniformly to the function a only under certain
conditions (say, if a is piecewise C1 ); to avoid the severe representation
problems that arise even in the classical case, it is convenient to restriet
the series developments to the pre-C*-algebra of srnooth functions on T2
(whereby the Fourier series converges to a both uniformly andin L 2 -norm).
Now a E C"" ("U" 2 ) if and only if the coefficients ars belong to the space S(7L. 2 )
of rapidly decreasing double sequences, i.e., those that satisfy

sup (1 + r 2 + s 2 )k lars1 2 < oo for all k E ~­


r,sE?l
12.2 Structure of NC tori and the Moyal approach 529

Definition 12.6. The noncommutative torus lf~ is the dense subalgebra of


A~ given by

lf~:={a= 2: arsUrV 5 :{ar 5 }ES(l 2 )}. (12.8)


r,selL

In particular, the unit 1 of A~ lies in lf~. The Ue group lf 2 acts on the algebra
A~ by (z, w) · urvs := zrws urvs, and lf~ is just the subalgebra of smooth
elements for this action. lt follows from Proposition 3.45 that lf~ is a unital
pre-C* -algebra.
The C*-algebra A~ comes equipped with a distinguished tracial state,
given on the dense subalgebra lf~ by T(a) := aoo. Clearly, T(1) = 1 and
it is easy to checkthat T(a*a) = Lr,s iarsi 2 ~ 0, with equality only if
a = 0. Weshall check later-in Proposition 12.10-that T is continuous for
the C*-norm, so it extends to a faithful state on A~. Bearing in mind that
A~ ""C(lf) ~cxl with unitary generators u = e 2 rricf> and v, we write elements
of lf~ as a =: Ls fsv 5 where fs = L:r ar 5 Ur is a Fourier expansion, so that

T(a) = aoo = J: fo(c/>) dc/>.


lt turns out that, when fJ is irrational, A~ is a simple C* -algebra, and
its tracial state T is unique. We postpone the proofs for a short while (see
Proposition 12.11 and Corollary 12.12 at the end of this section).
We now possess an isomorphism invariant that will allow us to classify
the C*-algebras A~, namely, the range of the trace Ton the set of all projec-
tors in A~. Since these C*-algebras are separable, these ranges are count-
able subsets of the interval [0, 1]. lt turns out that they give a complete
classification in the irrational case.
Of course, it is not obvious that A~ contains nontrivial projectors. When
fJ = 0, there are none; since AÖ "" C (lf 2 ), its only projectors are the constant
functions 0 and 1, because lf 2 is connected. On the other hand, for rational
nonintegral p I q, the algebra A~;q does have nontrivial projectors.
Exercise 12.3. Exhibita nontrivial projector in the algebra Ai 12 . <>

For irrational fJ, nontrivial projectors in A~ were first constructed by Rief-


fe! [390], building on a suggestion by Powersthat A~ contains selfadjoint el-
ements with disconnected spectra. Theseare now called the Powers-Rieffel
projectors. To construct such a projector, first consider any crossed pro-
duct of the form A = B ~a l; we can regard B as a subalgebra of A and cx
as its outer automorphism b ...... vbv- 1 , where v E Ais unitary and B, v
generate A. Rieffel's idea is to Iook for a projector of the form

p = hv- 1 + f + gv, where f,g, hEB. (12.9)


530 12. Tori

Now p* = vh* + f* + v- 1g* = oc(h*)v + f* + oc- 1 (g*)v- 1 , and


p2 = hoc- 1 (h)v- 2 + (hoc- 1 (j) +fh)v- 1 + (hoc- 1 (g) +f 2 +goc(h))
+ (f g + goc(j) )v + goc(g)v 2 ,
so that p 2 = p = p* irnplies that f is selfadjoint, g = oc(h)*, and
goc(g)=O, fg+goc(j)=g, gg*+oc- 1 (g*g)=j(1-j). (12.10)
Any solution (j,g) of (12.10) yields a projector in the crossed product
B ~(X 71., though such solutions rnay not be easy to find. Note, in passing,
that the last relation irnplies that 0 5 f 5 1 in B.
Proposition 12.4. Jf () ft 71., the algebra A~ contains a nontrivial projector p
lying in lr~.
Proof. We need to solve (12.9) in the case where B = C(lr), regarding its
rnernbers as periodic functions on ~ with period 1, where oc is given by
oc(g)(t) := g(t- 8). We can assurne that 0 < () 5 !.
because A~ ""A~+n for
any integer n and A~ "" Ai_ 0 . Then (12.10) sirnplifies to a systern of three
functional equations:
g(t) g(t- ()) = 0, (12.lla)
g(t) j(t- ()) = (1- j(t)) g(t), (12.llb)
lg(t) 12 + lg(t + ()) 12 = j(t) (1- j(t) ). (12.llc)
We clairn that this adrnits srnooth solutions fand g, whereby (12.9) defines
a projector p E lr~. Frorn (12.llb), j(t- ()) = 1- j(t) on the support of g,
and 0 5 j(t) 5 1 follows frorn (12.llc).
Suppose that we can arrange rnatters so that g(t) > 0 on (0, ()) while
g(t) = 0 on [(), 1]. Then (12.lla) is autornatic and (12.llc) irnplies that
j(t) = 0 or 1 for () 5 t 5 1- e. Also, (12.llb) shows that f is deterrnined
on the interval [1- e, 1] by its values on [0, ()], since
j(t + 1- ()) = j(t- ()) = 1- j(t) for 0 5 t 5 e. (12.12)
After these prelirninaries, we can solve (12.11). Let f decrease srnoothly
frorn 1 to 0 on the interval [0, ()], with allderivatives zero at both ends of
the interval; Iet j(t) := 0 on [(), 1- ()]; and let f be defined on the interval
[1- e, 1] by (12.12), having thus a srnoothincrease frorn 0 to 1 that rnatches
its decrease on [0, 8]. Then define g(t) := ~f(t)(l- j(t)) on [0, ()] and
g(t) := 0 on [(), 1] (a srnooth burnp function). The function h(t) := g(t + ())
satisfies h(t) = ~j(t)(1- j(t)) on [1- 8,1] and vanishes on [0,1- ()].
Therefore, p := hv- 1 + f + gv is the desired projector in lr~. D
Since there is considerable freedorn in drawing the graph of f on the
interval [0, ()], this projector p is of coursenot unique. But all such p are
hornotopic, so they define a unique dass [p] E Ko(A~).
12.2 Structure of NC tori and the Moyal approach 531

Corollary 12.5. IfO < () < 1, any Powers-Rieffel projector in A~ has trace
T(p) = e.
Proof. lt is clear from the construction of p that

T(p) =I: f(t) dt =I: f(t) dt +I: (1- f(t)) dt = e,


in the case that 0 < e : :; ~. Then T ( 1 - p) = 1 - e, and we can regard 1 - p as
a Powers-Rieffel projector in ALe. via the isomorphism A~ "" ALe induced
by v ..... v- 1 ; this effectively exchanges g with h and f with 1 - f in the
construction of Proposition 12.4. D

Theorem 12.6. lf e is an irrational real nurnber, the tracial state on A~


rnaps the set of all projectors in A~ onto (l + e7L.) n [0, 1].
Proof. Let u and v be unitary generators of A~ satisfying (12.6). For any
nonzero m E l, consider the closed subalgebra of A~ generated by um
and v. Since vum = e 2 rrimeumv, this subalgebra is isomorphic to A~e• and
the restriction of T to this subalgebra is the (unique) tracial state on A~e·
The Powers-Rieffel projector in A~e is then a projector Pm in A~ suchthat
T(Pm) = m()- l m() J, the fractional part of m8. (Recall that P-1 = 1- PI,
whose trace is 1- 8.) Now the numbers 0 = T(O), 1 = T(l) and { T(pm) :
m * 0} fill out (l + Bl) n [0, 1], so the rangeofT on projectors includes
this countable dense subset of [0, 1].
To see that no other numbers lie in this range, we use an embedding
result due to Pimsner and Voiculescu [374): there isanother C*-algebra Be
with a trace T', where T' maps projectors in Be into (l + Bl) n [0, 1], and
there is a unital embedding A~ ..... Be. That being the case, the restriction
of T' to A~ equals T by uniqueness, so the range of T on projectors is no
more than (l + Bl) n [0, 1].
The embedding is found by constructing two elements u, v E Be satis-
fying (12.6). This induces a morphism A~ -Be whose kernel is zero, since
A~ is simple.
The algebra Be is constructed from the continued fraction expansion of
the irrational number e. We assume, as we may, that () > 0. Then () =
limm-oo Pmlqm, where

e =: ro + - - -1 -1 - Pm := ro + ___1_ __
qm 1
ri+--- ri+---
1 1
rz+- ·.+-
rm

with all ri E ~. ri > 0 for j ~ 1, and Pm. qm E ~. (Take ro := l eJ, r1 :=


l (() - ro) -I J. etc.) The convergents Pm I qm obey the recurrence relations
Pm = YmPm-1 + Pm-2, (12.13)
532 12. Tori

and start with Po = ro, qo = 1, PI = r1ro + 1, ql = r1. Theserelations


can then be used to define an inductive system of finite-dimensional C*-
algebras { (Bm, cf>m) : m E l'ld, cf>m: Bm-1 - Bm} by

Bm := Mqm (C) Eil Mqm-1 (C),


cf>m(a Eil b) := Wm(a Eil··· Eil a Eil b)w~ Eil a,
Ym summands

where Wm E Mqm (C) are suitable unitary matrices.


In each Mqm (C) there are matrices Rm, Sm as in (12.7), for which
RmSm =Am SmRm with Am = exp(2rripm/qm);
let Vm := Rm Eil Rm-1 and Um:= Sm Eil Sm-1· Pimsner and Voiculescu [374]
showed how to choose the unitary matrices Wm so as to make the series
Lm llcf>m(Um-d- Um II and Lm llcf>m(Vm-d- Vmll converge; seealso [129,
VI.5]. Now define Be as the C*-inductive limit limm (Bm, cf>m) -recall Defini-
tion 3.11- so that there are morphisms jm: Bm - Be satisfying jm o cf>m =
jm-1· The sequences jm (Um) and jm (Vm) converge in norm to respective
unitaries u, v in Be satisfying vu = AUV where A = limm-oo Am = e 2 rrie;
they generate a subalgebra of Be isomorphic to A~.
The C*-algebra Be is an AF-algebra, being the inductive limit of finite-
dimensional C* -algebras. Moreover, since cf>m (1) = 1 in all cases, Be is
unital and jm ( 1) = 1 holds in every case. A tracial state T' of Be is deter-
mined by its restriction trm := T' o jm to each Bm, which must be of the
form
trm(a' Eil b') =Olm tra' + ßm trb' for a' E Mqm(C), b' E Mqm-1 (C).

The relations (12.13) show that


trm Um (a Eil b)) = (rmotm + ßm)tr a + Olm tr b,
so that YmOlm + ßm = Olm-1 and Olm = ßm-1 form ~ 2. Since Bl = Mr1 (C) Eil
C, the normalization T' ( 1) = 1 implies r 1a 1 + ß 1 = 1. In summary, T' is
determined by a sequence of nonnegative numbers {Olm} determined by
the recurrence formula
Olm+l = Olm-1- YmOlm (m ~ 2),
The starting value a 1 = () serves to define such a sequence and thereby
a tracial state T' on Be. Then every Olm lies in 7L + ()?L by induction. Now,
any projector in Be is unitarily equivalent to a projector in some Bm, by
Lemma 3.6, so the range ofT' on projectors is the union of the ranges of
the various trm. Any projector p in Bm is the sum of a projector in Mqm (C)
of rank Sm ::; qm and another in Mqm_ 1 (C) of rank Sm-1 ::; qm-1• for which
trm (p) = Olm Sm+ Olm+!Sm-1 E 7L + e?L. Thus the range ofT'' and a posteriori
that of T, is contained in 7L + e?L. o
12.2 Structure of NC tori and the Moyal approach 533

Exercise 12.4. Show that the initial condition a 1 = () is necessary and suf-
ficient to guarantee that C<m ~ 0 for all m. 0
Corollary 12.7. The C*-algebras A~, for irrational() with 0 < () < ~. are
mutually nonisomorphic.
Proof. If cp: A~ - A~. is a surjective isomorphism, and if T and T' are the
respective tracial states on A~ and A~., then T' o cp = T by uniqueness.
Therefore 71. + ()71. = 71. + ()'71.. If k, L, m, n are integers suchthat () = k + [()'
and ()' = m + ne, then k + Lm +Ln()= e, so Ln= 1 and L = n = ±1; since
m = ()' + () and 0 < ()' + () < 1, we get m = ()' - () E (- ~, ~), which forces
m = 0 and ()' = e. o
Of course, if () is irrational and ()' is rational, then A~ and A~. cannot be
isomorphic since only A~ has a trivial centre. The previous proof does not
apply directly to the case where both () and ()' are rational, but from the
presentation A~;q ""f(T 2 ,E) where Eisamatrix bundle of rank q 2 , it is
clear that A~;q "" A~. fq' only if q' = q. A more extensive analysis of such
bundles shows that isomorphism demands p' = p, too (251]. To sum up:
the 2-tori A~. for 0 5 e 5 ~. are mutually nonisomorphic.
Having determined the isomorphisms among the various 2-torus alge-
bras, we may inquire about automorphisms of a given A~. There are, of
course, the inner automorphisms; and there are also the hyperbolic auto-
morphisms, given on generators by

with a,b,c,d E 71.; since (ucvd)(uavb) = e 2rriO(ad-bcl(uavb)(ucvd), this


map extends to an automorphism of A~ if and only if ad - bc = 1. The
hyperbolic automorphisms can be thought of as a right action of the group
SL(2, 71.) on A~. This action preserves the smooth subalgebra 11}
~ Noncommutative 3-tori are pre-C*-algebras with three unitary genera-
tors; before introducing them, we adopt a few notational conventions. First,
we rewrite the basic commutation relation for 2-tori by setting u 1 := u,
u 2 := v, e1 2 := () and () 21 := -e. Then (12.6) becomes, for j, k = 1, 2:
UkUj = e2TTiOJkUjUk. (12.14)

The exponents determine a skewsymmetric matrix ( _0() ~). This prompts


us to change notation, writing () to denote this matrix.
Exercise 12.5. Let() be a real skewsymmetric n x n matrix with entries ()Jk·
Define G': 71.n X 71.n - T by
n
u(r,s) := e-rriB(r,s), where E>(r,s) := L YJ()JkSk. (12.15)
j,k=1
534 12. Tori

Show that u is a 2-cocycle for the group ln, suchthat u(r,s) = u(s,r)
and u(r, -r) = 1 for r,s E ln. 0
Any such skewsymmetric matrix () E Mn(~). n ;::: 2, determines a twisted
group C* -algebra A~ := C* (ln, u). To simplify the notation, we shall take
n = 3 in what follows, since the general case can be obtained by obvious
adaptations.
The algebraA~ is generated bythree unitaryelements u 1 , u 2 , u 3 , namely,
the delta functions in L 1 (~ 3 , u) corresponding to the standard generators
e 1 ,e 2 ,e 3 of l 3 . Replacing the integral in (12.5) by summation over l 3 , we
find that UJUk = u(ej,ek)8ej+ek• so that UkUJ = ä"(ej,ek) 2 uJuk; in other
words, the generators obey the commutation relations (12.14).
To prepare for Fourier expansions in such algebras, it is convenient [26,
470] to introduce a set of Weyl elements {ur : r E l 3 } of A~. Theseare
the images of { 8r : r E l 3 } under the inclusion L 1 (l 3 , u) c C* (l 3 , u).
Using (12.5), one can checkthat 8riei * 8r2e2 * 8r3e3 = nj<k u(rjej, Ykek) 8r
for any r E l 3 , and thus

(12.16)

These unitary elements obey the product rule

In particular, (ur)* = (u")- 1 =u-r for each r.


More generally, the Weyl elements of A~ are indexed by r E ln, with the
scalar factor in (12.16) replaced by exp{rri LJ<k rßJkrk}.
Proposition 12.8. The algebra A~ is isomorphic to an iterated crossed pro-
duct

Proof. We need only show that A~ "" B Xloc ~. where B "" AL is the C*-
subalgebra of A~ generated by u 1 and u2, and oc is the automorphism of B
given by conjugation with u3. Since U3UJU3 1 = e 2rri9j 3 u 1 for j = 1, 2, this
conjugation normalizes B. lt is clear that B and U3 tagether generate A~.
On the other hand, B Xloc l is generated by Band an extra unitary v such
that vbv- 1 = oc(b) for b E B. The obvious recipe 11(b) := b, 11(V) := u 3
determines a surjective morphism 11: B Xloc l - A~.
The circle group lr acts on both algebras by z · b := b, z · v := zv and
z · u 3 := zu 3 for b E B, z E lr. The fixed-point subalgebra of the action is B
in both cases, and 11 is lr -equivariant. Thus ker 11 is lr-invariant. If c E ker 11
is nonzero, then x := fr z · (c*c) dz is a nonzeropositive element of B Xloc l
that is fixed by the action of lr, so x E B; but then x = 11(X) = 0. Therefore
ker 11 = 0, so 11 is injective and provides the desired isomorphism. D
12.2 Structure of NC tori and the Moyal approach 535

Corollary 12.9. The algebra A~ is isomorphic to the universal C* -algebra


generated by three unitaries u 1, u2, u3, subject only to the relations (12.14)
forj,k=1,2,3. a
Exercise 12.6. lf 8 is a skewsymmetric 3 x 3 matrix with rational entries,
show that A~ is generated from C (lf 3 ) by iterated crossed products by finite
cyclic groups. Conclude that A~ "" r (E) for a suitable matrix bundle E - lf 3 ,
and that the centre of A~ is isomorphic to C (lf 3 ). 0

The C* -algebra A~ may be written as an n-fold iterated crossed product


by iZ:; it should be clear how to adapt the proof of Proposition 12.8 by in-
duction on n. The cases n = 1 and n = 2 yield, respectively, C XJC< il "" C(lf)
and C XJC< 1 il Xle<z il "" C(lf) XJC< 2 il ""A~ for a suitable cp.
~ The Weyl elements form the bases for Fourier expansions that give a
simple description of the elements of 3-tori.
Definition 12.7. The Schwartz space S(iZ: 3 ) is the vector space of rapidly
decreasing sequences indexed by iz: 3 , i.e., sequences { ar : r = (r 1 , r 2 , r 3 ) }
for which (1 + lrl 2 )k lar 12 is bounded for all k E ~.
The noncommutative torus lf~ is the unital dense subalgebra of A~ con-
sisting of (norm-convergent) sums a = Lrez3 ar ur with {ar} E S(iZ: 3 ),
analogously to (12.8). The Ue group lf 3 acts on A~ by z ·ur:= zrur, where
zr := z? z? z~3 E lf. The subalgebra of smooth elements for this action is
just lf~, which is thus a pre-C*-algebra.

Exercise 12.7. Show that {ar} E S(iZ: 3 ) if and only if the sequences {b~k)}
defined by b~kl := (1 + lrl 2 )kf 2 ar are square-summable for each k E N.
Conclude that S(iZ: 3 ) carries the topology of a Frechet space. 0

~ Each 3-torus comes equipped with a canonical normalized trace. The


following construction-which works for any n-shows, along the way, that
the trace on lf~ extends continuously to A~, as claimed earlier.
Proposition 12.10. On each C* -algebra A~ there exists a faithful tracial
state.

Proof. Elements of lf~ or, forthat matter, of L1 (iz: 3 ,a), can be written as
convergent series a = Lrez3 arur; on such elements we can define T(a) :=
a 0 . Clearly T(l) = 1 and

T(a*a) = T( L äsar<T(-s,r)ur-s) = L larl 2,


r,sez3 r

since a( -r, r) = 1; thus T is a normalized positive linear functional on


L 1 (iz: 3 , a), so it extends by continuity to a state on A~. Since T (a * a) > 0
for nonzero a E L 1 (iz: 3 , a), this state is faithful.
536 12. Tori

lf v = Ir Crur E lr~, then v* = Ir er u-r = Is C-sU 5 • Furthermore,


v*v = Im,r a(m, r) CrCr+m um, so v is unitary if and only if
(12.17)

lf a E lr~, then vav* = Ir,s,t CrasC-ta(r, s)a(r + s, t) ur+s+t, so unitarity


of v implies

T(vav*) = :2: Crascr+sa(r,s) = ao :2: lcr1 2 = T(a).


r,s r

By continuity, T(vav*) = T(a) for any unitary v E A~ and any a E A~. so


T is tracial. D

Exercise 12.8. The last step in the previous proof assumes that the unitary
elements in lr~ are densein the unitary group of A~. Prove this. 0

.,.. The trace T arises by averaging images of the Ue group action, already
introduced, on the twisted group C*-algebra. For that, it is immaterial
whether n = 2 or 3 (or more), so we consider the general case where 1rn
acts on Aß by z ·ur:= zr ur and the Weyl elements are given by the afore-
mentioned generalization of (12.16).
Recall that 1rn is just the dual group of 7Ln, consisting of all (continu-
ous) homomorphisms from 7Ln to lr, since each such homomorphism is
of the form r - zr for some z. More generally, on any twisted group C*-
algebra C* (G, a), any homomorphismp: G- 1r operates by multiplication
on U (G, a) and this operator extends continuously to C* (G, a).
Averaging the 1rn-action over Aß gives a linear operator E: Aß - Aß,
namely,
E(a) := Jln (z · a) dz
using the normalized Lebesgue measure on 1rn. E is called the conditional
expectation onto the fixed-point subalgebra of the torus action. It is clearly
idempotent, it satisfies E(1) = 1 and E(a*) = E(a)* for all a E Aß, and it
preserves positivity:

E(a*a) = Jln (z · a)*(z · a) dz ~ 0.


The term "conditional expectation" refers to the following property, shared
by projections that arise in probability theory:

E(E(a)bE(c)) = JJJ (zz 1 • a)(z · b)(zz2 · c) dz 1 dz 2 dz


= JJJ (z3 · a)(z · b)(z4 · c) dz3 dz4 dz = E(a)E(b)E(c),
12.2 Structure of NC tori and the Moyal approach 537

for a, b, c E A~. This identity implies the "generalized Schwarz inequality"


E(a)* E(a) ::5 E(a*a), since

0 ::5 E( (a- E(a))* (a- E(a))) = E(a*a)- E(a)* E(a).

Continuity of Eis then easy: if llall ::5 1, so that 1 - a*a ~ 0 in A~. then
1- E(a)* E(a) ~ 1- E(a*a) = E(1- a*a) ~ 0, and therefore IIE(a) II ::5 1;
we conclude that IIE(a) II ::5 llall for all a. When a = I r arur lies in lf~,
then
E(a) = I ar(J n zr dz)ur = ao 1 = T(a) 1.
YE?ln lr

By continuity, E(a) = T(a) 1 for all a E A~, and the range of Eis ( 1. This
shows, by the way, that the action of lfn on A~ is ergodic .
.,.. An important question is whether T is the only tracial state on A~. For
rational e, the answer is no. Consider, for instance, the 2-torus algebra
with the rational parameter p I q. To see that there are many tracial states
onA~fq• we use the isomorphismA~/q ""[(E) ofProposition 12.2. Consider
the map P: A~!q - C(lf 2 ) given by taking the matrix trace on the fibres of E
(normalized by the factor 1/q). This is a tracial operator: P(ab) = P(ba).
Then follow P by integration with respect to any positive Borel measure
on lf 2 of total mass 1; the composition is a tracial state.
On the other hand, we can expect less flexibility if the parameter matrix
e is irrational enough that the centre of the torus algebra is just ( 1. How
much irrationality is enough?
Definition 12.8. We say that the skewsymmetric real matrix e is quite ir-
rational if the lattice Ae generated by its columns is such that Ae + 71n is
densein ~n.
To put it another way, the representatives in the unit cell In of vectors
in Ae (after suitable integral translations) form a countable dense subset
of In. If Bk is the kth column of the matrix e, any vector in Ae is of the
form elsl + ... + BnSn for some s E 7ln. Exponentiation yields an element
Z(S) E lfn, where Z(S)j := exp(27TiLk (JjkSk); thus (} is quite irrational if
and only if { z(s): s E 71n} is densein lfn.
Exercise 12.9. Show that a 2 x 2 skewsymmetric real matrix e is quite ir-
rational if and only if 812 ft Q. o
Proposition 12.11. I(e is quite irrational, the tracial state T onA~ is unique.
Proof. Let T' be any tracial state on Aß. Then T' ( u 5 ur u -s) = T' (ur) for all
r, s E 71n. Since

for each r E it follows that u 5 au-s = z(s) · a for all a. Hence T'(z(s) ·
7ln,
a) = T'(a) for every s E 7ln. Foreach fixed a E A~. the set {z E lfn :
538 12. Tori

T' (z · a) = T' (a)} is closed in -u-n since z ..._ T' (z · a) is continuous, and is
dense since it contains every z(s) and 8 is quite irrational. Therefore,

T'(a) =I 11""
T'(z · a)dz = T'(E(a)) = T'(T(a) 1) = T(a).
Since a is arbitrary, we conclude that T' = T. D

Corollary 12.12. Jf 8 is quite irrational, the C* -algebra Aß is simple.


Proof. Let 1 be a closed ideal of A 0,and suppose that a E 1 is nonzero; then
a*a E 1 is positive and nonzero. From the previous proof, z(s) · a*a =
u 5 a*au-s E 1 for s E ?Ln, so that { z E -u-n: z · (a*a) E 1} is densein-u-n
since 8 is quite irrational; this set is closed in -u-n since 1 is a closed ideal
of A. Hence z · (a*a) E 1 for all z, and so

T(a*a) 1 = E(a*a) =I 11""


z · (a*a) dz

lies in 1. too. But T(a*a) > 0 since T is faithful; therefore, 1 E 1 and thus
1= Ae· D
In particular, if 8 E ~ is an irrational number, then the C* -algebras A~
are simple and carry a unique tracial state, as previously claimed .
.,.. The trace T can be regarded as an integral an the noncommutative torus
that reduces to the ordinary Lebesgue integral when 8 = 0. In fact, to any
a =Ir arur E T0, we can associate the periodic function

ä(c/>l, ... ,c/>n) := 2::arexp{2rri(rlcf>l + · · · +rncf>n)}.


r

This lies in C"'(Tn), since {ar} E S(7Ln). Then

T(a) = ao = f ä(c/>) dncf>. (12.18)


Jro.l]"
The element a E T 0 may be regarded as a quantization [130, 161] of the
function ä, where the Fourier-series exponentials e 2rrir·<t> are replaced by
the Weyl elements ur. In this context, (12.18) says that the expected value
of a in the tracial state T equals its classical expectation. In this way, the
noncommutative torus emerges as a leading example of a quantization of
the Moyal type: these and other quantizations are exhaustively explored by
Landsman in [308].
Indeed, in the even-dimensional (symplectic) case, the formula

(f Xn 1 2n
g)(u) = ( 2rr1'1) J.J j(u + v)g(v + w) exp( 2'ht s(v, w)) d 2nv d 2nw
plainly says that the Moyal product of two periodic functions is periodic,
with the same period. The matter is a bit less Straightforward than it seems,
12.3 Spin geometries on noncommutative tori 539

since one must first ascertain that the Moyal product is defined. This is
assured for the product of exponentials, these being elements of the Moyal
algebra :M.. With an adequate reinterpretation of the parameter 1't [484], one
immediately gets a correspondence between the product of elements a, b
in 1r0 and the Moyal product of ä, b. Moreover,

T(ab) = f ä(<j>)b(</>) dn</>.


Jro.IJ"
The Moyal product can be lifted to the C*-algebra level by means of tech-
niques that hark back to the classical paper [2 79]; that is to say, the torus al-
gebra is regarded as a subalgebra of the algebra generated by the Weyl ope-
rators of Section 3.5. Some subtle points of this identification are discussed
in [161]. For the odd-dimensional (presymplectic) case, the sequel [283]
to [279] is probably relevant.

12.3 Spin geometries on noncommutative tori


With the algebraic structure of the noncommutative tori now in place, we
turn to the construction of spin geometries on tori. For toral geometries,
the algebra will be one of the lr 0, and we must now find the other terms.
Since the noncommutative torus may be thought of as a quantization of
the commutative torus, we seek to identify a "Dirac K -cycle" over each 1r0
by suitably reinterpreting the formula [/) = -iyJ a1 that holds for e = 0
(when the ordinary torus carries the flat metric).
To start with, we need to exhibit the topological and differential struc-
tures of 1r0 as explicitly as possible. In principle, we already know how
to reconstruct any torus as a "noncommutative cell complex". Indeed, the
de Rham currents on an ordinary torus are given, via the HKRC theorem, by
the Hochschild cohomology classes of its algebra; moreover, Theorem 10.5
allows us to recover the de Rham homology of the space from the cyclic
cohomology of the algebra.
For example, the ordinary 2-torus lr 2 may be built up as a cell complex by
starting with a 0-cell (a point), attaching two 1-cells to form a pair of circles,
and then gluing a 2-cell (a plane sheet) to the two circles. These cells are
represented by four independent homology classes: one in Ho(lr 2 ), two in
H I(lr 2 ) and one in H 2 (lr 2 ). The Euler characteristic of the torus is then
computed as 1 - 2 + 1 = 0. The 3-torus 1r3 requires a 0-cell, three 1-cells,
three 2-cells attached independently to pairs of the 1-cells, and one 3-cell
to top it off.
The Hochschild cohomology of the 2-tori lr~ was computed by Connes
[86]; it depends strongly on the number-theoretic properties of the param-
eter (}. The cyclic and periodic cyclic cohomology modules are easier to
describe: it turnsout that HP 0 (lf~) = C2 and HP 1 (lf~) = C2 , indepen-
dently of e. The cyclic cohomology of tori for highernwas worked out by
540 12. Tori

Nest [357], who found that HP 0 (lr~) ""C 4 and HP 1 (lr~) ""C 4 , irrespective
of the skewsymmetric matrix () (so that the result may be read off from the
de Rham homology of the space 1r3 ). More generally, H pJ (lf~) "" c2"- 1 for
j = 0 and 1.
For our purposes, it is enough to exhibit a cyclic cocycle in each cohomol-
ogy dass. Since a cyclic 0-cocycle is a trace, unique up to multiples when ()
is quite irrational, HC 0 (lf~) is one-dimensional, with generator [T],\.
Cyclic 1-cocycles can be manufactured by using T and certain derivations
that are just the first-order partial derivatives acting on Fourier series.
Definition 12.9. The basic derivations <h, ... , 8n of the n-torus lf~ are de-
fined by

(12.19)
The property 8J(ab) = (8Ja) b + a (8Jb) is easily checked. It is clear that
T(8Ja) = 0 for all a.
These basic derivations commute; in fact, they span the abelian Lie alge-
bra of infinitesimal generators of the action of lfn on the C* -algebra A~,
and their common smooth domain is the pre-C*-algebra lf~.
lf 8 is any derivation of an involutive algebra into itself, then 8*: a -
(8(a*) )* is also a derivation. Thus, 8 is symmetric, i.e., 8* = 8, if and only
if (8a)* = 8(a*) for all a. Notice that the basic derivations (12.19) are
symmetric.
Definition 12.10. There are cyclic 1-cocycles 1/JI, ... , 1/Jn on lf~. defined by
1/lJ(ao,ad := T(ao8Jad. (12.20)
These are Hochschild cocycles because the 8 J are derivations:
biJJJ(ao,ai,az) = T(aoai8Ja2- ao8J(a1a2) + ao (8Jal)az) = 0.
and cyclicity also follows from the derivation property, since 1/JJ (a 1 , a 0 ) =
T(8J(aiao)- 8J(adao) = -1/JJ(ao,ad. They arenot cohomologous, since
1-coboundaries depend linearly on commutators.
Exercise 12.10. Show that the cocycles 1/J J are invariant under the action
of P, that is, 1/lJ(Z · ao, z · a!) = 1/lJ(ao, ad forzE lfn. 0
Note that (12.20) is a character of a one-dimensional cycle (lf~elf~. 8 J• T).
We can write down many other cyclic cocycles on lf~ that are characters of
higher-dimensional cycles over lr~. For K = { k 1 < · · · < kr} ~ {1, ... , n},
we define
1
1/JK(ao,ai, ... ,ar) := ( 2 rri)Y-l ~(-1)uT(ao8u(k!lal ... 8u(k,)ar),

where the sum ranges over all permutations of K. As in the previous exer-
cise, one checksthat each 1/JK is 1rn-invariant.
12.3 Spin geometries on noncommutative tori 541

Other cocycles arise by promoting lower-degree cyclic cocycles with the


periodicity operator S. For example,

ST(ao,ai,a2) oc T(aoala2),
S!JIJ(ao,ai, a2,a3) oc !JIJ(aoa1a2,a3)- !/lj(aoai, a2a3) + !/lj(ao, a1a2a3).

These do not contribute to the periodic cyclic cohomology HP"(lf~). which


is fully described by the dass es [ !J1 K]. When 0 = 0, we recover the de Rham
homology of lfn, in view of Theorem 10.6.
~ To each noncommutative torus there is a naturally associated Hilbert
space, given by the GNS construction associated with its state T.
Definition 12.11. Denote by J-fT the Hilbert space carrying the GNS repre-
sentation (1.22) associated to the tracial state Ton A~. Since T is faithful,
this is the completion of the vector space A~ in the Hilbert norm

!lall2 := ~T(a*a).
As in Section l.A, we shall write a E A~ as f! when regarded as a vector
in J-fT. The GNS representation of A~ on J-(T can then be expressedas

The vector ~T = 1 is a cyclic vector for the representation rrT (i.e., rrT(A~) ~T
is densein J-fT) and, since T is faithful, it is also a separating vector, that
is, f! = rrT(a)~T = 0 in J-fT implies a = 0 in A~.
Lemma 12.13. The mapping §..1 : f! - 8 1a, for a E lf~, extends to a closed
(unbounded) skewadjoint operator on :Fr;:
Proof. The derivation property, the symmetry and the vanishing of T o 81
combine to show that

T((8 1a)*b) = T(Dj(a*) b) = -T(a* (8jb))

for a, b E lf~. In other words, (§..1 (g_) I }!) = - (f! I §..1 (Q)) in J-fT. Writing the
closure of the operator as §..1 also, this means that §..j = -§..1. D
In the theory ofvon Neumann algebras (operator algebras that are closed
in the strong operator topology), algebras with a cyclic and separating vec-
tor have a special place. Suppose Ais a C*-algebra of operators acting on
a Hilbert space J-f, and that ~ E J-f is cyclic and separating. By von Neu-
mann's density theorem, the strong operator closure of A is just its bicom-
mutant A". The functional R ..... (~ I R~) defines a faithful state on A and
also on A". A fundamental theorem of Tomita [267, Thm. 9.2.9] states that
the commutant A' is isomorphic to the bicommutant via a certain antiuni-
tary intertwining operator ]o. This works in the following way: the densely
542 12. Tori

defined antilinear operator a~ ..... a * ~ turns out tobe closable, its closure S
has the polar decomposition 10 1::!,.11 2 where ~ isapositive selfadjoint opera-
tor and 1o is antiunitary with 15 = 1. The (antilinear) Tomita isomorphism
R ..... 1oR1o exchanges A' and A".
Consequently, the linear map ao ,_ 1oa* 1o is an isomorphism of the
opposite algebra A o with a strongly dense subalgebra of the commutant A'.
The positive operator ~ is the identity whenever a~ ..... a * ~ is isometric;
and in this case, 1o is just the obvious extension (by continuity) of this
map to an antilinear isometry of Jf. This happens whenever the state R .....
(~IR~} is tracial; in the case at hand, where the von Neumannalgebra is
rrT(A 0)'', it follows from

lla~TII~ = T(a*a) = T(aa*) = lla*~TII~.


Definition 12.12. The Tomita conjugation associated to the noncommu-
tative torus lf0is the antiunitary map 1o: J{T - J{T given by

1o(f!) := a*.

The representation rr;: (A 0) L(J-fT): ao ..... 1orrT(a*)1o acts through


0
-

right multiplication by elements of A 0 on J{T, that is,

rr;(a)f!. = 1orrT(a*)1ol!. = 1oa*b* = ba.


Exercise 12.11. Explain why (lf 0)o is isomorphic to lf~ 0 .

The Tomita conjugation implements the involution (10.83) on lf 0® (lf 0) 0


,

but does not determine a KRn-cycle over this involutive algebra, because
15 = + 1 irrespective of n mod 8. To get a conjugation with the correct
sign according to the table (9.45), we enlarge the Hilbert space J{T so as to
support a Clifford action of mtn.
Definition 12.13. Let J{ := J{T ® cN, where N = 2m (for n = 2m or 2m+ 1)
is the dimension of the irreducible spinor representation of G(+> (!Rtn ). Let
x := 1 ® c(y), where y is the chirality element of G(!Rtn)-acting trivially
when n is odd-and let C := 1o ® Co where Co is the antiunitary operator
on cN implementing the charge conjugation. In particular, for n = 2, the
Hilbert space is J{ = J{T Eil J{T and

c := (Ja -1o)
0 '
(12.21)

For n = 3, we use the same J{ and C, but x = 1.


The representation of the algebra on J{ is just the direct sum rr : = rrT ® 1
of N copies of the GNS representation. In matrix form, for n = 2 or 3,

rr(a) := (~ ~).
12.3 Spin geometries on noncommutative tori 543

Right multiplication by this matrix yields the corresponding representation


rr• of the opposite algebra T~ 9 • since Crr(b*)C- 1 = JorrT(b*)Jo ® 1 =
rr;(b•) ® 1.

~ We now seek to identify an operator Don J{ that can combine with the
data (T~, J{, C, x) to yield a spin geometry on the torus. We first examine
the case n = 2. The selfadjoint operator D must commute with C and
anticommute with x of (12.21). Therefore, it must be of the form

D = _. ( 0 Q12) (12.22)
1
o
-21 o ·
for suitable closed operators .Q 12 , .Q21 on JfT, satisfying .Q~ 1 -.Q 12 and
Jo.Q21 Jo = .Q 12 . The first-order condition [[D,rr(a)],Crr(b*)C- 1] = 0 for
a,b E T~. implies that the off-diagonal blocks of [D,rr(a)] are operators
on J{T that lie in the commutator rr;(T~ 9 )' which, by Tomita's theorem,
equals rrT(T~)''. Therefore, [D,rr(a)] E rr(T~)''.
The regularity and finiteness conditions now conspire to force [D, rr(a)]
to lie in rr (T~) itself. In this example, we find it useful to write the represen-
tation rr explicitly, even though it was omitted in (10.82). In consequence,

[D,rr(a)] = -i ([Qz~,a]
where o1 2, oz 1 arelinear endomorphisms ofT~; indeed, they are dertvations
ofT~. since rr(a)- [D, rr(a)] is a derivation.
Weshall also assume that the cyclic vector lis killed by eachoperator Qjko
i.e., .Q 12 (l) = .Q21 (l) = 0. It follows that .Q 12 (f!.) = [.Q 12 , a ](l) = 012 a for all
a E T~. and similarly for .Q21 . The selfadjointness of D now implies that
T((o1za)*b) = (o1za 112.) = -(gl oz1b} = -T(a*(oz1b)).

By setting a = 1 or b = 1, we find that T o o12 = T o oz 1 = 0, and thus

which implies that (o1za)* = 021 (a*) for all a. In other words, 021 = Of2 as
derivations.
It remains to classify the possible derivations oz 1 that can give rise to
geometries. An inner derivation a - [x, a] will not do, since D must be
unbounded in orderthat IDI- 1 be compact. It is known [49] that, if () is
irrational, any derivation ofT~ can be written uniquely as ocö1 + ßöz + ä
for some oc, ß E C, where ä is approximately inner (i.e., can be uniformly
approximated by inner derivations). Moreover, if the irrational number ()
satisfies the Diophantine condition that 11- e 2rrinO 1- 1 = O(nk) for some k
(those which do not form a nullset on the real line), then approximately
544 12. Tori

inner derivations are aetually inner [49, 84]. Thus, we are left with a21
tx<h + ß82.
This deeomposition of derivations depends on the invarianee of the basie
derivations 8 1,8 2 under the ergodie aetion of lr 2 on the torus lr~. Sinee,
by (12.19), they are clearly unbounded for the uniform norm, these are
outer derivations .
.,.. A similar analysis ean be used to find spin geometries on 3-tori. Starting
from the algebra lr~, with the same Hilbert spaee and eonjugation as in the
2-dimensional ease, but with x = 1 (no grading), we ean write

D = -i (Qu Q12) .
Q21 Q22

In dimension 3, D eommutes with C, so that

and thus .Q.22 = - ]oQu]o and Q12 = Jo.Q.2tfo.


The first-order property again shows that eaeh [.Q.1b a] lies in rrT (lr~)'',
and (10.82) then implies that [.Q.1ba] E rrT(lr~), for eaeh a. We eonclude,
as before, that
[D, rr(a)] = -i (aua a12a)'
a2la a22a

where the ajk are derivations of lr~ into itself.


Again we assume that the eyclie veetor llies in the kernel of eaeh .Q.1k; it
follows that .Q.1 k(~) = [.Q.Jb b](l) = aJkb. Selfadjointness of D yields

T((a1ka)*b) = (a1ka 1~> = -(f!.l ak1 b> = -T(a*(ak1 b)).


Setting a = 1 gives T 0 akj = 0, so the eondition

implies that (aJka)* = akJ(a*) for all a, that is, akJ = a;k. In partieular, au
and a22 are Symmetrie derivations.
Moreover, sinee

[Jo.Q.1 kfo,a*]~ = Jo(aJk(b*a))- a* Jo(aJkb*) = Jo(aJk(b*a)- (aJkb*)a)

= Jo(b* ajka) = (ajka)*b = (ajka)* ~.


we see that [Jo .Q.1kfo, a*] = (a1ka)*. Thus, the reality eondition yields a12 =
a~1 (redundantly) and also au = -a~2 ; eonsequently, a22 = -au.
Again, we use the deeomposition [49] of any derivation of lr~ into itself
a81 + ß82 + y83 + a, a
where is an inner derivation for generie 8, and
12.3 Spin geometries on noncommutative tori 545

ocl ß~ ;y E C. Leaving äasidel we end up with the following dass of operators:

D -
-
-.
t
(Qn
a (12.23)
-21
where s 11 s2 1S3 E su(2)1 i.e. 1they are traceless matrices satisfying Sj = -sj .
.,.. In higher dimensionsl one can proceed in the same way. For the alge-
bra lf~~ one obtains D = 2.}= 1 si §_j1 for certain skewhermitian matrices
si E MN((.). This operator is invariant under the action of lfn on L(Jf) 1

since z · ( 8ia) = 8i (z · a) in lf~. The following special case suggests itself


immediately andl as we shall seel defines a noncommutative spin geometry
on lf~ that deserves the name of Dirac geometry.
Definition 12.14. Let ;y 11... 1yn be the gamma matrices (8.20) 1implement-
ing an irreducible representation of n<+l(!Rn) on c.N. Weshall also write
yi rather than 1 ® yi for the corresponding operators on Jf = J{T ® c.N.
The Dirac operator on Jf is then defined as
(12.24)

The right hand side is essentially selfadjoint on the domain lf~ ® c.N and D
is its closurel a selfadjoint operator on Jf.
In particularl for n = 21

D- i ( O
- - §_1 + i§_2

and for n = 3 1

D = -i ( -83 -81 __
- s: i82)
3-
.
= -t(a1§.. 1 + a2§_2 + a3§_ 3 ).
§_1 + i§_2 u

In the commutative case e = 0 1 (12.24) corresponds to the Dirac operator


on the spinor bundle (dictated by the choice of Hilbert space Jf) for the
untwisted spin structure.
Proposition 12.14. I{D is the Dirac operator (12.24) 1 then IDI-n lies in L 1+
and is a measurable operator.
Proof. The square of the Dirac operator D = - i 2.i yi 8 i is the positive
operator D 2 = -(2.i 8J) ® 1N with N =2m. To diagonalize D 2 choose an
1 1

orthonormal basis { 'Ya : oc = 1 IN } of (.NI which yields an orthonormal


1 •••

basis of Jfby setting l/Jra :=ur ®'Ya. From the definition (12.19) ofthe 8j 1it
is immediate that D 2l/lra = 4rr 2 1rl 2l/lra- The subspace Jfr of Jf spanned
by { l/Jra : oc = 1 N} is invariant under D. The Operators Dl ID I and
1 ... 1

F := DIDI- 1 have the following restrictions to Jfr:

IDI = 2rr lrl~ (12.25)


546 12. Tori

(We define F := 1, by convention, on J-fo, the N-dimensional kernel of D.)


We can estimate the growth of the eigenvalues of IDI-s, for any s > 0,
just as in the first example of Section 7.5. From (7.78), mutatis mutandis,
we obtain

O'NR(IDI-s) = ~ L (2rrlrl)-s- 2m(2rr)-snn IR pn-s-Idp,


logNR logNR l:slri:,;R nlogR I

which vanishes for s > n, diverges for s < n, and is positive and finite for
s = n. Thus IDI-n lies in LI+ and is measurable, with
Tr+ IDI-n = 2mnn .
n(2rr)n
By adopting the normalization f T := n(2rr)n/(2mnn) Tr+ T for the non-
commutative integral, as in (9.39), we get f IDI-n = 1. D
Exercise 12.12. lf Dis the Operator given by (12.22) with a2I = IXÖI + ß8z,
show that n- 2 E LI+ if and only if the ratio ß!a is notareal number. o
Exercise 12.13. Show likewise thatthe operator IDI- 3 coming from (12.23)
lies in LI+ if and only if SI, s 2, s 3 are linearly independent in su(2). 0
Todefine a spin geometry on a torus lrö, we need a Hochschild n-cycle
to define the orientation dass. On an ordinary torus 1rn with angular Coor-
dinates </>I, ... , </>n, the normalized volume form is

d<f>I 1\ d</>z 1\ • • • 1\ d<f>n = (2rri)-nu;;I ... u2Iu1I du I 1\ duz 1\ • • • 1\ dun,

where u J := e 2rricb i are the unitary generators of lr~ = C"" (lrn). The corres-
ponding Hochschild n-chain in Cn(lrg) is obtained by applying the skew-
symmetrization operator (8.42):
in-m
C .
.= 1(2 ')n L"' ( - 1)(T Un-I •.. Uz-I UI-I ® Uu(l) ® Uu(2) ® ... ® Uu(n)·
n. 1Tt CTESn

with the same phase factor in-m as in Section 11.1. Using the commutativity
of 1rg, this can be rewritten as

(-i)m "' ( 1 )u -I -I -I
1(2 )n L - Uu(n) • • • Uu(2) Uu(l) ® Uu(l) ® Uu(2) ® • · · ® Uu(n),
n. 1T CTESn
(12.26)

which also makes sense as an n-chain in Cn(lrÖ).


lf .Jl is a unital algebra and .Jl ® .Jl o is regarded as an .Jl-bimodule via
a' (a ® bo )a" := a' aa" ® bo, then any n-chain ao ® ai ® • · · ®an E Cn (.Jl)
can be identified with (a 0 ® 1°) ® ai ®···®an and thereby regarded as
an n-chain in Cn (.Jl, .Jl ® .Jl o ). Thus, we may think of (12.26) as defining a
Hochschild n-chain over lrö with coefficients in lrö ® 11"~ 11 •
12.3 Spin geometries on noncommutative tori 547

Lemma 12.15. The n-chain (12.26) is a Hochschild cycle.


Proof. Write c' := n!(2rr)nim c; we must show that bc' = 0. Now bc' is a
sum of two kinds of terms. First, there are

( -1 )a { u;;~n) ... u;;b) ® Uu(2) ® · · · ® Uu(n)


+ ( -l)nu;;~n-1) ... u;;~l) ® Ua(l) ® · · · ® Ua(n-1)}.

coming from the first and last terms of the boundary of each summand
of (12.26). These two tensors can be matched by replacing a by A.a in
the second tensor, where A. is a cyclic permutation of {1, ... , n} with sign
( -1) n- 1 ; the second term in braces then gets a minus sign, and cancels
the first upon summing over all a. Next, the other summands in bc' yield
terms of the form
( - 1) 0"( - 1)k Ua(n)
-1 -1 -1
... Uu(2)Uu(l) ®Ua(l) ® · · · ®Ua(k)Uu(k+1) ® · · • ®Uu(n)·

The commutation relations (12.14) imply that

so these cancel in pairs upon skewsymmetrization. D

We can now represent this Hochschild cycle on .1-{, according to (10.85):

·-
1TD ( C ) .-
(-i)m ""< - 1)u Uu(n)
-1 -1 [D 'Ua(l) ] . . . [D 'Ua(n) ] .
n.1(2 rr )n L
a
... Uu(l)

Lemma 12.16. If D is the Dirac operator on "U" 0, then rrv(c) =X·


Proof. From the definition (12.24), [D, u 1 ] = -i 81(u 1 ) ® yJ = 2rr u 1 ® yi,
and thus
(-i)m
rrv(c) = --,- 2:<-1)0"1 ® ya(l) ... ya(n) = (-i)m1 ® y1 ... yn
n. a

= 1 ® c(y) =X· D

Wehave not computed the K-theory of the noncommutative tori, so we


shall not prove the Poincare duality property for these geometries. Never-
theless, a proof for 2-tori is given in [96], on the following basis. First of
all, since "U"~ is a pre-C*-algebra, its K-theory coincides with that of the C*-
algebra A~. lt is known [390] that Ko (A~) :::: 71. 2 , the two generators being [ 1]
and [p], where p is any Powers-Rieffel projector. Moreover, Kt(A~) :::: 71. 2
also, with generators [u] and [v] coming from the generating unitaries. The
pairings (10.87) on K0 (A~) and K1 (A~) are skewsymmetric integer-valued
forms that can be shown to satisfy ( [ p], [ 1]) = 1 and ( [v], [ u]) = 1, re-
spectively. The pairing on K.(A~) is the direct sum of these two and thus
is nondegenerate.
548 12. Tori

Putting it all together, we have constructed a Dirac geometry, of classical


dimension n, on each noncommutative n-torus lf~. When () = 0, we recover
the Dirac geometry on the ordinary torus with its untwisted spin structure .
.,. Exercises 12.12 and 12.13 suggest that the operators D of (12.22) and
(12.23) also give rise to geometries on 2- and 3-tori, for generic values of
their parameters. Since these operators are matrix combinations of the
basic derivations 8 1 , it can be expected that the same Hochschild cycle
c of (12.26) will provide the orientation dass for a corresponding spin geo-
metry, except perhaps for a normalization constant that we interpret as
the total volume of the geometry in question. We shall verify this explicitly
for the case n = 2.
Suppose, then that Dis given by (12.22) with o21 = oc8 1 + ß8 2 • For n = 2,
(12.26) becomes

i ( u -1 u -1 ® u1 ® u2- u -1 u -1 ® u 2 ® u 1) .
c := - Srr 2 2 1 1 2

-1 -1
( u2 u1 0 ) ( 0
0 u 2-1 u 1-1 ocu 1

and the representative of (2rri)- 2u1 1u2 1®U2 ®U1 is obtained by swapping
oc and ß. Therefore,

ä.ß _0 ocß ) = 4rr 2 (ocß-


- -
ocß) X·

This vanishes if and only if ocß is real, which is the case excluded by Exer-
cise 12.12.Noticethatrrv(c) = 4rr 2(oc{3-ä.ß)/8rr 2i = !J(oc[3).Inparticular,
(!J(oc[3))- 1c serves as orientation cycle for the spin geometry determined
by D. In other words, this spin geometry has total volume 1 I I!J(oc[3) 1.
Exercise 12.14. If s 1, s2, s 3 of (12.23) are linearly independent in su(2), and
if c is given by (12.26) with n = 3, find the total volume of the corresponding
spin geometry as the constant IK I such that rrv (K c) = 1. <>

The total volume may also be found with the noncommutative integral;
recall that our convention gives f IDI-n = 1 for the Dirac operators. Con-
sider now the case n = 2, and for convenience take 021 = 81 + T82 with
!JT > 0. In order to compute f v- 2 = 2rr Tr+ v- 2, we need to know the spec-
trum of v- 2. Now, D 2 = -(~ 1 + T~2 )(~ 1 + f~ 2 ) ® 12 on J-( = J-(T ® C2, with
eigenvalues 4rr 2 Im+ nTI 2 on the two-dimensional subspace with basis
1./Jroc with r = (m, n) E 71. 2 and ()( = 1, 2. Thus, v- 2 has a discrete spectrum
of eigenvalues (4rr 2)- 11m + nTI- 2, each with multiplicity two.
The Dixmier trace Tr+ v- 2 is the coefficient of logarithmic divergence
of the series z::n.n
Im+ nTI- 2, where the primed summation ranges over
12.4 Morita equivalence and crossed products 549

integer pairs (m, n) * (0, 0); we again ignore the kernel of D. Note that the
Eisenstein series
'
E2k(T) :=I ( 1 )2k'
m,n m+nT

converges absolutely for k > 1 but only conditionally for k = 1 [443), so the
series of absolute values certainly diverges. The following calculation [465)
shows that it diverges logarithmically, thereby confirming the two-dimen-
sionality of the geometry. By setting T =: 5 + it and pei<f> := m +in and
partially summing over lattice points for m 2 + n 2 5 R 2 , we get

f v-z = zrr_z_ lim _1_ I 1


4rr2 R-oo 2logR 1,;m2+n2,;R2 Im+ nTI 2
1 . 1 IR p dp Irr dcf>
= 2rr A~ logR 1 ()2 -rr (cos cf> + 5 sin cf>)2 + t2 sin2 cf>
1 Irr dcf> 1 ( 2rr) 1
= 2rr -rr (cos cf> + 5 sin cJ> )2 + t2 sin2 cf> = 2rr -t- = !JT ·
The last integral may be found by contour integration.
The result is inversely proportional to the area of the period parallel-
ogram of the elliptic curve ET with periods {1, T}, an analogy remarked
in [96). For the case 021 = oc81 + ß82, take T := ß/oc; the volume is

f D- 2 = loci- 2 I!JTI- 1 = I!J(ocß)l- 1.

12.4 Morita equivalence and crossed products


Wehaveseen that no two C*-algebras A~ are isomorphic if 0 5 e 5 How- i·
ever, those with rational parameter e = p/q are well-behaved subalgebras
of matrix algebras Mq (C (lr 2)), and one may immediately guess that they are
Morita-equivalent to C(lr 2), just like the full matrix algebras. This is indeed
the case, as we shall see, so allrational 2-torus algebras are lumped into
one Morita equivalence dass. We may then ask what Morita equivalences
hold when e is irrational.
Tobegin this enquiry, we need to construct C*-modules over the toral
C* -algebras. These algebras are crossed products, and much is known
about C*-modules over crossed products, due to the work of Rieffel. We
begin with an interesting result in harmonic analysis [388) that has direct
application to classifying the noncommutative tori.
Definition 12.15. Let G be an abelian Ue group. There are nonnegative inte-
gers p, q, r and a finite abelian group F suchthat G"" ~P x'l.q xlrr xF [439).
The Schwartz space S(G) consists of all the smooth functions ~: G - ([
that, tagether with all their derivatives, are rapidly decreasing in the vari-
ables corresponding to the factor ~P x 'l.q.
550 12. Tori

Suppose now that HandKare cocompact subgroups of G, that is, the


quotient spaces G/H and G/K are compact. Then H acts on C(G/K) by
translation, and likewise K acts on C(G/H). We use the coset notations
x := x+K andy := y+H, and write x-t := (x-t) +K, y-p := (y-p)+H
for the translates, where x,y E G, t EH and p E K. The actions are given
by ot.r{j)(x) := j(x- t) and ßp(g)(y) := g(y- p). This setup allows to
introduce two transformation-group C*-algebras,

A := C(G/K) XJ{)( H, B := C(G/H) >4ß K. (12.27)

Actually, commutativity of G is not essential; one can use any locally com-
pact group. In that case, K acts by left translations on the left coset space
G/H, while H acts by right translations on the right coset space K\G, and
one must take account of certain modular factors that convert left Haar
measures into right Haar measures. See [388] for the general case.
Proposition 12.17. The C*-algebras C(G/K) XJ{)( Hand C(G/H) >4ß Kare
Morita-equivalent.

Proof. Weshall indicate how 'E := S(G) carries the structure of a full pre-
C* B-A-bimodule that implements a Morita equivalence between the C*-
algebras A and B of (12.27).
The algebraic operations on A and B may be described in a less abstract
form than (12.1). A compactly supported function in Ce (H, C(G/ K)) is the
same thing as a continuous function f: H x (GI K) - I( with compact sup-
port; the operations (12.1) become

U1 * f2)(t,x) := Lfds,x) f2(t- s,x- s) ds,


f*(t,x) := f(-t,x- t),

where ds is a suitably normalized Haar measure on H. The corresponding


L1 -norm on this algebra is

II! III := f sup lf(t, x) I dt,


H xEG/K

and Ais its enveloping C*-algebra. The algebraBis analogously defined as


the completion of Cc(K x (G/H)) with variables (p,y).
On the bimodule 'E, the left action of B and the right action of .Jt are
given by

(g. ~)(y) := tg(p,y) ~(y- p) dp,


(~ · j)(x) := L~(x + s) j(s,x + s) ds.
12.4 Morita equivalence and crossed products 551

Indeed, by interchanging the order of integration, we get

(BI. (B2. ~))(y) := t tB1(p,y)B2(q,y- p) ~(y- p- q) dqdp


= ttBdp,y)B2(r-p,y-p)~(y-r)dpdr,
=((BI* B2). ~)(y),

*
and ((~ · fd · f2)(x) = (~ · (fi f2))(x) by a similar calculation.
Thus, 'Eisa B-A-bimodule. We introduce pairings on 'E by

(~I rJ)(t,.X) := t ~(X- q) rJ(X- q- t) dq,

{~I rJ}(p,y) := t ~(y + s) rJ(Y + s- p) ds.

Since x,y occurin theintegrals onlyin the combinations (x-q) and (y+s),
the first integral depends only on t and .X, while the second depends only
on p and y. The module properties (~I rJ) * f = (~I '7 · j) and B *{~I '7} =
{B · ~ I '7} are easily checked. The compatibility of the pairings follows from

[~ · ('71 ()](z) = t ~(z + s) (rJI ()(s,z + s) ds

= t t ~(z + s) rJ(z + s- q) ((z- q) dq ds

= t {~I rJ}(q, z) ((z- q) dq = [ {~ 117}. (](z), z E G.

Positivity of the pairing ( · I ·) depends on the following fact, proved


in [392]: the C*-algebra B contains an approximate unit Ba of the form
Ba= .Ld~ak I ~akl· For any ~ E 'E, this implies that

<~I~)= lim<Ba ·~I~)= lim:l:U~ak I ~ad ·~I~)


Cf. Cf. k

= limL(~ak · (~ak I~) I~)= limL(~ak I ~)*(~ak I~),


Cf. k Cf. k

so ( ~ I ~) is positive in A. Moreover, B = lima B *Ba = lima .LdB · ~ak I ~ak}


shows that 'Eis left-full. A similar argument, with A and B interchanged,
shows that {· I ·} is positive and that 'Eis right-full.
It follows that the ideal of B corresponding to End~0 ('E) is dense, so that
B ""End~ ('E). Theorem 4.26 now allows us to conclude that A ~ B. D

.,. Before applying this proposition directly, consider the closely related
case where K = {0} and Gis not compact (so K is not cocompact). Then
(12.27) reduces to A = C0 (G) ~a H and B := C(G/H); the proposition
remains valid and asserts that the Gelfand algebra of the coset space GI H
552 12. Tori

is Morita equivalent to the crossed product of Co ( G) by the action of H. This


foreshadow s the more general result that the appropriate noncommut ative
algebra of an orbit space is a crossed product by a group action [91, 11.7].
An important special case is the action of the group 71 on IR by transla-
tions, whose orbit space is the circle §1. This fits into our framework by
taking G = IR, H = 71 and K = {0}. We claim, then, that Co(IR) ~ 71 ~ C(§l ).
Rather than adapt Proposition 12.17 to this case, we give a direct construc-
tion of the equivalence B-A-bimodule.
Let A := C(§ 1 ) and B := C0 (1R) ~ 71, where 71 acts by e<n(X) := x- n. Let
'E be the Hilbert space of sequences, indexed by 71, of periodic functions
s ={Sn} on [0, 1], with norm

(12.28)

'Eisa right A-module under the obvious action [sg]n ( cJ>) := Sn (cJ>) g(cJ> ). A
commuting left action of C0 (1R) on 'Eis given by
[fs]n(cJ>) := f(cJ> + n) Sn(cJ>), f E Co(IR),

and 71 acts on 'E by "shifting the stack":

According to (12.4), this is a covariant representati on of (IR, 71, e<), and so


yields a left action of B on 'E.
To see this more clearly, we may write any element of the crossed product
as a sum f = Lm f m Um where f m E Co (IR). The product in B is then
f * h = ,2)J * h]nUn = L fmUm hkUk = L fm(Umhn-m Ui:t) Un,
n m,k m,n

or, alternatively,
[f * h]n(X) = Lfm(X)hn -m(X- m).
m

The left action of B on 'E is given by


[fsln(cJ>) = LUmUmS]n( cJ>) = Lfm<cJ> + n) Sn-m(cJ>).
m m

The B-module property follows from


[f(hs)]n(cJ>) = Lfm(cJ> + n) [hs]n-m(cJ>)
m

m,k
=LU* h]k(cJ> + n)sn-k(cJ>) = [(j * h)s]n(cJ>).
k
12.4 Morita equivalence and crossed products 553

As a right A-module, there is an obvious pairing on 'E:

(s I t)(cj>) := "'.sn(cf>) tn(cf>),


n

and it is obvious that (s I tg) = (s I t) g in A. A glance at (12.28) shows that


II (s I s) II = llls 111 2 • Moreover, 'Eis right-full. The left B-module pairing is less
evident:

{r I S}n(X) := Ym(X- m) Sm-n(X- m), where m := lxJ.

With cf> := x- m, the relation {fr I s} = f * {r I s} is seen from

{jr I S}n(X) = [fr]m(cf>) Sm-n(<f>) = "'.fk(cf> + m) Ym-k(cf>) Sm-n(<f>)


k
= "'.fk(x) Ym-k(<f>) Sm-n(cf>) = "'.fk(x) {r I S}n-k(X- k)
k k
= [f* {rls}]n(X).

The compatibility relation (4.10) for the pairings follows at once:

[ {r I s} t]m(cf>) =I. {r I s}n(<f> + m) tm-n(cf>)


n
=I. Ym(cf>) Sm-n(<f>) tm-n(cf>)
n
= Ym(<f>) (s I t)(cj>) = [r (s I t)]m(cf>).
Exercise 12.15. Show that 'E is left-full, and conclude that it is a Morita
equivalence bimodule for A and B. o
.,.. Proposition 12.17 also allows us to make the link to another important
dass of C*-algebras that codify the Kronecker flows on the 2-torus.
Definition 12.16. The Kronecker flow of angle 2rr0 on the 2-torus is the
following action of ~ on "U" 2 :

lf 0 is rational, this is effectively an action of "U" on -u- 2 • If 0 is irrational, the


orbits of this group action are homeomorphic to ~ and aredensein "U" 2 ; they
are leaves of a foliation. The crossed product F~ := C("U" 2 ) >4ß ~ is called the
Kronecker foliation algebra of the 2-torus.
Proposition 12.18. The Kronecker foliation algebra F~ and the toral alge-
bra A~ 116 are Morita-equivalent.
Proof. Consider the cylinder group G := "U"x ~. with two Ue subgroups, H :=
{1} x 7l and the helix K := { (e 2rris, Os) : s E ~ }. Clearly GI H "" "U" 2 , so that
C (GI H) >4ß K "" F~. On the other hand, GI K is identified to the circle "U" x {0}
554 12. Tori

=
that is transversal to K; it is clear that (e 2rrit, n) (e 2rri(t-n/el, 0) mod K, so
the subgroup H acts on T by translating by powers of e- 2rri/e, and therefore
C(G/K) :><Ia H:::: C("U") :><Ia 7L = A~ 11 e· 0

Proposition 12.19. The toral algebras A~ and ALe are Morita-equivalent.


Proof. In this case we take G = ~. H = 7L and K = 7LO; then both G/H and
G/K are circles, so the crossed products (12.27) are torus algebras. In the
case of A = C(G/K) :><Ia H, we may regard elements of C(G/ K) as periodic
functions on ~ with period e, and H acts by (anf)(t) := j ( t - n). Using
coordinates z := e2rrit!e, we identify j(t) with g(z) where g E C("U"), so
that (ang)(z) = g(e- 2 rrin!ez); we see that A ""ALe· On the other hand,
for B = C(G/H) :><lß K, elements of C(G/H) are periodic functions on ~
with period 1 and (ßmh)(s) = h(s- me), so that B :::: A~. The conclusion
follows from Proposition 12.17. 0

Corollary 12.20. The C* -algebras A~ and F~ are Morita-equivalent.


Proof. Wehave established that A~ ~ALe"" A~ 11 e ~ F~. 0

More Morita equivalences between the 2-torus algebras can be found by


combining Proposition 12.19 with the isomorphisms A~:::: ALn for n E 7L.
Let GL(2, 7L) denote the group of integral 2 x 2 matrices with determinants
± 1; it acts on ~ by linear fractional transformations,

( kl p) . e ·=. keW+q'
q
+P (12.29)

Exercise 12.16. Show that the two elements (~ ~) and (~ ~) generate


GL(2, 7L). Conclude that A~ ~ A~. whenever e and 0' belong to the same
orbit under GL(2, 7L). 0
These are in fact the only Morita equivalences between 2-tori. To see
that, we need two observations. The firstisthat any C* B-A bimodule 'E
implementing a Morita equivalence between unital C*-algebras A and Bis
a finitely generated projective A-module. For, since B :::: End~ ('E) by The-
orem 4.26, unitality of Bis the same as saying that h is A-compact (use
Lemma 4.24, with T = 1B). But now Proposition 3.9 implies that 'Eis finitely
generated and projective as a right A-module.
The second remark isthat a Morita equivalence between unital C*-alge-
bras gives a correspondence [390, Prop. 2.2) between their finite traces, i.e.,
positive multiples of tracial states.
Proposition 12.21. Ifthe unital C* -algebras A and Bare Morita equivalent
via a B-A bimodule 'E, there is a büection T - t between finite traces on A
and on B given by
t({s Ir})= T((r I s)). (12.30)
12.4 Morita equivalence and crossed products 555

Proof. Suppose that T is a finite trace on A. Since the bimodule 'E is full,
the formula (12.30) defines a unique positive linear functional on B. Since
B ""'End~('E) is unital, and the isomorphism takes {5 Ir} to l5)(rl, then
as in (3.6), we can find t1, ... , tn E 'E suchthat h = Lk=l {tk I tk}. Then
t ( 1s) = Lk= 1 T ( (tk I tk)), so t is a bounded linear functional on B.
The "row vector" t := (t1, ... , tn) is an element of ~ = 'E ®A nA, aB-
Mn (A)-bimodule that implements a Morita equivalence Mn (A) ~ B; also,
h = {tlt}. Now p := (tl t) is aprojectorinMn(A), since p 2 = (t I t(tl t)) =
(t I {t I t} t) = (t I t) = p; it is clear that p* = p. Define a morphism
cf>: B - Mn(A) by cf>(b) := (t I bt); notice that its multiplicativity comes
from c/>(bb') = (t I bb't) = (t I b{t I t}b't) = (t I bt(t I b't)) = c/>(b)c/>(b').
lt is injective since tcf>(b) = {t I t} bt = bt and so {tcf>(b) I t} = b. Also, if
x,y E ~. then

p(xly)p=(x(tlt) ly(tlt))=({xlt}t I {ylt}t)=c/>({tlx}{ylt}),


so that cJ> is an isomorphism from B onto pMn(A)p. In consequence, we
can use the trace T ® tr on Mn(A), i.e., T ® tr([aij]) := Lk=l T(akk), to
define a trace T' on B by T'(b) T ® tr(c/>(b)). We embed 'Ein~ by
5 - i:= (5,0, ... ,0). Then

T'({5 Ir})= T ® tr(t I {5 I r}t) = T ® tr(t I iCr.l t))


= T ® tr ( ( t I i) (.r. I t)) = T ® tr ( (.r. I t) ( t I i))
= T ® tr (.r. I t ( t I i)) = T ® tr (.r. I {t I t} i)
= T ® tr ( <.r. I i)) = T (( r I 5) ) ,
so that T' = t, making t a trace on B. By reading (12.30) from left to right,
every finite trace on B gives a unique finite trace on A in the same way. 0

Corollary 12.22. The 2-torus algebras A~ and A~. are Morita-equivalent if


and only if () and ()' lie in the same orbit of the action (12.29) of GL(2, 7L)
on ~-

Proof. The "if" part is Exercise 12.16. Notice that the orbit of 0 is precisely
the rational numbers Q, since if p, q are nonzero integers with greatest
common divisor 1, then we can find k, l E 7L suchthat kq - lp = 1, and
(12.29) shows that p I q lies in the orbit of 0.
For the "only if" part, let us suppose that A~ ~ A~. with 0, ()' both irra-
tional. By Proposition 12.11, their tracial states are unique. lf T is the tracial
state onA~, the formula (12.30) defines a finite trace t on A~., which must
be a positive multiple of the tracial state T' on A~., say, t = ,\ T' with
,\ > 0. Now, [p] - T ® tr(p), for p E Mn(A), defines an additive map
I: Ko (A~) - ~; by Theorem 12 .6, its range is the real subgroup 7L + 7L (). The
isomorphism cJ> of the previous proof preserves projectors and defines a
concrete isomorphism cJ> from Ko (A~.) to Ko (A~), which satisfies I o cJ> = f.
556 12. Tori

by construction. Therefore,
.\(l + U}') = f.(Ko(A~.)) = I(Ko(A~)) = l + UJ.
Thus, there are integers k, l, p, q suchthat k + UJ = ,\ and .\(p + q(}') = 1.
Therefore, (k + UJ)(p + qB') = 1; solving this equation for B' shows that
(}' E GL(2, l). D
Another characterization of Morita equivalence for 2-tori has been poin-
ted out by Landi, Lizzi and Szabo [307]. This isthat A~ !!. A~. if and only
if the AF-algebras Be and Be·, described in Section 12.2, are isomorphic.
Indeed, two irrational numbers e and B' are in the same GL(2, l)-orbit
under linear fractional transformations (12.29) if and only if their con-
tinued fraction expansions have a common tail [233, Thm. 175], that is,
rm(B) = rm+k(B') for some offset k and m large enough. The definition of
the AF-algebras as C*-inductive limits shows that this condition is in turn
equivalent to Be "" Be·.
For higher-dimensional tori, the classification up to Morita equivalence
is more involved. There is a generalization of the action (12.29) that pre-
serves Morita equivalence among noncommutative n-tori. Interpret k, l, p, q
in (12.29) as integral n x n matrices, and let SO(n, nll) be the discrete
group of such matrices with determinant + 1, satisfying kt l + zt k = 0 =
ptq + qtp and ktq + ztp = 1. Then if (} E Mn(~) is skewsymmetric, so
is (}' := (kB + p)(UJ + q)- 1 . Rieffel and Schwarz [399] have shown that, if
(UJ + q) is invertible for all elements of SO(n, nll), then A~ !!. A~..
~ A closely related enterprise is the construction of vector bundles over
tori, in other words, finitely generated projective modules for the algebras
lf~ or their C*-algebra completions; they all give rise to new spin geomet-
ries. By systematically examining the possible actions of lf~ on Schwartz
spaces of abelian groups, Rieffel constructed and classified such vector
bundles [394]; they areinstrumental in finding Morita equivalences. We re-
fer to [91, 117,465] for examples over 2-tori.
Morita-equivalent algebras have the same K-theory, by Theorem 4.30,
so we may ask how the K-groups of the tori lf~ differ as e varies. But
it turns out that their K-groups are all the same! A basic theorem about
crossed products by l, established by Pimsner and Voiculescu [373], is the
existence of a hexagonal exact sequence linking the K -groups of A and of
A X1a l; see [36] for this theorem and its proof. Since all tori are iterated
crossed products by l, from the obvious generalization of Proposition 12 .8,
Elliott [157] was able to compute the K-groups by induction on n. The up-
shot is that the K -groups do not depend on the parameter e and are isomor-
phic to the corresponding groups for commutative tori (see Exercise 3.20):
Ko(lf~) ""K0 (lfn) = l 2 "- 1 , KJ(lf~) ""K 1 (lfn) = l 2"- 1 •
Thus, tori provide a splendid illustration for the statement that Morita equi-
valence, while weaker than isomorphism, is stronger than K -equivalence.
13
Quantum Theory

In Part II, we expounded Shale-Stinespring theory and exhibited the ab-


stract charged fieldas the simplest example of a summable Fredholm mod-
ule, and thus a source of noncommutative geometry. Having come so far,
it would be a pity not to develop fermion quantum dynamics in external
fields, which comes straight from the spin representation. As on many oc-
casions in this book, we set to work here on an enterprise of translation; in
this case, to render in algebraic terms the quantization of wave equations
of the Dirac type. Most of the footwork has already been laid in Section 6.4.
The algebraic method has been araund for a lang while, since the pi-
oneering work of I. E. Segal at least, and always seemed a luxury for the
usual applications. In the path-breaking papers on noncommutative gauge
theory [103, 430], availability of the Moyal product has perhaps obscured
the need for this method; but the algebraic trend is unmistakable. The ta-
bles will be turned as quantum field theory becomes applied in a deeper
noncommutative manifold context. The algebraic reconstruction of spaces
able to sustain fermions, effected by Connes' spin theorem (Theorem 11.2),
provides an ideal ground for going further. This is why we chose to frame
our own study of the ultraviolet properties of quantum fields on noncom-
mutative geometries [4 70] fully within the Shale-Stinespring and Connes
theories. The unavailability, at present, of a noncommutative geometry de-
scription of Lorentzian spin manifolds will not hamper us.
The present excursion in physics includes a dabbling in renormalization
procedures, and thus serves as an extra motivation for the last chapter of
the book, wherein the Hopf algebra of rooted trees, related to renormaliza-
558 13. Quantum Theory

tion in quantum field theory, will be examined in the context of noncom-


mutative geometry.
We switch, in this chapter, to a style similar to that of physics books: no
formally stated "theorems" aretobe found. We are concerned with rigor-
ous results, nevertheless. Things are done from scratch; readers conversant
with Poincare invariance and the Dirac equation will be able to skip the first
sections, although they may be scanned anyway, for the notation. Familiar-
ity with distribution theory, as explained, for instance, in [170], is assumed
from the reader. Also, we shall continually refer back to Chapter 6, some-
times tacitly.

13.1 The Dirac equation and the neutrino paradigm


Let us denote by M4 the Minkowski space, i.e., (~ 4 .g) with g the Lorentz
bilinear form: if x = (x 0 , x), y = (y 0 , ji) are vectors in M4, their Lorentz
product is denoted by
(xy) := x 0 y 0 - x · ji = g(x,y).
Here x 0 = ct, where c is the speed of light in the vacuum, and t isatime
coordinate. Henceforth, units are taken so that c = h = 1.
The Poincare group 'P is by definition the group of transformations of
M4 leaving g invariant; it is then the semidirect product T4 XI 0 ( 1, 3), where
T4 denotes the subgroup of spacetime translations and 0 ( 1, 3) is called the
(full) Lorentz group; we write
(a,A) · (a',A') = (a+Aa',AA') for a,a' E T4 , A,A' E 0(1,3).
The Lorentz group has four connected components. Tobegin with, the de-
terminant of an element of 0(1,3) can be +1 or -1. Examples of Lorentz
transformations with negative determinant are

-1
-1

the space-reflection and time-reversal transformations. Now, although PT


has determinant + 1, it cannot be continuously joined to the identity. In
effect, if A = (A~), fromg(Ax,Ax) = g(x,x) for x = (1,Ö) E M4 we infer
(A8)2 = 1 + (AÖ)2 + (AÖ)2 + (AÖ)2,

implying that the sign of A8 must be constant on any component. There are
then three noteworthy subgroups of 0(1,3): the proper Lorentz transfor-
mations generated by 50 0 (1, 3) and T; the orthochronous Lorentz transfor-
mations generated by 500 (1, 3) and P; and SO(l, 3), the subgroup of or-
thochorous Lorentz transformations, generated by S0o(l,3) and PT. The
13.1 The Dirac equation and the neutrino paradigm 559

intersection of any two of these is the group of restricted Lorentz trans-


formations S0o(1, 3). That SOo(l, 3) = {A : detA = signA8 = 1} is a
consequence of its maximal compact subgroup being 50(3). For the same
reason, rr1 (S0o(1, 3)) = 7Lz. Therefore each of the mentioned groups has a
simply connected two-sheeted covering.
The noninvariance of particle physics under space-reflection and time-
reversal teaches us that the relevant group is SOo(1, 3) -or more pre-
cisely, its double cover Spin0 (1, 3). Before continuing, note that although
Spin(l, 3) ""Spin(3, 1), the respective double covers Pin(l, 3) and Pin(3, 1)
of 0 ( 1, 3) and 0 (3, 1) are nonisomorphic. This can be seen as follows: let
±A(P) and ±A'(P) cover P in Pin(l,3) and Pin(3,1), respectively. Then
{ ±1, ±A(P)} "" 7Lz x 7Lz, whereas { ±1, ±A' (P)} "" 71 4 . Much was made of
this curiosity in [71]. Typical elements of 50 0 (1, 3) are rotations (5.26b):

(t', x') = (t, cos 1/1 x + (1 - cos IJI)( ii · x) ii + sin 1/1 ii "x) =: RtJJn x;
and boosts L<:n• given by

( t', x' ) = ( t cosh ' + (ii · x) sinh ', x + (cosh ' - 1 )( ii · x) ii + t sin ' ii) .
Wehave adopted for both the "active" viewpoint. Any restricted Lorentz
transformation is the product of a rotation and a boost [416].
~ To deal with fermions, we naturally invoke Clifford algebra and Dirac
operator theory. Indeed, most of what was said in Chapters 5 and 9 applies
to spaces with Lorentz forms. The complex Clifford algebra G(~ 4 .g) ""
M 4 (() of course coincides with the Clifford algebra associated to an Eu-
clidean form. There is no particularly natural complex structure on M 4 ,
and so we omit the details of the Fock space construction and choose to
think of the spinor space S in abstract terms, just as ( 4 • The spin connec-
tion is almosttrivial in this flat-space context, as the Levi-Civita connection
adapted to g has null Christoffel symbols, and the Dirac operator, defined
as in (9.20) -except for a customary change of sign- is written in cartesian
coordinates:
.- ty J.la J.l -- t'( Yo~
l/) ·-. at + Y-~)
ax -· ·::~
-. tyr,
where yJJ = c(dxJJ), as usual, fulfil
(13.1)

and we have introduced Feynman's "slash" notation for general vectors:


1t := (yk). Of course, there is no Ionger any question of D being selfadjoint.
The (free) Dirac wave equation is then:

i~IJI = 0, or, more generally, i~IJI = mtjJ. (13.2)

Exactly as in Section 5.2 for the positive definite case, one defines the
spin group Spin(1, 3) c (1(~ 4 ). As advertised, we are mainly interested in
560 13. Quantum Theory

its neutral component Spin0 (1, 3). Equation (13.2) is not so natural from
the Standpoint of the spin representation. For then S is no langer irre-
ducible, rather it splits into two irreducible spaces s+, s- (again we forego
an explicit construction and think of them just as copies of C2 ). We shall
see that the representations of Spin0 ( 1, 3) on s+, s- both give isomor-
phisms of Spino(1, 3) with SL(2, ((), generalizing the well known isomor-
phism Spin(3) == SU(2) that we reviewed in Chapter 5. For that, let us
introduce
ä" := (1z, cf).
(Given any 4-vector x, we write fc for the image of x under spatial reflec-
tion.) Then
tr all ä" v = 2gllv,
form the known properties of the Pauli matrices. Define

X:= x 0 1 + x · cf = (ax)
for x E M4 ; this X is a general selfadjoint matrix, with x = ~ tr ä" X, and
detX = (xx). Consider now AXAt, for A E SL(2,((). It is immediate that
AXAt is still selfadjoint and detAXAt = (xx), therefore there is an ele-
mentA(A) in the (restricted) Lorentz group suchthat AXAt = (a[A(A)x]).
In fact,
A(A)~ = ~ tr (TII AavA t.
lt is also clear that A .... A(A) is a homomorphism; its surjectivity is proved
with the explicit formulae:

exp( -~1/l ii · cf) x exp( ~IJJ ii · cf) = R!Jin x,


exp( ~~ ii · cf) x exp( ~~ ii · cf) = Lz;n x,
the first of which is (5.26a), which can be obtained also by direct computa-
tion, using

(ay)(az) = (aw), where w = (y 0 z 0 + y. i,y 0 i + z 0 y + iy Ai).

To the decomposition of Lorentz transformations into products of rota-


tions and boosts, there corresponds simply the polar decomposition in
SL(2, ((). Finally, if A(AI) = A(Az ), then A1 = ±Az.
The group SL(2, (() has two inequivalent representations on C2 , namely
the self-representation, and also A .... (At) -l. The latter equivalent to the
conjugate representation, since

The two representations are respectively called D~· 0 and D 0 ·~, as they
belang to a family of irreducible representations DI·7-, with dimensions
13.1 The Dirac equation and the neutrino paradigm 561

(n + 1)(m + 1). lt is not difficult to checkthat

.,. Fields occurring in quantum field theory are thought tobe (quantized
versions of) spinor fields transforming according to the representations
D~·7- -or direct sums of those. The simplest type is the scalar field, which
transforms according to the trivial representation D 0 •0 of SL(2, ([):

But there are no scalar differential operators of the first order. Nature,
moreover, does not seem to like second-order equations for the description
of matter. lnstead, matter particles are fermions, described at the classical
level by first-order equations. The simplest possibilities will transform ac-
1 I
cording to D2· 0 and D 0 ·2; they correspond to the description of (massless)
neutrino fields. The (Weyl) neutrino wave equations are respectively given
by

i(ao)tJh = ( i :t - iä . a~) l/lt = o, (13.3a)

and

·(-::.)
t au lJlR = (·a ·- · OX
tot + ta a) l/JR = 0 , (13.3b)

where the unit 2 x 2 matrix multiplying the time derivative is understood.


They are indeed elegant. The subscripts L, R are explained as follows. The
Fourier-transformed equations are

(E ± ä · p)rPt.R = 0,
and the projection of the spin along the direction of motion ("helicity")
is negative for the equation originally with the minus sign, and positive
for the other. The neutrinos found so far in nature have negative helicity
(and the antineutrinos positive helicity). In contradistinction to (13.2), no
mass term is possible -it would break the invariance of the equations. This
could be seen as a shortcoming, all the more so since nowadays neutrinos
are believed to possess a (small) mass. However, mass should perhaps be
regarded as the result of an interaction, rather than an intrinsic property
of the particles. Indeed, as mentioned at the beginning of this Part IV, in
the Standard Model of particle physics, the fundamental objects are chiral
fermions, described in first approximation by equations such as (13.3), with
very different gauge couplings for the fields lJlR and l/JL, so that one can
conclude that "the left and right-handed fermion fields are fundamentally
independent entities, mixed to form massive fermians by some subsidiary
562 13. Quantum Theory

process" [3 71]. In other words, all particles are born massless, and fields are
seen to propagate inside the null cone as a result of repeated scattering -as
explained in [368, §5.12]. This is what Marshak [328] termed the "neutrino
paradigm" of modern particle physics.
On the other hand, in this book we only consider the mathematical de-
scription and problems of a more venerable, nonerural model quantum
theory, namely quantum electrodynamics (QED), for which the mass of
fermians is a given parameter; thus the Dirac operator is sufficient for our
purposes. We can getan explicit representation for it on s+ es- "" C 4 , com-
patible with the foregoing treatment, by supposing that t.JIR, t.JIL interact via
a mass term. This yields the system of equations
·- · - a) t.JIL =
(t.a- - ta mt.J!R ·- · - a) t.JIR
(t.a- + ta mt.J!L
at a.x · at a.x =
·
or

(13.4)

where t.J1 := ( ~~). That leads us to introduce the Weyl-Dirac gamma ma-
trices:
0 all) Jl = 0,1,2,3,
;yll := ( ä'll 0 '

that is,

1z)
0 ' . (
;yi ·= ai
0 }=1,2,3.

We see that (13.1) and (13.2) are satisfied. In this way, the operator QJ is
realized as
o
(w+ w-)
o ·
where QJ±: f"" (M4, S±) - ["" (M4, s+ ). It will came as no surprise that the
chirality element c(;y) := i;y 0 ;y 1 ;y 2 ;y 3 (here customarily denoted by ;y5 ) is
represented by 2 (~ -~J·
It remains to check that there exist implementors S (A) such that
S(A) (;ya) S(A)- 1 = ([A(A);y]a). (13.5)

Take

then
13.1 The Dirac equation and the neutrino paradigm 563

and (13.5)follows; the proofthat Spino(l, 3) ""SL(2, () is thereby achieved .


.. lt is always handy to have formulae for the solutions of the Dirac equa-
tion. Let p be a 4-vector satisfying (pp) = m 2 > 0 and p 0 > 0, in which
case p is called forward timelike, and let E(p) := ~lpl2 + m2, so that
p := (E(p), p). Note the identities

(ap)(ä"p) = (ap)(ap) = (pp) = m 2 ,


where agairr the unit 2 x 2 matrix is understood. In order to solve the equa-
tion (13.4), we write

lfl(x) =: u(p)e-iE(plt+ip·x, or lfl(x) =: v(p)e+iE<p>t-if-x.

Then
((~;) (~::/) u(p) = (,P- m)u(p) = 0,
whose solution is
u(p) = C(p)(~i)·
where ~ derrotes an arbitrary normalized 2-spinor. Then

The normalization is fixed by ut(p)u(p) = 2E(p)IC(p)l 2. We thus adopt


C(p) = (2E(p))- 1 12. Less tachygraphically,

1 ((.JE+m+.JE-mä·p/lpi)~)
u(p)= J4E <v'E+m-.JE-mä·p/lpi)~ ·
We now turn our attention to the v(p). Notice that

m
( (ä"p) (ap))
m v(p) = (,P + m)v(p) = 0,

whose solution is

v(p) = vh(-~~).
with 1J denoting another arbitrary normalized 2-spinor. More explicitly,

1 ( (.JE+m+.JE-mä·p/lpl)~)
v(p)=J4E -(.JE+m+.JE-mä·p/ipi)~ ·

Note also that


ut(p)v(p) = vt(p)u(p) = 0.
564 13. Quantum Theory

Introduce u(p) := ut(p)y0 = (f,t.j(fp ~t .JUP), and define v(p) anal-


ogously. Then

u(p)u(p) = ";, v(p)v(p) = - ";, u(p)v(p) = o.

There are closure relations also:

(ap)) = ,; + m
m 2E '

where {u 1 ( p), u 2 ( p)} is an orthonormal basis of 2-spinors, and

Other important closure relations, involving the ut(p) and vt(p) are:

where Ls= 1,2 ~s~st = (~ n has been used; and similarly,

(13.6b)

A general solution of the Dirac wave equation is written:

1/J(X, t) = (27T)-3/2 f (IJJ+s(15)us(p)e-ipx + I/J-sü1)vs(p)e+ipx) d3p,


(13.7)

where summation over the spin index is understood.

13.2 Propagatars
If we are given a differential operator L, acting on functions on spacetime,
and if we know the solution K (x, x') of the inhomogeneaus problern

LK(x, x') = 8 4 (x- x'),

(K (x, x') is called the "propagator" for L), then the solution of the inhomo-
geneaus problern with source p, namely,

Lcf>(x) = p(x),
13.2 Propagators 565

is in principle afforded by

cf>(x) = cf>h(x) +I K(x,x')p(x')d x', 4

where cf>h is a solution of the homogeneaus equation Lcf>(x) = 0, perhaps


further determined by boundary conditions. We propose to solve by the
propagator method the free Dirac equation with a "source",

Denote by S (x, x') the generic solution of the equation


(i~- m)S(x,x') = 8 4 (x- x').
Assuming the boundary conditions are translation invariant, as is the equa-
tion itself, then S(x,x') = S(x- x'). We now use the Fourier method.
Writing

the equation becomes


(J?- m)S(p) = 1,
that is,
S( ) = ,; + m
P p2 _ m2

in relativistic notation. This expression, however, is ill-defined over the


"mass hyperboloid" p 2 - m 2 = 0. That is to say, there is a choice of many
different distributions under that cryptic formula. Their inverse Fourier
transforms differ by solutions of the homogeneaus equation. The propa-
gator is given formally by

S(x- x')
1
:= (2rr)4
I eip(x-x')
p2- m2 (,; + m) d4p
. I
-(t~ + m)
eip(x-x')
m2- p2 (2rr)4
d4p

I .---· I
=

d3 p dpo /2rr
e-ipo<t-t')
= -(i~ + m) etp(x-x) (2rr)3 E(p)2- (p0)2

=: -(i~ + m)Ll(.X- .X', t- t'). (13.8)

One is then led to work out Ll, which actually has a physical significance of
his own: it is the propagator for a Klein-Gordon particle, that is, the generic
solution of
566 13. Quantum Theory

where w 2 denotes the positive operator -(a;ax) 2 + m 2 • Note the simple


poles at p 0 = + E ( p) of the integral on p 0 • There is thus a choice of contours
that will give rise to a zoo of different propagators. The contributions of
the residues at the poles are respectively given by

+-t-·_ e±iE(p)(t-t')
2E(p) .

We try to dictate the contours from physical considerations. Tobegin


with, we think of the retarded propagator ~ret· lts contour must run above
the poles, as then the "lemma of the big circle" in the residue calculus
ensures that the integral vanishes for t < t'. We obtain:

~
ret
(x _ x') = O(t _ t') . f
t
2(2rr)3
d3-
eip· (x-x'> (e-iE(p)(t-t'l _ eiE(p)(t-t'l) __!!_
E(p)'

where (} is the Heaviside function. Analogously, one can define the ad-
vanced propagator ~adv. by prescribing that the contour goes below the
poles. lt is clear that

~adv(X- x') = ~ret(X'- X).

Plowing right through the poles, we get the half-advanced, half-retarded


propagator:

K(x- x') := ~~ret(X- x') + ~~adv(X- x').

These real functions are more than sufficient for the needs of classical theo-
ries, like classical (retarded) electrodynarnics, Wheeler-Feynman electrody-
narnics, and linearized gravity, in the massless case. They can be rewritten
ala Feynman as
~ ( - - _,
ret,adv X X ,t
_ ')---
t - ( 2 rr)4
1
f e
ip·(x-x'ld3-
P f lpl2 + m2 _
e
-ip 0 (t-t') d
p
(pO ± iE)2 ·
0

That is, instead of playing with the contour, we imagine that the poles are
moved below or above the real line by an amount E, and then take the
lirnit E I 0. For quantum fields, however, we need propagators obtained
by making different runs araund the poles: they now carry the names of
founders of quantum field theory. The single most important one is the
Feynman propagator, which is obtained by going below the firstpole and
above the second. In relativistic notation,

~F(X- x') = (2rr)4


1 f m2-eip(x-x')
p2- iE d4p.

Indeed, the pole corresponding to positive frequencies

E = +~lpl 2 + m 2 - iE = E(p)- iE
13.2 Propagatars 567

lies below the realE -axis, while the pole corresponding to negative frequen-
cies

is located above it .
.,.. The difference between two propagators, loosely called a propagator
also, is a distributional solution of the homogeneaus wave equation. A very
important "propagator" of this type is the jordan-Pauli function

tl.jp(X- x') := tl.ret(X- x')- fiadv(X- x')


__1_
- (2rr)3
Je ip·(x-x'l sinE(p)(t- t') d 3
E(p)
_
p,

which is then the integral kernel of sin w(t- t') I w, a solution of the Klein-
Gordon equation characterized by ti1p(O,x) = 0 and attl.fp(t,x)it=O =
8(x). At this point, we pause to reflect that all propagators can in fact
be obtained from circuits turning (clockwise) separately araund the two
poles, giving rise, by definition, to the distributions ti +, ti-, called Wight-
man functions. Salutions of the homogeneaus wave equation, like the ti +,
ti- themselves, correspond to bounded closed contours araund the poles.
Clearly, tl.Jp = ti + + ti- with
_ 1· J e+iE(p)(t-t')±ip·(x-x')
d3 -
A
'-'
±(
X- X
')
= (2rr)3 2E(p) p. (13.9)

We alsosingleout the Schwinger function, corresponding to a figure-eight


circuit araund the poles:

tis(x-x') = 2(tl.p-K)(x-x') = (ß+ -tl.-)(x-x').

This is i times the integral kernel of cos wt I w, certainly also a solution of


the Klein-Gordon equation. Coming back to the "true" propagators, we find
that
tl.ret,adv = ±O(±t)tl.jp = ±O(±t)(ß+ + tl.-).

By the same token,

K =! signt tl.Jp =! signt (tl.+ + tl.-).


Also,

(13.10)

One can define as well the Dyson propagator:


568 13. Quantum Theory

corresponding to a contour that runs over the first pole and below the sec-
ond. A prodigious aspect of Feynman and Dyson propagators is the free-
dom to rotate the contour ("Wiek rotation") that allows computing them by
an integral on Euclidean space.
In fact, we need only compute one integral, as obviously .1- and .1 + are
complex conjugates. Let us do it, for training, for the massless case, in
which .1 is customarily replaced by D in the notation. Weshall use several
well-known distributional identities. First of all,

O(u) = -2
i Joo -,
e-ili.u
- . di\, (13.lla)
TT -oo + lE
t\

and, by derivation,

8(u) = O'(u) = -21 Joo e-ili.udi\, (13.llb)


TT -oo
that amounts to a trivial residue calculation. Next,

- 1- . = P..!.. ± irr8(a), (13.llc)


a + tE a
where P denotes, as usual, the principal part distribution; and finally, equa-
tion (7.51).
Now, abbreviating r := I.X- .X' I, we compute

v+(x- x') = . I
t
2(2rr) 3
eip·(x-x')e-ilpl(t-t') -t
d3 ~
lpl
= i foo lple-ilpl<t-n d1p1
2(2rr) 2 Jo
Jl
-1
eilplrcosB dO

= _1_ Joo 0( IPI) e-ilpl(t-t') (eilplr _ e-ilplr) dl PI


8rr2r -oo
= -- i - Joo dlpl Joo --.-e-ilpl(t-t')
e-ili.lpl
(eilplr _ e-ilplr) di\
16TT r -oo -oo i\ + lE
3

= __i_3_ Joo ~ Joo e-ilpl(li.+t-t'-r)- e-ilpl(li.+t-t'+r) dlßl


16TT Y -oo i\ + lE -oo
= _i_ Joo 8(,\- (t'- t + r))- 8(,\- (t'- t - r))~
8rr2r -oo ,\ + tE
_ _t _ · [P 1 _p 1 ]
- 8rr2r t' - t +r t' - t - r
+ - 1-[ö(t'- t + r)- 8(t'- t - r)].
Brrr
In summary, on using (7.51), we find that

+ '). 8((x-x') 2 )_ i
D-(x-x =signt 4 TT +P 4 2( ') 2 •
TT X-X
13.2 Propagators 569

lt follows that

DF(x-x')=8((x-x')2)_p i -i 1
4rr 4rr2(x-x')2 = 4rr2 (x-x')2-iE.

Also,

Drer(x- x') = B(t- t') 8((x- x')2) = B(t- t') 8(ix- .X' I- (t- t'))
2rr 4rrr '
on using (7.51) again, and

Dactv(X- x') = ß(t'- t) Ö( (X Z-TTx') 2 ).

The half-advanced, half-retarded propagator

8((x- x') ) 2
D( X - X ') = --'--'--:------'-....:...
4TT
is the real part of the Feynman propagator; historically, here is the link be-
tween the Wheeler-Feynman formulation of classical electrodynamics and
Feynman's formulation of quantum electrodynamics. For the Jordan-Pauli
propagator, we get
8((x- x') 2 )
DJp(x- x') = signt Zrr

All propagators considered are invariant under (restricted) Lorentz trans-


formations. One can actually write them in terms of Lorentz scalars only.
On using (13.8), one obtains explicit expressions for S, in the massless
case, in terms of the normal derivative of 8(x - x') with respect to the
light cone. We shall not bother to write them. Nor shall we attempt to give
more explicit formulas here for the t1 or massive S in x-space, since one
nearly always works in momentum space. The most important thing, in
this context, are the support properties: the leading singularities of the t1
on the lightcone are the same than that of the D, independently of mass.
Note that the behaviour of S on the lightcone will be even worse than that
of t1. We do recommend [384] for a thorough study of the distribution t1 +,
regarded as an oscillatory integral.
The propagators f1F, l1s, t1v, t1 + and f1- have in common that their imagi-
nary parts do not vanish, even outside the lightcone; whereas f1ret, t1actv. l1, "K
vanish at spacelike separations. It is plain that, for propagation in more ge-
neral manifolds than Minkowski space, the definition of the formerwill be
troublesome. We cannot refrain from quoting Steve Fulling [192] at length:
"t1ret. f1actv and hence l1, "K can be characterized by their support properties
in spacetime. The spectral representation plays no essential role in its def-
inition. All that is needed is the basic theorem ab out existence of solutions
of hyperbolic differential equations: the solutions are uniquely defined by
570 13. Quantum Theory

their Cauchy data and finite propagation speed. In contrast, ßs, ßp have
definitions that hinge on the decomposition of solutions into positive and
negative frequency parts [ ... 1there is no obvious, natural definition of the
functions of the second type if the dynamics is not independent of time.
This is a fundamental problern for the quantum theory of such systems
[ ... 1 we realize that God made ßret. ßadv. ß, A, but all the others have
been the work of man". Wise words, tobe heeded even in "safe" Minkowski
space.
As we shall see, the "the decomposition of solutions into positive and
negative frequency parts" is equivalent to the choice of the complex struc-
ture J for quantizing (the space of solutions of) the wave equation.

13.3 The classical Dyson expansion in QED


To deal with external field problems in quantum theory, two main strategies
recommend themselves. For fields that satisfy the Shale-Stinespring crite-
rion, one approach is to compute the classical scattering matrix Sc1, and
then to obtain the quantum scattering matrix as its rigorous quantization
p(Sc~). It is difficult to see how divergences could arise in this procedure.
The other is to apply to the linear problern the formal methods routinely
used in the nonlinear problern (which, after the development of the renor-
malization program, have had enormous practical success). This gives rise
to "bubble" divergences related to vacuum polarization effects. By compar-
ing methods, one can reach a better comprehension of the meaning of some
Feynman diagrams and their associated divergences, and one can hope to
guess what their proper definition in the noncommutative context should
be.
In both methods, the Dyson perturbative expansion plays a central role.
There is an intermediate strategy: to quantize first rigorously at the infin-
itesimallevel, using the infinitesimal spin representation, and then to try
to perform the Dyson expansion in Fock space. This turns out to be equi-
valent to a refinement of the first strategy, in which a precise definition
of the phase of the quantum scattering matrix is sought. Such is the path
followed in the next sections.
Consider a wave equation in Hilbert space

i :t c/>(t) = H(t)c/>(t), cJ>(s) = 1./J. (13.12)

We solve it, following more or less the treatment of [3841. Let us, for the mo-
ment, assume boundedness of H(t), as weil as differentiability or at least
strong continuity in t. Then (13.12) is equivalent to the integral equation

c/>(t) = 1.fJ- i J: H(u)cJ>(u) du


13.3 The classical Dyson expansion in QED 571

and thus has the solution cp(t) = U(t,s)tp, where the operators U(t,s) are
jointly strongly continuous in ( t, s) and satisfy

U(t,s) = 1- i J: H(u)U(u,s) du. (13.13)

Iteration of this equation yields the Dyson expansion:

The series converges absolutely in norm and can be regarded as a product


integral [141]:
U(t,s) = n t
e-iH(u)du
s

(that would become an ordinary exponential exp(f; -iH(u) du) if the ope-
rators H(u) at different times were to commute). The inverse U(s, t)
u- 1 ( t' s) satisfies the integral equation

U(s, t) = 1 + i f U(s, u)H(u) du.

lf the H(t) are selfadjoint, then the U(t,s) are unitary. Weshall assume
this henceforth. The two-parameter family U ( t, s) satisfies

U(t, t) = 1, U(t, s)U(s, r) = U(t, r). (13.15)

By definition, U is then called a unitary propagator.


Exercise 13.1. Check the unitary propagator properties (13.15) from the
Dyson expansion (13.14). <>
In quantum practice, H(t) is not bounded; instead,

H(t) = Do + V(t),

where Do is typically a time independent, "free" Hamiltonian Dirac operator


(defined precisely below); in particular, Do is unbounded selfadjoint, and
the interaction V ( t) is typically a gauge potential, fulfilling the previous
hypothesis of boundedness and strong continuity in t. Following Dyson,
we pass to the "interaction representation" by defining
V(t) := eiDotv(t)e-iDot.

Let fJ be the propagator that solves the problern

fJ(t,s) = 1- i J: V(u)fJ(u,s) du. (13.16)


572 13. Quantum Theory

Then
U(t,s) := e-iDotfJ(t,s)eiD0 s

solves weakly the original problem:

i :tU (t, s) = Wo + V ( t)) U (t, s).

We say "weakly", because it is unclear that U(t,s)I.Jl will be in the domain


of D0 even if 1./J is, and so D 0 U ( t, s) may not make sense. However, if
V(t)Dom(Do) c Dom(Do) and II Wo. V(t)]ll is a bounded function oft,
there is a strong solution. Those conditions are often met in practice .
.,. Before going on, we rewrite the iteration solution in a form more suited
to the needs of quantum field theory. The time-ordered product of several
time-dependent operators is denoted by the prefix T, which rearranges the
factors with arguments decreasing from left to right. For instance,

we can analogously use Heaviside functions to enforce this instruction in


more complicated cases. Since T[V(t!)V(tz) ... V(tn)] is necessarily sym-
metric in the time variables, we render the iteration solution of (13.16) as

(13.17)

The doubtful reader might be convinced by the following calculation:

The second term, by relabelling t1 - tz, can be rewritten as

For our scattering problem, by the standard argument of "in" and "out"
states [483], the classical (or "first-quantized") scattering matrixSei equals
fJ ( oo, - oo). Often V ( t) will be a multiplication operator with compact sup-
port in time, so then Sei = V(T, -S) for large enough T and S. For conver-
gence of the series, we need then to assume that Coo IIV(t)ll dt < oo .
.,. The next question is whether Sei will be implementable, in the sense of
Chapter 6. We must now declare what the complex structure defining the
13.3 The classical Dyson expansion in QED 573

vacuum state is. We naturally choose] := iF, with F = DoiDol- 1 =: P+-


P_, where P± project, respectively onto the states of positive and negative
frequency (we "fill up the Dirac sea"). In Chapter 6, we used the notation] =
i (E+ - E_), but the present notation for the projectors is more convenient
here.
It is relatively easy to see that, were we to assume that II [F, V ( t)] ll2 < oo
for all t and C'oo II[F,V(t)]ll2dt < oo, then the propagator would beim-
plementable at all times. This rarely happens in practice, but we need not
worry. The paper [177] established that the vacua associated to the values
of the propagator at different times have no relativistically invariant mean-
ing. Hence the "retreat" into scattering theory makes perfect sense. The only
thing we need care about at this momentisthat [F, Sci1 be Hilbert-Schmidt,
so that Sc1 will be quantizable. There have been many papers giving suffi-
cient conditions for that, in particular by Ruijsenaars in the late seventies.
Strang results in that respect are however quite recent [313]. The tech-
nique by Langmann and Mickelsson consists in modifying the propagator
by conjugation with a time-rlependent unitary family of operators T(t), de-
pending on the interaction locally in time, i.e., T(t) = 1 when V(t) = 0, so
that the scattering matrix is not altered. They show that, with the proper
choice of T(t), and provided there is an integer p for which V(t) and all its
temporal derivatives V(kl (t) up to order p fulfil the suggestive condition
that IDoi-PV(kl be Hilbert-Schmidt, then the new propagator

can be quantized. The V(kl (t) are required to be bounded and to have
bounded commutators with D0 , plus suitable falloff conditions when the
space is not compact, the V(kl being multiplication operators defined on IRl. n.
Weshall assume, then, that our external fieldissuch that Sc1 is quantiz-
able. One can submit that external fields not complying with this condition
are unphysical, in that they would be quickly destroyed by the backreaction
of the quantum field. We turn to our main concern, which is to quantize
or irnplement Sc1 according to the precepts of Chapter 6, and to examine
the resulting quantum expansion in different scattering Situations, with an
eye to the Feynman rules. For that, however, we need to make precise the
nature of V(t). Quantum electrodynamics describes the coupling of Dirac
field to the electromagnetic field; we next review the basic setting.
Introduce the hermitian matrices

oc- := y o-y = (a -ao_),


0

so that the Dirac equation is rewritten as


574 13. Quantum Theory

in "Hamiltonian" form. In view of the comment after Theorem 9.15, ap-


plied to ~ 3 • the operator Do is selfadjoint. We consider "minimal coupling"
Hamiltonians of the form

H(t) = Do + e(A 0 (t)- ä · .Ä(t)) =: Do + V(t),

where e denotes the electromagnetic coupling constant and A 0 (t), .Ä(t) are
real c-number functions. Note the form

V(t) = ey 0 ~.

with the usual notation A := (A 0 ,.Ä) for the electromagnetic vector poten-
tial and ~ := y~ A~, so that, in covariant form,

i~tfJ = m(/J + e~tfJ.

In momentum (i.e, Fourier-transformed) space, Do is just given by the


matrix multiplication operator

Do(k) = ä · k + ßm,

and V becomes a convolution operator:

[V(t)f](k) = (2rr)- 3 i 2 JV(t, k- k')j(k') d k'.


3 (13.18)

Let us also introduce V(k 0 , k) by

(13.19)

The matrix operator Do(k) is already diagonalized, since

D 0 (k)u(k) = E(k)u(k), and D 0 (k)v(k) = -E(k)v(k).

These equalities can be checked by direct computation, but it is perhaps


simpler to note that

- -
E(k) + Do(k) = ((ak)
m
m )
(äk)
-
and E(k)- -
Do(k) = ((äk)
-m

and then (E(k) + Do(k))u(k) = 2E(k)u(k) and (E(k)- Do(k))v(k)


2E(k)v (k), by (13.6), which also implies

P+(k) = E(k)±f!o(k) andso D 0 (k) =E(k)(P+(k)-P_(k)). (13.20)


- 2E(k)
13.4 The Rules 575

~ Now write Sc1 =: I~=O S:n for the iteration formula corresponding to
the QED scattering problem. Explicitly, in view of (13.17) and (13.18), S:n is
given as an integral kernel in momentum space by

S.n(i( k') ·= (-i)n


. ' . (2rr)3n/2
Joo-oo · · · Joo-oo I··· IeiDo(k)t V(t1 ' k- ki)tJ(t1- t2)
1

X e-iDo(kJl(tJ-t2)V(t2, k1- k2) ... (}(t1- tn)e-iDo(kn_Jl(tn-1-tn)


-
X V(tn, kn-1- k-, )e -iDo(k'Hn d 3 k1
-
... d 3 kn-1
-
dtn ... dt1.
We arenot entirely happy with this expression, however, because it does
not exhibit the Poincare invariance of the classical scattering matrix. So we
rewrite S:n in covariant form. Using the formula

(}(t)e-iDo(k)t = 2~ I Sret(k)yoe-ikot dko, (13.21)

and integrating with respect to the time variables, with the help of (13.19)
we obtain

S:n(k, k') = -i(2rr)- 2n+ 1enyO I··· I J.(k- kdSret(k1)J.(k1- k2) ...

X Sret(kn-1)J.(kn-1- k') d 4 k1 ... d 4 kn-1• (13.22)


where k 0 = ±E(k) and k' 0 = ±E(k') are understood.
That the retarded propagator appears in the explicit expression for the
classical scattering matrix is no surprise. It remains to prove (13.21). Recall
that
~+m
Sret(k) := k2- m2 + ikoE·
We invoke the formulae (13.lla) and (13.20) to compute:
-
(}(t)e-iDo(k)t = -
i J"" e-i(k 0 +P+E-P-E)t
dko
2rr -oo k 0 + iE
i I"" P+e-ikot p _e-ikot o
= 2rr -oo k 0 - E + iE + kO + E + iE dk ·
We used the old trick of Section 1.5, to exponentiate the projectors. Substi-
tutingbackP± = (E±Do(k))/2E,thisgives

_!_ Ioo ko + ä . k + ßm e-ikot dko


2rr -oo (k0)2 - E2 + ikOE '
which is precisely (13.21).

13.4 The Rules


To quantize the Dirac field now, one has only to apply the general theory
developed in Chapter 6, in the most direct way; the quantization machinery
576 13. Quantum Theory

and the "first quantized" input are largely independent of one another -
this is, of course, the beauty of the Shale-Stinespring theory. The Dirac field
is an instance of a charged field: there exists a given complex structure, the
charge Q := i, that commutes with all relevant operators. Thus we think
of V as the Hilbert space J{ of solutions of the free Dirac equation, regarded
as a real vector space with the symmetric bilinear form

We identified the complex structure defining the Fock vacuum as i times


the sign of the free Hamiltonian Dirac operator. Let {<f>k} and {I.Jlk} denote
orthonormal bases for J{+ and J{-, respectively. Among the quantized
currents, we shall need the quantum (ree Hamiltonian

i,j i,j

which is a positive operator, and the quantum interaction Hamiltonian

i,j i,j

(13.23)
i,j i,j

The outcome of the discussion in Section 6.4 is recalled here for ease
of reference. In QED the quantum charge is conserved, i.e., charged vacua
do not occur in our external field problem. The regular quantum scattering
matrix is of the form

J.i(Scl) =: S = (Om I Üout) :expdA(l):,

where (Om I :expdA(l): Üin) = 1 and

is the unique operator solving the equation I= S- 1- (S- 1)P_I.


To ascertain the pertinence of I, we show, following [213), that natural
physical quantities are expressed immediately in terms of this operator.
The simplest of these are the one-pair amplitudes. There are four of them.

1. For "electron" (i.e., particle) scattering from initial state <Pi to final
state <f>t:

Sfi := (bt(<f>J)Oin I sbt(<f>düin) = (Om I Dom) (<f>J I0 +h+)<f>i>·


(13.24a)
13.4 The Rules 577

2. For "positron" (i.e., antiparticle) scattering frominitial state «/Ji to final


state «1J 1:

Sfi := (dt(«<JJ)Oin I sdt(«JJdOin) = (Oin I Oour) («/Ji I (1-L-)«<JJ).


(13.24b)

3. For creation of an electron-positron pair in respective states c/J, (jJ:

4. For annihilation of an electron-positron pair in respective states cjJ, (jJ:

In each case, the first scalar product is in Fock space, whereas the last is
the original product in Jf. We verify them carefully in turn. For the first,

(bt(c/JJ)Om I sbt(c/JdO!n)
= (Oin I Üout) (Oin I b(c/Jj)ebtJ+_dt:ebti++b-dLdt:edL+bbt(cPi)Üin)
= (Üin I Üout) (Üin I ebtJ+_dtb(c/Jj):ebti++b-dLdt:edL+bbt(c/JdOin)
+ (Üin I Üout) (Oin I dtULcPJ)ebtJ._dt:ebth+b-dLdt:edL+bbt(cPi)Üin),

on using (6.48a); the term containing dt vanishes simply by taking it to the


left. Using now (6.48b), the remaining nontrivial commutation gives

(Om I Sb((l + IL)c/JJ)bt(c/JdOm)


= (Om I Oout) ((1 + I!+>cPJ I cPi) = (Oin I Oour) (cPJ I (1 + I++)cPi>·
For the second,

(dt(«<JJ)Om I sdt(«/Ji)Om)
= (Üin I Üout) (Üin I d(«/Jj)ebtJ+_dt:ebti++b-dL-dt:edL+bdt(«JJdOin)

= (Oin I Üout) (Üin I ebt[+_dtd(«/Jj):ebti++b-df__dt:edL+bdt(«JJdOin>


+ vanishing term.
The remaining nontrivial commutation gives

(Om I Sd((L_ -l)«JJJ)dt(«JJdOm) = (Om I Oour) («/Ji I (L_ -1)«/JJ).

For the third,

(bt(cjJ)dt(«JJ)Üin I S01n) = (Om I Üom) (Om I d(«jJ)b(cjJ)ebtJ+_dtOm)

= (üin I Oour> (üin I d(«JJ)ebt[+_dt dt uLc~J>Oin>

= (Om I Üout) (Om I ebtJ+_dtd(«jJ)dt(JLc/J)Om) + vanishing term,


578 13. Quantum Theory

by means of tricks already learned. We are left with

Finally,

(Üin I S bt ( </> )dt ( !Jl)Üin)


= (Oin I SdU-+<f>)dt(!Jl)Üin) + vanishing term = (Üin I Ü0 ut) (!Jll L+<f>).
So (13.24) has been completely verified. The conclusion isthat an explicit
expression for I is of foremost importance. Such an expression is necessar-
ily perturbative, and yields the Feynman rules (the Rules, for short) for QED
in external fields. To see what is involved here, let us rewrite the expansion
for Sei in the form

Sei- 1 = n~l S:n = J_')Ooo O(t!) dt1 + J:oo L00

00
O(t!)B(tl - tz)O(tz) dtz dt1

+ J:oo J:oo J:oo O(t!)B(tl- tz)O(tz)B(tz- t3)0(t3) dt3 dt2 dt1 + · · · .


Here, precisely,
O(t) = -i eiDotV(t)e-iDot
(2rr)3/2 ,

but the form of 0 will not matter for a while. Now, we cantend that

I= I I:n = f O(t!)dt1

f
n~l

+ O(ti)(P+B(ti- tz) -P_B(tz- tl))O(tz)dtzdtl

+ JO(td[P+B(ti- tz)- p_e(tz- ti)]O(t 2)

x [P+B(tz- t3)- p_ß(t3- tz)]O(t3) dt3 dtz dt1 + ···

lt is enough to check order by order. Firstly, S:I - 1:1 = ((Sei - 1 )P-Ih = 0.


At the next order,

At the third order, the integrand of 5:3 - 1:3 is found to be

8(t1- tz)O(t!)P+O(tz)P_Q(t3) + B(tz- t3)0(t!)P_O(tz)P+O(t3)


+ 8(t1- tz)B(tz- t3)0(t!)P_O(tz)P_O(t3)
(13.25a)
13.4 The Rules 5 79

whereas the integrand of ((Sei- 1)P_/h is

O(ti)ß(t1- t2)0(t2)P_O(t3)
+ O(ti)P_O(t2)[P+ß(t2- t3)- p_ß(t3- t2)]0(t3), (13.25b)

so, using the old trick that

(i.e., + 1 for the t i in decreasing order, -1 for increasing order, 0 otherwise),


the expressions (13.25a) and (13.25b) coincide. At all orders, equality fol-
lows from the identity [408]:

n-1
-I I O(t1)[E1PE 1 ]8(EI(t1- t2))0(t2) ... [En-1PEn-l]
i=1 E;=±

n-1
= I I ß(t1- t2) ... ß(ti-1- ti)ß(Ei+1 (ti+1 - ti+2)) ...
i=1 Ej=±,j*i
X ß(En-dtn-1- tn))0(ti)PE 1 ••• PE;- 1
X 0 (tdP- 0 (ti+dEi+1pEi+l . · · En-1PEn-1 ·

1t is enough now to pull together (13.10), which of course is equally valid


for Dirac propagators, and the covariant expansion (13.22) of Sc1 in terms
of the retarded propagator, to conclude that

l:nU(, k') = -i(2rr)- 2n+ 1eny 0 I··· I J,.(k- k1)SF(k1)J,.(k1- k2) ...

X SF(kn-dJ.. (kn-1 - k') d 4 k1 ... d 4 kn-1, (13.26)

(where again, k0 = ±E(k) and k' 0 = ±E(k') are understood). Presto! The
Feynman propagators have materialized in the quantum expansion!
.,. lt will not have escaped the reader's attention that (13.26) can be thought
of as an iteration solution for an equation that we symbolically write as

Furthermore, the successive terms of the expansion carry the correspond-


ing powers of the coupling constant e. As shown by Ruijsenaars some
time ago [409], from (13.26) one can rederive the Rules universally used
to compute transition amplitudes in quantum field theory. This will ac-
complish the translation process we referred to at the beginning of this
chapter. We anticipate, however, that there is no line-by-line translation,
580 13. Quantum Theory

due to the quirks of the formal theory. Before tackling that, there is the
unfinished business of giving an explicit expression for the preexponential
factor (Üin I Oout) of the quantum scattering matrix. We would also like to
express (the absolute value of) the vacuum persistence amplitude in the
form 00

n=O

Recall that I (Om I 0 0 m) 12 = det(J- s+_sL ). We can use the formula


T Ak)
det(1- A) = exp(Trlog(1- A)) = exp(- L --7<:---
oo
=: exp(- L
oo
~k ).
k=l k=l
(13.27)
for A selfadjoint of norm less than 1. Let Am be the eigenvalues of A,
counted with multiplicity; then the second equality is spelled out as fol-
lows:
det(l- A) = n
1- Am(A) = exp(2:Iog(1- Am));
m m
therefore, interchanging the order of summation:

yields (13.27). lt is worth recalling that for nonselfadjoint A, these manip-


ulationspass muster, too: this is proved in [437], for instance, essentially
by use of Hadamard's factorization theorem for entire functions. Note that
Udskii's theorem ~m Am (A) = Tr A, a far from trivial identity in that case,
is an infinitesimal version of (13.27).
In order to reorganize the series on the right of (13.27), consider the
expansion
L znbn := exp
oo (
- L -k-
zkuk)
oo
,
n=O k=l
where the right hand side is found by substituting zA for A in (13.27). lt is
clear that b 0 = 1. Differentiation of both sides gives

i
n=O
zn(n + 1)bn+l =- ( i
k=O
zkbk) ( i z ul+l),
l=O
1

which is a recurrence relation for the bn terms:


1 n
bn = - - L U!bn-l.
n l=I
so, for example,

b1 = -U!. b2 = !<uf- U2), b3 = -~(uf- 3UIU2 + 2u3) ...


13.4 The Rules 581

0
0
(13.28)

Notice that we have actually found that

JJ
TrA n-1
1 ( TrA 2 TrA
Tr(ArA) = n! det : :
Tr A" Tr A"- 1

Exercise 13.2. Suppose A is a 3 x 3 matrix. Give a formula for Tr A 4 in


terms of the traces of its three first powers. 0
This process is redolent of Fredholm theory, andin fact :expdA(l): is
essentially (though not quite) a Fredholm resolvent [291], as pointed out by
Neuman [358] and Salam and Matthews [4ll]long ago, in the framework
of the standard formal theory.
Recall that, in our case, CTn = Tr(S+_st_)". This can be reexpressed in
terms of the I matrix, as I (Om I 00 ur) 12 = det(J + J+_JL )- 1. Note that 0"1 is
the total probability for the production of an electron-positron pair. Similar
formulae obtain for I (Om I Oour) I =: .L:7=o bn, just replacing O"n by !O"n. We
wish to rewrite 00 00

I (Oin I Oout) I = 2:: bn = 2:: e"dn.


n=O n=O
1t is clear that
1 t (13.29)
do = 1, dz = - 2e2 Tr(J+-:1 J+-:1 ).

In fact, dzn+ 1 = 0 for all n ; :; 0. This is the farnaus theorem of Furry,


which comes straight from the existence of a real structure. In effect, this
"unbounded Lorentz K-cycle" has all the algebraic underpinnings of KRi-
cycle theory, and so, as is well known to physicists, there exists a con-
jugation operator C on the space of solutions of the Dirac equation. Be-
cause of the Lorentz signature, the properties of C correspond to the entry
6 = -2 = 1 - 3 of the table in Section 9.5. Therefore,

and
c~c- 1 = -~.
582 13. Quantum Theory

In fine,
CH(e)c- 1 = -H( -e),

from which follows

Using (6.45), we then get

I(Otn I Oour}l(-e) = det 112 (S __ s!_)(-e)


= det 112 (CS++sLc- 1)(e) = I(Otn I Oour}l(e).

Exercise 13.3. Work out the recurrence relation for the dn in terms of the
1+-:i and I!-:j· 0

~ We turn to the formal theory. The everyday method of calculation in


quantum field theory was introduced by Feynman on an intuitive basis.
Shortly afterwards, Dyson gave a derivation, which consisted in directly
applying the expansion to a formally defined quantum scattering matrix.
The place of the operator H ( t) is thereby taken by a quantum interaction
Hamiltonian -usually obtained from formally quantizing some Hamilton-
ian expression of Lagrangian field theory. The nontrivial part of such a
quantum Hamiltonian is usually of the form

I :Jf(x, n d 3 x,

where the Hamiltonian density :J{ is a scalar under Poincare group trans-
formations. Therefore,

S= 00-on I
2:- (
1- •••
I T[:Jf(xi) ... :Jf(xn)]d 4 Xn .•. d
4
X1. (13.30)
n=O n.

This is manifestly Poincare-invariant, except for the time-ordering. A suffi-


cient condition for Poincare invariance is
[:Jf(x),:Jf(x')] = 0,

for spacelike separations, (x- x' ) 2 > 0, a farnaus causality condition. Now,
ingeneral formal field theory, when evaluatingT[:Jf(xi):Jf(x2) ... :Jf(xn) ],
we can separate the terms with, say, m vertices whose field operators are
fully contracted among each other; these are the "vacuum fluctuation" dia-
grams. Denote them by Tvac. and by Text the ones that have "connection to
the outside" [215]. We can write
13.4 The Rules 583

and so the two factors decouple:

Actually, the first sum starts at k = 2.


To obtain an explicit expression for J{ (x), we introduce the formal (free)
fermion fields 'Y(x, t), prospectively related to their "rigorous" cousins of
Chapter 6 by

'Yt(f) := eiHot'Y(j)e-iHot =:I f(x)'Y(x, t) d3x,

for an element f in the spinor Hilbert space. lt is then said that 'Y is an
"operator-valued distribution". Explicitly,

where

j(x) = (2rr)-3/2 I U+s(p)us(p)eip·x + f-s(p)vs(p)e-ip·x) d3p;

compare (13.7). Then

for any orthonormal systems {f1 }, {Bk} on J{+ and J{_, respectively. By
the same token, the adjoint field is

'Yt (X, t) = (2rr) -3/2 I(b} (p)ust (p )eipx + ds (p)vst (p )e-iPX) d3 p.

The analogue of the commutation relations is clearly

{'Y(x, n. 'Yt <x', n} = 8(x- x').


More important is the Dirac adjoint 'Y (x) : = 'Yt (x) y 0 •
Exercise 13.4. Verify that

['Y(x), \f(x') ]+ = -iS1p(x- x'). 0

In our case, in view of V(t) = ey 0 J,., the Hamiltonian density is given by

Jf(x) = e :'Y(x)J,.'Y(x):. (13.31)


584 13. Quantum Theory

On integrating with respect to the space variables, one straightforwardly


obtains (13.23). Allfermion interactions of the Standard Model of particle
physics are of this type, with the classical gauge field A replaced by a gauge
boson field. (Remember that in noncommutative geometry, the Higgs field
is also regarded as a gauge field.) This expression is potentially trouble-
some, because in interacting field theories normal ordering sits uneasily
with spacetime locality; but it is rigorously defined for the external field
problem.
We do not soldier on with the derivation of the Rules for calculating the
S-matrix elements from (13.30) and (13.31). There is a wealth of texts which
cover this. The reader may consult [260,371,410,436,483,498], among the
relatively recent ones; also, reference [455] is original in deriving the Rules
for chiral fermians on their own. The Rules involve a graphical descrip-
tion of the factors, in terms of vertices and lines; they give the gist of the
so-called Wiek theorem, by which the time-ordered product of a set of ope-
rators is decomposed into the sum of the corresponding contracted normal
products. In particular, pairing of a field with a Dirac adjoint field in dif-
ferent interactions gives the Feynman propagator-pairings of fields and
adjoint fields in the same interaction are excluded by normal ordering.
Ruijsenaars' strategy to compare the "formal" Rules with the rigorous
expansion obtained in this chapter is now plain: the (externally) connected
diagrams should correspond to the second factor in the expression S =
(Om /Oout) :exp dA(!): and the vacuum diagrams should yield the first factor.
The first identification succeeds completely. In fact, one checks that

dAU:n) = (-ie)n J\f(xi)~(xi)SF(XI- X2)~(X2) ...


SF(Xn-1- Xn)~(Xn)'I'(Xn) d 4 nx,
coinciding with the expression associated with Text· It is also straightfor-
ward, though a bit laborious, to verify combinatorially that the expression
:expdA(J): coincides with
00
(-0 1
~-l!-
1
J
Text[.1f(xi) ... .1f(xL)]d 4 X1 ... d
4
XI.

when the Hamiltonian density is given by (13.31). So far, so good: for con-
nected diagrams the rules extracted from the rigorous procedure are the
ordinary Rules.

13.5 The quantum Dyson expansion


On the other hand, the following expression associated to Tvac immediately
spells trouble:

-~ Jtr{~(XI)SF{Xl- X2) .. . ~(Xk)SF(Xk- XI)) d 4Xk ... d 4 x1


13.5 The quantum Dyson expansion 585

diverges at least for k = 2, whatever the configuration of A may be. (This


corresponds to the bubble diagram

leading to vacuum polarization.) But we know that there are fields J,. for
which I(Om I Oour} I, andin particular a2, is perfectly finite!
Dyson's procedure, in the light of the present framework, can be inter-
preted as a (not one hundred percent successful) direct attempt to obtain
the spin representation operator S from the infinitesimal spin representa-
tion for V. In fact, Tvac[Jf(xd .. . Jf(xk)] formally corresponds to dets __
at order k, as Ruijsenaars [409] was able to show. While I(Om I 00 ur}l 2 =
det s __ s!_, we cannot wrtte (Om IOour} = det s__ and (Oin IOour}* = det s!_,
say, because both expressions are divergent; if we could, the spin represen-
tation would be (redefined so as to become) a homomorphism, which we
know it is not.
However, let us take a closer look. The Feynman integral above, for k = 2,
representing the firstnontrivial contribution to log(Om I Oour}, is certainly
divergent. But the usual unitarity argument (see [371, §7.3], for instance)
connecting the real part of (Oin I Oout} at second order to the probability of
pair creation holds in the formal theory. Therefore, despite appearances,
the real part of this integral must be finite! And so it is, as Feynman fully
knew back in the forties. The complete (residue) computation can be found
in [31, p. 508], and leadsback to our previous result. The crux of the matter
is that the integration actually takes place only on the intersection of two
mass hypersurfaces.
The imaginary part, contributing to the phase of the scattering matrix,
diverges in the Feynman procedure. One could say, so what? The phase
does not intervene in the computation of the transition amplitudes. But,
although we avoided the issue for a lang while, the phase of the quantum
scattering matrix does matter: the (quantum) current density is modified
with respect to the free field Situation by the vacuum polarization effect
(that bears on the photon propagator in the nonlinear theory), and the in-
teracting current density is found by functional derivation of S with respect
to the gauge potential, in which the phase intervenes. Thus, it is imperative
to find a convergent expression for the phase.
Since we possess up to a phase factor the quantized scattering opera-
tor, rigorously derived by means of the spin representation, it should be
possible for us to obtain the Dyson expansion in Fock space by integrating
carefully the cocycle associated to that projective representation. This we
proceed to do now, following our article [206]; the validity of the treatment
is not restricted to the charged field case, but applies to Shale-Stinespring
theory in full generality.
586 13. Quantum Theory

Assume, for the moment, that e'oo


II [F, V(t)] ll2 dt < oo, so the propagator
U(s, t) is implementable at all times; if need be, weshall assume that the
support of V ( t) is compact in spacetime. Let Veven. Vodd denote respectively
the parts of V that commute and anticommute with the complex structure
defining the Fock space (they were called V+, V_ in Chapter 6). Consider
Jl(U(s, t)). We recall here (6.30),

Jl{U)Jl(U') = c(U, U')Jl(UU'),

where the cocycle c is (near the identity of the implementable orthogonal


group) given by
c(U, U') = exp(iargdet 112 (1- Tu· Tu))= exp(iargdet 112 (p(j 1Puu·Pü1)).

In the present case,


Jl(U(s, t))Jl(U(t, r)) = c(s, t, r)Jl(U(s, r)), (13.32)
with an obvious notation, and c(t, t, r) = c(s, t, t) = c(s, t, s) = 1.
On the other hand, arguing exactly as with (6.33), we see that
0
i~
I ~
p(U(s, ~ ~
t)) = dA(V(t)) =: V(t).
uS s=t

We seek to redefine p(U(s, t)) by multiplying it by a phase factor eiB(s,t),


so that the new quantum propagator U(s, t) := eiB(s,t) Jl(U(s, t)) fulfils
U(t, t) = 1, U(s, t)U(t, r) = U(s, r). (13.33)
If we manage that, then 8( +oo, -oo) will have every right tobe called the
phase of the quantum scattering operator S. Let c(s, t, r) =: exp(i~(s, t, r) ).
Differentiation of (13.32) gives
~
V(t)Jl(U(t, r)) = i~
01 ~ 0 ~
c(s, t, r)Jl(U(t, r)) + i~p(U(t, r)).
uS s=t ut

Just as in Section 6.3, we redefine

U(s, t) := exp(i I: oo'A I.\=T ~('A, T, t) dT) p(U(s, t)),


verifying

i :s V(s, t) = V(t)U(s, t). (13.34)

Equation (13.34) is the quantized version of (13.13) and sports the same
kind of solution, the quantum Dyson expansion:

U(s,t) = 1+ 2: (-i)n is V(si) isl V(s2) · · · iSn-l V(sn)dsn ... ds2ds1.


00

n=l t t t
(13.35)
13.5 The quantum Dyson expansion 587

Let us pause a moment to argue this. The claim may look dubious, because
the V ( t) are now unbounded operators. They are, however, fairly tarne ones.
Let Em denote the projector on the states containing at most m particles.
Then, in view of (13.23), V(t)Em = Em V(t) is a bounded operator from Em
into Em+2· Introduce the norm, mentioned in Section 6.3,

IIIV(t)lll := IIVeven(t)ll + IIVoctct(t)112·

By continuity and uniform boundedness, IIIV(t) III ::5 a(s, r) holds for some
finite function a(s, r), when r ::5 t ::5 s. lt is not difficult to checkthat there
are constants Cm such that

IIV(t)Emll ::5 Cma(s,r).

Consult [218, 362] for precise analysis of these bounds, encompassing both
the boson and fermion cases. On the other hand, from the integral equa-
tion (13.16),
IIIU(s,r)lll ::51+ (s- r)a(s,r).
Putting both inequalities together, we get the estimate

with the result that the series in (13.35) indeed converges to a unitary ope-
rator for s- r small enough. Equation (13.33) then follows from (13.35)
and allow us to extend the validity of the last conclusion-and, in turn,
its own domain of validity. More detail on this is found in the important
paper [311].
Notionally, U( +oo, -oo) is the same object as (13.30). In view of (6.31), we
can compute the missing phase, with the result that

O(s, t) = -~ J: Tr[ 0°.\ I


A=T fu<A.Tl• Tu<T,tJ] dT.

To see this, use the standard identities

:t (argz(t)) = ~ :t (logz(t)) and :t (logdetA) = Tr(A- 1 ~~ ).


which imply

oo.\ IA=T ~(.\, T, t) = -~~ Tr( Tu(T,t) :.\ I.\=T fu(.\,T)).


The commutator form appears because det 112 (1 - TU(T,t) fu(.\,T)) has com-
plex conjugate det 112 (1- Tu(.\,T)TU(T,t)), as explained in Section 5.5. The
588 13. Quantum Theory

trace of this commutator is not zero because it is taken in Fock space-


recall (6.36), for the charged field case.
Note, before continuing, that

: IS=t (}(s, t) =
vS
0
-::. 1
VI\
IA=s ~(A,s,s) = o. (13.36)

In other words, there is no contribution from the coincidence points of


fu(A,T) and Tu(T,t>· This will prove tobe a crucial remark.
In the charged field case, 0°" IA=Tfu(A,Tl = iVoctd (T). Thus, we arrive at

(13.37)

which is our basic formula. This expression has a nice interpretation in

r
noncommutative geometry, as it is rewritten

(}(s, t) = ~ T(Voctct(A), T(A, t)) dA,

where T is Connes' Chern character of Section 8.4.


Note that (13.37) differs from zero only at second order in perturbation
theory. At that order,

e:2(s,t) =2
1 fst Tr(V+_(T)U-+:dT,t)-
~ ~ ~ ~
u+-:dT,t)V_+(T))dT,

with an obvious notation. Since U: 1(T, t) = -i ftT V(A) dA, we set out to
compute

e:2<s.t> = -~ J: Tr[v+-<T>(f V(A)dA)- ( f V(A)dA)v-+<T>] dT.


The total phase (}:2 := (}:2 ( oo, -oo) at this approximation is then

e:2 = -~ J~oo J~oo e<tl- t2HV+-<ti>V-+<t2)- v+-<t2)v_+(ti)) dt1 dt2.


It should be clear that, at the same order of approximation, this is precisely

-~!J(Üin I J~ooJ~oo T[V(tl)V(t2))dt1dt2 Üin).


where T denotes the time-ordered product. The technical condition that
e'ooII [F, V ( t)] ll2 dt < oo can be dropped here, as Langmann and Mickels-
son's trick, mentioned in Section 13.3, applies.
Next, in order to compare with the bread-and-butter of quantum field
theorists, we calculate the phase in QED. Working like in the derivation
of (13.31), the surviving integral is recast as

:3-z Je2 (}(tl- t2) tr[..t\(xi)s-(xi- x2).-t\.(x2)S+(x2- x!)


- 4\ (XI)S+ (XI - X2 )4\. (X2 )S- (X2 - XI)) d 4X! d 4X2.
13.5 The quantum Dyson expansion 589

The first thing to remark is that, since


s- (x)yv s+ (-X) -s+ (x)yv s- (-X) = Sjp(X)yv s+ (-X) -s+ (x)yv Sjp( -X),

and S1p has support inside the lightcone, then the integrand has support
inside the lightcone. That allows one to substitute for 8(t 1 - t2) the more
covariant expression 8( (v (x1 -x2))) =: x<x1 -x2 ), where v is an arbitrary
timelike vector, which can thus be varied at will. Let
pllv(x) := tr[yllS-(x)yvs+(-x)- yllS+(x)yvs-(-x)],
pv (x) := X(X)Filv (x).

I
Then
e= e22 (2rr) 2!J Pll (k)Aj.l(k)Av(k) d 4 k,

where we take into account that A( -k) = A(k), because A(x) is real. For-
mally, pv = (2rr)- 112x * pllv in momentum space, with * denoting ordi-
nary convolution; then, in view of (13.lla),

k - -i8(k)
X( ) - .J2Ti(k0- iE)'
(13.38)

when choosing a framein which k = (k 0 , Ö). As mentioned in Section 8.2,


this gives a dispersion relation for Fllv.
We said "formally" because it is time to declare that xFilv does not ex-
ist: because of the singularities on the lightcone mentioned at the end of
Section 13.2, multiplication of the Heaviside function by the product of
two S-distributions is undefined. Instead, since x<x)Filv (x) makes sense
for x * 0, one defines xFilv as some extension or regularization [165] of
the latter distribution. Distinct regularizations of this quantity will be seen
to differ by linear combinations of the delta function at the origin and its
derivatives up to order two, i.e., by polynormals in k of degree at most two
in momentum space. Regularization of distributions takes the place here
of the standard renormalization prescriptions.
Let us see what pllv (k) is, first. We look at the Fourier transform of
tr[yllS+(x)yvs-(x)]. Using (7.51), (13.8), and (13.9), this is expressed by

- ( 2 ~) 4 I tr[yll(p + m)yv(q- m)]e(p 0 )8(q 0 )8(p 2 - m 2) (13.39)


1
x 8(q2- m2)84(k- p- q) d4q d4p =:- (2rr)4 yvll(k).

Moreover,
tr[(p + m)yll(q _ m)yv] = 4 (pllqv + qllpv _ ((pq) _ m2)gllv).

With that, one integral in (13.39) is immediately disposed of with the help
of the 8 4-function. The other integral is easily done, with the help of the
590 13. Quantum Theory

remaining 8-functions, again by choosing a framein which k = (k0,Ö) so


that FJ.Iv can be regarded as a function of only one variable; this is found
in many books [260,413]:
pv (k) = 2; ( k~~v _ gJ.IV)

X [k 2(1 + y(k 2))(1- 2y(k 2)) 112ß(1- 2y(k 2))ß(k 0)],


where y(k 2) :=2m 2fkz.
Wehave been led formally to compute x * FJJv, where FJJV(k) equals
1 (kJ.IkY -gJ.lv)[k 2 (1+y(k 2))(1-2y(k 2))11 2B(l-2y(k 2 ))sign(k 0)]
6(2rr)3 k2 '
that indeed behaves as a polynomial of degree two at high momentum
transfer. lt is clear from (13.llc) and (13.38) that
~FJ.Iv(k) = iFJ.Iv(k).

The correct (unique) recipe to regularize the imaginary part of x FJJv is *


selected by prescribing that the result FJJv vanishes, together with deriva-
tives up to order two, at zero momentum. This kills the delta function at
the origin and its derivatives in configuration space, which otherwise would
give a nonzero contribution to a ;as Is=tB(s, t), contradicting (13.36).
The prescription that FJ.Iv have a zero of the third (indeed, fourth) order at
k = 0 is also natural on physical grounds, from the following strong heuris-
tic argument. On invoking the Maxwell equations (in the Lorentz gauge)
All (k) = JJ.I (k) 1k 2 to conjure up the source J of the classical field, one gets

e; (2rr) 2 ~ f pv(k)AJ.I(k)Av(k)d k - 2~~ f(j(k)](k))G(k)d k,


4 = 4

where
G(k) := : 2 (1 + y(k 2))(1- 2y(k 2 )) 112ß(1- 2y(k 2)) sign(k 0).
The continuity equation (J(k)k) = 0 has also been employed to simplify the
result. This simple expression exhibits only gauge-invariant variables. Now,
(classical) gauge transformations are not implementable in 1 + 3 dimensions
(see the discussion in the next section); and in the linear theory, this is the
only source of divergence difficulties. We expect, then, tobe able to express
the phase in a similar way:

e = - 2~~ J(J(k)J(k) )H (k) d 4 k,


where H(k) is regular. On using the subtracted dispersion relation (at the
origin) between the real and imaginary parts of FJJv:
!JFJ.lY(kO) = !JFJ.IY(O) + !)j:J.IV(O)kO + i!J.fJ.lY(O)(k0)2
(k0)3 Joo ~pv(1;)
- ----rr- p -oo ~3 (~- kO) d~'
13.5 The quantum Dyson expansion 591

plus the advertised prescription, and taking into account that G is odd, we
are finally led to a relation between G and H of the simple form
H(k) = .!_pf"" (1 + y(.\))(1- 2y(.\))112 d.\.
TT J4m2 .\(.\ - k2)

The restriction to timelike k is removed by analytical continuation.


Exercise 13.5. Express() in terms of other gauge-invariant variables, to wit,
(the Fourier transformed) electric and magnetic fields. 0

The above integral for H(k) can also be easily carried out. Making the
change of variable.\ =:4m 21(1 - v 2), we get
3 fl v2 - v4 13
H(k) = rrk2 Jo v2 -1 + 4m21k2 dv.
One then considers integrals of the form

·- il
In.- vndv
0 V
2
- OC
,

for n = 2,4, .... Clearly, In= 1l(n- 1) + ocin-2· Therefore, they are all
referred to

1 {~oc- 1 ' 2 log((fo -l)l(fo + 1)), if oc > 1,


Io =
Jof -v 2dv- oc = ~oc- 1 ' 2 log((1- fo)l(1 +.ja)), ifO < oc < 1,
(-oc)- 112 arctan(11.,J=a), if oc < 0.

In the second case, the integral acquires an imaginary part, which corre-
sponds to the function G we already know. Here oc = 1 - 4m 2I k 2, and the
three regions correspond respectively to k 2 < 0, k 2 > 4m 2 and 0 < k 2 <
-4m 2. The final result is immediate and againfound in many books on
quantum fields, und er the guise of the "renormalized vacuum polarization"
functional; we omit it, noting only that there are two possible singularities
to investigate. At k 2 = 0, the function H is perfectly smooth. At k 2 = 4m 2,
the onset of the absorptive part, H has a cusp.
Wehave finally obtained a closed expression for

(Üin I Oout) = I (Om I Oout) I eitl


in QED at the first significant order. Recall that we had, at second order in
the coupling constant:

I(Om I Oour) I "" 1 - ~P ""exp( -P 12),


where P, the probability of pair creation from the vacuum at that order, is
given from (13.29)-glancing back at the computations already made-by

p = 1~~ I (](k)](k))G(k) d 4 k.
592 13. Quantum Theory

In summary,

(Oin I Oom) ""exp(- 2~~ J(](k)j(k))(G(k) + iH(k)) d 4 k) =: exp(iW).

Here W is the effective action at order e 2 •


Strangely enough, the computation just performed does not appear to
have been pushed to the finish line before [206], although the tools have
been there since the seventies at least [28,409]. The authors [213,468] and
also Langmann and Mickelsson [311,313] came quite close in the nineties.
Now, the "bubble" diagram is essentially the same as the one-loop va-
cuum polarization or "photon self-energy" diagram in full QED (see, in this
respect, [264, pp. 195-196]). The readerwill have noticed that, in order to
avoid pitfalls, we reorganized the calculation in the same way as is done
for vacuum polarization in the Epstein-Glaser renormalization procedure
in [413]. This last reference is a remarkable book which goes a lang way
in popularizing that procedure. Here, of course, we knew all along that
the phase is finite. Thus, the contention by Scharf and followers, that the
Epstein-Glaser renormalization procedure is the more fundamental one, is
vindicated to some extent.
To summarize: in addition to the standard renormalization procedure,
there are in principle three (of course equivalent) methods to compute the
phase of the quantum scattering matrix in the linear theory. The first is the
use of (subtracted) dispersion relations. The second consists of importing
to the linear theory procedures of nonlinear quantum field theory that pur-
port to show 5log(Oin I0 0 u 1 ) and related quantities as explicitly finite; that it
is to say, the Epstein-Glaser procedure. The quantum scattering matrixwas
found lang ago with this method by Bellissard [28] in the boson case and
by Dosch and Müller [145] for QED of static fields; but they stopped short
of computing the phase. The third treatment is the one developed here,
whereby one proceeds to the Dyson expansion in Fock space, integrating
the cocycle associated to the spin representation.

13.A On quantum field theory on noncommutative


manifolds
The need for renormalization is partly tied to the failure of local formu-
lae to capture all the nonlocal aspects of quantization. Contrary to naive
expectations long held [79], it is not true that the presence of divergences
or finiteness of quantum theories is primarily associated to commutativity
or noncommutativity of geometry. Yang-Mills theories on noncommuta-
tive manifolds with physical dimension of the kind studied in this book are
ultraviolet divergent. That was pointed out first by Filk [182] in the frame-
work of the Moyal product and by the authors in a general, nonperturbative
13.A On quantum field theory on noncommutative manifolds 593

context in [470] -written when we were unaware of the elegant work by


Filk. Other early investigations, by Chaichian, Demichev and Presnajder [7 4]
about the "quantum plane", by Perez-Martin and Sänchez-Ruiz [330] on the
U(1) gauge theory on "noncommutative ~ 4 " and by Krajewski and Wulken-
haar [294] on the same theory on the noncommutative torus, reached the
same conclusion.
By December 1999, the number of papers dealing with quantum field
theory on noncommutative manifolds was stillabout a dozen (consult the
reference list in [344]); since then, it has had almost exponential growth.
A main reason is that noncommutative Yang-Mills theories of the Moyal
type (i.e., gauge theories on noncommutative manifolds whose algebra law
can be expressed as a Moyal product on a commutative manifold) crop up
naturally in string theory, as was discovered in [103]. Although neither an
experimentally established theory nor a fully rigorous branch of mathe-
matics, string theory enjoys great popularity. The first to discuss the non-
commutative geometry of strings were Fröhlich and Gaw~dski [191]. The
article [430] surveyed the subject of noncommutative Yang-Mills theories
and gave a great impulse to it. Let us just say here that there is a limit in
which the quantum string dynalllies on a flat background with a Neveu-
Schwarz B-field and with D-branes becomes effectively a field theory that
can be regularized in two different ways, and the transformation from ordi-
nary (in general, nonabelian!) to "noncommutative" Yang-Mills theory cor-
responds perturbatively to the relation between these regularizations. The
point is made in [430] that only U(N) gauge groups are obtained. Further-
more, in the region of "large noncommutativity", tachyon condensation in
open strings gets a particularly simple description [489].
Thus most papers on the subject matter of this section employ the Moyal
product. Note that the Moyal product rule corresponds locally to the com-
mutation rules

for the coordinates. A model, with a different physical rationale, which


leads to Moyal rules, with ()iJ interpreted as a Lorentz tensor, is the "quan-
tum spacetime" of [143]. Of course, the ordinary Moyal product is not co-
variant, but this can be fixed by a slight modification of the latter [396]. At
any rate, the covariant case is of little interest in string theory.
Filk's argument for the prevalence of ultraviolet divergences in theories
of the Moyal type proceeds by looking at Feynman diagrams in momentum
space. The standard integrands appear there to be multiplied by trigono-
metrical functions of the momenta, due to the modification of the interac-
tion vertices. A distinction is made between "planar diagrams", i.e., the ones
that get multiplied by phase factors dependent only on external momenta
(and thus, as Filk recognized, have the same degree of divergence as their
commutative Counterparts), and nonplanar diagrams. The dependence of
phase factors on internal momenta "softens" the integrals, so the latterare
594 13. Quantum Theory

in principle better behaved than their commutative Counterparts. Now, as


explained in [217], some nonplanar diagrams may have phases that vanish
when the external momenta satisfy particular relations; that gives rise to
nasty clashes of ultraviolet andinfrared problems that tend to spoil renor-
malizability. Other pathologies of (at least some) theories of the Moyal type
are pointed out in [203,209]. (We hasten to mention that the distinction be-
tween planar and nonplanar diagrams is of course very old; for instance,
one can see it in action, in a context not entirely unrelated to the present
one, in [204].)
The dust has not settled enough for finer distinctions to be made, so
we shall say no more here; except that apparently nothing is known about
theories not (or not explicitly) of the Moyal type -although they are in prin-
ciple approachable by a combination of the external field method of [470]
and functional integration over noncommutative gauge fields.
We turn, then, to our own argument for the prevalence of ultraviolet di-
vergences. (To dispel any possible confusion, we want to emphasize that
we deal with theories on noncommutative manifolds of well-defined posi-
tive (classical) dimension, not with 0-dimensional proxies for them, which
are automatically finite; see [216] and references therein for that line of
work.) The argument is first couched, for simplicity, in terms of tori (and
thus related to the Moyal product as weil); but then we show that the same
conclusion is unavoidable in the framework of any noncommutative geo-
metry in the sense of Part III. To make it watertight, we call on the theory
of Chapters 6 and 10 and this present chapter.
As an indicator of the ferocity of the ultraviolet divergences of a quan-
tum theory on an (in general, noncommutative) spin manifold, we take the
degree of summability of the corresponding Fredholm module. This was
defined, in Section 8.2, as the least integer n for which all [F, a] E .[n+l;
let us call it the quantum dimension of the theory. (To leave aside geo-
metrical complications extraneous to the analytical problern at hand, think
of a simple U ( 1) model.) If the quantum dimension is greater than one,
gauge transformations will not be unitarily implementable in the sense of
Shale-Stinespring. Now, Lemma 10.18 teils us at once that the quantum
dimension cannot be larger than the classical dimension.
So far, so good. However, is not difficult to check by direct calculation
that the quantum dimension of tori in fact equals their classical dimension.
Let us choose n = 3, for definiteness. Let v = :Lr arur E lr~ be a uni-
tary element; recall that the unitarity can be expressed in terms of the
relation (12.17) among the coefficients ar. Choose an orthonormal basis
{ 1/J"F: r E ~ 3 } for the Hilbert space J{ = J{T $ J{T, which diagonalizes D 2 ;
with respect to this basis, F is given by IYI- 1 r · ä on the two-dimensional
subspace spanned by 1/J~ and IJJ;: (according to (12.25), where the yi = CYJ
are just the Pauli matrices). Therefore, the matrix entries of the operator
13.A On quantum field theory on noncommutative manifolds 595

A = [F, v] are given by

r + §) . a- r . ä)
[F ] +
, v 1/Jr: = -7 o-(s, r)as
"" _ _ ( <
IY + 51 -VI +
1/Jr+S"

Ta obtain the Schatten dass of A, we must determine the finiteness of


the p-norm IIAIIp := (Tr(A* A)Pi 2 ) 11 P, whichis ingeneral hard to compute.
A simpler alternative is to calculate lilA III P := (Ir IIAIJJt IIP + IIAI/1;:: IIP) l/p.
However, thesearenot equivalent norms unless p = 2. It is known [201]
that IIAIIp :::;; IIIAIIIp if 1 :::;; p :::;; 2, whereas IIIAIIIp :::;; IIAIIp if p ;::; 2. Thus,
in general, for p > 2 the divergence of IIIAIIIp implies that Art LP, but not
conversely.
For the particular case A = [F, u 5 ], with u 5 being any Weyl element, this
does not matter, since A* Ais diagonal in the chosen basis. Indeed,

[F ,us] * [F ,us] 1/Jr:=o-r+s,s,o-s,r


+ ( - - :;"\ <- -) ( < + s} . ä
IY+sl r r . ä ) 2 1/Jr:
-VI +

( (r + s}. r) +
= 2 1 - IY + sllrl 1/Jr: •
since o-(r + s, s} o-(s, Y) = lo-(r, §) 12 = 1 by (12.15) and the skewsymmetry
of e. Thus,

We conclude that [F, u 5 ] E LP if and only if Ir,.o IYI-P converges, if and


only if ft p 2 -P dp converges, if and only if p > 3. In other words, the
quantum dimension of lr~ is 3.
This property is not peculiar to tori. Assurne that the classical dimension
of a spin geometry is n. If the quantum dimension were lower, it would still
be an integer of the same parity, being the summability degree of an even
or odd Fredholm module. Suppose it is n- 2. By Proposition 10.7, the Chern
character T}! is cohomologous to STp- 2, in the notation of Definition 10.6.
However, the cochainmap S, as constructedin Section 10.1, takes the cyclic
cocycle Tp- 2 into a Hochschild coboundary ST}!- 2. If c denotes the n-cycle
fulfilling rro(c) = x, then Connes' Hauptsatz (Theorem 10.32) shows that
Tr+ IDI-n = Tr+ xrro(c) IDI-n = ST}!- 2(c) = 0,

which is not possible in classical dimension n. In summary, the quantum


dimension is not lower than n. To avoid this conclusion, one mustrelax the
596 13. Quantum Theory

conditions for noncommutative (spin) geometries; but then one faces the
task of showing in what sense the new geometries would be able to sustain
fermions.
The validity of our argument is not restricted to the case in which the
gauge field is treated as an external classical source; for the nonlinear the-
ory can be quantized by the dual way of treating the fermians with Fock
space techniques, and with the gauge bosons by functional integration on
the space of classical configurations. So there is nowhere to hide. This does
not detract from the interest of quantum field theory on noncommutative
manifolds, which has revealed other fascinating ways in which the noncom-
mutative world begs to differ from the commutative one.
Returning to the difficulties with local formulae, we realize that the main
problern of quantum field theory is extraordinarily akin to the fundamen-
tal and formidable problern of computing the full nonlocal Connes' Chern
character (and nor merely the Hochschild dass, as in Chapter 10) by local
formulae, involving the Dirac operator. Then it is not at all surprising that
Connes and Moscovici's methods for doing so have the flavour of renor-
malization [113]. This also makes less than surprising, in retrospect, that
essentially the same regulating Hopf algebra has emerged, in recent times,
in both fields. Such is the theme of the last chapter of this book.
14
Kreimer-Connes-Moscovici Algebras

The main purpose of this chapter is to discuss the Hopf algebras intro-
duced by Connes and Moscovici in connection with the index problern for
K-cycles on foliations [114), and the ones introduced by Kreimer in connec-
tion with perturbative renormalization theory [296). Both types of Hopf al-
gebras were originally found as organizing principles to simplify some com-
putations. This is not unexpected, in view of our discussion at the end of
Chapter 1. They coalesce in the concept of the "(extended, Connes-Kreimer)
Hopf algebra of rooted trees" HR, which is the subject of Section 14.1; we
examine it in some depth from several angles, and identify an important
subalgebra HcM·
A very important theme, running throughout the chapter, is duality. We
take the Ieisure to prove the most important preliminary result for dual-
ity in Hopf algebra theory, to wit, the Milnor-Moore theorem [340), in Sec-
tion 14.2. Before doing so, we introduce the Grossman-Larson Hopf algebra
of rooted trees, and we prove in Section 14.4 that it pairs with HR.
The link between Hopf algebras and renormalization in quantum field
theory, which was the original insight by Kreimer, has now been made more
functional and precise by the introduction of Hopf algebras of Feynman
diagrams [1 07, 108, 208). Insufficient perspective at the time of writing pre-
vents a full-blooded treatment ofthismatter here; however, we give abrief
introduction to it in Section 14.5. The "prehistoric" versionofthat link, in
terms of the Connes-Kreimer algebra of rooted trees, apart from its role
in noncommutative geometry proper, bears close relations with topics in
applied mathematics, and keeps its normative character in quantum field
theory [297); this is why HR remains the main subject.
598 140 Kreimer-Connes-Moscovici Algebras

Section 14.6 explains how the Hopf algebra HcM arises as the natural
infinitesimal Hopf action on the crossed product .J\. of a smooth function
algebra by a group of diffeomorphismso In Section 14.7, we briefly explore
the cyclic cohomology of Hopf algebras, opening the way for computing
characteristic classes on .J\. by transfer from cyclic cohomology classes of
the Hopf algebrao

14.1 The Connes-Kreimer algebra of rooted trees


In Section 1oB we have given an introduction to Hopf algebras, where our
notation for this chapter is establishedo We now need to add some basic
remarks about duality that did not find their place thereo
From Definition 1.22, it is obvious that a coalgebra is obtained by just
reversing arrows in the definition of an algebrao Taking the dual space of
a vector space is a process that reverses arrows, so one might expect that
algebras and coalgebras are the duals of one anothero Unfortunately, this
works smoothly only in one directiono If V is a vector space over IF and
V* := Hom(V, IF) denotes its (algebraic) linear dual space, the transpose of
a linear map cf>: V - W is the linear map cf>t: W* - V* given by f .... f o cf>o
Comparing the diagrams (1.25) and (1.26) with (1.28) and (1.29), it is clear
that if (C, ~. E) is a coalgebra, then C* is an algebra with multiplication
~t (actually, the restriction of ~t to C* ® C*) and unit Et 0But, if (A, m, u)
is an arbitrary algebra, A * ® A * is in general a proper subalgebra of (A ®
A) *, so the image of m t : A * - (A ® A) * need not be included in A * ®
A *, and therefore need not define a coproducto Nevertheless, if A is finite-
dimensional, then A * is a coalgebra with coproduct m t and counit u t 0
Therefore, the linear dual of a bialgebra need not itself be a bialgebra,
nor need the dual H* of a Hopf algebra H be a Hopf algebrao Even so,
given an arbitrary Hopf algebra (H, m, u, ~. E, S), one can find the largest
subspace H# of H* for which m t (H#) !;;;; H# ® H# 0This subspace is called
the finite dual (or Sweedler dual) of H, and consists of those f E H* that
vanish on a left ideal of Hof finite codimension -see [347] for other equi-
valent definitionso One can prove that (H#, ~t, Et, mt, ut, st) is also a Hopf
algebra [347,446]0
(The desire to go beyond the finite dual is a powerful motivation to
replace the category of Hopf algebras by some other category allowing
stronger duality properties; in particular, a theorem that the bidual alge-
bra coincides with the original one under favourable circumstanceso For in-
stance, one may wish to extend the Pontryagin duality theorem for compact
abelian groupso The category of Hopf C*-algebras [463] has a subcategory
fulfilling this requirement: these are the "reduced C*-algebraic quantum
groups" of Kustermans and Vaeso See [301], where an isomorphism with
14.1 The Connes-Kreimer algebra of rooted trees 599

the bidual is constructed, which generalizes Pontryagin duality to locally


compact nonabelian groups and beyond.)
We approach the Connes-Kreimer algebra of rooted trees by considering
a universal cohomological problern [106]. Let B be a bialgebra. We define an
n-cochain as a linear map L: B - B~m, and its coboundary as the (n + 1)-
cochain
n
bL(x) := (id®L) o Ll(x) + I(-1)iLli oL(x) + (-l)n+ 1 L(x) ® 1, (14.1)
i=l
where Lli means that the coproduct Ll is applied on the ith factor of L(x).
Denoting a map of the form x ...... N(x) ® 1 by N ® 1, we find that
n
b 2 L = (id® 2 ®L) 0 (id ®Ll) 0 Ll + I ( -1) i Lli+l 0 (id ®L) 0 Ll
i=l
n+l
+ (-l)n+l(id®L ® 1) oLl+ I (-1)jLlj o (id®L) oLl
j=l
n+l n n+l
+ I I(-1)j+iLljoLlioL+ I<-l)n+j+lLljo(L®1)
j=l i=l j=l
n
+ (-l)n(id®L ® 1) oLl+ I<-on+i(Lli oL) ® 1-L ® 1 ® 1
i=l
= (id® 2 ®L) o (Ll ® id) oLl - Llt o (id ®L) o Ll
n
+ I (Llj o Llj o L- Llj+l o Llj o L) + Lln+l o (L ® 1)- L ® 1 ® 1
j=l
= (Ll ® L- Ll 1 o (id ®L)) oLl+ L ® Ll(l)- L ® 1 ® 1 = 0.
In this calculation, the nontrivial cancellations arise from Ll(l) = 1 ® 1 and
Llj o Llj = Llj+l o Llj by coassociativity. We conclude that b 2 = 0, so we
indeed get a cohomology.
To see what is involved here, let us pretend in the notation that B* equals
the finite dual bialgebra ß!l under the transposed operations. We may take
advantage of its counit E' = ut to regard B* as an B*-bimodule in a non-
trivial way:
a'·a·a":=a'aE'(a") forall a,a',a"EB*.
Nowrt: (B*)®n-+ B* determinesaHochschildcochain<PL E cn-l(B*,B*)
by <Pdat .... 'an) := rt (al ® ... ®an). Since E' is the transpose of the unit
map of B, (14.1)yields (ao® ···®an. bL(x)) = (b<PL(a 0 , ... , an), X), where
n
bqJL(ao, ... ,an) = ao<PL(at, ... ,an) + I (-1)jq;L(ao, ... ,ajaj+l·····an)
j=l
+ q;L(ao, ... , an-d E' (an),
600 14. Kreimer-Connes-Moscovici Algebras

which is none other than the Hochschild coboundary for the given B*-
bimodule. The calculation of b 2 L = 0 is just the transpose of the familiar
calculation that the square of the Hochschild coboundary vanishes!
On that basis, one can denote by Z} (B*) the set of 1-cocycles for (14.1)
and by H} (B*) the corresponding cohomology group.
Example 14.1. Perhaps the simplest example of a pair (B, L) where L E
H} (B*) is given by the algebra of polynomials H 1 on one generator, say
8, with the counit defined as the constant term of the polynomial, and the
coproduct given by
tl8 = 8 ® 1 + 1 ® 8,
extended as an algebra morphism. An easy induction shows that

Now, if Lo: H 1 - IF is a 0-cochain, then

bLo(8k) = (id®Lo) o tl(8k)- Lo(8k) = i~ G) 8k-iLo(8i)- Lo(8k)

= ±(
j=l
k .)Lo(8k-J)81.
k- J

Thus, the 1-coboundaries are those linear maps L 1 : H 1 - H 1 suchthat


the degree of the polynomial L 1 P is at most the degree of P, for every
polynomial P. On the other hand, Iet L: H 1 - H 1 be defined by L(8k) =
(k + 1) -l8k+ 1 , extended by linearity (so L corresponds to integration of the
polynomial). Since

It follows that L E Zj(H 1*) .

.,. lt turnsout that among all the pairs (B, L), consisting of a commutative
bialgebra B and a 1-cocycle L E Z} (B*), there is one (unique up to isomor-
phism) so that, given another such pair (B', L'), there is a unique bialgebra
14.1 The Connes-Kreimer algebra of rooted trees 601

morphism p: B - B' making the following diagram commute:

B~B
I I
PI IP (14.2)
't' L' 't'
B'-B'.

The solution happens tobe a Hopf algebra; by Proposition 1.24, if B' is also
a Hopf algebra, then p is a Hopf algebra morphism. To prove this claim, we
introduce a Hopf algebra of rooted trees.
Definition 14.1. A rooted tree is a finite, connected, simply connected, one-
dimensional simplicial complex with a distinguished vertex called the root
of the tree.
lt is easier to handle a more pictorial version, so we canthink of a rooted
tree as a finite set of n points, called vertices, joined by oriented lines;
the other adjectives mean that all the vertices have exactly one incoming
line, except the root which has only outgoing lines, so that there is a unique
path of lines and vertices (a branch) that join the root with any other vertex.
Moreover, we assume that the lines do not intersect.
A vertex with no outgoing line is a leaf. To avoid unnecessary repetition,
we shall actually be working with isomorphism classes of trees. To explain
how this goes, we need some more terminology: the fertility of a vertex is
the number of its outgoing lines or children, and its length (with respect
to the root) is the number of lines that make up the unique positive path
joining it to the root. Two rooted trees are isomorphic if the number of
vertices with given length and fertility is the same for all possible choices
of lengths and fertilities. Fora given rooted tree T, we denote by V(T) the
set of its vertices and by E(T) the set of outgoing lines from its root.
In particular, there is only one rooted tree t 1 with a single vertex, and also
only one with two vertices -call it t 2 • With three vertices, there are two iso-
morphism classes distinguished by the number of outgoing lines from the
root, we call them t31 and t32· For n = 4, there are 4 isomorphism classes,
but the criterion used for three vertices does not teil them apart; even so,
we shall denote by t 41 the rooted tree where all vertices have fertility 1 (a
stick), by t42 and t43 the four-vertex trees with 2 or 3 outgoing lines from
the root respectively (a hook and a claw), and by t 44 the tree whose root
has fertility 1 and whose only vertex with length 1 has fertility 2 (a biped):
602 14. Kreimer-Connes-Moscovici Algebras

All four 4-vertex trees are shown in the diagram; in each case the root is
marked with a <> symbol. We follow the custom of making such banging
gardens, with the root always at the top of each tree.
A simple cut c of a tree T is a subset of its lines (selected for deletion)
such that the path from the root to any other vertex includes at most one
line of e; the cardinality le I ofthissimple cut is the number of such lines.
Deleting the cut lines produces Iei + 1 subtrees; the component containing
the original root is denoted Re ( T): let us call it the trunk. The remaining
branches also form rooted trees, where in each case the new root is the
vertex immediately below the deleted line; Pc ( T) denotes the set of these
pruned branches. Here, for instance, are the possible simple cuts of t42:

The set of nontrivial simple cuts of a tree Twill be denoted by C(T);


we exclude as trivial the "empty cut" e = 0, for which R0(T) = T and
P0 ( T) = 0. There is one more trivial cut f (not a subset of lines of T)
called the "full cut", defined by declaring that RJ(T) = 0 and Pf(T) = T.
lt is sometimes handy to allow trivial cuts; we write C'(T) := C(T) u {0},
and C"(T) := C'(T) u {j}.
We denote by #T the number of vertices ofT.
Definition 14.2. The algebra of rooted trees HR is the commutative alge-
bra generated by symbols T, one for each isomorphism dass of rooted
trees, plus a unit 1 corresponding to the empty tree; the product of trees
is written as the juxtaposition of their symbols.
The counit E : HR - IF is the linear map defined by E(1) := 11F and
E(Tl T2 ... Tn) = 0 if T1, ... , Tn are trees. We also define a map ß: HR -
HR ® HR on the generators, and extend it as an algebra homomorphism, as
follows:

ß1 := 1 ® 1; ßT := T ® 1 + 1 ® T + L Pc(T) ® Rc(T). (14.3)


CEC(T)

Notice that Rc(T) is a single tree, whereas Pc(T) is the product of the Iei
subtrees pruned by the cut e. lt is clear that we can shorten (14.3) to

ßT = T® 1 + L Pc(T) ®Rc(T),
cEC'(T)

since the term 1 ® T corresponds to the empty cut; or even

ßT = L Pc(T) ® Rc(T),
CEC"(T)
14.1 The Connes-Kreimer algebra of rooted trees 603

by regarding T ® 1 as being produced by the full cut.


For instance, Ll(t42) is given by

or, in symbols,

Ll(t42) = t42 ® 1 + 1 ® t42 + t1 ® t32 + tz ® tz + t1 ® t31 + tzt1 ® t1 + ti ® tz.


(14.4a)

The other examples of generators with at most 4 vertices are

Ll(ti) = tl ® 1 + 1 ® ti.
Ll(tz) = tz ® 1 + 1 ® tz + t1 ® t1,
Ll(t31) = t31 ® 1 + 1 ® t31 + tz ® t1 + t1 ® t2,
Ll(t32) = t32 ® 1 + 1 ® t32 + 2tl ® t2 + ti ® tl.
Ll(t4d = t41 ® 1 + 1 ® t41 + t31 ® tl + t2 ® t2 + tl ® t31'
Ll(t43) = t43 ® 1 + 1 ® t43 + 3tl ® t32 + 3ti ® t2 + ti ® tl,
Ll(t44) = t44 ® 1 + 1 ® t44 + t32 ® t1 + 2t1 ® t31 + ti ® tz. (14.4b)

The sprouting of a new root is precisely the morphism L we are looking


for.
Lemma 14.1. Let L: HR- HR be the linear map given by

(14.5)

where T is the rooted tree obtained by conjuring up a new vertex as its root
and extending lines from this vertex to each root of T1 , ..• , Tk. Then

Ll o L = L ® 1 + (id ®L) o Ll. (14.6)

For instance,

Proof. If 1 is a subset of { 1, ... , k }, we denote its complement by ]'. Let


a = T1 ... Tk. then
k
Ll(a) = Ll(TI) ... Ll(Tk) = n(TJ ® 1+ 2:: Pc1 (TJ) ®Rc1 (TJ)) = 2::Si,
j=l CjEC'(Tj) i

where each Si is a product of k terms, one from each j. Fora fixed i, let
1 = {j 1 , ••• , jr } be the set of subindices for which the corresponding term
604 14. Kreimer-Connes-Moscovici Algebras

is of the form Ti® 1, then for j E ]' the factor in the product is of the form
PcJ (Ti) ® RcJ (Ti).
Let -l!i be the line that joins the root ofT with the tree Ti, and let ci the
simple cut of T given by

Ci:= {-f!jp···•-f!ir} U u
jE]'
Cj.

lt is clear that Si - ci establishes a bijective correspondence between the


summands in ß(a) and the simple cuts ofT; moreover, for each i,

and so

ß(L(a)) -L(a) ® 1 = L PcdT) ®Rc;(T) = (id®L) o ß(a). D


ciEC'(T)

Lemma 14.1 is our main workhorse; once we prove that the set of rooted
trees is indeed a Hopf algebra, the Hochschild equation (14.6) means that
L is a 1-cocycle, so that [L] EH} (H;>; and [L] * 0 is clear since L(l) = t 1 ,
whereas, for any 0-cochain L 0 ,

boLo(1) = (id ®Lo) o ß(1)- Lo(1) 1 = (id ®Lo)(1 ® 1)- Lo(1) 1 = 0.

Corollary 14.2. The map ß is a coproduct.

Proof. Let Hn be the polynomial subalgebra of HR generated by the symbols


of rooted trees with at most n vertices. lt is enough to prove that (id ®ß) o
ß = (ß ® id) o ß on each Hn. Since ß is defined recursively, we do this by
induction.
Since L(l) = t1, then, by Lemma 14.1,

(14.7)

so ß is coassociative on H 1 . Assurne that ß is coassociative on Hn and let


T be a rooted tree with n + 1 vertices. Then T = L(a) where a = T1 ... Tk
and T1 , ••. , Tk are the branches obtained by removing all the outgoing lines
from the root (this operation is a left inverse for L with a one-dimensional
kernel). Using Lemma 14.1 repeatedly and coassociativity on Hn,

(id ®ß) 0 ß(T}


= L(a) ® 1 ® 1 + (id ®L) o ß(a) ® 1 + (id ®id ®L) o (id ®ß) o ß(a)
= L(a) ® 1 ® 1 + (id ®L) o ß(a) ® 1 + (id ®id ®L) o (ß ® id) o ß(a)

= L(a) ® 1 ® 1 + (id ®L) o ß(a) ® 1 + (ß ® id) o (id ®L) o ß(a)

= (ß ® id) o ß(T). D
14.1 The Connes-Kreimer algebra of rooted trees 605

Definition 14.3. The factorial T! of a tree T is recursively defined by t1! := l


and (L(Tl ... Tk))! := #L(T1 ... Tk) T1! ... Tk!. For instance, t41! = 24 (the
relation T! = (#T)! always holds for sticks), t 42 ! = 8, t 43! = 4 and t 44 ! = 12 .

.,. Now we solve the universal problem.


Theorem 14.3. The pair (HR,L), with L defined in Lemma 14.1, is the solu-
tion of the universal problern ( 14.2).

Proof. Let B' be a bialgebra and L' a 1-cocycle in Z} (B' *). If T is a rooted
tree, let c the simple cut that contains all the outgoing lines from the root.
Then T = L(Pc(T)). We define p: HR- B' on the generators by

p(T) := L'(p(Pc(T))),

and extend it as an algebra homomorphism. By definition, p satisfies (14.2).


Since p is defined recursively, we prove by induction that it is a coalgebra
homomorphism. Since L' E Z} (B' *), then

~'(L'(1)) = L'(1) ® l + (idB' ®L') o ~'(l) = L'(l) ® l + 1 ® L'(1),

which yields

~' o p(ti) = ~'(L'(1)) = L'(1) ® 1 + 1 ® L'(l)


= (p ® p)(tl ® 1 + 1 ® td = (p ® p) 0 ~(ti),

so ~, o p = (p ® p) o ~ on H 1 . Assurne this equation holds on Hn and let


T be a rooted tree with ( n + 1) vertices; then Pc ( T) E Hn. Thus, using that
both L' and L are 1-cocycles, the inductive hypothesis and (14.2),

~' o p(T) = ~' o L'(p(Pc(T)))


= L'(p(Pc(T)) ® 1 + (idB' ®L') o ~'(p(Pc(T)))
= p(L(Pc(T))) ® l + (idB' ®L') o (p ® p) o ~(Pc(T)))
= p(L(Pc(T))) ® 1 + (p ®poL) o ~(Pc(T)))
= (p ® p)(L(Pc(T)) ® 1 + (idj}; ®L) o ~(Pc(T)))
= (p ® p) o ~(L(Pc(T))) = (p ® p) o ~(T).

On the other hand, if a E B', using (1.28) and bL' = 0,

L'(a) = (E' ®idB·) o ~'(L'(a)) = (E' ®idB·HL'(a) ® 1 + (idB' ®L') o S(a))


= (E' ® idB' )(L' (a) ® 1) + (idF ®L') o (E' ® idB') o ~, (a)
= (E' ®idB·)(L'(a) ® 1) + (idF®L')(l ®a)

= E'(L'(a)) +L'(a),
so that E'(L'(a)) = 0. Thus, E' o p(T) = E' o L'(p(Pc(T))) = E(T) for all T,
so p is counital. 0
606 14. Kreimer-Connes-Moscovici Algebras

.,. So far, we have described HR as a bialgebra. We introduce the antipode


S: HR - HR by exploiting its very definition as the convolution inverse of
the identity-see Definition 1.26-via a geometric series:

S := id*- 1 = (u o E- (u o E- id))*- 1
= U oE+ (U o E-id)+ (U o E- id)* 2 + · · · . (14.8)

For convenience, we abbreviate 11 := u o E-id. Note that I'J(T) = - T if T is


any rooted tree.
Lemma 14.4. If T is a rooted tree with n vertices, the geometric series ex-
pansion of S ( T) has at most n + 1 terms.

Proof. We claim that 17*m (T) = 0 if m is greater than the number ofvertices
ofT. This is certainly true for t 1 , using (14.7) and 17(1) = 0. Assurne that it
holds for all trees with n vertices. Let T be a rooted tree with n + 1 vertices;
using (14.3), we find that

17*(n+2>(T) = 11* 17*(n+U(T) = m o (17 ® 17*(n+U) o ~(T)


=mo(17®17*(n+ 1>)(T®1+1®T+ L Pc(T)®Rc(T)).
cEC(T)

By the inductive hypothesis, the third term is zero; the first and second
terms vanish because 17(1) = 0. D

As an immediate corollary, we obtain that S, extended as an algebra ho-


momorphism (since HR is a commutative algebra), is indeed an antipode.
a;
Moreover, if a E Hn, ~(a) = Li 1 1 ® a;;. ~(a;;> = Li 2 1 i 2 ® a;;iz and a;
in general ~(a;;, ... ,ik) = Lik+l a;], ... ,ik+I ® a;;, ... ,ik+I, then, for k ~ 1,

where

if a~l} ,. .. ,lj. = 1 or a~' .


lt ,. .. ,lj
= 1'
otherwise,

b"z., ... ,ti. ·=


·
{0
a~' .
11 ,. .. ,lj
if a'-'1} ,. .. ,lj. = 1'
otherwise.

Thus, we can compute S directly from the coproduct. For instance, us-
ing (14.4), we get

(14.9a)
14.1 The Connes-Kreimer algebra of rooted trees 607

and

S(t42) = -t42 + (t1t32 + t~ + t1t31 + 2tftz)- (5tftz + 2tf) + 3tf


= -t42 + t1t32 + t~ + t1t31- 3tftz + tf. (14.9b)

Similarly, we obtain

S(t31) = -t31 + 2t1t2- tf,


S(t32) = -t32 + 2t1t2- tf,
S(t41) = -t41 + 2t1t31 + t~- 3tftz + tf,
s<t43) = -t43 + 3t1t32- 3trtz + tf,
s<t44) = -t44 + t1t32 + 2t1t31- 3trtz + tf.

The definition (14.8) for the antipode is due to the authors [179]. We
briefly turn to the definitions given by the discoverers. The equation m o
(S ® id) o ~(T) = 0 and (14.3) suggest that one may define the antipode
recursively, by

SB(T) := -T- L SB(Pc(T)) Rc(T), (14.10)


CEC(T)

For instance,

which gives (14.9b) again, assuming that (14.9a) also holds for SB.
Proposition 14.5. IfT is any rooted tree, then S(T) = SB(T).

Proof. The statement holds, by a direct check, if T has 1, 2 or 3 vertices. If


it holds for all rooted trees with at most n vertices and if T is a rooted tree
with n + 1 vertices, then

S(T) = 1J(T) + i
j=l
11*i * 1J(T} = -T + m o ( i
j=l
TJ*i ® 1J) o ~(T)

=-T+moi1J*i®1J(T®1+1®T+ L Pc(T)®Rc(T))
j=l CEC(T)
n
= -T- L L TJ*i(Pc(T))Rc(T)
cEC(T) j=l

= -T- L SB(Pc(T)) Rc(T) = SB(T),


CEC(T)

where the penultimate equality uses the inductive hypothesis. D


608 14. Kreimer-Connes-Moscovici Algebras

The subscript B in the previous version of the antipode is to remind us


of the similarity with Bogoliubov's recursive formula [38) for the renormal-
ization of Feynman diagrams with subdivergences; a nonrecursive formula
for the same purpose is the famous forest formula [500) of Zimmermann
(see Section 14.5). To the forest formula corresponds the following nonre-
cursive version of the antipode:

Sz(1) := 1, Sz(T) :=- L (-1)ldlpd(T)Rd(T),


dED(T)

where D ( T) is the set of all cuts, not necessarily simple, including the empty
CUt.

Exercise 14.1. Prove that, for an arbitrary rooted tree T,

S(L(T)) = -L(T) -S(T)tl- L S(Pc(T))L(Rc(T)). (14.11)


CEC(T)

Conclude that S(T) = Sz(T) by showing that Sz also satisfies (14.11). 0


Exercise 14.2. Prove by a direct calculation that S 2 = idHR· 0
Exercise 14.3. Let PI. p 2 , p 3 , ... be the Iist of prime numbers in ascending
order. If T1, ... , Tr are rooted trees, prove that the map f defined recur-
sivelyby
n
r
j(ti) := 1 and j(L(TI. .. . , Tr)) := PJ(T1 J
j=l

is a bijection between the set of rooted trees and the positive integers. 0

~ How many elements does HR contain, up to a given order in the vertices,


say n - 1? There are as many of these as there are trees of order n, so it is
enough to count trees. This is a venerable subject. The following recurrence
relation for the number tn of rooted trees with n vertices was apparently
first given by Cayley [72) in 1857:

(14.12)

In particular, t5 = 9, t6 = 20, t7 = 48, ts = 115, and so on. In the next


section, following Grossman and Larson [219), we give a derivation of this.
Cayley's formula has never been summed in closed form. Much easier to
count are heap-ordered trees, which are rooted trees with a total ordering
of the vertices, increasing along the way from the root to any vertex. There
are (n - 1)! heap-ordered trees with n vertices.
We can now introduce a natural operator that allows us to find an impor-
tant subalgebra of HR.
14.1 The Connes-Kreimer algebra of rooted trees 609

Definition 14.4. The natural growth operator N: HR - HR is the unique


derivation defined, for each tree T with vertices v 1 , ••. , Vn, by

N(T) := T1 +Tz+···+ Tn.

where each Tj is obtained from T by adding a leaf to v i.


In particular,

N(t1) = tz,
N 2 (ti) = N(tz) = t31 + t32,
N 3 (ti) = N(t31 + t32) = t41 + 3t4z + t43 + t44·
For example, the last equality is computed like this:

N(Y + / \ )
•I • •
= t!

+ 3 ./'"· +
•I
/~'\
•. •
+ y
•/ ' \•
by sprouting new leaves from each vertex of t31 and of t3z.
From (14.4),

tl(N(ti)) = N(tt) ® 1 + 1 ®N(ti) + t1 ® t1,


tl(N 2 (ti)) = N 2 (ti) ® 1 + 1 ® N 2 (ti) + ti ® t1 + tz ® t1 + 3tl ® tz,
tl(N 3 (t 1)) = N 3(ti) ® 1 + 1 ® N 3 (ti) + N 2 (tl) ® t1 + 4tz ® tz
+ 6t1 ® N 2(tl) + ?ti ® tz + 3tztl ® t1 + tr ® t1.
The number of times that the tree T appears in N#T - 1 ( t 1 ) is designated
CM(T) by Kreimer in [298], who called it the "Connes-Moscovici weight"
ofT. For instance, CM(t42) = 3. lt is found, in [52], that
(#T)!
CM(T) = T! a(T),

where a (T), the symmetry factor of a tree T, is the number of isomorphic


trees that can be generated by permutation of the branches. For instance,
a(t43) = 3!, so CM(t43) = 4!/4 · 3! = 1. By construction, CM(T) is the
number of heap-ordered trees with shape T. The number of summands in
Nn- 1 (tt) is therefore

L CM(T) = (n- 1)!. (14.13)


#T=n

~ We now extend the Hopf algebra HR into a new Hopf algebra iiR by
adding two extra generators X and Y. The first one implements the natural
growth derivation:
[X, T] := N(T),
610 14. Kreimer-Connes-Moscovici Algebras

It is clear that #(TI Tz) = #T1 +#Tz, so# introduces a l-grading on HR, and
we can add a second generator satisfying

[Y, T] := (#T) T.

The Jacobi identity forces [Y, X] = X (the affine Lie algebra relation), as
follows:

[[Y,X], T] = [[Y, T],X] + [Y, [X, T]] = (#T) [T,X] + [Y,N(T)]


= -(#T) N(T) + (#T + 1) N(T) = N(T) = [X, T].

We must declare the coproduct of X and Y. The following exercise shows


that Y must be primitive. On the other hand, X is not primitive; the defini-
tion of L\(X) anticipates the result of Proposition 14.6 below:

L\(Y) := Y ® 1 + 1 ® Y,
L\(X) :=X® 1 + 1 ®X+ 81 ® Y,

where we write 81 := t1. For later convenience, we also abbreviate 8n+l :=


Nn (ti) for n = 1, 2, 3, ... ; notice that 8n+l = [X, 8n].
Exercise 14.4. Show that L\([Y, T)) = [Y ® 1 + 1 ® Y,L\(T)] for any rooted
me~ o
To motivate the next proposition, note that if L\ is to be a morphism of
algebras and L\ (8n) = Lj aj ® aj, then

L\(8n+l) = [L\(X),L\(8n)J
= 2)X,aj] ® aj + :Z::aj ® [X,aj] + [81 ® Y,t18n]
j j

= LN(aj) ® aj' + L aj ® N(aj') + [81 ® Y, L\8n]


j j

= (N ® id)L\(8n) + (id ®N)L\(8n) + [81 ® Y, t18n].

Weshall verify this relation for every rooted tree, and thus, by the derivation
property, for any element of HR.
Proposition 14.6. Foreach rooted tree T, the following relation holds:

L\N(T) = (N ® id)L\(T) + (id ®N)L\(T) + [81 ® Y, L\T].

Proof. By (14.3),

L\N(T) = N(T) ® 1 + 1 ® N(T) + L L Pci (Tj) ® Rci (Tj ). (14.14a)


j CjEC(Tj)

Let v 1, ... , Vn be the vertices ofT and f1, ... , fn the outgoing lines that are
attached by N to each Vj. Then we write C(Tj) = Aj w Bj w Dj wEj, where
14.1 The Connes-Kreimer algebra ofrooted trees 611

CJ e AJ if -l!J is not in CJ and VJ is in the trunk RcJ (TJ ), CJ e BJ if -l!J is not


in cJ but v J is in a pruned branch of TJ, D J consists of the cut {-1!J}, and
CJ e EJ if -l!J e CJ and CJ \ {-l!J} is not empty. The rightmost term in (14.14a)
becomes
L( L + L + L + L )PcJ(Tj) ®RcJ(Tj). (14.14b)
j CjEAj CjEBj CjEDj CjEEj

Foreach c e C(T), we define CJ e AJ by CJ := c n L(TJ) where L(TJ) is the


set of lines in TJ· Clearly, c ...... (c1, ... , Cn) matches the indices of the first
sum in (14.14b) with those on the right hand side of
(id ®N)ß(T) = 1 ® N(T) + L Pc(T) ® N(Rc(T)).
cEC(T)

Moreover, Pc (T) ®N(Rc(T)) = LJ PcJ (TJ) ®RcJ (TJ ); so the two sums agree.
Similarly, since
(N ® id)ß(T) = N(T) ®1 + L N(Pc(T)) ® Rc(T),
cEC(T)

the same argument shows that the sum on the right hand side agrees with
the second sum of (14.14b).
On the other hand, since DJ consists of the cut {-l!J},
[81 ® Y, T ® 1 + 1 ® T] = [81, T] ® Y + 81 ® [Y, T]
= n81 ® T =L L PcJ(Tj) ® RcJ(Tj).
j CjEDj

Finally, since HR is commutative,


[ 81 ® Y, L Pc(T) ® Rc(T)] = L 81Pc(T) ® [Y,Rc(T)] (14.15)
cEC(T) CEC(T)

= L #(Rc(T)) 81Pc(T) ® Rc(T).


cEC(T)

Now Iet K c { 1, ... , n} be the integers for which Vk E Re (T). To each


c e C(T) we associate #(Rc(T)) cuts, namely ck := (c u {-l!k}) n L(Tk).
Notice that Pq(Tk) ®Rq(Tk) = 81Pc(T) ®Rc(T) for each k E K, so the last
sum of (14.14b) equals the right hand side of (14.15). D
In particular, ß extends as an algebra morphism.
Corollary 14.7. The algebra generated by X, Y and all 8n is a Hopf sub-
algebra of fiR. The algebra generated by all 8n only is a Hopf subalgebra
~~. B

._ The larger of these Hopf subalgebras coincides with the simplest of the
real Hopf algebras introduced by Connes and Moscovici in [114] in another
context (about which we shall have more to say in Section 14.6). It is worth
a stand-alone definition.
612 14. Kreimer-Connes-Moscovici Algebras

Definition 14.5. Let HcM be the universal enveloping algebra of the Lie
algebra !!JcM linearly generated by elements X, Y, and {An : n = 1, 2, 3, ... },
subject only to the relations

[X, An] = An+1. and [An, Am] = 0. (14.16)

The Jacobi identity shows that [Y,X] =X, as before.


As a universal enveloping algebra, HcM has already a Hopf algebra struc-
ture; in particular, we let E: HcM - ~ be the usual counit of 'U(!!JcM ). How-
ever, we consider instead a new coproduct ll: HcM - HcM ® HcM which
identifies HcM with the aforementioned Hopf subalgebra of HR. Namely,
we define ll on the generators by

ll(Y) := Y ® 1 + 1 ® Y,
ll(AI) :=Al® 1 + 1 ® A1,
ll(X) :=X® 1 + 1 ®X+ A1 ® Y, (14.17)

and extend it as an algebra homomorphism.


Exercise 14.5. Show that (ll®id) oll and (id ®ll) oll agree on the Lie algebra
fJcM; conclude that the algebra map ll is coassociative, and hence that HcM
is a bialgebra. o
Exercise 14.6. Show by direct calculation that

ll(A2) = A2 ® 1 + 1 ® A2 + A1 ® A1,
ll(A3) = A3 ® 1 + 1 ® A3 +At® A1 + A2 ® A1 + 3A1 ® A2,

and compute ll(A4).


Exercise 14.7. Explain why the antipode S of HcM must satisfy

and S(An+d = [S(An),S(X)]. Prove that the linear operator S on !!JcM de-
termined by these relations extends to an algebra antihomomorphism of
*
HcM into itself, satisfying id *S = S id = u o E; thus, HcM is a Hopf algebra
with antipode S. 0
In summary: three algebra generators, X, Y and A1 and the relations
(14.16) and (14.17) uniquely specify a Hopf algebra. 1t follows from Propo-
sition 14.6 that this Hopf algebra HcM may be identified with the natural
growth subalgebra of HR. Even so, it remains a healthy exercise to directly
check the compatibility of simple cuts with natural growth.
More trivial subalgebras of HR are known: the algebra of trees whose
vertices have fertility not exceeding a given n is a subalgebra of HR; for
n = 1, this produces a cocommutative subalgebra.
14.2 The Grossman-Larson algebra of rooted trees 613

14.2 The Grossman-Larson algebra of rooted trees


There is another Hopf algebra structure on the set of rooted trees, quite
different from the one considered by Connes and Kreimer. It was intro-
duced in the late eighties by Grassman and Larson [219,220]. Weshall now
describe it, postponing until the next section the matter of how these two
Hopf algebras are related.
As before, we shall be dealing with isomorphism classes of rooted trees.
Given two rooted trees Sand T, we call M(S, T) the set of maps from E(S)
to V(T); for each a E M(S, T), we form a new tree Ta from T by hanging
the line e (and the attached subtree of S) from the vertex a(e), for each
e E E(S). The product m(S, T) = S · T is then defined as the formal sum
of the Ta as a runs over M(S, T).
Clearly, t1 is the unit for this product: if S = t 1, E(S) is empty and T
remains unchanged, whereas if T = t 1, hanging the subtrees of S from the
new root t1 just produces another copy of S. Moreover, tz · tz = t31 + t32,
and

tz · t31 = t41 + t42 + t44, t31 · tz = t41 + t42,


tz · t32 = 2t42 + t43, t32 · tz = 2t4z + t43 + t44.

In particular, this product is not commutative. Note that if #E(S) = k and


#V(T) = n, then S · T is a sum of nk terms.
Foreach subset I = {e1, ... , er} of E(T), let T1, ... , Tr be the subtrees
ofT that are pruned by deleting the lines in I, and write TI := L(T1 T2 ... Tr ),
using the notation of the previous section. In other words, TI is the subtree
of T obtained by cutting the lines of E(T) that are not in I. Write also
T0 := t 1 • A coproduct Li is now defined by

Ll(T) := L TI® TI·.


I,;E(T)

where I' := E(T) \I. In other words, the coproduct is given by all possible
splittings of the outgoing lines from the root into two groups. Notice that
Li (t 1 ) = t 1 ® t 1. If T * t 1, then by displaying separately the terms where I
is empty or full, we can write

Ll(T) = T ® t1 + t1 ®T + L TI® Tr. (14.18)


0*IcE(T)

In particular,

Ll(tz) = t2 ® t1 + t1 ® tz, Ll(t31) = t31 ® t1 + t1 ® t31,


Ll(t4J} = t41 ® t1 + t1 ® t41, Ll(t44) = t44 ® t1 + t1 ® t44,
614 14. Kreimer-Connes-Moscovici Algebras

while

~(t32) = t32 ® tl + tl ® t32 + 2 t2 ® t2.

~(t42) = t42 ® tl + tl ® t42 + t31 ® t2 + t2 ® t3I.

~(t43) = t43 ® tl + tl ® t43 + 3 t2 ® t32 + 3 t32 ® t2.

A counit l is defined by l(ti) := 1 and l(T) := 0 for any rooted tree T * t1.
and extended by linearity.
Proposition 14.8. The vector space HeL generated by all rooted trees, with
the product m and the coproduct ~. is a cocommutative bialgebra.

Proof. To check associativity of the product, first we observe that

(R · S) · T = I
UEM(R,S)
Su · T = I I
UEM(R,S) pEM(Tcr,Tl
Tp, (14.19a)

R · (S · T) = I
IJEM(S,T)
R· T 11 = I
iJEM(S,T)
I
vEM(R,T~)
Tv. (14.19b)

Let v denote the root of S. Given (a,p) as in (14.19a), we define f.J as the
restriction of p to E(S), and Iet v(e) := p(e) if a(e) = v, or v(e) := a(e)
otherwise. Conversely, given (f.J, v) as in (14.19b), we define a(e) := v if
v(e) E V(T), and if not, a(e) := v(e) E V(S) by regarding S as a subtree
of T11 • Since E(Su) = E(S) u { e E E(R): a(e) = v }, we may define p(e) :=
f.J(e) if e E E(S), and p(e) := v(e) otherwise. In this way we establish a
one-to-one correspondence between the summands of (14.19a) and those
of (14.19b), showing that the product is associative.
Next, ~ is coassociative, since
(id®~) o ~(T) = 2: TI® T1 ® h = (~ ® id) o ~(T),
where the sum runs over all possible disjoint splittings E(T) =I ltl] ltl K.lt
is clearly cocommutative, too. The counit property follows from
(id ®l) o ~(T) = Tl(td + t1l(T) + L Til(Tr) = T,
e>"'IcE(T)

forT* t 11 and similarly (l ® id) o ~(T) = T.


lt remains to show that ~ is an algebra homomorphism. For this, note
that

~(S · T) = L ~(Tu)= L L Tier® TI;,., (14.20a)


UEM(S,T) UEM(S,T) Icr<;;E(Tcr)

whereas

~(S) • ~(T) = L SI · TJ ®Sr · Tr = L L L Tp ® TT.


I,] I,] pEM(SJ,TJ) TEM(S1•,T1 •)
(14.20b)
14.2 The Grossman-Larson algebra of rooted trees 615

Now, to each pair (a-,Iu) as in (14.20a) we associate the sets I:= E(S) n
Iu and 1 = E(T) n Iu, and also the maps p, T defined as the respective
restrictions of a- to I and I'. Conversely, given I, 1. p, and T, as in (14.20b),
we define a- E M(S, T) by a-(e) := p(e) or T(e) according as e EI or e EI'.
We also assign the set
Iu := 1 u { e EI: p(e) is the root of TJ }.
In this way we establish a one-to-one correspondence between the sum-
mands of (14.20a) and those of (14.20b), so that 3. is a homomorphism of
algebras. 0
Weshall shortly see that Her is in fact a Hopf algebra. Note that a rooted
tree whose root has only one child is a primitive element of the bialgebra
Her; thus, L(T) is primitive for eachrooted tree T. Indeed, P(Her) is exactly
the range of the linear map L. To see that, one can define a linear map
J1: Her® Her - Her such that, if S, T are two rooted trees, then J1 (S ® T) is
the rooted tree formed by identifying their roots. Clearly, J1(3.(T)) = zr T
whenever E(T) has r elements; whereas J1(3.(A)) = J1(A® t 1 + t 1 ®A) = 2 A
for all A E P(Her). Since all trees are linearly independent in Her, it follows
that P(Her) is spanned by the trees S whose root has one child only.
This remark, with the help of the Milnor-Moore theorem (Theorem 14.14
below), provides us with a proof [219] of Cayley's counting formula (14.12).
Proposition 14.9. Let tn denote (in this proposition) the number of rooted
trees with n vertices; then

Proof. The identification of P(Her) as the range of L shows that tn is also


the nurober of primitive trees with n + 1 vertices, counting the root.
Such trees span a subspace Pn of P(Her); indeed, P(Her) is the di-
rect sum of these subspaces Pn, for n = 1, 2, .... In the next section, we
shall see that the Milnor-Moore theorem provides an isomorphism Her ""
'U(P (Her)).
Thus tn+l is the dimension of the subspace of 'U(P(Her)) spanned by
products of primitive trees up to a certain degree. The Poincare-Birkhoff-
Tt
Witt basis of 11 (P (Her)) consists of monamials 1 • • • T;P, where the prim-
itive trees T 1 , ••• , Tp are selected from an ordered basis of P(Her) obtained
by concatenating bases of P1 , P2, P3, ... (in that order). If S is a tree in Pk and
T is a tree with l+ 1 vertices, then S · T is a sum of trees with k + l+ 1 vertices.
Tt
If the monomial 1 • • • T;P contains mk terms from Pk for k = 1, ... , r,
then the nurober of vertices in each tree appearing in the product is n + 1,
where m 1 + 2m2 + · · · + rmr = n.
Therefore, tn+l is the sum, over all such partitions of the integer n, of
the amount of monamials composed by concatenating monamials of de-
gree mk from each Pk. Since the latter may be selected independently, their
616 14. Kreimer-Connes-Moscovici Algebras

k
total is the product over each of the number of monamials 1 • • • 1 of T/ T/
degree ii + · · · + iL = mk with the Ti chosen from a basis of Pk. But
(mk;;;:-
1

dimPk = tk. so that number is 1); the recursion formula (14.12) is

established. D

With the help of the concept of "k-primitive element"l beautiful gener-


ating functions (and asymptotic formulae) for this tn sequence are given
by Broadhurst and Kreimer in [52]. lt is easy to see that if the vertices of
the trees (root excepted) are "decorated" by k symbolsl then the number of
trees is given by

~ The reader may be curious to know more about trees and their applica-
tions. Let us mention that Cayley [72] found a combinatorial expression for
differentials in terms of rooted treesl by considering a system of ordinary
differential equationsl in !Rn say:
1

X(t) = j(x(t)); X(to) = Xo.

In componentsl this is

k = 11... 1n.

or we may consider the equivalent integral systeml

x(t) = xo + Jtto j(x(s)) ds. (14.21)

Repeated derivatives of x are computed by using the chain rule; for in-
stancel

xk(t) = J}Ji~

xk(t> = JkJ!'Ji2
)i )2
k. JhJi2
+ 1)!)2 I

-x·k(t> = fkfhJ!2Jh
)J )2 )3
k. JhJi2Ji3
+ 31)!)3 )2
k . . JhJhJh + 1)JkJh.
+ 1)!)2)3 )2)3
JhJi3 I

where f}1 •.. ir := (JYjijoxi 1 ••• oxir(x(t)). To each summand on the right
hand sidel we associate a rooted tree whose vertices are Iabelied by the
indices which appearl with k representing the root. For each term f}, .. ·Jr I

the subindices j1 Iabel vertices growing out of the ith vertex. For examplel
14.2 The Grossman-Larson algebra of rooted trees 617

In particular, we see that


xk - N(td = t2. we write x = Nf,
xk- N 2<td = t31 + t32, we write x = N 2 f,
x·k - N 3 <td = t41 + 3t42 + t43 + t44, we write x· = N 3f.
(Since both d/dt and N satisfy the Leibniz rule, it is clear that d/dt corre-
sponds to the natural growth operator N.)
Thus, the solution of (14.21) is, at least formally, given by the develop-
ment
(t- to)#T
x(t) = Xo + ~ (#T)! CM(T) Tf(to), (14.22)

where CM(T) is the Connes-Moscovici weight ofT. To see (14.22) in action,


consider the example corresponding to f = exp, to = 0, xo = 0, that is,

x(t) = J: exp(x(s))ds.

The solution of this nonlinear equation is obviously x(t) = -log(1- t) =


Ln tn /n, so it is given, in terms of (14.22), by
t#T
~ (#T)! CM(T),

using (or proving, according to point of view) the equation (14.13).


In general, it can be said that "trees allow simple and systematic manage-
ment of calculations involving higher derivations, a fact which was known
to Cayley in the middle of the nineteenth century" [220]. An interesting ap-
plication to nonlinear differential equations with polynomial coefficients is
indeed made in [221].
Now, HR is a commutative algebra, so in principle it falls within the
precinct of Tannaka-Krein duality. Indeed, the set of characters cjJ: HR - IR
is a group GR, with the convolution product given by (1.35). Keep in mind
that cp- 1 = cp o S. Reciprocally, one canthink of HR as the algebra of co-
ordinate functions on GR. Weshallsee in the next section that a notion of
Lie algebra for GR can be given, so that HR and the Hopf enveloping alge-
bra of that Lie algebra are in duality. lt is shown in detail in a most useful
paper by Brouder [56], reinterpreting previous work by Ruteher [60], that
any character cp E GR corresponds to the formula
(t- to)#T
x(t) = xo + ~ (#T)! CM(T) T! c/J(T) T f(to),

that in turn represents a Runge-Kutta approximation method for solving


the original flow equation (we are making a long story short). We agree then
to call GR the Ruteher group.
618 14. Kreimer-Connes-Moscovici Algebras

14.3 The Milnor-Moore theorem


Our main goal now is to describe the dual Hopf algebra of HR. Since this
algebra is commutative, its dual would be a cocommutative Hopf algebra:
we therefore start with a description of such algebras.
Definition 14.6. We say abialgebraBis connected if it has a filtration com-
patible with its bialgebra structure having only scalars in degree 0; that is,
if there is a family {Bn}nEN of increasingly nested vector subspaces of B
suchthat
(a) Bo = 1F1 and B = UnEN Bn;
(b) BnBm ~ Bn+m for all n, m;
(c) tl. (Bn) ~ L.r=o Bk ® Bn-k for all n.
The name harks back to the topological origins of the theory: the coho-
mology Hopf algebra H"(G, ~) of a compact Lie group G has this property
when G is connected. Indeed, it is known that this cohomology is that of a
product of odd-dimensional spheres. A "topological" argument for that is
beautifully made in [44,45]. It is not difficult to see that a connected finite-
dimensional algebra with only one generator must be N [x], where x is of
odd degree k; this is nothing but H(§k, ~) with k odd. Therefore, H"(G, ~)
is an exterior algebra with generators in odd dimensions [435).
In some treatments, for instance in [347), a bialgebra is called connected
if its "coradical", namely the sum of all simple subcoalgebras (those that
have no proper subcoalgebras), is one-dimensional. The coradical turns out
tobe the bottom piece of a filtration (called the coradical filtration), which
is compatible with the coproduct of any coalgebra, and with the algebra
structure too in the case of a bialgebra.
Example 14.2. The Grossman-Larson bialgebra Her is connected: for let
Bn be the vector space generated by the rooted trees with at most n + 1
vertices. Since the product of two rooted trees, with m + 1 and n + 1 vertices
respectively, is a sum of trees with n + m + 1 vertices, the fibration is
compatible with m. Furthermore, as the coproduct L\ is given by all possible
splittings of the outgoing lines from the root into two groups, the filtration
is also compatible with the coproduct.
Example 14.3. For any Lie algebra g, the Hopf algebra 11 (g) is connected. lts
canonical filtration is given by the subspaces 'Un generated by the products
of at most n elements of g; this is compatible with the coproduct since
n
ß(X1 ... Xn) = n(XJ ® 1 + 1 ®XJ)
j=l
n
= I I (Xu(l) .. . Xu(k)) ® (Xu(k+l) .. . Xu(n)),
k=O UES(k)
14.3 The Milnor-Moore theorem 619

where S(k) denotes the set of "shuffle" permutations of {1, ... , n} such
that a(l) < · · · < a(k) and a(k + 1) < · · · < a(n).
The Milnor-Moore theorem [340] asserts that any cocommutative and
connected bialgebra is an enveloping algebra of a Ue algebra. For its proof,
we follow the argument of [42, Il.l.6]. We first establish a few properties of
primitive elements.
Lemma 14.10. Let B be a connected bialgebra. If a E Bn, then

ß(a) = a® 1 + 1 ®a +X, where x E Bn-1 ®Bn-1· (14.23)

Moreover, B 1 s;; IF1eP(B), where P(B) is the Lie algebra ofprimitive elements
ofB.

Proof. Since the filtration is compatible with the coproduct, we can write
ß(a) = b ® 1 + 1 ® c + y for some b,c E Bn and y E Bn-1 ® Bn-1· By the
counit property (1.29), a = (id®E)(ß(a)) = b + E(c)l + (id®E)(y), so that
a-b E Bn-1· Similarly, a-c E Bn-1· Thus, x := (a-b)® 1 + 1 ® (a-c) + y
lies in Bn-1 ® Bn-1·
Notice now that IFl nP(B) = 0, since (1.29) implies E(a) = 0 for a E P(B).
Thus, if a E B1r we can write ß(a) = a® 1 + 1 ®a+.\(1 ® 1) for some .\ E IF;
in consequence, a + .\1 is primitive, and thus a = -.\1 + (a + .\1) lies in
a E IF1 eP(B). D

In view of the previous lemma, it will be very convenient to consider the


map ß' defined on B by

ß' (a) := ß(a) - a ® 1 - 1 ® a.

In other words, ß' is the restriction of ß as a map ker E- ker E ® ker E. We


write B' := ker E and B~ := Bn n ker E. Note that aisprimitive if and only if
it lies in the kernel of ß'. Primitivity is a dual concept to indecomposability:
an element of B is indecomposable if it belongs to the cokernel of ker E ®
ker E - ker E. The set of indecomposable elements is algebraically of the
form ker E ®B IF, where B acts on the right through E.
lt should be clear that when B is a connected bialgebra, then B = Bo e
kerE = IFl e ker E. Indeed, if a E Bn. then by Lemma 14.10 and the counit
property, a = a + E(a)l + (id ®E)(x), and the last term lies in Bn-1 ®
E(Bn-d 1; we conclude by induction that E(a) = 0. Now one can use the
equation m o (id ®S) o ß = u o E to define S recursively by S(a) := -a +
m o (id ®S) o ß' (a), just as we have already donein (14.10).

Exercise 14.8. Show that if P(B) n B' 2 = {0}, then the connected bialgebra
B is commutative. 0
Exercise 14.9. Show that the map ß' is coassociative, and that it is cocom-
mutative when ß is cocommutative. 0
620 14. Kreimer-Connes-Moscovici Algebras

Proposition 14.11. Let rr: B - B' be the projection defined by rr ( a) : = a -


E(a) 1; then (TT ® TT) o ß = ß' o TT.

Proof If ß(a) = Lj aj ® aj, then Lj ajE(aj') = a = Lj E(aj)aj', and thus

(rr ® rr)(ß(a)) = 2>} ® aj'- E(aj)aj ® 1- 1 ® E(aj)aj + E(ajaj)l ® 1


j

= ß(a) - a ® 1 - 1 ® a + E(a) 1 ® 1,
= ß'(a)- E(a)ß'(l) = ß'(rr(a)},

since ß'(1) = -1 ® 1. D

Since rr(a) = a when a E B' and since rr(Bn) = B~, Lemma 14.10 and
Proposition 14.11 entail that

(14.24)

Given any connected bialgebra B, the universal property of enveloping


algebras gives an algebra morphism ] : 'U (P (B)) - B which extends the
canonical injection j: P(B) - B. The following Iemma shows that it is ac-
tually a bialgebra morphism.
Lemma 14.12. IfB is a bialgebra and g is any Lie algebra, and if tJJ: g - P (B)
is a Lie algebra homomorphism, then the unique extension 'I' of tJJ to 'U(g)
is a morphism of bialgebras.

Proof By definition, 'I' is an algebra homomorphism; we must show that it


is also a coalgebra morphism. Since tjJ(X) is primitive for each X E g,

('I'® 'l')(ß-u(g)(X)) = tjJ(X) ® 1 + 1 ® tjJ(X) = ßB('I'(X)).

Thus, ('I' ® 'I')


o ß-u(g) and ß 8 o 'I' are two algebra homomorphisms from
'U(g) toB® B that agree an g, so by the universal property of 'U(g), they
are equal. Similarly, EB o 'I' = E-u(g) since both are algebra homomorphisms
that agree (indeed, vanish) on g. D

Recall the symmetric algebra S(V) of a vector space V; it may be consid-


ered as the universal enveloping algebra 'U(V) where V is given the trivial
Ue bracket -see Section l.B. As such, it is a graded bialgebra: S(V) =
EB~; 0 sn(V), where S0 (V) = IF, S 1 (V) =V, and sn(V) is the vector space
generated by homogeneaus monamials of degree n. In fact, sn(V) is the
vector space spanned by the powers { vn : v E V}. When g is a Lie al-
gebra, there is an isomorphism of graded vector spaces (not of algebras)
p: S(g)- 'U(g), given by the symmetrization map:
14.3 The Milnor-Moore theorem 621

Notice that for powers of vectors, this reduces to p (Xn) = xn. The relation

~'U(g)(p(Xn)) = ~'U(g)(Xn) = ±(~)xk


k=O
®Xn-k = (p ® p)(~s<g>(Xn))
shows that p is a coalgebra isomorphism.
Lemma 14.13. IfC is a coalgebra, and if-F: S(V)- Cis a coalgebra mor-
phism whose restriction to IF e V is injective, then f is injective.

Proof. Let Sn := EBk=O sk (V); with this filtration, S (V) is a connected bial-
gebra. We shall prove inductively that -Fis" is injective. By hypothesis, this
is true for n = 1. Assume, then, that f is injective on Sn. and consider
x E Sn+ I suchthat f(x) = 0. Then

Since ~~<V> (x) E Sn ®Sn, in view of (14.24), it follows that ~~(V) (x) = 0,
so that x is primitive. Because S(V) is in particular an enveloping algebra,
Lemma 1.21 implies that x E S 1 (V) = V; consequently x = 0, and we
conclude that f is injective on Sn+ I· 0

Theorem 14.14 (Milnor-Moore). If B is a connected cocommutative bialge-


bra, then 'U(P(B)) ""'B.

Proof. We must show that the bialgebra morphism ]: 'U(P(B)) - Bis an


isomorphism. ByLemma 14.13, the coalgebra morphism] o p: S(P(B)) - B
is injective, and therefore ] is injective, too.
lt remains to prove that J is surjective. Let {adiei be a basis oftheUe
algebra P(B); then the ordered products ar = a~: ... a~:, with each Yj E ~.
form the corresponding Poincare-Birkhoff-Witt basis of 'U(P(B)), and so
the elements Ar := (1 /r!) ](ar) form a basis for the image of ].
We know that B = IF1 + U;:;'= 1 B~ since Bis connected. Thus it is enough
to prove that B~!:;;; im] for all n E ~- We do so by induction on n.
By Lemma 14.10, this claim is true for n = 1. Assurne now that B~ !:;;; im]
for some n and Iet a E B~+l· Then, as remarked after Proposition 14.11,
~~(a) E B~ ® B~. By the induction hypothesis,

~~(a)= L ,\(r,s)Ar®N
r,s'*'O

for some ,\(r, s) E IF; the sumruns over nonzero multiindices r,s E ~~ with
• 1y many nonzero entnes.
fi mte " s·mce U'U(P(B))
A (
ain) = 4-..k=O
,n (n) k aik ® ain-k , so
that
622 14. Kreimer-Connes-Moscovici Algebras

Y,S,t*O

(id®ß_B}(ß_B(a)) = L ,\(r,s + t)Ar ®A 5 ®At.


Y,S,t*O

The coassociativity of ß_B entails

,\(r + s, t) = A(r,s + t), (14.25a)

whereas the cocommutativity gives

,\(r,s) = ,\(s,r). (14.25b)

It follows from the relations (14.25) that i\(r, s) depends only on the sum
r+s. Indeed, ifr+s = t+u, we canfindmultiindices k, l, m, n with r = k+l,
s = m + n, t = k + m and u = l + n; for this "Riesz decomposition" can
be performed in each coordinate separately, since it is clearly feasible with
nonnegative integers. This means that, if k * 0, then

i\(r,s) = i\(k + l, m + n) = i\(k, l + m + n) = i\(k + m, l + n) = i\(t, u),


as claimed; while i\(r, s) = i\(l, m + n) = i\(l + n, m) = i\(u, t) = i\(t, u) if
k = 0. Therefore, we can write i\(r,s) =: p(r + s). In particular,

ß_B(a) = L p(r + s) Ar® N = L p(t) ß_B(At),


Y,S*Ü ltl;;,;2

hence b := a- Liti;;,; 2 Jl(t)At is primitive, and on that account b = j(b) =


](b). Therefore, a = b + :L 1t 1;;,; 2 p(t) At lies in im]. D

Thus, a connected cocommutative bialgebra is automatically a Hopf al-


gebra. In particular, the Grossman-Larson bialgebra Her is a Hopf algebra.
Naturally, when one drops the connectedness, the Hopf algebra will no
Iongerbe an enveloping algebra; but, strikingly, Kostant [290] generalized
this theorem by proving that any cocommutative Hopf algebra is the smash
product -see Definition 1.29- of a group algebra and an enveloping alge-
bra [446]. Fora modern treatment, see [347, §5.6].

14.4 Duality in Hopf algebras


For our present purposes, it is convenient to formulate duality in a rather
weak fashion, as a bilinear pairing between two Hopf algebras.
14.4 Duality in Hopf algebras 623

Definition 14.7. We say that two bialgebras BandCarein duality if there


is an IF-bilinear form ( ·, ·) on B x C such that, for all b, b' E B and c, c' E C,

(bb',c) = (b ® b',ßc(c)), (b,c) = Ec(c),


(b, cc') = (ßB(b),c ® c'), (b, 1c) = EB(b). (14.26a)

If B and C are Hopf algebras, we also require compatibility with the an-
tipodes:

(SB(b),c) = (b,Sc(c)). (14.26b)

The duality is called separating if (b, c) = 0 for all c E C implies b = 0,


and (b, c) = 0 for all b E B implies c = 0.
The conditions (14.26a) mean that the maps cp: B- C* and tfJ: C- B*
defined by cp(b): c ,_ (b,c) and tfJ(c): b ,_ (b,c) are algebra homomor-
phisms. The duality is separating when the homomorphisms cp and tfJ are
injective. When B* and C* are actually bialgebras -as happens, for in-
stance, in finite-dimensional cases- (14.26a) implies that cp and tfJ are coal-
gebra morphisms, too.
Exercise 14.10. Verify all the Statements of the previous paragraph. 0
Example 14.4. Let G be a compact Ue group with neutral element 1 and let
g be its Ue algebra. By regarding the elements of g as left-invariant vector
fields on G, the Hopf algebras 'U(g) and 'R.(G) are in dualityvia the following
bilinear pairing:

(X,j) := Xj(l) = dd
t
It=O j(exp tX), for XE g, f E 'R.(G).

and more generally, by (XI .. . Xn,f> := XI(··· (Xnf>)(1); also, (1,j) :=


j(1). Weshall check the conditions (14.26) only on the generators of U(g).
Since X E g acts as a derivation,

(X,jh) = Xj(l)h(l) + j(l)Xh(1) =(X® 1 + 1 ® X)(j ® h)(1 ® 1)


= (ßX)(j ® h)(1 ® 1) = (ßX,j ® h).
Since 1 E 'R.(G) is a constant function, (X, 1) = 0 = E(X). Moreover,

(X®Y,ßj) = dd I dd I (ßj)(exptX®expsY)
t t=O 5 s=O

= dd I dd
t t=O S s=O
I
j(exptXexpsY) = dd
t
It=O (Yj)(exptX)
= X(Y j)(e) = (XY,j).
624 14. Kreimer-Connes-Moscovici Algebras

Since (l,f} = j(l) = E(j) also, 'U(g) and R(G) are in duality as bialgebras.
Finally, since S(X) = -X,

(S(X),f} = dd I j(exp(-tX)) = dd I S(j)(exptX) = (X,S(j)),


t t=O t t=O

so they are in duality as Hopf algebras. Since representative functions are


realanalytic[287], (D,f) = OforallD E U(g)impliesthatfhasavanishing
Taylor series at 1 and thus f = 0. On the other band, if (D, f> = 0 for all
f, the left-invariant differential operator determined by Dis null and thus
D = 0 in 'U(g). Therefore, the duality is separating. Moreover, it follows
from Lemma 1.27 that representative functions belang to the Sweedler dual
of 'U(g). In fact, R(G) ""'U(g)~ if g is semisimple.
lf H is a real skewgroup, then, by the Tannaka-Krein theorem, H ""
R(§'(H)). Indeed, if cp E §'(H), then mt(cp)(a®b) = cp(ab) = cp(a)cp(b) =
(cp ® cp)(a ® b), so cp belongs to the Sweedler dual H~ and is group-like.
The same calculation shows that group-like elements are algebra homo-
morphisms, and thus §'(H) = G(H~). the group of group-like elements
of H~!
Definition 14.8. A linear functional o: H - IF on a Hopf algebra His called
a derivation if O(ab) = o(a)E(b) + E(a)O(b) for all a, b E H. (The ter-
minology is justified, in the light of Definition 8.1, by regarding IF as an
H-bimodule via h · i\ · k := E(h)i\E(k).) Notice that 8(1) = 0 necessarily.
Also,
(o,ab) = O(a)E(b) + E(a)o(b) = (8 ® 1 + 1 ® o,a ® b),
so that mt (8) = o ® 1 + 1 ® o. Thus, o belongs to H~ and is primitive there.
Thesederivations form a vector subspace Derc(H) of P(H~).
Exercise 14.11. Show that the commutator [8, y] := (8 ® y- y ® o) o ß of
two derivations is a derivation. Conclude that Derc (H) is a Lie subalgebra
of P(H~). 0

Proposition 14.15. Let H be a connected Hopf algebra with filtration {Hn},


and Iet 8 E Derc(H). Then on+ 1(a) = 0 for all a E Hn.

Proof. Write ß 2 := (ß ® id) o ß = (id ®ß) o ß and, more generally, denote


by ßn: H - H®(n+l) the n-fold iteration of the coproduct. The convolution
power on+ 1 is defined as (8 ® · · · ® o) o ßn. This is immediate when n = 1,
and inductively
on+2 := (0 ® on+1) 0 ß = o®(n+2) 0 (id®ßn) 0 ß = o®(n+2) 0 ßn+1.

Since Ho = IF1 by connectedness, o(a) = 0 for a E Ho. Suppose a E H1;


then, by Lemma 14.10, ß(a) = a ® 1 + 1 ® a + i\ 1 ® 1 for some scalar i\,
and thus o 2(a) = (0 ® o)(a ® 1 + 1 ® a + i\1 ® 1) = 0 since 8(1) = 0.
lnductively, if a E Hn and (14.23) holds, then on+ 1(a) = (8 ® on) (x) = 0
since XE Hn-1 ® Hn-1· 0
14.4 Duality in Hopf algebras 625

In view also of Example 14.4, one should think of derivations as "infini-


tesimal characters". Therefore, when His connected and 8 E DerE(H), we
define 4> := exp 8 in H* by the usual exponential series: for each a E H,
the series exp 8 (a) has only a finite number of terms. Moreover, since 8 is
primitive,

so that 4> is a group-like element of H~, that is, a character of H. As a


consequence, a type of Campbell-Baker-Hausdorff formula can be written
for the product of characters. In particular, the simplest cases of the CBH
formula for H~ suggest that GR is nilpotent; the proof of that, however, is
beyond the scope of this book.
.,.. By now, there is a strong motivation to look for the dual of the Connes-
Kreimer Hopf algebra of rooted trees as a suitable enveloping algebra. It is
easy to construct a Hopf algebra in duality with HR, since we have a canon-
ical system of generators for HR: we associate to each rooted tree T the
derivation ZT: HR - IF defined by ZT(Tl ... Tk) := 0 unless the monomial
T1 ... Tk isT itself, in which case ZT(T) := 1. We denote by ij ~ DerE(H) the
linear span of the elements ZT·
The commutator, in DerE(H), of two such derivations can be computed
explicitly as follows. Let CI(T} denote the set of simple one-line cuts of
the tree T and let C2 ( T) be all its other simple cuts. Notice that Pc ( T) is a
single tree if and only if c E C1 (T). Now,

ß(T) = T ® 1 + 1 ® T + L Pc(T) ® Rc(T) + L Pc(T) ® Rc(T).


CECdTJ CEC2(TJ

Since ZT kills constants and monamials which are products of two or more
trees, the product of two such derivations satisfies

(ZRZs, T) :=(ZR® Zs,ß(T)) = L ZR(Pc(T)) Zs(Rc(T)).


CEC1(TJ

Each summand on the right hand side vanishes unless both Pc ( T) = R and
Rc(T) = S. Therefore,
(ZRZs, T) = n(R,S, T),

where n(R, S, T) is the number of (one-line) cuts c ofT suchthat Pc (T) = R


and Rc(T) = S. Thus, ([ZR, Zs], T) = n(R, S, T)- n(S,R, T). Evaluation of
the derivation [ZR, Zs] on a product T1 ... Tk of two or more trees gives
zero, since each E ( T1 ) = 0. Therefore,

[ZR,Zs] = L(n(R,S,T) -n(S,R,T))ZT, (14.27)


T
626 14. Kreimer-Connes-Moscovici Algebras

where the sum runs over all trees T that can be obtained by grafting R
onto S or vice versa. This is the Lie bracket introduced by Connes and
Kreimer in [106].
It is immediate from (14.27) that fj is a Lie subalgebra of Deri'(H). More-
over, as derivations, the elements Zr areprimitive in H~. This shows that
the coproduct on 'U(fj) induced by its inclusion in H~ coincides with the
universal algebra coproduct. In this way, 'U(fj) is in (separating) duality
withHR.
~ The promised "Lie algebra of the group GR ", then, is fj. However, envelop-
ing algebras are somewhat awkward to work with. Now, it has been pointed
out by Panaite [363] that the duality between 'U(fj) and HR can be given a
nice and concrete description as a matehing of trees: 'U(fj) is precisely the
Grossman-Larson Hopf algebra of rooted trees!
Theorem 14.16 (Panaite). The Hopf algebras Her and'U(fj) are isomorphic.
Proof. As a connected, cocommutative bialgebra, Her is, by the Milnor-
Moore theorem, isomorphic to the enveloping algebra of its Lie algebra
of primitive elements, P(Her). Consider the linear map !J1: fj - P(Her)
defined by !Jl(Zr) := L(T). It is bijective, since the trees of the form L(T)
span the primitive elements of Her.
The map !J1 is actually a Lie algebra homomorphism. First of all,

[!/l(ZR), !/l(Zs)] = [L(R),L(S)] := L(R) · L(S)- L(S) · L(R).

The root of L(R) has only one outgoing line, and cutting this line leaves the
subtree R, so L(R) · L(S) is the sum of trees formed by attaching R to a
vertex of L(S). If that vertex is the root of L(S), the resulting tree is L(RS).
Otherwise, we obtain a tree of the form L(T), where T itself is formed by
attaching R to some vertex of S; if c is the cut of T that severs this joining
line, then Pc(T) =Rand Rc(T) = S. Therefore,

L(R) · L(S) = L(RS) +I n(R, S, T) L(T).


T

It follows that

[!Jl(ZR), !Jl(Zs)] = L(RS)- L(SR) + I<n(R,S, T)- n(S,R, T)) L(T)


T

= I<n(R,S,T) -n(S,R,T))!Jl(Zr) = !Jl([ZR,Zs]).


T

In summary, !J1 is a Lie algebra isomorphism. By the universal property


of enveloping algebras, the canonical extension 'I': 'U(fj) - 'U(P(Her)) is
an isomorphism of algebras. Lemma 14.12 then shows that 'I' is a bialgebra
isomorphism, and Proposition 1.24 ensures that it is indeed a Hopf algebra
isomorphism. 0
14.5 Hopf algebras of Feynman diagrams 627

At this point, the reader should try his hand at defining and establishing
the main properties of a dual algebra for HcM, in particular the existence
of a surjective Lie algebra morphism from fj to such a dual algebra [106].

14.5 Hopf algebras of Feynman diagrams


This section tries to take account of developments still taking place while
this book was essentially written within the framework of the previous
sections. For more detail, we refer to the original papers [107, 108, 208].
Readers unfamiliar with Feynman diagrams may wish to skip this section,
anyway.
Consider the integrand formally corresponding to any given Feynman
graph r appearing in a perturbative quantum field theory. If we dilate all
loop momenta in the integrand by a common factor .\, the integral will be-
have, as .\ - oo, like some integer power w(f) of .\, called the superficial
degree of divergence of the diagram. To each graph we associate the family
of its proper connected subdiagrams (precise definitions are given below).
A theorem in renormalization theory asserts that (the Feynman integral cor-
responding to) r is convergent if w(;y) < 0 for all elements ;y ofthat family;
note that r can be divergent even though w ([) < 0, when it contains super-
ficially divergent subdiagrams. The main idea of renormalization theory is
to associate a "counterterm" to each superficially divergent (sub)diagram,
in order to obtain a finite result by subtraction.
We begirr by explaining some graph terminology which we need. A graph,
or diagram, of the theory is specified by a set of vertices and a set of lines
among these vertices; externallirres are attached to only one vertex each,
internal lines to two. Diagrams with no external lines are not taken into
account.
A diagram is connected when any two of its vertices are joined by lines,
of course. Given a graph r, a subdiagram ;y of r is specified by selecting
two or more vertices of r, as well as a subset of the lines in r that join
those vertices. Clearly, the external lines for ;y include not only some of
the original incident lines but also some internallirres of r which are not
irrtemal to ;y. The empty subset 0 will be admitted, exceptionally, as a
subdiagram of r; and the whole graph r is considered to be a subdiagram
of itself, too. The connected components of r are its maximal connected
subdiagrams. A diagram is proper when the removal of a single internal
line would not increase the nurober of its connected pieces; otherwise, it is
called improper. An improper graph is the union of its proper components
plus subdiagrams containing a single line.
When working in configuration space [208], it suffices to consider sub-
graphs, instead of the more general subdiagrams just defined. A subgraph
of a proper graph is a subdiagram containing all the propagators that join
628 14. Kreimer-Connes-Moscovici Algebras

its vertices in the whole graph (i.e., a "full" subdiagram); as such, it is de-
termined solely by the vertices. A subgraph of an improper graph r, other
than 0 and r itself, is a proper subdiagram each of whose components is
a subgraph with respect to the proper components of r: in other words, a
product of subgraphs. We write y ~ r if and only if y is a subgraph of ras
here defined: this will be the important concept for us.
Given y ~ r, the quotient graph or cograph r IY is defined by shrinking
y in r to a vertex, that is to say, y (bereft of its externallines) is considered
as a vertex of r, and alllines in r not belanging to this amputated y belang
to r I y. The graphs r and r I y have the same external structure.
Let j(f) denote the distribution or "unrenormalized integral" corres-
ponding to a diagram r. Weshall assume that r is connected, since
f(YI 1.tJ • • • 1.tJ Yn) = f(yi) · · .f(Yn)
whenever Y1, ... , Yn are disjoint graphs. lf r is superficially divergent, and
r is not primitive (that is to say, it does have subdivergences), we Iet
Rrf([) := (1-Sr)Rrf(f),
where S denotes here the subtraction procedure and the operation Rr re-
flects the renormalization of the subdivergences present in j(f). Now, Bo-
goliubov's recursive formula [38] for Rr is known tobe
Rrf([) := - L (SyRyj(y)) j(f ly), (14.28)
0!";y!;;[

where the sum is taken over all proper, superficially divergent, not neces-
sarily connected subgraphs, -S0 = 1 and j(f IY) is just defined by the
splitting j(y)j(f ly) := j(f). Then Ry and also C(y) := -SyRyj(y) are
recursively defined.
A graph without subdiagrams of the kind just defined is called primitive.
This recursive formula for renormalization gives rise, for a proper con-
nected diagram r, to the following nonrecursive formula:

Rrf([) = (1 + L n (-Sy))J<n.
E([) yEE([)
(14.29a)

where E(f) runs over all nonempty sets whose elements y are proper, di-
vergent, not necessarily connected subdiagrams made of subgraphs of r
(possibly including r itself), and YI, Y2 E E(f) implies that either y 1 ~ Y2 or
Y2 ~ YI· When Y1 ~ Y2. the order of the subtractions is · · · Sy2 • • • Sy, · · ·.
When the counterterm map C verifies
C(yl 1.t1 • • • 1.t1 Yn) = C(yi) ... C(yn),

the previous formula can be rewritten as

Rrf([) = (1 + L n
F([)yEF([)
(-Sy))J<n. (14.29b)
14.5 Hopf algebras of Feynman diagrams 629

with F(f) now denoting nonempty sets whose elements y are proper, di-
vergent, and connected subgraphs of r, possibly including r itself, and
:YI, :Yz E F ([) implies that :YI ~ :Y2 or :Y2 ~ :YI or eise that :YI and :Y2 are
disjoint; the order of subtractions is as before. In other words, we then
sumover forests in the sense of Zimmermann [500].
We are now ready to prove the compatibility of the formulae a Ia Bogo-
liubov with the Hopf algebra picture. First, we observe that if y s; y' s; r,
then y' IY is naturally interpreted as a subdiagram of r jy; moreover, it is
obvious that

(f/y)/(y'/y) c:::.[jy'. (14.30)

The algebra H is defined as the polynomial (hence commutative) algebra


generated by the empty set 0 and the connected Feynman graphs that are
(superficially) divergent and/or have (superficially) divergent subdiagrams,
with set union as the product operation; thus, 0 is its unit element.
We only need to define the coproduct on connected diagrams. The co-
product of a graph r is defined to be

~r := 2: y ® r 1y. (14.31)
0s;ys;f

The sum extends over all divergent, proper, not necessarily connected sub-
diagrams of r which are either subgraphs or products of subgraphs, such
that each piece is divergent, including the empty set and r itself (exception-
ally, since r need be neither divergent nor proper). If y = 0, we write that
term instead as 1 ® f; when y = r, we write its term as r ® 1.
If r has no nontrivial subdiagrams, then this graph will be primitive in
both senses, that of quantum field theory and the Hopf algebra sense.
For explicit, pictorial calculations of coproducts, we refer to [208] for <1>4
theory; the appendix of [108] exhibits other examples in <I>~ theory.
The counit E : H - Cis given by E ([) : = 0 for all graphs and E ( 0) := 1. All
the Hopf algebra properties are easily verified, except for coassociativity,
which is of course the sine qua non; but that is quite straightforward.
Proposition 14.17. The algebra of graphs H is a Hopf algebra.

Proof. We deal only with coassociativity, leaving the rest to the reader. We
need to show that

(~ ® id)(~(f)) = (id ®~) (~([) ),

for every connected graph r. Using (14.31), this putative equality may be
rewritten as

y ® y" ® (f /)1)/:y".
630 14. Kreimer-Connes-Moscovici Algebras

We must then show that, for each subgraph y of r,

2: y' I y ® r I y' = 2: y" ® (f IY)Iy".

Choose y' so that y ~ y' ~ f; then 0 ~ y' IY ~ f /y. Reciprocally, to every


y" ~ r I y corresponds a y' such that y ~ y' ~ r and y' 1y y". The
relation (14.30) now provides the required matching. D

It follows directly from Definition 1.25 and (14.31) that characters of H


are convolved by

g *f ([) = g([) + 2: g(y) j([ IY) + j([).


0o;;yo;r

Assuming that the linear map C is multiplicative, the renormalization for-


mula
Rrf([) =: R(f) = C(f) + 2:
C(y) j(f ly) + j([)
0o;;yo;r

is simply recast, on dropping the variable tag r, as

Bogoliubov's renormalization maps R are characters of the Hopf algebra of


graphs H. They are the Hopf convolution of the (unrenormalized) Feynman
graph homomorphism and a counterterm homomorphism. This opens the
door to the application of the Lie-theoretical methods of (co)commutative
Hopf algebra theory, developed in the earlier sections, to the renormalized
theory [108]. In particular, R-maps differing in the choice of subtractions
are related by elements of a huge "renormalization group" sitting inside
the group of characters of H.
Multiplicativity of C was proved in the context of dimensional regular-
ization by Connes and Kreimer in [107] and for Epstein-Glaser renormal-
ization [162] by Lazzarini and one of the authors in [208].
In principle, by systematizing the methods of [78] in the Hopf algebra
framework, one could extend this construction to field theories on non-
commutative spaces, in the sense of Section 13.A.

14.6 Hopf algebras and diffeomorphism groups


The natural growth operatorNon trees, discussed in Section 14.1, gives
rise to a Hopf subalgebra of the extended Hopf algebra HR of rooted trees.
At the end of that section, we indicated how this subalgebra may be char-
acterized abstractly, as a Hopf algebra, by commutator and coproduct rela-
tions on three algebra generators. The latter, here denoted HcM. originally
14.6 Hopf algebras and diffeomorphism groups 631

appeared in connection with a local index computation in the noncommu-


tative geometry of foliations [114]. Although the theory of foliations lies
outside the scope of this book, we shall now briefly indicate how they give
rise to the Hopf algebra in question.
One wishes to study the geometrical invariants of the action of a group
of diffeomorphisms (or, more generally, a pseudogroup of local diffeomor-
phisms) on an oriented manifold M. This is conceptually an important
first step in dealing with gravity in a noncommutative geometric frame-
work [97,98]. We confine the discussion to the one-dimensional case, where
the Hopf algebra structure already emerges quite clearly.
Consider the problern of trying to build a noncommutative geometry on
an orbit space M /G, where the homomorphism oc: G - Diff+ (M) gives
the group action by orientation-preserving diffeomorphisms. The difficulty
that M /G may have a poor topology is solved by replacing the C*-algebra
Co(M /G) with the crossed product C(M) ~ocG. (As discussedin Sections 4.5
and 12.4, in favourable cases these algebras are Morita-equivalent.) In the
smooth category, we may use the algebra .Jlo := C"" (M) ~oc G-see below.
A more se~ious difficulty is how to construct a suitable representation for
this algebra; this would require a covariant representation of (C"" (M), G, oc)
on some Hilbert space L2 (M, f.1), but if G is large M might not carry any G-
invariant measure.
The second problern is surmounted by the following procedure [99, 114].
One can replace M by its oriented frame bundle F+ - M, lift the action
of G to the frame bundle, find an invariant measure v on F+ for the lifted
action öc, and represent on L 2 (F+, v) the crossed product
(14.32)
where C;' (F+) denotes the compactly supported smooth functions on F+.
(When dimM > 1, one should actually use the bundle of Riemannian met-
rics over M, which is the quotient of the frame bundle under a natural fi-
brewise action of the group SO(n); but it is handier to deal with the frame
bundle first and restriet to SO(n)-invariant quantities afterwards.) lt then
becomes possible to build a noncommutative spin geometry in this frame-
work [99]. The Hopf algebra we seek describes the effect of the symmetries
of the frame bundle on the algebra .Jl.
The one-dimensional example consists in taking M tobe the line ~ or the
circle §1 (our longstanding assumption of compactness does not matter
much here), with linear or angular coordinate x, say. Theorientedframe
bundle is the (trivial) principal bundle P = ~ x ~ + or § 1 x ~ +, with vertical
coordinate y =e-s. The whole group G = Diff+ (~) or Diff+ (§ 1 ) acts on F+
by
(j)(x,y) := (1/J(X),IJ/(x)y) = (.X,ji),
or, alternatively, (j)(x, s) := ( !JJ(x), s -log !JJ' (x) ). It is clear that the 2-form
y- 2 dy 1\ dx = es ds 1\ dx
632 14. Kreimer-Connes-Moscovici Algebras

is invariant under each (j), so it defines aG-invariant measure 11 on p+.


The vertical vector field Y := o/os = yo/oy in J€(F+), which generates
the fibrewise action of the group ~ + on the frame bundle, is invariant und er
the lifted action of t/J. Indeed, (j)*Y = t/J'(x)yo/oy = yo/oy = Y. The
canonical1-form oc := y- 1 dx is also invariant:
$* oc = y- 1 dx = y- 1 dx = oc.
On the other hand, the 1-form w := y- 1 dy is not invariant, since
~ t/J" (x)
t/J*w = y- 1 dy + t/J'(x) dx = y- 1 dy + t/J'(x)- 1 dejJ'(x).
This transformation rule is characteristic of a principal connection 1-form;
vector fields satisfying w(X) = 0 are called horizontal.
The vector field X := y oI ox is horizontal and also fulfils oc(X) = 1. It
is not G-invariant; indeed, oc($; 1X) = oc(X) = 1 by invariance of oc, while
$*w($; 1X) = w(X) = 0, implying
~-1 a 2 t/J"(x) a .
t/J* X= Y ox - Y t/J'(x) oy =.X- hi/JY, (14.33)

where hi/J(x,y) := y t/J"(x)/t/J'(x).


The crossed product algebra (14.32) can be described by using a faithful
covariant representation-on L 2 (F+ ,y- 2 dy dx), say. Therefore, we may
regard .JI. as the linear span of elements fUt, where f E c; (F+) and
{ Ul/1 : t/J E G} is a group of unitary operators; we use the adjoint operators
for computational convenience. The covariance property
UI/JfUt = jl/1, where ji/J(x,y) := j(lj)- 1 (x,y)),
shows that the product of two such elements is given by

uut><ou~> = Jwtoui/J)Utu~ = J<o o (j))U~I/1'


that is, by the smash product (1.3 7). This rule also makes sense if G consists
only of local diffeomorphisms, provided that
supp(j(g o lj))) c Dom(/J n (/J- 1 (Dom</J) ~ Dom(</Jt/J).
To each vector field Z E *(F+) we can associate a linear operator on .JI.,
also denoted by Z, by setting
zuut> := <Zf> ut. (14.34)
Proposition 14.18. The operator Y corresponding to the vertical vector field
is a derivation; whereas X, corresponding to the horizontal vector field, sat-
isfies
X(ab) = X(a)b + aX(b) + AI(a)Y(b) for all a, b E .JI., (14.35)
where A1 is also a derivation on .JI..
14.6 Hopf algebras and diffeomorphism groups 633

Proof. For any vector field Z, we find that

zuutgul) = Z(f(g o $))ul!JI = (Zf)(g o $)Ul!JI + JZ(g o $)Ul!JI


= (Zf)Ul gUl + fUl (Z(g o $) o qJ- 1 )Ul
= (Zj)Utgul + JUt $*Z(g)Ul.
Since $* Y = Y, the operator Y is a derivation on .Jl:

On the other hand,

Using again the invariance of Y and (14.33),

and thus

Therefore,

where

(14.36)

To find the behavior of .\ 1 on products, we first note that

The invariance of oc then implies that

which establishes that .\ 1 is a derivation:

AdfUtgul) = ($*hct> + h!JI)f(g o $)Ul!JI


= f ((hcpg) o $)Ul!JI + h!Jifutgul
=tut hcpgul + h!Jifut gul. o
634 14. Kreimer-Connes-Moscovici Algebras

We examine the commutators of X, Y and A1. First,

[Y,,\I](jUt) = j(YhiJI)Ut,
and YhiJI = hiJI, thus [Y,,\d = A1. Similarly, [X,,\d(JUt) = j(XhiJI)Ut. so
,\z := [X, ,\ 1 ] is the operator of left multiplication by XhiJI. In general, there
is a family of operators An inductively defined by An+ I := [X, An], where
one multiplies by h'J := xn-l (hiJI ), so that [Am, An] = 0. From X= y o jox
and 1/J"(x)/l/J'(x) = ojox(logljJ'(x)), wegetat once

hn n dn l ' ) (14.37)
1/1 = y dxn og ljJ (x .

lt follows that Y h'J = nh'J and thus [ Y, An] = n,\n. Moreover, from the
relation [y ojoy,y ojox] = y ojox and (14.34), it follows that [Y,X] =X
as operators on 5'1. •
.,.. At this point, we have constructed an action on 5'1. of the algebra gene-
rated by the operators X, Y and A1, which camplies with the commutator
relations (14.16). It is therefore a representation of HcM. as an algebra, by
operators on 5'1..
Proposition 14.18 says that this is actually also a Hopf action: the primi-
tive elements Y and A1 of HcM act by derivations, and the Leibniz rule for
the action of Xis precisely (14.35). The remark after Definition 1.28 shows
that it suffices to verify the Leibniz rules on these algebra generators.
Exercise 14.12. In any case, check directly, from ,\ 2 = [X, ,\I), that

,\z(ab)=,\z(a)b+a,\z(b)+,\I(a),\I(b), forall a,bE.JI.. 0

It remains to characterize the antipode in this context. Thinking of dif-


feomorphisms ljJ of ~ (modulo affine diffeomorphisms, if one wishes) as
flows at the origin recasts them as elements of the character group. Then,
in view of (14.36) and (14.37), there is a functional of the generators An on
Diff+ ( ~) given by

Then l/J(An) = ljJ- 1 (5(,\n)) yields the antipode in HcM· Therefore, yet an-
other method for computing antipodes in HcM consists in expressing ljJ (x)
as a series with coefficients given by the An, using (13.28), making a series
reversion [240, 302] to obtain the series corresponding to ljJ- 1 (x), and dif-
ferentiating again to get the new coefficients .
.,.. Over higher-dimensional manifolds, this Hopf algebra construction may
also be carried out [114, 496]. If dimM = n, the frame bundle will carry
n 2 independent vertical vector fields Yj, generating the fibrewise action of
GU(n, ~); and there willbenhorizontal vector fields Xk. determined by
14.7 Cyclic cohomology of Hopf algebras 635

ai (Xk) = 8{ and w}(Xk) = 0, where the ai := (y- 1 )~ dxll are components


of a canonical1-form, and w} := (y - 1 )~ ( d y'j + r~ßY j dxß) are components
of a connection 1-form. (There is some trouble due to the nonflatness of
this connection in general: the matter is discussed in [99].) The coproduct
relations for the Xk, acting on the algebra 5\ of (14.32), take the form

The operators Xi, Y( and ,\~i generate a Hopf algebra H~~ with a tauto-
logical Hopf action on 5\. In [496], Wulkenhaar explains how these higher-
dimensional Connes-Moscovici algebras can also be described in terms of
rooted trees, by decorating the vertices with Iabels corresponding to the
indices of .\~j· We refer tothat paper for the details.

14.7 Cyclic cohomology of Hopf algebras


The Hopf algebra HcM and its higher-dimensional counterparts arose, as
described in the previous section, in the course of dealing with certain natu-
ral endomorphisms of the crossed product algebras of foliations. The (gen-
eralized) Leibniz rules obtained on Iifting vector fields to endomorphisms
of those algebras indicate dearly that Hopf actions are involved, and the
coproduct is in fact dictated by those Leibniz rules.
The next step in the Connes-Moscovici program was to use these Hopf ac-
tions to efficiently compute local index formulae. There is a spectral triple
(5\,Jf,D) naturally associated with a homomorphism a: G- Diff+(M),
where 5\ is the algebra (14.32) and DIDI = FD 2 is a second-order hypoel-
liptic differential operator on the frame bundle over M; for its description,
we refer to [113]. One can then set up the corresponding local index for-
mula, as stated in Theorem 10.33; to compute the cocyde components 'Pn
of (10.78), one must first replace D by Q = DIDI, taking advantage of the
homotopy t .... DIDit which leaves the result unchanged, and then evalu-
ate the nonzero terms in the resulting formula. It turns out that the num-
ber of terms involved is very large: several thousand for even the simplest
cases [114]! In an (ultimately successful) effort to shrink the calculation
to manageable proportions, Connes and Moscovici proved that the desired
dass [<p],\ in HC(5l) could be identified as a canonical image of a cydic
cohomology dass over the Hopf algebra.
Here we describe briefly how the cydic cohomology of a Hopf algebra H
is constructed, and how an action on a Hopf H-module algebra 5\ allows
us to build a "characteristic map" from HC(H) to HC(5l). Foramore
complete account, one may consult [116].
We ought to start with the "unimodular" situation wherein the algebra 5\
comes equipped with a distinguished trace <p (we reserve the Ietter T for
636 14. Kreimer-Connes-Moscovici Algebras

the upcoming cyclic operator), which is invariant under the Hopf action,
that is, <p(h · a) = E(h) <p(a) for all h E H, a E .Jl.. Our chief example,
however, does not comply with these requirements. It does indeed possess
a distinguished trace, given by

qJ(j) := f
F+
f dp,

where 11 is the invariant measure on F+. However, it turns out that

cp(h · a) = 8(h) <p(a), (14.38)

where 8: HcM - (( is the character of HcM defined on generators by 8 ( Y) :=


1, 8(X) = 8(.\ 1 ) := 0. (Note that X and .\1 are commutators.) Weshall say
that the linear functional qJ is 8-invariant if (14.38) holds.
The character 8 on HcM (which, after all, is the algebra 'U(gcM) with a
new coproduct) plays a role analogous to that of the modular function of
a locally compact group. On the other hand, when H is a locally compact
quantum group in the sense of Kustermans and Vaes [300], there isanother
candidate forthat role, namely the "modular element" CT EH, a group-like
element which satisfies a relation of the form

<p(ab) = cp(b(CT · a)), for all a,b E .Jl., (14.39)

where qJ is now no Ionger a trace, but only anH-invariant positive linear


functional on .Jl. (the Haar integral). We shall call qJ a CT-trace if (14.39)
holds. (These are closely related to the KMS-functionals of von Neumann
algebra theory [267, 366].) Consistency with (14.38) requires that 8(CT) = 1,
as is seen by taking b = 1.
We thereby arrive at the following definition [115], which codifies the
concept of modulus for Hopf algebras.
Definition 14.9. A modular pair (8, CT) for a Hopf algebra H consists of a
character 8: H - IF and a group-like element CT E G (H), suchthat 8 ( CT) = 1.
On convolving u o 8 with the antipode S of H, we obtain an algebra an-
tihomomorphism Sö: H - H given by

Sö(h) := ((u o 8) * S)(h) = LJ 8(hj) S(h'j).


called a "twisted antipode" [114, 123].
Exercise 14.13. Verifythat Sö *id = u o8, and that EoS8 = 8 and 8 oS8 = E.
Show also the coproduct relation

where ß(h) = LJ hj ® h'j, as usual. 0


14.7 Cyclic cohomology of Hopf algebras 637

The mapping S8 behaves in many ways like an antipode, although neither


S nor S8 need be involutive. For instance,

If cp is a <5-invariant functional, then

Lj cp((hj · a)(h'j · b)) = cp(h ·(ab))= 8(h) cp(ab)


= cp(a 8(h)b) = Lj cp(a (S8(hj)h'j · b)),

*
where the last equality comes from S8 id = u o 8. Replacing h'j · b by b
and hj by h leads to an alternative formulation of 8-invariance, equivalent
to (14.38) if .J\ is unital, which uses the twisted antipode:

cp((h · a) b) = cp(a (S8(h) · b)),

a Hopf-algebra formulation of "integration by parts" .


.,.. Cyclic cohomology on the Hopf algebra His defined by building a suit-
able double complex BC", with cochain maps band B, as in Section 10.1.
The formal definitions run exactly parallel to what was done there for al-
gebras. First, we define cn(H) := H®n for n = 1, 2, ... and C0 (H) := IF.
Then we take BCPq := CP-q(H) for p ~ q ~ 0 and BCPq := 0 other-
wise. On C (H) = EB;;; 0 cn (H) we introduce maps b of degree + 1 and
B of degree -1, defined so that b 2 , B2 and bB + Bb all vanish; the desired
cyclic cohomology HC (H) is then just the cohomology of the total com-
plex (Tot• BC, b + B).
To produce maps b and B with these properties, all we need to do is to
gift C (H) with the structure of a 1\-module, where J\ is the cyclic category.
That is to say, we must find maps

8i: cn-l(H)- cn(H), CTj: cn+l(H)- cn(H), T: cn(H)- cn(H),


(14.40)

for i, j = 0, 1, ... , n, satisfying the relations (10.10) and (10.12) which spec-
ify a functorial image of J\. Then we define b: cn- 1 (H) - cn(H) and
B: cn+ 1 (H)- cn(H) by the same formulae as in Section 10.1:
n
b ==I (-1)i8i.
i;Q

where N := Lk;o( -l)nkTk is the cyclic antisymmetrizer on cn(H). The


task, then, is to find a good set of maps (14.40); these will depend on a
modular pair (8, er).
lt is easy enough to make C (H) a cosimplicial object by introducing
the maps 8i and CTj, using the coalgebra structure of Hand the modular
638 14. Kreimer-Connes-Moscovici Algebras

element a only. If n ~ 1, we define

8o(h1 ® . . . ® hn-1) := 1 ® h1 ® •.. ® hn-1,

8i(h 1 ® · · · ® hn- 1) := h 1 ® · · · ®ß(hi) ® · · · ® hn- 1, i = 1, ... ,n -1,


8n(h1 ® . . . ® hn-1) := hl ® . . . ® hn-1 ® a.

For n = 0, with C0 (H) = IF, it suffices to define

8o(1) := 1, 81 (1) := a.

The relation (10.10a), namely, 8J8i = 8i8J-I for i < j, is trivial unless
j = i + 1, whereupon it expresses the coassociativity of ß and the group-
like nature of 1 and a. The operators aJ are defined using the counit: if
n = 0, we define a 0 (h) := E(h) for h EH; otherwise,
aj(hl ® . . . ® hn+l) := hl ® .•. ® hlE(hj+l) ® . . . ® hn+l

= E(hj+l) hl ® . . . ® hl ® hj+2 ® . . . ® hn+l.

Exercise 14.14. Verify the remaining relations in (10.10) among the 8i and
a1, using the coalgebra properties of H. 0

One might think that the obvious cyclic permutation of the tensor factors
in H'Fm would yield maps T making C" (H) into a A-module; and so indeed it
does, provided that one replaces H®n by H®(n+l). Butthis is clearlynotwhat
we need, since the algebra structure of H, the antipode Sand the character 8
are never used. Detailed consideration of Hopf actions in [114) led instead
to the following proposal for the cyclic endomorphism of cn(H):

(14.41)

where ßn-I: H - H®n is the (n- 1)-fold iterated coproduct. If we write


ßn-I (k) = Ir k~ ® k~ ® · · · ® k~nl, taking account of the coassociativity,
then
ßn- 1 (S0 (k)) = IrS(k~nl) ® • • · ®S(k~) ®So(k~),
with S applied to all factors but the last.
The cyclicity property Tn+I = idcn is not obvious, to say the least. In fact,
even when n = 1, it imposes a requirement on the modular pair (8, a):

h = T2 (h) = T(So(h) a) = So(So(h) a) a


= a- 1 s~(h)a = a- 1so(a- 1So(h)).

Following [115, 116), we say that a modular pair (a, 8) is in involution if

(14.42)
14.7 Cyclic cohomology of Hopf algebras 639

This is a necessary condition for the maps T, defined by (14.41) for n =


1, 2, ... , to satisfy the relations (10.12) and thus to determine a cyclic struc-
ture. Remarkably, it turnsout that it is also a sufficient condition [115].
To see how the full structure of the Hopf algebra and its involutive mod-
ular pair are needed to establish the cyclic structure, we give the calculation
for n = 2. We require a couple of auxiliary identities. First, for any h EH,

LiS 2 (h'j) So(hj) = Lr S 2 (h~') 8(h~) S(h~) = Lr 8(h~) S(h~S(h~'))


= Lj 8(hj) S(E(h'j) 1) = Lj 8(hj E(h'j)) 1 = 8(h) 1.
(14.43)
Secondly, linearity of S 0 implies that

S~(h) = LjSo(8(hj)S(h'j)) = Lj8(hj)SoS(h'j). (14.44)

Lemma 14.19. The endomorphism Ton H ® H satisfies T 3 = 1.

Proof. Let h, k E H; we must show that T 3 (h ® k) = h ® k. First of all,

T(h ® k) = ß(So(h))[k ® a] =Li S(h;') k ® So(h;) a.

Since ß(S(l)k) = Li,j S(l'j) k; ® S(lj) k;', a second application ofT gives

T 2 (h ® k) = Li,r S(k;')S 2 (h~)So(h~) a ® So(k;)soS(h~') a


= Li,i S(k;') a ® 8(hj) So(k;)soS(hj') a

= Li,j S<k;') a ® So(k;)s~(h) a


=Li S(k;') a ® So(k;) ah,

where we have used (14.43), (14.44) and the involution property (14.42), in
turn. A third application of T then yields
T 3 (h ® k) = Lr S(S(k~) a) So(k~) ah ® So(S(k~') a) a

= Lr (T-I S 2 (k~) So (k~) ah ® {T-l SoS (k~') (T

=Li 8(kj) h ® a- 1 SoS(kj') a


= Lj h ® a- 1 so(8(kj) S(k'j)) a

=h ® a- 1 s~(k) a =h ® k.

Herewe have used (14.43) again and the involution property, to finish. D

.,.. We can now establish the correspondence between the cyclic cohomolo-
gies of H and its Hopf module algebra .Jl which arises from any 8-invariant
a-trace cp. Foreach n = 1,2, ... , we define a linear map y<P: cn(H) -
cn (.Jl, .Jl *) as follows:

Y<P[h 1 ® · · · ® hn](ao, .. . , an):= cp(ao (h 1 · a1) ..• (hn ·an)). (14.45)


640 14. Kreimer-Connes-Moscovici Algebras

It is easy to checkthat Y<P intertwines the maps 8i and a 1 defined on the


two cochain complexes. For instance,
Y<P[8n(h 1 ® · · · ® hn)](ao, ... ,an+d
= y<P[h 1 ® • · • ® hn ® a](ao, ... , an+d

= <P(ao (h 1 · a!) ... (hn ·an) (a · an+I))


= <P(an+lao (h 1 • a!) ... (hn ·an))
= 8n(Y<P[h 1 ® • · • ® hn])(ao, ... ,an+d·

The matehing of the cyclic actions is accomplished by noting that


Y<P[T(k ® h 2 ® • · · ® hn)] = Y<P[~n-l(Sö(k))(h 2 ® · · • ® hn ® a)]
= 2..1 Y<P[~n- 2 (S(kj'))(h 2 ® • · • ® hn) ® Sö(kj) a].

We apply this to (ao, ... , an). abbreviating b := (h 2 • a1) ... (hn ·an_ I); the
8-invariance and a-tracial property of <P conspire to give
2..1 <P(ao (S(kj') · b) Sö(kj) · (a ·an))= 2..1 <P(kj · (aoS(kj') · b) a ·an)
=Ir <P(an (k~ · ao) (k~ S(k~') · b))
= 2..1 <P(an (kj · ao) E(kj')b) = <P(an (k · ao) b)
= <P(an (k · ao) (h 2 • a!) ... (hn · an-d),

and thus Y<P[T(k ® h 2 ® • • · ® hn)] = T(y<P[k ® h 2 ® · • · ® hn]).


In retrospect, the origin of (14.41) becomes clear: by reading the last
calculation backwards, we see that T is predetermined by the condition
that Y<r o T = T o Y<P for any <P·
Definition 14.10. The map Y<r: HC (H) - HC (.J\.) induced by (14.45) is
called the characteristic map associated to the 8-invariant a-trace <P·
The result which ties all these threads together asserts that the cyclic
cohomology classes entering in the statement of the local index theorem
lie, in fact, in the range of a particular characteristic map. This is Propo-
sition IX.3 of [114]. This opens a new road toward index theory, through
Hopf algebra territory. Apart from the founders, a few adventurous souls
have already travelled that road. To mention one instance, Perrot [3 70] has
used these methods to compute explicitly the Chern character of the K-
cycle corresponding to the local conformal action of a discrete group on a
Riemann surface.
The cyclic cohomology has been computed for several known Hopf alge-
bras; some early results are surveyed in [116]. For the motivating case of
algebras obtained by the actions of diffeomorphism groups, with a = 1, it
is known [114] that the periodic cyclic cohomology of H~~ can be identi-
fied with the Gelfand-Fuchs cohomology of the Lie algebra of formal vector
fields on IR n. Much more remains to be learned in the years to come.
References

[1] R. Abraham, j. E. Marsden and T. Ratiu, Manifo/ds, Tensor Analysis, and Ap-
plications, Springer, Berlin, 1988.
[2] T. Ackermann, "A note on the Wodzicki residue", j. Geom. Phys. 20 (1996),
404-406.
[3] M. Adler, "On a trace functional for formal pseudo-differential operators and
the symplectic structure of the Korteweg-de Vries type equations", Invent.
Math. 50 (1979), 219-248.
(4] P. M. Alberti and R. Matthes, ''Connes' trace formula and Dirac realization of
Maxwell and Yang-Mills action", math-ph/9910011 (v2), Leipzig, 1999.
[5] E. Alvarez, J. M. Gracia-Bondia and C. P. Martin, "Parameter constraints in a
noncommutative geometry model do not survive Standard quantum correc-
tions", Phys. Lett. B 306 (1993), 55-58.
[6] E. Alvarez, J. M. Gracia-Bondia and C. P. Martin, "Anomaly cancellation and
the gauge group of the Standard Model in NCG", Phys. Lett. B 364 (1995),
33-40.
[7] 0. Alvarez, I. M. Singer and P. Windey, "Quantum mechanics and the geo-
metry of the Weyl character formula", Nucl. Phys. B 337 (1990), 467-486.
[8] H. Araki, "Bogoliubov automorphisms and Fock representations of canonical
anticommutation relations", in OperatorAlgebrasand Mathematical Physics,
P. E. T. j0rgensen and P. S. Muhly, eds., Contemp. Math. 62 (1987), 23-141.
[9] W. B. Arveson, "Notes on extensions of C*-algebras", Duke Math. j. 44 (1977),
329-356.
[10] W. B. Arveson, "The harmonic analysis of automorphism groups", Proc.
Symp. Pure Math. 38 (1982), 199-269.
642 References

[11] M. F. Atiyah, "K-theory and reality", Quart.]. Math. 17 (1966), 367-386.


[12] M. F. Atiyah, K-theory, Benjamin, NewYork, 1967.
[13] M. F. Atiyah, "Bott periodicity and the index of elliptic operators", Quart.].
Math. 19 (1968), 113-140.
[14] M. F. Atiyah, "Global theory of elliptic operators", in Proceedings of the In-
ternational Symposium on Functional Analysis, Univ. of Tokyo Press, Tokyo,
1969; pp. 21-30.
[15] M. F. Atiyah, R. Bott and A. Shapiro, "Clifford modules", Topology 3 (1964),
3-38.
[16] M. F. Atiyah and I.M. Singer, "The index of elliptic operators. III", Ann. Math.
87 (1968), 546-604.
[17] M. F. Atiyah and I. M. Singer, "Index theory for skewadjoint Fredholm ope-
rators", Publ. Math. IHES 37 (1969), 305-326.
[18] S. Baaj and P. Julg, "Theorie bivariante de Kasparov et operateurs non
bornees dans les C*-modules hilbertiens", C. R. Acad. Sei. Paris 296 (1983),
875-878.
[19] C. Bär, "The Dirac operator on space forms of positive curvature", ]. Math.
Soc. Japan 48 (1996), 69-83.
[20] A. P. Balachandran, G. Bimonte, G. Landi, F. Lizzi and P. Teotonio-Sobrinho,
"Lattice gauge fields and noncommutative geometry", ]. Geom. Phys. 24
(1998), 353-385.
[21] R. G. Bartle and L. M. Graves, "Mappings between function spaces", Trans.
Amer. Math. Soc. 72 (1952), 400-413.
[22] M. S. Bartlett and]. E. Moyal, "The exact transition probabilities of quantum-
mechanical oscillators calculated by the phase-space method", Proc. Cam-
bridge Philos. Soc. 45 (1949), 545-553.
[23] P. Baum and R. G. Douglas, "Index theory, bordism and K -homology", in Ope-
rator Algebras and K-Theory, R. G. Douglas and C. Schochet, eds., Contemp.
Math. 10 (1982), 1-31.
[24] F. Bayen, M. Flato, C. Fronsdal, A. Lichnerowicz and D. Sternheimer, "Defor-
mation theory and quantization. II: Physical applications", Ann. Phys. (NY)
111 (1978), 111-151.
[25] R. W. Beals and P. C. Greiner, Calculus on Reisenberg Manifolds, Princeton
Univ. Press, Princeton, NJ, 1988.
[26] E. Bedos, "An introduction to 3D discrete magnetic Laplacians and noncom-
mutative 3-tori",]. Geom. Phys. 30 (1999), 204-232.
[27] W. Beer, "On Morita equivalence of nuclear C*-algebras",]. Pure Appl. Alge-
bra 26 (1982), 249-267.
[28] ]. Bellissard, "Quantized fields in interaction with external fields I. Exact
solutions and perturbative expansions", Commun. Math. Phys. 41 (1975),
235-266.
[29] ]. Bellissard, A. van Elst and H. Schulz-Baldes, "The noncommutative geo-
metry of the quantumHall effect",]. Math. Phys. 35 (1994), 5373-5451.
References 643

[30) S. K. Berberian, Lectures in Functiona/ Analysis and Operator Theory, Spring-


er, Berlin, 1974.
[31) V. B. Berestetskii, E. M. Lifshitz and I. P. Pitaevskii, Quantum Electrodynamics
(volume 4 of the Landau and Lifshitz Course of Theoretical Physics), Perga-
mon Press, Oxford, 1980.
[32) F. A. Berezin, The Method of Second Quantization, Academic Press, New York,
1966.
[33) N. Berline, E. Getzler and M. Vergne, Heat Kerneis and Dirac Operators,
Springer, Berlin, 1992.
[34) R. Bhatia, Matrix Analysis, Springer, Berlin, 1997.
[35) L. C. Riedenharn and J. D. Louck, Angular Momentum in Quantum Physics:
Theory and Applications, Addison-Wesley, Reading, MA, 1981.
[36) B. Blackadar, K -theory for Operator Algebras, 2nd edition, Cambridge Univ.
Press, Cambridge, 1998.
[37) D. P. Blecher, "On Morita's fundamental theorem for C*-algebras", math/
9906082, Houston, 1999; to appear in Math. Scand.
[38) N. N. Bogoliubov and D. V. Shirkov, Introduction to the Theory of Quantized
Fields, Wiley, New York, 1980.
[39) P. ]. M. Bongaarts and J, Brodzki, "Möbius transformations in noncom-
mutative conformal geometry", Commun. Math. Phys. 201 (1999), 35-60;
physics/971 0005.
[40) J,-B. Bost, "Principe d'Oka, K-theorie et systemes dynamiques non commu-
tatifs", Invent. Math. 101 (1990), 261-333.
[41) R. Bott and L. W. Tu, Differential Forms in Algebraic Topology, Springer,
Berlin, 1982.
[42) N. Bourbaki, Groupes et Algebres de Lie, Hermann, Paris, 1972.
[43) J.-P. Bourguignon and P. Gauduchon, "Spineurs, operateurs de Dirac et vari-
ations de metriques", Commun. Math. Phys. 144 (1992), 581-599.
(44) L. J. Boya, "The geometry of compact Lie groups", Rep. Math. Phys. 30 (1991),
149-162.
[45) L. J, Boya, "Representations of simple Lie groups", Rep. Math. Phys. 32 (1993),
351-354.
[46) L. J, Boya and M. Byrd, "Clifford periodicity from finite groups", J, Phys. A
32 (1999), L201-L205; math-ph/9902013.
[47] J,-P. Brasselet, A. Legrand and N. Teleman, "Hochschild homology of algebras
of controlled functions", preprint, Ancona, 1998.
[48) 0. Bratteli, "Inductive limits of finite dimensional C* -algebras", Trans. Amer.
Math. Soc. 171 (1972), 195-234.
[49) 0. Bratteli, G. A. Elliott and P. E. T. ]0rgensen, "Decomposition of unbounded
derivations into invariant and approximately inner parts", J, reine angew.
Math. 346 (1984), 166-193.
644 References

[50) R. Brauerand H. Weyl, "Spinors in n dimensions", Amer.]. Math. 57 (1935),


425-449.
[51) A. Brill and M. Noether, "Die Entwicklung der Theorie der algebraischen Func-
tionen in älterer und neuerer Zeit", Bericht der DMV 3 (1894), 107-566.
[52) D. ]. Broadhurst and D. Kreimer, "Towards cohomology of renormalization:
bigrading the combinatorial Hopf algebra ofrooted trees", hep-th/0001202,
Milton Keynes and Mainz, 2000.
[53) J. Brodzki, An Introduction to K -theory and Cyclic Cohornology, Polish Scien-
tific Publishers (PWN), Warszawa, 1998.
[54) T. Bröcker and T. tom Dieck, Representations of Cornpact Lie Groups,
Springer, Berlin, 1985.
[55) R. Brooks and C. S. Gordon, "lsospectral families of conformally equivalent
Riemannian metrics", Bull. Amer. Math. Soc. 23 (1990), 433-436.
[56) C. Brouder, "Runge-Kutta methods and renormalization", Eur. Phys.]. C 12
(2000), 521-534; hep-th/9904014.
[57] L. G. Brown, P. Green and M. A. Rieffel, "Stahle isomorphism and strong
Morita equivalence of C*-algebras", Pac.]. Math. 71 (1977), 349-363.
[58) P. Budinich and A. Trautman, The Spinorial Chessboard, Springer, Berlin,
1988.
[59) H. Bursztyn and S. Waldmann, "*-Ideals and formal Morita equivalence of
*-algebras", math/0005227, Berkeley, CA, 2000.
[60) J. C. Butcher, "An algebraic theory of integration methods", Math. Comp. 26
(19 72), 79-106.
[61) M. Cahen, S. Gutt and M. de Wilde, "Local cohomology of the algebra of coo
functions on a connected manifold", Lett. Math. Phys. 4 (1980), 157-167.
[62) A. P. Calderön, "The analytic calculation of the index of elliptic equations",
Proc. Natl. Acad. Sei. USA 57 (1967), 1193-1194.
[63) R. Camporesi and A. Higuchi, "On the eigenfunctions of the Dirac operator
on spheres and real hyperbolic spaces", J. Geom. Phys. 20 (1996), 1-18.
[64) A. Cannas da Silva and A. Weinstein, Geometrie Models for Noncornrnutative
Algebras, Berkeley Math. Lecture Notes 10, Amer. Math. Soc., Providence, RI,
1999.
[65) A. L. Carey, C. A. Hurst and D. M. O'Brien, "Automorphisms of the canonical
anticommutation relations and index theory",]. Funct. Anal. 48 (1982), 360-
393.
[66) A. L. Carey and D. M. O'Brien, "Abscence of vacuum polarization in Fock
space", Lett. Math. Phys. 6 (1982), 335-340.
[67) A. L. Carey and D. M. O'Brien, "Automorphisms oftheinfinite dimensional
Clifford algebra and the Atiyah-Singer mod 2 index", Topology 22 (1983),
437-448.
[68) A. L. Carey and]. Palmer, "Infinite complex spin groups",]. Funct. Anal. 83
(1989), 1-43.
References 645

[69] J. F. Cariftena and H. Figueroa, "Geometrie formulation of higher-order La-


grangian systems in supermechanics", Acta Appl. Math. 51 (1998), 25-58.
[70] J. F. Cariftena, C. Lopez and E. Martinez, "Sections along a map applied to
higher-order Lagrangian mechanics", Acta Appl. Math. 25 (1991), 127-151.
[71] S. Carlip and C. DeWitt-Morette, "Where the sign of the metric makes a dif-
ference", Phys. Rev. Lett. 60 (1988), 1599-1601.
[72] A. Cayley, "On the theory of the analytical forms called trees", Phil. Magazine
13 (1857), 172-176. Also in Collected Mathematical Papers of Arthur Cayley,
Vol. 3, Cambridge Univ. Press, Cambridge, 1889; pp. 242-246.
[73] M. Cederwall, G. Ferretti, B. E. W. Nilsson and A. Westerberg, "Schwinger
terms and cohomology of pseudodifferential operators", Commun. Math.
Phys. 175 (1996), 203-220.
[7 4] M. Chaichian, A. Demichev and P. Presnajder, "Quantum field theory on non-
commutative spacetimes and the persistence of ultraviolet divergences",
Nucl. Phys. B 567 (2000), 360-390; hep-th/9812180.
[75] A. H. Chamseddine and A. Connes, "Universal formula for noncommutative
geometry actions: Unification of gravity and the Standard Model", Phys. Rev.
Lett. 77 (1996), 4868-4871.
[76] A. H. Chamseddine and A. Connes, "The spectral action principle", Commun.
Math. Phys. 186 (1997), 731-750.
[77] V. Chari and A. Pressley, A Guide to Quantum Groups, Cambridge Univ. Press,
Cambridge, 1994.
[78] I. Chepelev and R. Roiban, "Renormalization of quantum field theories on
noncommutative IRd, 1: Scalars", J. High Energy Phys. 0005 (2000), 037; hep-
th/9911098.
[79] S. Cho, R. Hinterding, J. Madore and H. Steinacker, "Finite field theory on
noncommutative geometries", hep-th/9903239, Chechon, Korea, 1999; to
appear in lnt. J. Mod. Phys. D.
[80] E. Christensen and A. M. Sinclair, "A survey of completely bounded opera-
tors", Bull. London Math. Soc. 21 (1989), 417-488.
[81] F. Cipriani, D. Guido and S. Scarlatti, "A remark on trace properties of K-
cycles", J. Oper. Theory 35 (1996), 179-189.
[82] L. Conlon, Differentiable Manifolds: A First Course, Birkhäuser, Boston, 1993.
[83] A. Connes, "C*-algebres et geometrie differentielle", C. R. Acad. Sei. Paris
290A (1980), 599-604.
[84] A. Connes, "Spectral sequence and homology of currents for Operator al-
gebras", Tagungsbericht 42/81, Mathematisches Forschungszentrum Ober-
wolfach, 1981.
[85] A. Connes, "Cohomologie cyclique et foncteurs Ext"", C. R. Acad. Sei. Paris
296 (1983), 953-958.
[86] A. Connes, "Noncommutative differential geometry", Publ. Math. IHES 39
(1985), 257-360.
646 References

[87) A. Connes, "Entire cyclic cohomology of Banach algebras and characters of


9-summable Fredholm modules", K-Theory 1 (1988), 519-548.
[88) A. Connes, "The action functional in non-commutative geometry", Commun.
Math. Phys. 117 (1988), 673-683.
[89) A. Connes, "Trace de Dixmier, modules de Fredholm et geometrie Rieman-
nienne", Nucl. Phys. B (Proc. Suppl.) 58 (1988), 65-70.
[90) A. Connes, "Compact metric spaces, Fredholm modules, and hyperfinite-
ness", Ergod. Thy. & Dynam. Sys. 9 (1989), 207-220.
[91) A. Connes, Noncommutative Geometry, Academic Press, London and San
Diego, 1994.
[92] A. Connes, "Geometry from the spectral point of view", Lett. Math. Phys. 34
(1995), 203-238.
[93) A. Connes, private communication, 1995.
[94) A. Connes, "Noncommutative geometry and reality", ]. Math. Phys. 36 (1995),
6194-6231.
[95) A. Connes, Cours au College de France, Paris, ]anuary- March 1996.
[96] A. Connes, "Gravity coupled with matter and foundation of noncommutative
geometry", Commun. Math. Phys. 182 (1996), 15 5-176.
[97] A. Connes, "Brisure de symetrie spontanee et geometrie du point de vue
spectral", Seminaire Bourbaki, 48eme annee, Expose 816, 1996; ]. Geom.
Phys. 23 (1997), 206-234.
[98) A. Connes, "Noncommutative differential geometry and the structure of
spacetime", in Cyc/ic Cohomology and Noncommutative Geometry, ]. ]. R.
Cuntz and M. Khalkali, eds., Fields Institute Communications 17, Amer.
Math. Soc., Providence, RI, 1997; pp. 17-42.
[99] A. Connes, "Hypoelliptic operators, Hopf algebras and cyclic cohomology",
in Algebraic K -theory and its Applications, H. Bass, A. 0. Kuku and C. Pedrini,
eds., World Scientific, Singapore, 1999; pp. 164-205.
[100] A. Connes, "Trace formula in noncommutative geometry and the zeroes of
the Riemann zeta function", Selecta Math. 5 (1999), 29-1 06; math/9811 068.
[101] A. Connes, "A short survey of noncommutative geometry", ]. Math. Phys. 41
(2000), 3832-3866; hep-th/0003006.
[102) A. Connes, private communication, 2000.
[103) A. Connes, M. R. Douglas and A. Schwartz, "Noncommutative geometry
and Matrix theory: compactification on tori", JHEP 2 (1998), 003; hep-
th/9711162.
[104) A. Connes and N. Higson, "Deformations, morphismes asymptotiques et K-
theorie bivariante", C. R. Acad. Sei. Paris 311 (1990), 101-106.
[105) A. Connes and M. Karoubi, "Caractere multiplicatif d'un module de Fred-
holm", C. R. Acad. Sei. Paris 299 (1984), 963-968.
[106) A. Connes and D. Kreimer, "Hopf algebras, renormalization and non-
commutative geometry", Commun. Math. Phys. 199 (1998), 203-242; hep-
th/9808042.
References 64 7

[107] A. Connes and D. Kreimer, "Renormalization in quantum field theory


and the Riemann-Hilbert problern 1: the Hopf algebra structure of graphs
and the main theorem", Commun. Math. Phys. 210 (2000), 249-273; hep-
th/9912092.
[108] A. Connes and D. Kreimer, "Renormalization in quantum field theory and
the Riemann-Hilbert problern II: the ß-function, diffeomorphisms and the
renormalization group", hep-th/0003188, IHES, Bures-sur-Yvette, 2000.
[109] A. Connes and J. Lott, "Particle models and noncommutative geometry",
Nucl. Phys. B (Proc. Suppl.) 18 (1990), 29-47.
[110] A. Connes and H. Moscovici, "Transgression du caractere de Chern et coho-
mologie cyclique", C. R. Acad. Sei. Paris 303 (1986), 913-918.
[111] A. Connes and H. Moscovici, "Cyclic cohomology, the Novikov conjecture
and hyperbolic groups", Topology 29 (1990), 345-388.
[112] A. Connes and H. Moscovici, "Transgression of the Chern character of finite-
dimensional K-cycles", Commun. Math. Phys. 155 (1993), 103-122.
[113] A. Connes and H. Moscovici, "The local index formula in noncommutative
geometry", Geom. Func. Anal. 5 (1995), 174-243.
[114] A. Connes and H. Moscovici, "Hopf algebras, cyclic cohomology and the
transverse index theorem", Commun. Math. Phys. 198 (1998), 198-246;
math/98061 09.
[115] A. Connes and H. Moscovici, "Cyclic cohomology, Hopf algebras and the
modular theory", math/9905013, IHES, Bures-sur-Yvette, 1999.
[116] A. Connes and H. Moscovici, "Cyclic cohomology and Hopf symmetry",
math/0002125, IHES, Bures-sur-Yvette, 2000.
[117] A. Connes and M. A. Rieffel, "Yang-Mills for noncommutative two-tori", in
Operator Algebras and Mathematical Physics, P. E. T. ]0rgensen and P. S.
Muhly, eds., Contemp. Math. 62 (1987), 237-266.
[118] R. Coquereaux, "Modulo 8 periodicity of real Clifford algebras and particle
physics", Phys. Lett. B 115 (1982), 389-395.
[119] R. Coquereaux, G. Esposito-Farese and G. Vaillant, "Higgs fields as Yang-Mills
fields and discrete symmetries", Nucl. Phys. B 353 (1991), 689-706.
[120] R. Coquereaux, A. 0. Garcia and R. Trinchero, "Associated quantum vector
bundles and symplectic structure on a quantum plane", math-ph/9908007,
CPT,Luminy, 1999.
[121] R. Coquereaux and E. Ragoucy, "Currents on Grassmann algebras", J. Geom.
Phys. 15 (1995), 333-352.
[122] G. Corach and A. R. Larotonda, "Stahle range in Banach algebras", J. Pure
Appl. Algebra 32 (1984), 289-300.
[123] M. N. Crainic, "Cyclic cohomology of Hopf algebras and a noncommutative
Chern-Weil theory", math/9812113, Utrecht, 1998.
[124] J. J. R. Cuntz, "K-theory and C*-algebras", in Algebraic K-theory, Number
Theory, Geometry and Analysis, A. Bak, ed., Lecture Notes in MathemaUes
1046, Springer, Berlin, 1984; pp. 55-79.
648 References

[125] j. j. R. Cuntz and N. Higson, "Kuiper's theorem for Hilbert modules", in


Operator Algebras and Mathematical Physics, P. E. T. )0rgensen and P. S.
Muhly, eds., Contemp. Math. 62 (1987), 429-434.
[126] j. j. R. Cuntz and 0. Quillen, "Algebra extensions and nonsingularity", j.
Amer. Math. Soc. 8 (1995), 251-289.
[127] H. L. Cycon, R. G. Froese, W. Kirschand B. Simon, Schrödingeroperators with
applications to quantum mechanics and global geometry, Springer, Berlin,
1987.
[128] M. Dädärlat, "C*-algebras and applications", in Representation Theory of
Groups and Algebras, j. Adamset al., eds., Contemp. Math. 145 (1993), 393-
421.
[129] K. R. Davidson, C* -algebras by Example, American Mathematical Society,
Providence, Rl, 1996.
[130] S. Oe Bievre, "Chaos, quantization and the classicallimit on the torus", in Pro-
ceedings ofthe XIVth Workshop on Geometrical Methods in Physics, Bialowieza
1995, Polish Scientific Publishers (PWN), Warszawa, 1998.
[131] j. A. Oieudonne, Foundations of Modern Analysis, Academic Press, New York,
1969.
[132] j. A. Dieudonne, Treatise on Analysis, Vol. 2, Academic Press, New York,
1970.
[133] j. A. Dieudonne, Treatise on Analysis, Vol. 3, Academic Press, New York,
1972.
[134] j. A. Dieudonne, Elements d'Analyse, Vol. 9, Gauthier-Villars, Paris, 1982.
[135] j. A. Dieudonne, A History ofAlgebraic and Differential Topology 1900-1960,
Birkhäuser, Boston, 1989.
[136] j. Dixmier, "Existence de traces non normales", C. R. Acad. Sei. Paris 262A
(1966), 1107-1108.
[137] j. Dixmier, Les C*-algebres et leurs Representations, Gauthier-Villars, Paris,
1969.
[138] j. Dixmier, Algebres Enve/oppantes, Gauthier-Villars, Paris, 1974.
[139] A. Dold, Halbexakte Homotopiefunktoren, Lecture Notes in Mathematics 12,
Springer, Berlin, 1966.
[140] A. Dold, Lectures on Algebraic Topology, Springer, Berlin, 1980.
[141] J. D. Dollard and C. N. Friedman, Product Integration with Applications to
Differential Equations, Addison-Wesley, Reading, MA, 1979.
[142] W. F. Donoghue, Distributionsand Fourier Transforms, Academic Press, New
York, 1969.
[143] S. Doplicher, K. Fredenhagen and j. E. Roberts, "The quantum structure of
spacetime at the Planck scale and quantum fields", Commun. Math. Phys.
172 (1995), 187-220.
[144] R. S. Doran and V. A. Belfi, Characterizations of C* -Algebras: the Gelfand-
Naimark Theorems, Marcel Dekker, New York, 1986.
References 649

[145) H. G. Dosch and V. F. Müller, "Renormalization of quantum electrodynamics


in an arbitrarily strong time independent external field", Fortschr. Phys. 23
(1975), 661-689.
[146) R. G. Douglas, C* -algebra Extensionsand K -homology, Princeton Univ. Press,
Princeton, NJ, 1980.
[147) T. Dray, "The relationship between monopole harmonics and spin-weighted
spherical harmonics",]. Math. Phys. 26 (1985), 1030-1033.
[148) M. Dubois-Violette, "Derivations et calcul differentiel non commutatif", C. R.
Acad. Sei. Paris 307 (1988), 403-408.
[149) M. Dubois-Violette, R. Kernerand]. Madore, "Gauge bosons in a noncommu-
tative geometry", Phys. Lett. B 217 (1989), 485-488.
[150) ]. Dugundji, Topology, Allyn and Bacon, Boston, 1966.
[151) S. Echterhoff, Crossed Products with Continuous Trace, American Mathemat-
ical Society, Providence, Rl, 1996.
[152) E. G. Effros, "Some quantizations and reflections inspired by the Gelfand-
Naimark theorem", in C*-A/gebras: 1943-1993: A Fifty Year Ce/ebration, R.
S. Doran, ed., Contemp. Math 167, 99-113; Amer. Math. Soc., Providence, Rl,
1994.
[153) E. G. Effros and F. Hahn, Locally Compact Transformation Groups and C*-
a/gebras, Amer. Math. Soc., Providence, Rl, 1967.
[154) S. Eilenberg and N. Steenrod, Foundations of Algebraic Topology, Princeton
Univ. Press, Princeton, NJ, 1952.
[155) E. Elizalde, G. Cognola and S. Zerbini, "Applications in physics of the mul-
tiplicative anomaly formula involving some basic differential operators",
Nucl. Phys. B 532 (1998), 407-428; hep-th/9804118.
[156) G. A. Elliott, "On the classification of inductive Iimits of sequences of
semisimple finite dimensional algebras",]. Algebra 38 (1976), 29-44.
[157) G. A. Elliott, "On the K-theory of the C*-algebra generated by a projective
representation of a torsion-free discrete abelian group", in Operator Alge-
brasand Group Representations 1, G. Arsene, ed., (Pitman, London, 1984),
pp. 157-184.
[158) G. A. Elliott and D. E. Evans, "The structure of the irrational rotation C*-
algebra", Ann. Math. 138 (1993), 477-501.
[159) G. A. Bliott, T. Natsume and R. Nest, "The Atiyah-Singer index theorem as
passage to the classicallimit in quantum mechanics", Commun. Math. Phys.
182 (1996), 505-533.
[160) K. Bsner, "Elektroschwaches Modell und Standardmodell in der nichtkom-
mutativen Geometrie", Diplomarbeit, Universität Marburg, 1999.
[161) G. G. Emch, "Chaotic dynamics in noncommutative geometry", in Quantiza-
tion, Coherent States and Poisson Structures, A. Strasburger et al., eds., Polish
Scientific Publishers (PWN), Warszawa, 1997.
[162) H. Epstein and V. Glaser, "The role of locality in perturbation theory", Ann.
Inst. Henri Poincare A 19 (1973), 211-295.
650 References

[163) E. Ercolessi, G. Landi and P. Teotonio-Sobrinho, "K·theory of noncommuta-


tive lattices", K-Theory 18 (1999), 339-362.
[164) R. Estrada, "The Cesaro behaviour of distributions", Proc. Roy. Soc. London
A 454 (1998), 2425-2443.
[165) R. Estrada, "Regularization of distributions", Int. J. Math. & Math. Sei. 21
(1998), 625-636.
[166) R. Estrada and S. A. Fulling, "Distributional asymptotic expansions of spec-
tral functions and of the associated Green kernels", Electron. ]. Diff. Eqns.
1999:7 (1999), 1-37; funct-an/9710003.
[167) R. Estrada, ]. M. Gracia-Bondia and ]. C. Värilly, "On asymptotic expansions
of twisted products", ]. Math. Phys. 30 (1989), 2789-2796.
[168) R. Estrada, ]. M. Gracia-Bondia and ]. C. Värilly, "On summability of distri-
butions and spectral geometry", Commun. Math. Phys. 191 (1998), 219-248;
funct-an/9702001.
[169) R. Estrada and R. P. Kanwal, "Regularization, pseudofunction and Hadamard
finite part", ]. Math. Anal. Appl. 141 (1989), 195-207.
[170] R. Estrada and R. P. Kanwal, Asymptotic Analysis: a Distributional Approach,
Birkhäuser, Boston, 1994.
[171) R. Exel, "A Fredholm operator approach to Morita equivalence", K-Theory 7
(1993), 285-308.
[172) B. V. Fedosov, "Index of an elliptic system on a manifold", Funct. Anal. Appl.
4 (1970), 312-320.
[173) B. V. Fedosov, Deformation Quantization and Index Theory, Akademie Verlag,
Berlin, 1996.
[174) B. V. Fedosov, F. Golse, E. Leichtman and E. Schrohe, "The noncommutative
residue for manifolds with boundary", J. Funct. Anal. 142 (1996), 1-31.
*
[175] J. M. G. Fell and R. S. Doran, Representations of -Algebras, Locally Compact
*
Groups and Banach -Algebraic Bund/es, Vol. 1, Academic Press, New York,
1988.
[176) B. Felsager, Geometry, Partie/es and Fields, Springer, Berlin, 1998.
[177) H. Fierz and G. Scharf, "Particle interpretation for external field problems in
QED", Helv. Phys. Acta 52 (1979), 437-453.
(178) H. Figueroa, "Function algebras under the twisted product", Bol. Soc. Paran.
Mat. 11 (1990), 115-129.
(179) H. Figueroa and J. M. Gracia-Bondia, "On the antipode of Kreimer's Hopf
algebra", hep-th/9912170, San Jose, 1999.
[180) H. Figueroa, J. M. Gracia-Bondia, F. Lizzi and J. C. Värilly, "A nonperturbative
form of the spectral action principle in noncommutative geometry", J. Geom.
Phys. 26 (1998), 329-339; hep-th/9701179.
[181) H. Figueroa, ]. M. Gracia-Bondia and J. C. Värilly, "Moyal quantization with
compact symmetry groups and noncommutative harmonic analysis", J.
Math. Phys. 31 (1990), 2664-2671.
References 651

[182) T. Filk, "Divergences in a field theory on quantum space", Phys. Lett. B 376
(1996), 53-58.
[183) P. A. Fillmore, A User's Guide to Operator Algebras, Wiley, New York, 1996.
[184) G. B. Folland, Harmonie Analysis in Phase Space, Princeton Univ. Press,
Princeton, 1989.
[185) G. B. Folland, "Harmonie analysis of the de Rham complex on the sphere", J.
reine angew. Math. 398 (1989), 130-143.
[186) M. Frank, "The Standard Model-the commutative case: spinors, Dirac ope-
rator and de Rham algebra", math-ph/0002045, Leipzig, 2000.
[187] M. Frank and D. R. Larson, "A module frame concept for Hilbert C*-
modules", Contemp. Math. 247 (2000), 207-233.
[188) D. S. Freed, "An index theorem for families of Fredholm operators parame-
trized by a group", Topology 27 (1988), 279-300.
[189) D. S. Freed, "Review of 'The heat kernel Lefschetz fixed point formula for the
Spine Dirac operator' by j. J. Duistermaat", Bull. Amer. Math. Soc. 34 (1997),
73-78.
[190) T. Friedrich, Dirac-Operatoren in der Riemannschen Geometrie, Vieweg,
Braunschweig/Wiesbaden, 1997.
[191) j. Fröhlich and K. Gaw~dzki, "Conformal field theory and geometry of
strings", CRM Proceedings and Lecture Notes 7, CRM, Montreal, 1994; pp. 57-
97.
[192) S. A. Fulling, Aspects o(Quantum Field Theory in Curved Space-Time, Cam-
bridge Univ. Press, Cambridge, 1989.
[193) S. A. Fulling and R. A. Gustafson, "Some properties of Riesz means and
spectral expansions", Electron. j. Diff. Eqns. 1999:6 (1999), 1-39; physics/
9710006.
[194) I.M. Gelfand and M. A. Naimark, "On the embedding of normed rings into
the ring of operators in Hilbert space", Mat. Sbornik 12 (1943), 197-213.
[195] M. Gerstenhaber and S. D. Schack, "Algebras, bialgebras, quantum groups
and algebraic deformations", in Deformation Theory and Quantum Groups
with Application to Mathematical Physics, J. Stasheff and M. Gerstenhaber,
eds., Contemp. Math. 134, (1992), 51-92.
[196) E. Getzler and A. Szenes, "On the Chern character of a theta-summable Fred-
holm module", j. Func. Anal. 84 (1989), 343-357.
[197] P. B. Gilkey, Invariance Theory, the Heat Equation, and the Atiyah-Singer
Index Theorem, 2nd edition, CRC Press, Boca Raton, FL, 1995.
[198) H. Gillet, "Comparing algebraic and topological K-theory", in Higher Alge-
braic K -theory: an Overview, Lecture Notes in MathemaUes 1491, Springer,
Berlin, 1992; pp. 55-99.
[199) C. Godbillon, Elements de Topologie Algebrique, Hermann, Paris, 1971.
[200) I. C. Gohberg and M. G. Krein, Introduction to the Theory of Linear Nonsel(ad-
joint Operators, Amer. Math. Soc., Providence, RI, 1969.
652 References

[201] I. C. Gohberg and A. S. Markus, "Some relations between eigenvalues and


matrix elements of linear operators", Mat. Sbornik 64 (1964), 481-496.
[202] ]. N. Goldberg, A. ]. Macfarlane, E. T. Newman, F. Rohrlieh and E. C. G. Sudar-
shan, "Spin-s spherical harmonics and ä",]. Math. Phys. 8 (1967), 2155-2161.
[203] ]. Gomis and T. Mehen, "Space-time noncommutative field theories and uni-
tarity", hep-th/0005129, CalTech, Pasadena, CA, 2000.
[204] A. Gonzälez-Arroyo and M. Okawa, "Twisted Eguchi-Kawai model: a reduced
model for large-N lattice gauge theory", Phys. Rev. D 27 (1983), 2397-2411.
[205] ]. M. Gracia-Bondia, "Generalized Moyal quantization on homogeneaus sym-
plectic spaces", in Deformation Theory and Quantum Groups with App/ica-
tion to Mathematical Physics,]. Stasheff and M. Gerstenhaber, eds., Contemp.
Math. 134 (1992), 93-114.
[206] ]. M. Gracia-Bondia, "The phase of the scattering matrix", Phys. Lett. B 482
(2000), 315-322; hep-th/0003141.
[207] ]. M. Gracia-Bondia, B. lochum and T. Schücker, "The Standard Model in non-
commutative geometry and fermion doubling", Phys. Lett. B 416 (1998), 123-
128; hep-th/9709145.
[208] ]. M. Gracia-Bondia and S. Lazzarini, "Connes-Kreimer-Epstein-Glaser renor-
malization", hep-th/0006106, Marseille and Mainz, 2000.
[209] ]. M. Gracia-Bondia and C. P. Martin, "Chiral gauge anomalies on noncom-
mutative I~Y', Phys. Lett. B 479 (2000), 321-328; hep-th/0002171.
[210] ]. M. Gracia-Bondia and ]. C. Värilly, "Algebras of distributions suitable for
phase-space quantum mechanics. I",]. Math. Phys. 29 (1988), 869-879.
[211] ]. M. Gracia-Bondia and]. C. Värilly, "Phase-space representation for Galilean
quantum particles of arbitrary spin",]. Phys. A: Math. Gen. 21 (1988), L879-
L893.
[212] ]. M. Gracia-Bondia and ]. C. Varilly, "The metaplectic representation and
boson fields", preprint CPP/91/21, Austin, Texas, 1991.
[213] ]. M. Gracia-Bondia and ]. C. Varilly, "QED in external fields from the spin
representation", ]. Math. Phys. 35 (1994), 3340-3367.
[214] P. Green, "The local structure of twisted covariance algebras", Acta Math.
140 (1978), 191-250.
[215] W. Greinerand]. Reinhardt, Field Quantization, Springer, Berlin, 1996.
[216] H. Grosse, C. Klimcik and P. Presnajder, "On finite 4D quantum field theory
in noncommutative geometry", Commun. Math. Phys. 180 (1996), 429-438.
[217] H. Grosse, T. Krajewski and R. Wulkenhaar, "Renormalization of noncommu-
tative Yang-Mills theories: a simple example", hep-th/0001182, Wien, 2000.
[218] H. Grosse and E. Langmann, "A super-version of quasi-free second quanti-
zation: I. Charged particles", ]. Math. Phys. 33 (1992), 1032-1046.
[219] R. Grassman and R. G. Larson, "Hopf-algebraic structure offamilies of trees",
]. Algebra 126 (1989), 184-210.
[220] R. Grassman and R. G. Larson, "Hopf-algebraic structure of combinatorial
objects and differential operators", Israel]. Math. 72 (1990), 109-117.
References 653

[221] R. Grossman and R. G. Larson, "Solving nonlinear equations from higher-


arder derivations in linear stages", Adv. Math. 82 (1990), 180-202.
[222] A. Grossmann, "Parity operators and quantization of 8-functions", Commun.
Math. Phys. 48 (1976), 191-193.
[223] A. Grossmann, G. Loupias and E. M. Stein, "An algebra of pseudodifferen-
tial operators and quantum mechanics in phase space", Ann. lnst. Fourier
(Grenoble) 18 (1968), 343-368.
[224] A. Grothendieck, Produits Tensoriels Topologiques et Espaces Nuc/eaires,
Memoirs of the Amer. Math. Soc. 16, Providence, Rl, 1966.
[225] B. Grünbaum and G. C. Shephard, Tilings and Patterns, W. H. Freeman, New
York, 1987.
[226] V. W. Guillemin, "A new proof of Weyl's formula on the asymptotic distribu-
tion of eigenvalues", Adv. Math. 55 (1985), 131-160.
[227] V. W. Guillemin and S. Sternberg, Supersymmetry and Equivariant de Rham
Theory, Springer, Berlin, 1999.
[228] A. A. Hallack, "0 trac;o de Dixmier e o teorema do trac;o de Connes", M. Sc.
thesis, Universidade Federal do Rio de Janeiro, 1998.
[229] P. R. Halmos, A Hilbert Space Problem Book, Springer, Berlin, 1982.
[230] M. Hamermesh, Group Theory, Addison-Wesley, Reading, MA, 1962.
[231] G. H. Hardy, Divergent Series, Clarendon Press, Oxford, 1949.
[232] G. H. Hardy and]. E. Littlewood, "The Riemann zeta function and the theory
of the distribution of primes", Acta Math. 41 (1918), 119-196.
[233] G. H. Hardy and E. M. Wright, An Introduction to the Theory of Numbers,
Clarendon Press, Oxford, 1960.
[234] R. Häußling, "The su(211) model of electroweak interactions and its connec-
tion to noncommutative geometry", Notesofatalk given at the workshop
"The Standard Model of elementary particle physics from a mathematical-
geometrical viewpoint", Hesselberg, March 1999.
[235] R. Häußling, N. A. Papadopoulos and F. Scheck, "su(211) symmetry, algebraic
Superconnections and a generalized theory of electroweak interactions",
Phys. Lett. B 260 (1991), 125-130.
[236] M. Hazewinkel, Formal Groups and Applications, Academic Press, New York,
1978.
[237] M. Hazewinkel, "lntroductory recommendations for the study of Hopf alge-
bras in mathematics and physics", CWI Quarterly 4 (1991), 3-26.
[238] S. Helgason, Differential Geometry, Lie Groups, and Symmetrie Spaces, Aca-
demic Press, New York, 1978.
[239] ]. W. Helton and R. E. Howe, "Integral operators: traces, index, and homol-
ogy", in Proceedings of a Conference on Operator Theory, P. A. Fillmore, ed.,
Lecture Notes in MathemaUes 345, Springer, Berlin, 1973; pp. 141-209.
[240] P. Henrici, "An algebraic proof of the Lagrange-Bürmann formula", j. Math.
Anal. Appl. 8 (1964), 218-224.
654 References

[241) E. Hewitt and K. A. Ross, Abstract Harmonie Analysis I!, Springer, Berlin,
1970.
[242) N. Higson, "Algebraic K-theory of stable C*-algebras", Adv. Math. 67 (1988),
1-140.
[243) N. Higson, "On the K-theoryproof ofthe index theorem", in Index Theoryand
Operator Algebras,]. Fox and P. Haskell, eds., Contemp. Math. 148 (1993),
67-86. .
[244) N. Higson, G. G. Kasparov and ]. D. Trout, "A Bott periodicity theorem for
infinite dimensional Euclidean space", Adv. Math. 135 (1998), 1-40.
[245) P. ]. Hilton and U. Stammbach, A Course in Homological Algebra, Springer,
Berlin, 1971.
[246) M. W. Hirsch, Differential Topology, Springer, Berlin, 1976.
[247) N. Hitchin, "Harmonie Spinors", Adv. Math. 14 (1974), 1-55.
[248) G. Hochschild, La Structure des Groupes de Lie, Dunod, Paris, 1968.
[249) G. Hochschild, B. Kostant and A. Rosenberg, "Differential forms on regular
affine algebras", Trans. Amer. Math. Soc. 102 (1962), 383-408.
[250) G. Hochschild and ].-P. Serre, "Cohomology of Ue algebras", Ann. Math. 57
(1953), 591-603.
[251) R. H0egh-Krohn and T. Skjelbred, "Classification of C*-algebras admitting
ergodie actions of the two-dimensional torus", j. reine angew. Math. 328
(1981), 1-8.
[252) L. Hörmander, "The Weyl calculus of pseudodifferential operators", Com-
mun. Pure Appl. Math. 32 (1979), 359-443.
[253) L. Hörmander, The Analysis of Linear Partial Differential Operators III,
Springer, Berlin, 1986.
[2 54) A. Horn, "On the singular values of a product of completely continuous ope-
rators", Proc. Natl. Acad. Sei. USA 36 (1950), 374-375.
[255) R. Howe, "Quantum mechanics and partial differential equations",]. Funct.
Anal. 38 (1980), 188-254.
[256) W. Hurewicz and H. Wallman, Dimension Theory, Princeton Univ. Press,
Princeton, NJ, 1948.
[257) D. Husemoller, Fibre Bund/es, Springer, Berlin, 1975.
[258) D. Husemoller, Lectures on Cyclic Homology, Tata Institute/Springer, Mum-
bai, 1991.
[259) B. Iochum and T. Schücker, "Yang-Mills-Higgs versus Connes-Lott", Com-
mun. Math. Phys. 178 (1996), 1-26.
[260) C. ltzykson and j.-B. Zuber, Quantum Field Theory, McGraw-Hill, New York,
1980.
[261) K. Jänich, "Vektorraumbündel und der Raum der Fredholm-Operatoren",
Math. Ann. 161 (1965), 129-142.
[262) A. jaffe, A. Lesniewski and K. Osterwalder, "Quantum K-theory: I. The Chern
character", Commun. Math. Phys. 118 (1988), 1-14.
References 655

[263] A. Jaffe, A. Lesniewski and j. Weitsman, "Pfaffians on Hilbert space", J, Funct.


Anal. 83 (1989), 348-363.
[264] J, M. Jauch and F. Rohrlich, The Theory of Photons and Electrons, Springer,
Berlin, 1976.
[265] P. Jordan and E. P. Wigner, "Ober das Faulische Aquivalenzverbot", Z. Phys.
47 (1928), 631-651.
[266] R. V. Kadison and J, R. Ringrose, Fundamentals of the Theory of Operator
Algebras I, Academic Press, Orlando, FL, 1983.
[267] R. V. Kadison and J, R. Ringrose, Fundamentals of the Theory of Operator
Algebras II, Academic Press, Orlando, FL, 1986.
[268] W. Kalau and M. Walze, "Gravity, noncommutative geometry and the Wodz-
icki residue", j. Geom. Phys. 16 (1995), 327-344.
[269] I. Kaplansky, Rings ofOperators, W. A. Benjamin, New York, 1968.
[270] M. Karoubi, "Les isomorphismes de Chem et de Thom-Gysin en K-theorie",
Seminaire Henri Cartan, Fascicule 2, numero 16 (1965).
[271] M. Karoubi, K·Theory: An Introduction, Springer, Berlin, 1978.
[272] M. Karoubi, "Homologie Cyclique et K-theorie", Asterisque 149 (1987), 1-
147.
[273] G. Karrer, "Einführung von Spinoren auf Riemannschen Mannigfaltigkeiten",
Ann. Acad. Sei. Fennicae Ser. AI Math. 336/5 (1963), 3-16.
[274] G. Karrer, "Darstellung von Cliffordbündeln", Ann. Acad. Sei. Fennicae Ser.
AI Math. 521 (1973), 3-34.
[275] G. G. Kasparov, "Topological invariants of elliptic operators. I: K -homology",
Math. USSR Izv. 9 (1975), 751-792.
[276] G. G. Kasparov, "Hilbert C*-modules: theorems of Stillespring and Voicul-
escu", J, Oper. Theory 4 (1980), 133-150.
[277] C. Kassel, "Le residu noncommutatif", Seminaire Bourbaki 708 (1989).
[278] C. Kassel, Quantum Groups, Springer, Berlin, 1995.
[279] D. Kastler, "The C*-algebras of a free boson field", Commun. Math. Phys. 1
(1965), 175-214.
[280] D. Kastler, Cyclic cohomology within the differential envelope, Hermann,
Paris, 1988.
[281] D. Kastler, "The Dirac operator and gravitation", Commun. Math. Phys. 166
(1995), 633-643.
[282] D. Kastler, "Lectures on Clifford bundles and generalized Dirac operators",
Lecture notes, Universidade de Madeira, Funchal, 1995.
[283] D. Kastler and M. Mebkhout, "Revisiting the Mackey-Stone-von Neumann
theorem: the C*-algebra of a presymplectic space", CPT-90/P.2450, CPT,
Luminy, 1990.
[284] T. Kato, Perturbation Theory for Linear Operators, Springer, Berlin, 1976.
656 References

[285) M. Khalkali, "A survey of entire cyclic cohomology", in Cyclic Cohomology


and Noncommutative Geometry, j. ]. R. Cuntz and M. Khalkali, eds., Fields
Institute Communications 17, Amer. Math. Soc., Providence, RI, 1997; pp. 79-
89.
[286) A. A. Kirillov, Elements de la Theorie des Representations, Mir, Moscow, 197 4.
[287) A. W. Knapp, Representation Theory of Semisimple Groups, Princeton Univ.
Press, Princeton, NJ, 1986.
[288) M. Kontsevich, "Deformation quantization of Poisson manifolds I", q-alg/
9709040, IHES, Bures-sur-Yvette, 1997.
[289) M. Kontsevich and A. L. Rosenberg, "Noncommutative smooth spaces", in
The Gelfand Mathematical Seminars, 1997-1999,1. M. Gelfand and V. S. Re-
takh, eds., Birkhäuser, Boston, 1999; math/9812158.
[290) B. Kostant, unpublished. (But see R. A. j. Matthews, Eur. j. Phys. 16 (1995),
172-176.)
[291) G. Kowalewski, Einführung in die Determinantentheorie einschließlich der
Fredholmschen Determinanten, Chelsea, New York, 1948.
[292) T. Krajewski, "Classification of finite spectral triples", j. Geom. Phys. 28
(1998), 1-30; hep-th/9701081.
[293) T. Krajewski, "Geometrie non commutative et interactions fondamentales",
these de doctorat, Universite de Provence, 1999; math-ph/9903047.
[294) T. Krajewski and R. Wulkenhaar, "Perturbative quantum gauge fields on the
noncommutative torus", Int. j. Mod. Phys. A 15 (2000), 1011-1030; hep-
th/9903187.
[295) 0. S. Kravchenko and B. A. Khesin, "Central extension of the Lie algebra of
(pseudo)differential symbols", Func. Anal. Appl. 25 (1991), 83-85.
[296) D. Kreimer, "On the Hopf algebra structure of perturbative quantum field
theories", Adv. Theor. Math. Phys. 2 (1998), 303-334; q-alg/9707029.
[297] D. Kreimer, "On overlapping divergences", Commun. Math. Phys. 204 (1999),
669-689; hep-th/9810022.
[298) D. Kreimer, "Chen's iterated integral represents the operator product expan-
sion", Adv. Theor. Math. Phys. 3 (1999), 3; hep-th/9901099.
[299) N. H. Kuiper, "The homotopy type of the unitary group of Hilbert space",
Topology 3 (1965), 19-30.
[300) j. Kustermans and S. Vaes, "A simple definition for locally compact quantum
groups", C. R. Acad. Sei. Paris 328 (1999), 871-876.
[301) j. Kustermans and S. Vaes, "Locally compact quantum groups", preprint,
Cork and Leuven, 1999; to appear in Ann. Sei. Ec. Norm. Sup.
[302) j.-L. Lagrange, "Nouvelle methode pour resoudre les equations litterales par
la moyen des series", Mem. Acad. Royale des Seiences et Belles-lettres de
Berlin 24 (1770), 2 51-326.
[303) E. C. Lance, Hilbert C* -modules, Cambridge Univ. Press, Cambridge, 1995.
[304) G. Landi, An Introduction to Noncommutative Spaces and their Geometries,
Lecture Notes in Physics: Monographs m51, Springer, Berlin, 1997.
Referenees 657

[305] G. Landi, "Deeonstrueting monopoles and instantons", math-ph/9812004,


Univ. of Trieste, 1998; to appear in Rev. Math. Phys.
[306] G. Landi, "Projeetive modules of finite type and monopoles over § 2 ", math-
ph/9905014, Univ. of Trieste, 1999; to appear in j. Geom. Phys.
[307] G. Landi, F. Lizzi and R. j. Szabo, "From large N matriees to the noneommu-
tative torus", hep-th/9912130, Copenhagen, 1999.
[308] N. P. Landsman, Mathematical Topics between Classical and Quantum Me-
chanics, Springer, Berlin, 1998.
[309] N. P. Landsman, "Lie groupoid C*-algebras and Weyl quantization", Com-
mun. Math. Phys. 206 (1999), 367-381; math-ph/9903039.
[310] S. Lang, Algebra, 3rd edition, Addison-Wesley, Reading, MA, 1993.
[311] E. Langmann, "Coeycles for boson and fermion Bogoliubov transformations",
j. Math. Phys. 35 (1994), 96-112.
[312] E. Langmann, "Noneommutative integration ealculus", j. Math. Phys. 36
(1995), 3822-383 5.
[313] E. Langmann and j. Miekelsson, "Scattering matrix in external field prob-
lems", j. Math. Phys. 37 (1996), 3933-3953.
[314] H. B. Lawson and M. L. Miehelsohn, Spin Geometry, Prineeton Univ. Press,
Princeton, NJ, 1989.
[315] A. Lichnerowiez, "Spineurs harmoniques", C. R. Aead. Sei. Paris 257A (1963),
7-9.
[316] E. H. Lieb and M. Loss, Analysis, Graduate Studies in Mathematies 14, Amer-
iean Mathematieal Soeiety, Providenee, Rl, 1997.
[317] F. Lizzi, G. Mangano, G. Miele and G. Sparano, "Constraints on unified gauge
theories from noneommutative geometry", Mod. Phys. Lett. A 11 (1996),
2561-2572.
[318] F. Lizzi and R. j. Szabo, "Duality symmetries and noneommutative geo-
metry of string spacetimes", Commun. Math. Phys. 197 (1998), 667-712;
hep-th/9707202.
[319] j.-L. Loday, Cyclic Homology, Springer, Berlin, 1992.
[320] j.-L. Loday and D. Quillen, "Homologie eyclique et homologie de l'algebre
des matrices", C. R. Aead. Sei. Paris 296 (1983), 295-297.
[321] j.-L. Loday and D. Quillen, "Cyclie homology and the Lie algebra homology
of matriees", Comment. Math. Helv. 59 (1984), 565-591.
[322] V. I. Ma~aev, "A dass of eompletely eontinuous operators", Sov. Math. Dokl.
2 (1961), 972-975.
[323] j. Madore, An Introduction to Noncommutative Differential Geometry and its
Physical Applications, Cambridge Univ. Press, Cambridge, 1999.
[324] S. Majid, Foundations of Quantum Group Theory, Cambridge Univ. Press,
Cambridge, 1995.
[325] Yu. I. Manin, "Algebraie aspeets of nonlinear differential equations", j. Sov.
Math. 11 (1979), 1-22.
658 References

[326) Yu. I. Manin, Quantum Groups and Noncommutative Geometry, Publ. Centre
de Recherehes Mathematiques 1561 Universite de Montreal, Montreal, 1988.
[327] ]. Marion and K. Valavane, "Good spectral triples, associated Lie groups
of Campbell-Baker-Hausdorff type and unimodularity", math-ph/9903037,
CPT, Luminy, 1999.
[328) R. E. Marshak, Conceptual Foundations of Modem Partide Physics, World Sci-
entific, Singapore, 1993.
[329) C. P. Martin, ]. M. Gracia-Bondia and ]. C. Värilly, "The Standard Model as
a noncommutative geometry: the low energy regime", Phys. Reports 294
(1998), 363-406.
[330) C. P. Martin and D. Sänchez-Ruiz, "The one-loop UV divergent structure of
U(l) Yang-Mills theory on Noncommutative 11~:!'', Phys. Rev. Lett. 83 (1999),
476-479; hep-th/9903077.
[331) V. Mathai and D. Quillen, "Superconnections, Thom classes and equivariant
differential forms", Topology 25 (1986), 85-110.
[332) T. Matsui, "The index of scattering operators of Dirac equations", Commun.
Math. Phys. 110 (1987), 553-571.
[333) T. Matsui, "The index of scattering operators of Dirac equations.II", J. Funct.
Anal. 94 (1990), 93-109.
[334) P.]. .r1cCann, "Geometry and the integerquantumHall effect", in Geometrie
Analysis and Lie Theory in Mathematics and Physics, A. L. Carey and M. K.
Murray, eds., Cambridge Univ. Press, Cambridge, 1997.
[335) R. Meyer, "Analytic cyclic cohomology", Ph. D. thesis, Universität Münster,
1999; math/9906205.
[336) ]. Mickelsson, Current A/gebras and Groups, Plenum, New York, 1989.
[337) ]. A. Mignaco, C. Sigaud, A. R. da Silva and F.]. Vanhecke, "The Connes-Lott
program on the sphere", Rev. Math. Phys. 9 (1997), 689-718.
[338) ]. A. Mignaco, C. Sigaud, A. R. da Silva and F. ]. Vanhecke, "Connes-Lott model
building on the two-sphere", hep-th/9904171, UFRJ, Rio de Janeiro, 1999, to
appear in Rev. Math. Phys.
[339) ]. W. Milnor, "On spaces having the homotopy type of a CW-complex", Trans.
Amer. Math. Soc. 90 (1959), 272-280.
[340) ]. W. Milnor and J. C. Moore, "On the structure of Hopf algebras", Ann. Math.
81 (1965), 211-264.
[341) ]. W. Milnor and]. D. Stasheff, Characteristic Classes, Princeton Univ. Press,
Princeton, NJ, 1974.
[342] ]. A. Mingo, "K-theory and multipliers of stable C*-algebras", Trans. Amer.
Math. Soc. 299 (1987), 397-411.
[343) ]. A. Mingo and W. ]. Phillips, "Equivariant triviality theorems for Hilbert
C*-modules", Proc. Amer. Math. Soc. 91 (1984), 225-230.
[344) S. Minwalla, M. V. Raamsdonk and N. Seiberg, "Noncommutative perturbative
dynamics", hep-th/9912072, Princeton, NJ, 1999.
References 659

[345) A. S. Mishchenko, "Banach algebras, pseudodifferential operators, and their


application in K-theory", Russ. Math. Surveys 34 (1979), 77-91.
[346) A. S. Mischenko and A. T. Fomenko, "The index of elliptic operators over
C*-algebras", Math. USSR Izv. 15 (1980), 87-112.
[347] S. Montgomery, Hopf Algebrasand their Actions on Rings, CBMS Regional
Conference Series in MathemaUes 82, American Mathematical Society, Prov-
idence, Rl, 1993.
[348) K. Morita, "Duality for modules and its applications to the theory of rings
with minimum condition", Sei. Rep. Tokyo Kyoiku Daigaku A 6 (1958), 83-
142.
[349) H. Moscovici, "Cyclic cohomology and local index computations", in the Pro-
ceedings of the EMS Summer School on Noncommutative Geometry and Ap-
plications, Monsaraz and Lisboa, September 1997; P. Almeida, ed., to appear.
[350) H. Moscovici and F. Wu, "Index theory without symbols", in C* -Algebras:
1943-1993: A Fifty Year Celebration, R. S. Doran, ed., Contemp. Math. 167
(1994), 305-351.
[351) ]. E. Moyal, "Quantum mechanics as a statistical theory", Proc. Cambridge
Philos. Soc. 45 (1949), 99-124.
[352) G.]. Murphy, C* -algebras and Operator Theory, Academic Press, San Diego,
CA, 1990.
[353) S. B. Myers and N. Steenrod, "The group of isometries of a Riemannian ma-
nifold", Ann. Math. 40 (1939), 400-416.
[354) M. S. Narasimhan and S. Ramanan, "Existence of universal connections",
Amer. J. Math. 83 (1961), 563-572.
[355) T. Natsume and R. Nest, "Topological approach to quantum surfaces", Com-
mun. Math. Phys. 202 (1999), 65-87.
[356) Y. Ne'eman, "Irreducible gauge theory of a consolidated Salam-Weinberg
model", Phys. Lett. B 81 (1979), 190-194.
[357) R. Nest, "Cyclic cohomology of crossed products with 71.", J. Funct. Anal. 80
(1988), 235-283.
[358) M. Neuman, "Fredholm structures in positron theory", Phys. Rev. 83 (1951),
1258.
[359) ]. von Neumann, "Die Eindeutigkeit der Schrödingerschen Operatoren",
Math. Ann. 104 (1931), 570-578.
[360) ]. von Neumann, "Charakterisierung des Spektrums eines lntegralopera-
tors", Actualites Sei. Ind. 229 (1935), 38-55.
[361) E. T. Newman and R. Penrose, "Note on the Bondi-Metzner-Sachs group",].
Math. Phys. 7 (1966), 863-870.
[362) ]. T. Ottesen, Infinite Dimensional Groups and Algebras in Quantum Physics,
Lecture Notes in Physics: Monographs m27, Springer, Berlin, 1995.
[363) F. Panaite, "Relating the Connes-Kreimer and Grossman-Larson Hopf alge-
bras built on rooted trees", math/0003074, IMRA, Bucharest, 2000.
660 References

[364] M. Paschke, "Ober nichtkommutative Geometrien, ihre Symmetrien und et-


was Hochenergiephysik", Ph. D. thesis, Universität Mainz, 2000.
[365] M. Paschke and A. Sitarz, "Discrete spectral triples and their symmetries", J.
Math. Phys. 39 (1998), 6191-6205.
[366] G. K. Pedersen, C* -algebras and their automorphism groups, Academic Press,
London, 1979.
[367] G. K. Pedersen, Analysis Now, Springer, Berlin, 1989.
[368] R. Penrose and W. Rindler, Spinors and Spacetime I: Two-spinor Calcu/us,
Relativistic Fields, Cambridge Univ. Press, Cambridge, 1987.
[369] D. Perrot, "BRS cohomology and the Chern character in noncommutative
geometry", Lett. Math. Phys. 50 (1999), 135-144; math-ph/9910044.
[370] D. Perrot, "A Riemann-Roch theorem for one-dimensional complex group-
oids", math-ph/0001040, CPT, Luminy, 2000.
[371] M. E. Peskin and D. V. Schroeder, An Introduction to Quantum Field Theory,
Addison-Wesley, Reading, MA, 1995.
[372] P. Petersen, Riemannian Geometry, Springer, Berlin, 1998.
[373] M. V. Pimsner and D. Voiculescu, "Exact sequences for K-groups and Ext-
groups of certain cross-product C*-algebras", J Oper. Theory 4 (1980), 93-
118.
[374] M. V. Pimsner and D. Voiculescu, "Imbedding the irrational rotation C*-
algebra into an AF-algebra", J. Oper. Theory 4 (1980), 201-210.
[375] R. J Plymen, "Spinors in Hilbert space", Math. Proc. Camb. Phil. Soc. 80
(1976), 337-347.
[376] R. J Plymen, "The Weyl bundle", J Funct. Anal. 49 (1982), 186-197.
[377] R. J Plymen, "Strong Morita equivalence, spinors and symplectic spinors", ].
Oper. Theory 16 (1986), 305-324.
[378] R. J Plymen and P. L. Robinson, Spinors in Hilben Space, Cambridge Univ.
Press, Cambridge, 1994.
[379] S. C. Power, Hanke/operators on Hilben space, Longman Scientific & Techni-
cal, Harlow, Essex, UK, 1982.
[380] A. Pressley and G. B. Segal, Loop Groups, Clarendon Press, Oxford, 1986.
[381] M. Puschnigg, "A survey of asymptotic cyclic cohomology", in Cyclic Coho-
mology and Noncommutative Geometry, J J R. Cuntz and M. Khalkali, eds.,
Fields Institute Communications 17, Amer. Math. Soc., Providence, RI, 1997;
pp. 155-168.
[382] A. 0. Radul, "Ue algebras of differential operators, their central extensions
and W-algebras", Func. Anal. Appl. 25 (1991), 25-39.
[383] M. Reed and B. Simon, Methods of Modem Mathematical Physics, I: Functional
Analysis, Academic Press, New York, 1972.
[384] M. Reed and B. Simon, Methods of Modem Mathematical Physics, II: Fourier
Analysis, Se/( Adjointness, Academic Press, New York, 1975.
References 661

[385] M. Reeder, "On the cohomology of compact Lie groups", Enseign. Math. 41
(1995), 181-200.
[386] A. Rennie, "Commutative geometries are spin manifolds", math/9903021
(v2), Adelaide, 1999.
[387] M. A. Rieffel, "Induced representations of C* -algebras", Adv. Math. 13 (197 4),
176-257.
[388] M. A. Rieffel, "Strong Morita equivalence of certain transformation group
C*-algebras", Math. Arm. 222 (1976), 7-22.
[389] M. A. Rieffel, "Unitary representations of group extensions: an algebraic ap-
proach to the theory of Mackey and Blattner", Adv. Math. Suppl. Studies 4
(1979), 43-82.
[390] M. A. Rieffel, "C*-algebras associated with irrational rotations", Pac.]. Math.
93 (1981), 415-429.
[391] M. A. Rieffel, "Morita equivalence for operator algebras", Proc. Symp. Pure
Math. 38 (1982), 285-298.
[392] M. A. Rieffel, "Applications of strong Morita equivalence to transformation
group C*-algebras", Proc. Symp. Pure Math. 38 (1982), 299-310.
[393] M. A. Rieffel, "Dimension and stable rank in the K-theory of C*-algebras",
Proc. London Math. Soc. 46 (1983), 301-333.
[394] M. A. Rieffel, "Projective modules over higher-dimensional noncommutative
tori", Can. J. Math. 40 (1988), 257-338.
[395] M. A. Rieffel, Deformation Quantization (or Actions of IRd, Memoirs of the
Amer. Math. Soc. 506, Providence, Rl, 1993.
[396] M. A. Rieffel, "On the operator algebra for the spacetime uncertainty re-
lations", in OperatorAlgebrasand Quantum Field Theory, S. Doplicher, R.
Longo, ]. E. Roberts and L. Zsid6, eds., International Press, Cambridge, MA,
1997; pp. 374-382; funct-an/9701011.
[397] M. A. Rieffel, "Metrics on states from actions of compact groups", Doc. Math.
3 (1998), 215-229; math/9807084.
[398] M. A. Rieffel, "Metrics on state spaces", Doc. Math. 4 (1999), 559-600;
math/9906151.
[399] M. A. Rieffel and A. Schwarz, "Morita equivalence of multidimensional non-
commutative tori", Int. J. Math. 10 (1999), 289-299; math/9803057.
[400] P. L. Robinson and J. H. Rawnsley, The metaplectic representation, Mpc struc-
tures and geometric quantization, Amer. Math. Soc., Providence, Rl, 1989.
[401] ]. Roe, Elliptic operators, topology and asymptotic methods, Pitman Research
Notes in MathemaUes 179, Longman Scientific & Technical, Harlow, Essex,
UK, 1988.
[402] ]. Rosenberg, "The role of K-theory in noncommutative algebraic topology",
in Operator Algebras and K-Theory, R. G. Douglas and C. Schochet, eds.,
Contemp. Math. 10 (1982), 155-182.
[403] ]. Rosenberg, Algebraic K -theory and its Applications, Springer, Berlin, 1994.
662 References

[404) j. Rosenberg, "The algebraic K-theory of operator algebras", K-Theory 12


(1997), 75-99.
[405) A. Royer, "Wigner function as the expectation value of a parity operator",
Phys. Rev. A 15 (1977), 449-450.
[406) W. Rudin, Real and Complex Analysis, McGraw-Hill, New York, 1966.
[407] W. Rudin, Functional Analysis, McGraw-Hill, New York, 1973.
[408) S. N. M. Ruijsenaars, "Charged particles in extemal fields. I. Classical theory",
j. Math. Phys. 18 (1977), 720-737.
[409) S. N. M. Ruijsenaars, "Charged particles in extemal fields. II. The quantized
Dirac and Klein-Gordon theories", Conunun. Math. Phys. 52 (1977), 267-294.
[410) L. H. Ryder, Quantum Field Theory, 2nd edition, Cambridge Univ. Press, Cam-
bridge, 1996.
[411) A. Salam and P. T. Matthews, "Fredholm theory of scattering in a given time
dependent field", Phys. Rev. 90 (1953), 690-695.
[412) H. H. Schaefer, Topological Vector Spaces, Macmillan, New York, 1966.
[413) G. Scharf, Finite Quantum Electrodynamics: The Causa[ Approach, 2nd edi-
tion, Springer, Berlin, 1995.
[414) G. Scharfand H. P. Seipp, "Charged vacuum, spontaneous positron produc-
tion and all that", Phys. Lett. B 108 (1982), 196-198.
[415) R. Schatten, Norm Ideals of Completely Continuous Operators, Springer,
Berlin, 1960.
[416) F. Scheck, Mechanics, Springer, Berlin, 1994.
[417) R. Schelp, "Fermion masses in nonconunutative geometry", hep-th/9905047,
Austin, TX, 1999.
[418) C. Schochet, "Topological methods for C*-algebras, III: Axiomatic homol-
ogy", Pac. j. Math. 114 (1984), 399-445.
[419) E. Schrödinger, "Eigenschwingungen des sphärischen Raumes", Conunenta-
tiones Pontificia Academia Scientiarum 2 (1938), 321-364.
[420) E. Schrohe, "Nonconunutative residues, Dixmier's trace, and heat trace ex-
pansions on manifolds with boundary", in Geometrie Aspects of Partial Dif-
ferential Equations, B. Booss-Bavnbek and K. Wojciechowski, eds., Contemp.
Math. 242 (1999), 161-186; math/9911053.
[421) E. Schrohe, M. Walze and j.-M. Warzecha, "Construction de triplets spectraux
a partir de modules de Fredholm", C. R. Acad. Sei. Paris 326 (1998), 1195-
1199; math/9805063.
[422) T. Schücker and j.-M. Zylinski, ''Connes' model building kit", j. Geom. Phys.
16 (1995), 207-236.
[423) R. T. Seeley, "Complex powers of an elliptic operator", Proc. Symp. Pure Math.
10 (1967), 288-307.
[424) G. B. Segal, "Equivariant K-theory", Publ. Math. IHES 34 (1968), 129-151.
[425) G. B. Segal, "Unitary representation of some infinite dimensional groups",
Conunun. Math. Phys. 80 (1981), 301-342.
References 663

[426] I. E. Segal, "lrreducible representaUons of operator algebras", Bull. Amer.


Math. Soc. 53 (1947), 73-88.
[427] I. E. Segal, "A nonconunutative extension of abstract integration", Ann. Math.
57 (1953), 401-457.
[428] I. E. Segal, Review of "Noncommutative Geometry", Bull. Amer. Math. Soc. 33
(1996), 459-466.
[429] N. Seiberg and E. Witten, "Electromagnetic duality, monopole condensation,
and confinement in N = 2 Supersymmetrie QCD", Nucl. Phys. B 426 (1994),
19-52.
[430] N. Seiberg and E. Witten, "String theory and nonconunutative geometry", j.
High Energy Phys. 9 (1999), 032; hep-th/9908142.
[431] H. P. Seipp, "On the S-operator for the external field problern in QED", Helv.
Phys. Acta 55 (1982), 1-28.
[432] J.-P. Serre, "Modules projecUfs et espaces fibres ä fibre vectorielle", Seminaire
Dubreil 23, 1957-58.
[433] D. Shale, "Unear synunetries of free Boson fields", Trans. Amer. Math. Soc.
103 (1962), 149-167.
[434] D. Shale and W. F. Stinespring, "Spinor representations of infinite orthogonal
groups", J. Math. Mech. 14 (1965), 315-322.
[435] S. Shnider and S. Sternberg, Quantum Groups: (rom Coalgebras to Drinfeld
Algebras, a Guided Tour, International Press, Boston, 1993.
[436] W. Siegel, Fields, hep-th/9912205, Stony Brook, 1999.
[437] B. Simon, "Notes on infinite determinants of Hilbert space operators", Ad-
vances in MathemaUes 24 (1977), 244-273.
[438] B. Simon, Trace Idealsand their Applications, Cambridge Univ. Press, Cam-
bridge, 1979.
[439] B. Simon, Representations of Finite and Compact Groups, Graduate Studies
in MathemaUes 10, American MathemaUcal Society, Providence, Rl, 1996.
[440) G. Skandalis, "Kasparov's bivariant K-theory and applications", Expo. Math.
9 (1991), 193-250.
[441] R. M. Solovay, "A model of set theory in which every set of the reals is
Lebesgue measurable", Ann. Math. 92 (1970), 1-56.
[442) E. H. Spanier, Algebraic Topology, Springer, Berlin, 1966.
[443) j. Stalker, Complex Analysis, Birkhäuser, Boston, 1998.
[444] A. A. Suslin and M. Wodzicki, "Excision in algebraic K-theory", Ann. Math.
136 (1992), 51-122.
[445) R. G. Swan, "Vector bundles and projecUve modules", Trans. Amer. Math.
Soc. 105 (1962), 264-277.
[446] M. E. Sweedler, Hopf algebras, Benjamin, New York, 1969.
[447] j. L. Taylor, "Banach algebras and topology", in Algebras in Analysis, J. H.
Williamson, ed., Springer, Berlin, 1975.
664 References

[448) J. L. Taylor, "Topological invariants of the maximal ideal space of a Banach


algebra", Adv. Math. 19 (1976), 149-206.
[449) M. E. Taylor, Pseudodifferential Operators, Princeton Univ. Press, Princeton,
NJ, 1981.
[450) M. E. Taylor, Partial Differential Equations, 3 volumes, Springer, Berlin, 1996.
[451) N. Teleman, "From index theory to noncommutative geometry", in Quan·
turn Symmetries: Lectures of the LXIV Ecole d'Ete de Physique Theorique, Les
Hauches 1995, A. Connes, K. Gawedzki and J. Zinn-Justin, eds. (Elsevier, Am-
sterdam, 1998), pp. 787-843.
[452) N. Teleman, "Microlocalisation de l'homologie de Hochschild", C. R. Acad.
Sei. Paris 326 (1998), 1261-1264.
[453) B. Thaller, The Dirac Equation, Springer, Berlin, 1992.
[454) E. G. F. Thomas, "Characterization of a manifold by the *-algebra of its
C"" -functions", Mathematics Institute, University of Groningen, unpublished,
1997.
[455] R. Ticciati, Quantum field theory for mathematicians, Cambridge Univ. Press,
Cambridge, 1999.
[456] J. Tomiyama, Invitation to C* -algebras and Topological Dynamics, World Sci-
entific, Singapore, 1987.
[457] A. Trautman, "Spin structures on hypersurfaces and the spectrum of the
Dirac operator on spheres", in Spinors, Twistors, Clifford Algebrasand Quan-
tum Deformations, Z. Oziewicz et al, eds., (Kluwer, Dordrecht, 1993), p. 25.
[458] F. Treves, lntroduction to Pseudodifferential and Fourier Integral Operators
I, Plenum Press, New York, 1980.
[459) E. V. Troitskii, "Contractibility of the full general linear group of the C*-
Hilbert module -l' 2 (A)", Func. Anal. Appl. 20 (1986), 301-307.
[460) J. D. Trout, "Asymptotic morphisms and elliptic operatorsover C*-algebras",
K-Theory 18 (1999), 277-315; math/9906098.
[461) B. L. Tsygan, "The homology of matrix Ue algebras over rings and the
Hochschild homology", Russ. Math. Surveys 38 (1983), 198-199.
[462) W. J. Ugalde, "Operadores de Dirac en fibrados de base esferica", M. Sc. thesis,
University of Costa Rica, 1996.
[463) S. Vaes and A. Van Daele, "Hopf C*-algebras", math/9907030, Leuven, 1999;
to appear in Proc. London Math. Soc.
[464) F. J. Vanhecke, "On the product of real spectral triples", Lett. Math. Phys. 50
(1999), 157-162; math-ph/9902029.
[465) J. C. Värilly, "An introduction to noncommutative geometry", in the Pro-
ceedings of the EMS Summer School on Noncommutative Geometry and Ap-
plications, Monsaraz and Lisboa, September 1997; P. Almeida, ed., to appear;
physics/9709045.
[466) J. C. Värilly and J. M. Gracia-Bondia, "Algebras of distributions suitable for
phase-space quantum mechanics. II. Topologies on the Moyal algebra", J.
Math. Phys. 29 (1988), 880-887.
References 665

[467] J. C. Värilly and j. M. Gracia-Bondia, "The Moyal representation for spin",


Ann. Phys. (NY) 190 (1989), 107-148.
[468] J. C. Värilly and J. M. Gracia-Bondia, "S-matrix from the metaplectic repre-
sentation", Mod. Phys. Lett. A 7 (1992), 659-667.
[469] J. C. Värilly and J. M. Gracia-Bondia, ''Connes' noncommutative differential
geometry and the Standard Model", J. Geom. Phys. 12 (1993), 223-301.
[4 70] j. C. Värilly and J. M. Gracia-Bondia, "On the ultraviolet behaviour of quantum
fields over noncommutative manifolds", Int. J. Mod. Phys. A 14 (1999), 1305-
1323; hep-th/9804001.
[471] J. C. Värilly, J. M. Gracia-Bondia and W. Schempp, "The Moyal representation
of quantum mechanics and special function theory", Adv. Appl. Math. 11
(1990), 225-250.
[472] M. Vergne, "Representations of Lie groups and the orbit method", in Emmy
Noether in Bryn Mawr, B. Srinivasan and j. Sally, eds., Springer, Berlin, 1983;
pp. 59-101.
[473] M. Vergne, "Geometrie quantization and equivariant cohomology", in ECM:
Proceedings of the First European Congress of Mathematics, A. Joseph et al,
eds., Progress in Mathematics 119, Birkhäuser, Boston, 1994; pp. 249-295.
[474] D. Voiculescu, "Some results on norm-ideal perturbations of Hilbert-space
operators", J. Oper. Theory 2 (1979), 3-37.
[475] D. Voiculescu, "On the existence of quasicentral approximate units relative
to normed ideals 1", J. Funct. Anal. 91 (1990), 1-36.
[476] A. Voros, "An algebra of pseudodifferential operators and the asymptotics
of quantum mechanics", J. Funct. Anal. 29 (1978), 104-132.
[477] S. Wang, "Quantum symmetry groups of finite spaces", Commun. Math. Phys.
195 (1998), 195-211.
[478] F. Warner, Foundations of Differentiable Manifolds and Lie Groups, Scott,
Foresman, Glenview, IL, 1971.
[479] j.-M. Warzecha, "Von Fredholmmoduln zu Spektralen Tripein in Nicht-
kommutativer Geometrie", Ph. D. thesis, Universität Mainz, 1997; Shaker,
Aachen, 1998.
[480] N. Weaver, "Lipschitz algebras and derivations of von Neumann algebras",
J. Funct. Anal. 139 (1996), 261-300.
[481] N. E. Wegge-Olsen, K -theory and C* -algebras -a friendly approach, Oxford
Univ. Press, Oxford, 1993.
[482] C. A. Weibel, An Introduction to Homological Algebra, Cambridge Univ. Press,
Cambridge, 1994.
[483] S. Weinberg, The Quantum Theory of Fields I, Cambridge Univ. Press, Cam-
bridge, 1996.
[484] A. Weinstein, "Symplectic groupoids, geometric quantization and irrational
rotation algebras", in Seminaire sud-rhodanien aBerkeley (1989), P. Dazord
and A. Weinstein, eds., Springer, Berlin, 1991; pp. 281-290.
666 Referenees

(485] A. Weinstein, "Groupoids: unifying internaland external symmetry", Notiees


Amer. Math. Soe. 43 (1996), 744-752.
[486] N. ]. Wildberger, "Finite eommutative hypergroups and applieations from
group theory to eonformal field theory", Contemp. Math. 183 (1995), 413-
434.
(487] ]. 0. Winnberg, "Superfields as an extension of the spin representation of
the orthogonal group",]. Math. Phys. 18 (1977), 625-628.
[488] E. Witten, "Monopoles and four-manifolds", Math. Res. Lett. 1 (1994), 769-
796.
[489] E. Witten, "Noneommutative taehyons and string field theory", hep-th/
0006071, CalTeeh, Pasadena, CA, 2000.
[490] M. Wodzieki, "Loeal invariants of speetral asymmetry", Invent. Math. 75
(1984), 143-178.
[491] M. Wodzieki, "Noneommutative residue. Chapter 1: Fundamentals", in K-
theory, Arithmetic and Geometry, Yu. I. Manin, ed., Leerure Notes in Mathe-
maUes 1289 (Springer, Berlin, 1987), pp. 320-399.
[492] M. Wodzieki, "Cyclie homology of differential operators", Duke Math. ]. 54
(1987), 641-647.
[493] M. Wodzieki, "Cyclie homology ofpseudodifferential operators and noneom-
mutative Euler dass", C. R. Aead. Sei. Paris 306 (1989), 321-325.
[494] ]. Wolf, "Essential selfadjointness for the Dirae operator and its square",
Indiana Univ. Math. ]. 22 (1972), 611-640.
[495] S. L. Woronowiez, "Compaet quantum groups", in Quantum Symmetries, A.
Connes, K. Gaw~dski and ]. Zinn-Justin, eds. (Les Houehes, Session LXIV,
1995), Elsevier Seienee, Amsterdam, 1998; pp. 845-884.
[496] R. Wulkenhaar, "On the Connes-Moseoviei Hopf algebra assoeiated to the
diffeomorphisms of a manifold", math-ph/9904009, CPT, Luminy, 1999.
[497] ]. Xia, "Geometrie invariants of the quantumHall effeet", Commun. Math.
Phys. 119 (1988), 29-50.
[498] F.]. Ynduräin, Relativistic Quantum Mechanics and Introduction to Field The-
ory, Springer, Berlin, 1996.
[499] K. Yosida, Functional Analysis, Springer, Berlin, 1971.
[500] W. Zimmermann, "Convergenee of Bogoliubov's method of renormalization
in momentum spaee", Commun. Math. Phys. 15 (1969), 208-234.
Symbol Index

f, 258, 326 A 0B, 33


J", 203 A ®B, 33
f,297,492, 546 A ®h B, 33,46
*,423 A ®,., B, 33, 113
#,210,272,6 02 A ® B, 189, 210
II · llp. 312, 595 a x 11 b, 116
111·111. 65, 161, 313 [A,B]+. 17
( · I · ), 65, 159, 252, 423, 551, 553 [A,B], 111, 130
(· 1·), 66,423,500 3tv. 466
{·I·}, 159,189,500 ,551,553 (A, G, a), 523
«·I·>>. 185 A ~,. G, 524
A#H,47
10}, 122, 242 (.JI., 3{, D), 401, 450, 481
(Oin I Oour}, 245, 580, 585 (.J1.,3f,D,C ,x),401,405, 483,485
(.JI., 3{, F), 327
A+, 4, 136
AI, 16
Ax, 18,27,91,13 4,469
a 1 (v), aj(v), 187
At, 393
.Jtk(M), 253, 325, 371,423
A±±•238,246
.J1. k (M, E), 253
,!., 574
.JI.±(M), 427
(Ao,At •... ,An}v, 451
An, nA, 66
.JI.'' .JI."' 483, 542
.JI., 322 An,347, 363,442
.J\. ,61,353,481 ,488
0 A~. 189
g_, 541 A~, 526, 533
a(A, B), 223 As. 85, 112
I.JI.I"(V), 257 atsa, aTa, atTat, 201, 220
668 Symbol Index

B, 365,431,436,439,447,637 Cf,432
B0 ,364,430,498 er (.Jl, .J\ * ). 432
b,348,349,430,637 ICP 1 ' 75
b', 345, 430 l(poo. 56, 75
B~,2oo C;'(U), 298
Boo, 112 Curr1 (V), 235
Be·. 434, 637 C(X), 4
b(cp), bt(cp), 241, 576 C0 (Y), 4
BK,204
B,~. 432 D, 387, 545
Be, 531,556 w. 387,406,411,487,506,562
ßY, 13 w~. 397
dt, 424
C, 189,380,399,481,487,505,581 4,321
c, 387 6,35, 193,259,438,599,602
c( · ), 172, 371, 502, 503 6+, 6-, 193, 567
1(,4 8, 130,150,322,466,482,636
( 00 , 75 21,613
{, {x, 19 a. 19
c,488,490,493, 546 ~. 559
Coo. 314 da,320
CB, 23 w(A). 512
Cb(Y), 13 6(AJ,69
cc·, 432 a"'a, 118
Cc(G-A), 523 6actv, 6reto 566
Cf,9 ac, 355, 364
C'P,355,499 6v, 6F, 6Jp, 6s, 566-567
c(:y), 3 72 Der(.Jl, T), 320, 350
C*(G), 525 Der,(H), 624
C*(G,a), 526,533 dg, 388,492,505
ch,339,340,447,514 6H, 261,425
x. 174,212,330,378,397,488 dH(.\.), 273
c(h, k), 236 dH(X, X;.\.), 278
X(M), 361, 428 8i. 437,637
Ch"(D), 451 Diff+(§ 1 ), 240,631
~n (D, V), 454 divX, 260
ch~(D), 457 DJ, 262,298
O(M), 370 81, ~1 • 540-541
o<+)(M), 373 'Dk(M), 362
CL(n), 190 6k(M), 357
Clp,qo 175, 192, 399 dA(A), 196
l(l(V), 173, 214 6Ls.260,425
o~(v), 174 8(.\.- H), 274
Cl(V,g), 173 v!nn. 418
C""(M), 9, 12,136,324,357,370,442 6n, 438
CM(T), 609,617 DomH, 273, 393
Cn (.Jl), 345 Dom""(H), 276
C" (.Jl, .J\ * ), 430 'D' (~). 275
C" (.Jl, T), 350 6 5 ,394,411,413,511
Symbol Index 669

vg. 298 y, 179,192,215


d((jl), dt((jl), 241, 577 [(E), [(M,E), 56, 370
Oz, Bz, 413 f 0 (E), 59
f""(E), f""(M,E), 60,339, 370
e, 574 >'<P•639
'E,65,83 G(H), 40, 624
'E, 160, 338 (J(H), 43

'E<' 73, 75 ria• 382
E, 35 Bii• grs, 252
E(·), 172,371,423 ri~· 256, 409
Ea,385,409 yi,yj. 191,333,384,40 1,507,545
'EA, 65 GL(Jf), 143
'E ® .Jl., 357 GLoo(.J\.),94, 131,343
Eq,('E), 60, 73, 95 y~', 562

'E ®A J', 81, 159 GR,617


E(g, u), 114 grad, 252, SOS
EH(i\.), 273 Gr(CN), 53, 64
'E(m), 77, 419 f(U,E), 52, 57
EndA ('E), 70, 83
End~('E), 71,84,89, 147,161,164 H, 67,374
End~0 ('E), 71, 89 H•, 598,624
EndA(J{A), 91 .1{±,238,397
E(T), 601, 613 .1{"",467,482,4 89, 500
'E' (U), 298 H 0 , 576
Ex, 5, 10, 63 17.122,606
expy,360 11, 115
J{A, 71, 147
F,326,446, 545,595 J{ ®A, 67
P,631 HcM. 612, 634
J', 309, 331 HC"(.Jl.), 432,436,444
IF,34,96 HE 0 (.Jl.), HE 1 (.Jl.), 449
<I>, <1>', 314-315 HJ(B*), 600
cf>',61 HeL. 614
cpc, 362 HHn(.Jl.), 345
cp~. 470,493 HH"(.Jl.), 349,430,436
f*E, 52 HfR(M), 442
IFG,40 H~R (M), 342, 426
J'J(V), 186, 215 11KK', 205, 424
[cph, 432,442 H 1 (M,7L), 19
FredA, 156 H2(M, ll.), 20, 377
FredA ('E, }'), 146, 149 H 2(M,ll.2), 378, 506
Fred(Jf), 142 H 3 (M,7l.), 20, 375
FredX, 157 H" (.Jl., .Jl. * ), 350
fr,208,217 Hn (.Jl., 'E), 345
F~,553 HomA ('E,J'), 70, 79
Horn~ ('E, }'), 71
g, 172,252,503 Hom±(V, W), 211
y, 5,10,485,487 HP 0 (.Jl.), HP 1 (.Jl.), 444, 540
g,37,420 HR, 602
670 Symbol Index

H 5 , 274, 281 L(Jf), 8, 33, 310


J{To 214, 541 L 11 (u,v,w), 116
Jf(x), 582 limn-wo 290
L2·'(M), 423
/,365,436 A~,558
/±±•245,248, 576-578 L 2(M,S), 389
L(·), 172,371,423,454 LP,285, 311,328,482
LD, 320 LP+,316,481
idA, 11 LP-,316,464,496
indexF, 142, 145 L(p,q>, 317
Lx, 253, 360 ArA, 196
NV, 172, 196
],43, 184,573 Lx,253,361
Ja, 542
JJ(V), 226, 234 m,34, 320
Jw, 186 M4, 558
:M, :ML, :MR, 117
)(,24,27,33,84,85,99,310 J1, 192,235
Je-, 316,317 {1, 195,235,383
K, 191,399 M(A), 5
~.p,t,f, 559 :M(A), 14, 84
K 0 (A), 95, 138, 152, 157,485, 556 Map+ (X, Y), 11
K1(A), 128,480,485,556 Mcp, 9
K2(A), 125 (M,g), 252
Ko(.Jl), K 1(.Jl), 400, 485 Mn(A), 66, 81, 93, 162
Je® A, 85 Moo(.Jl), 94
K~ 1 g(.Jl), 94, 101, 344 M~, 189
Kf1g(A), 131 Mpc(V), 182
Koc/J,95,97 Mrt(B,A), 376
KG(M), 422 M(S, T), 613
JC(H), 374
kj,496 N, 365, 430, 609
Ko(M), K0 (M), 100-101 N, 242
k~,496 V,253, 335,387,419
K", 338 v•, 426
JC(IR"), JC'(IR"), 117, 275 v,490, 500
K~0 P(A), 92 [n], 438
K~0 P(A), 128 v:E, 385
VB,254,262, 382
L, 599,603 v9 ,258,423, 507
L 1 , 285, 311 lv9 1, 258
L 1 +,286,316,317,398 NH (.\), 272
L~+,288,316 N(p), 509
A,439,637 (v, S), 378, 504
.\,350,365,430 (v,S,C),381,505
AA, 196 Vx, 253, 382
.f(y), 388
L1(G-A), 523 n, 76,186,208,215,242
L2(G-Jf), 525 n·, 326
Symbol Index 671

0 15'\, 320, 335 resA, 267


n· 5'\, 322 ResH, 282
n!bA, 321 Rcf>, 60
n;bA, 325, 346 RjU,307
<9c(l!1l."), <'JM(IJ1l."), 117 R(G), 422
w([), 627 n(G), 34, 42, 623
OJ(V),225,2 36,327 Rr, 628
Oj(V), 236 Rr, 628
OJ(V), 221, 223 Rj.332
ok. 101 Rjl. 256
On, 119, 269 R,,(T), 273
Opi), 467 RljJn• 194
Op(p), 299
Q(u), 113 S,436,440, 541,606
O(V,g), 180 5,256,270, 395,414,511
5,5',430
P, 558 s±±•239, 576
rro, 137, 157 5,372,487
rr', 481 s<,378
P(A 5 ), 86, 91 S±, 397
P(B), 38, 619 S,246,247, 576
PB. 77, 122 §1' 240, 390, 631
Pc(T), 602 § 2 , 75,292,408, 515
7TD,484, 547 a,35, 173,267,636
1Tcf>, 31, 541 {1, 560

PfB,204 ~A. ~4J. 24


PuJM,21 SB, Sz, 607
Ph.qh,225 Sc!. 572
P"h 1 , 228 S(D), 492
Pic(A), 74, 376 Sö,636
Pin(V), 225 at(D), 502
P1, 185 Sd(U), 298
1Tj, 187,215 SF, 579
P(M), 267, 349 S(G), 549
1Tn(X, *), 17, 137 Uj, 215
p $ q, 91 Uj, 437,637
5k(T),248,2 84,311,464
Q, 173, 238, 396 Sk(V), 200,217
Q, 242 a;.,(T), 285
Qn. 115 ~M.iM,24,26, 108
Q.(J-f), 133, 141 S:n. 575
Qoo (5\), 94, 343 §", 109,121,270 ,292
Un(T),285,3 12,464
R,339,384 SOJ(V), 234
n. 264, 266, 361 SO(V), 180
nt. 496 so(V), 182, 195
p, 506 a(P), 301
r(a), 7, 111, 137 Spin(n), Spinc(n), 381
Rc(T), 602 Spin(V), 181
672 Symbol Index

Spinc(V), 181, 192, 198 V*, 598


Sret• 575 V(A), 91, 95
S(IR"), 116, 135 vc,61
S'(IR"), 116, 298, 306 Vectr(M), 50, 56
CT(r,s), 525, 533 VJ, 184
S · T, 613 V(T), 601, 613
Str, 202, 212, 330 V(t), 576, 586
SU(2), 193,416
SUJ(V), 199 WJ, 185
S(V), 38, 620 w(M), 421
S(iE" ), 528, 535 Wz(M), 381
WresA, 267, 282, 506
ITI, 310 wresx A, 267, 507
T!, 605
T[ ... ], 572 (X,*), 17
11'2 ,334,398, 527 Jt(M), 136, 325, 337
T, 179,438,483,529,535,637,638 Jt(M, IR), 337
T,Ch~(D), 457 X /Y, 11
Text[ ... ], Tvac[ ... ], 582 [X, Y]+, 17
TP,445,460,470,493 (xy), 558
0,254,337, 526,533 X v Y, X A Y, 23
eh. 174,214,232
(9 1 , ... ,9"), 382,423 y+,4,23
9"',385 <tJII, 3 78
O(t), 566 'Yd (M), 266, 299
T(kl, 468 ylo;,., 415
TJ.(T), 286 tJIL, tJIR, 561
11'",291, 536 'I'co (M), 'I'ci (M), 266
11'~.529, 535
'I'-co (M), 266
Tot• BC, 435, 637 tJIN, tJis, 409
Tot• CC, 432, 435 'I'(x), \f(x), 583
Tr,285,312, 329 'I'(tJI), 'I't(tJI), 249
Tr', 329, 471
Tr+,288, 546 Zq,,23
(H(s), 281
Trw,288,464,465,470,493
T(V), 37
zeR(M), 442
zr.432
U(oo), 121, 225 Zr,625
'U(A), 91
{U 01 }, 29, 65, 155
'U(g), 37,42,139,618,623
U(J-f), 143
Uj, 80
Ul (J-f), 240, 244
UJ(V), 185,198,226,234
UK, 196
u(p), 564
ur, 534
U(s, t), 586
Subject Index

(1, 1)-periodicity, 177,402 anticommutator, 173, 328


4-vector, 558 antiderivation, 211
antipode,39,606,634
.J\-bimodule, 81, 320 twisted, 636
absolute value of an operator, 310 antiskew operator, 200, 217, 239
A-compact operator, 71, 84, 89, 146 antiunitary operator, 189, 399, 481,
A-compact projector, 86 505
action functional, 270, 492, 507 approximate unit, 29, 65, 497
acyclic complex, 344, 353, 358, 433 arc length, 390
adjoint, 14, 70, 393 asymptotic expansion, 118, 275
adjointable operator, 70, 84 asymptotic morphism, 110, 130
AF-algebra, 96, 129, 532, 556 Moyal, 120, 122
affine connection, 254 attachment, 21, 109
A-finite rank module, 91 augmentation, 153
A-finite rank operator, 71 augmented algebra, 136
A-Fredholm operator, 146, 164 automatic continuity, 12
Alexander-Spanier cohomology, 325 automorphism, 15, 523
algebra of classical symbols, 26 7, 349 Bogoliubov, 174,181,214
algebra of rooted trees, 602, 614 automorphism group, 11
algebraic K-theory, 94, 131
algebraic tensor product, 32 Banach algebra, 4, 27, 135, 448, 469
.J\-linear map, 79 involutive, 6, 27, 138, 523
amplitude, 298 Banach-Lie group, 221, 225
annihilation operator, 122, 187, 427 Banach module, 69
anomalous commutator, 222 Banach space, 312
anticommutation relation, 173 Banach-space interpolation, 317, 46 7
canonical, 241 bar resolution, 353, 362, 433
674 Subject Index

Berezin integral, 203, 479 C*-bundle, 158


Berezin quantization, 201 C*-cross-norm, 24
Berezin-Patodi formula, 203 C*-dynamical system, 523
bialgebra, 36, 599 Cech cocycle, 19, 50, 374
Bianchi identity, 256, 339 Cech cohomology, 18, 50, 325, 378
bicommutant, 483, 542 cell complex, 108, 420, 539
bimodule, 81, 599 central extension, 181, 194,223,236
bivector, 182 Cesara mean, 286, 474
Bochner-Weitzenböck formula, 396 Cesaraorder at infinity, 275
Bockstein homomorphism, 19 chain homotopy, 344, 431
Bogoliubov automorphism, 174, 181, chain map, 344, 362
214,232,378,488 character, 5, 617, 630, 636
Borel functional calculus, 129 of a cycle, 350, 352
Borel-Weil construction, 420 characteristic map, 639
Bott element, 122, 133, 189 charge, 238
Bott map, 124 quantized, 242
Bott periodicity, 189, 192 charge anomaly, 243
Bott projector, 77, 122, 342, 514 charge conjugation, 175, 181, 191,399
bra operator, 378 charged field, 238, 328, 576
branch of a rooted tree, 601 charged vacuum, 244
Bratteli diagram, 96 Chern character, 339, 340, 350, 441,
Brauwer degree, 77 445,447,450,470,493, 514,
bubble diagram, 585, 592 595
bundle equivalence, 50, 53, 55 Chern isomorphism, 343, 485
bundle map, 50, 370 children of a tree vertex, 601
bundle morphism, 52, 505 chiral fermions, 562
Butcher group, 617 chirality element, 179, 562
Christoffel symbol, 256, 382, 409
Calder6n's formula, 144, 243 isotropic, 410
C*-algebra, 4, 27, 523 C*-inductive Iimit, 96, 99, 532
contractible, 16, 23 classical dimension, 481, 488
elementary, 167 classical symbol, 301
nonunital, 14, 97, 128 Clifford algebra, 172, 214, 333, 504,
nuclear, 33 559
simple, 214, 529, 538 Clifford bundle, 3 70, 504
stable, 85 Clifford connection, 385, 512
O"-unital, 15, 29, 148, 157 Clifford group, 190
transformation group, 524, 550 Clifford module, 371
Calkin algebra, 133, 141 irreducible, 3 72
cancellation semigroup, 95 closed operator, 393
canonical anticommutation relations, closed range, 141, 146
241 C*-module, 65, 83, 159, 189
Cartan identity, 253, 265, 361 full, 65, 159
Cartan subalgebra, 420 selfdual, 73
cartesian product, 23 C*-norm, 4, 24, 65, 84, 524
Cauchy integral formula, 134 coadjoint orbit, 114, 421
Cauchy principal-value integral, 330 coalgebra, 35, 598
Cayley's counting formula, 608, 615 coassociativity, 35
C* -bimodule, 160 coboundary,223, 599
Subject Index 675

cocommutativity, 3 5 Hermitian, 338, 383


cocompactsubgroup, 550 Levi-Civita, 254, 337, 382
cocycle, 236, 525 spin, 383
codifferential, 424 tensor product, 336, 385
cofunctor, 9, 53 torsion-free, 254, 426
full, 10 universal, 336
halfexact, 103, 109 connection 1-form, 262, 632, 635
cograph, 628 connection Laplacian, 259, 262, 394
cohomology of a bicomplex, 433 Connes bicomplex, 434, 637
coincidence limit of a kernel, 277 Connes boundarymap, 431,452,498
collapsing map, 11, 17 Connes' character formula, 479,493,
colurnn vectors, 80 498
commutant, 483, 542 Connes' distance formula, 390, SOS
commutationrelation, 198,221,249, Connes-Moscovici algebra, 612
383, 526, 533 Connes-Moscovici weight, 609, 617
commutative superalgebra, 356 continuedfractionexpansion, 531,556
commutator, 173, 269, 625 continuous field
commutator ideal, 344 locally trivial, 68
compact group, 44, 54, 623 of C*-algebras, 69, 167
compact Hausdorff space, 4 of elementary C* -algebras, 3 73
compact üe group, 416, 420 of Hilbert spaces, 67, 373
compact operator, 85, 310, 373 continuous linear functional, 29
compact pair, 11, 21, 103 contractible C* -algebra, 2 3
compact perturbation, 141, 150, 155 contractible chain complex, 344,433
compact resolvent, 273 contractible space, 17, 53, 143, 156
compact space, 4, 15 contracting homotopy, 344, 361, 433
compactification, 13, 15 contraction, 172,253, 371
one-point, 4, 11 contragredient operator, 228
Stone-Cech, 13, 15 convolution, 38, 523, 630
complemented submodule, 70, 86 coordinate functions, 494
complete symbol, 507 coproduct,35, 599,604,612,613
complex projective space, 75 corner of a C* -algebra, 165
complex structure, 184, 226, 238, 240, correlation of operators, 450
327, 570, 573 cotangent bundle, 264, 305, 349
complexification, 61, 173 counit, 35, 599, 602
conditional expectation, 536 counterterm map, 628
conditional trace, 329 counting function, 272, 316
cone algebra, 23, 26 covariant representation, 524, 527,632
cone of a morphism, 2 5 Cramer's rule, 205, 206, 260
conjugate space, 160, 380 creation operator, 187, 427
conjugation operator, 381, 403, 413, cross-norm, 24, 32
487,505,581 crossed product, 167, 524, 529, 550,
connected bialgebra, 618 631
connectinghomomorphism, 105,126, iterated, 534, 556
130,150,436 current, 242
connection, 56, 253, 335 current group, 235
affine, 254 curvature, 338
Clifford, 385, 512 Riemannian, 255, 339
dual, 426 scalar, 256, 270, 395, 414, 511
676 Subject Index

cutoff, 307, 471 Dirac geometry, 487, 545, 548


CW-complex, 23, 108, 137, 343 Dirac operator, 261, 387, 406, 411,
cycle over an algebra, 326, 350 487, 506, 545, 559
cyclic category, 439, 637 equivariant, 417
cyclic cocycle, 350, 352,432,498, 540 generalized, 387, 512
cyclic cohomology, 344, 432 spectrum, 419
analytic, 446 Dirac spinors, 397
asymptotic, 446 dispersion relation, 589
entire, 449 distance, 272, 390, 391, 492
of Hopf algebras, 63 7 geodesic, 388
periodic, 444 distribution, 9, 275, 298, 360
cyclic permuter, 350, 365, 430, 438, operator-valued, 583
638 periodic, 275
cyclic skewsymmetrizer, 365, 430 divergence formula, 392
cyclic vector, 31, 216, 541 divergence of a vector field, 259
Dixmier ideal, 286, 316, 317
de Rham boundary, 355, 364, 365, Dixmier trace, 251,288,464,465,470,
442 549
de Rham cohomology, 325, 340, 426, Dixmier trace dass, 398
490 Dixmier-Douady dass, 373
de Rham complex, 325, 355,423,442 domain of an operator, 393
de Rham current, 355, 362 dual A-module, 57, 73, 75, 482
dosed, 499 dual Banach space, 315
de Rham homology, 442, 444 dual of a Hopf algebra, 598, 623
deformation of C* -algebras, 69, 111 Duhamel equation, 455
deformation retract, 89, 91, 137 Dunford integral, 28, 134, 281
degeneracy operators, 430 , Dyson expansion, 5 71
dense ideal, 27 quantum, 586
density, 257
dequantization, 114, 123 effective action, 592
derivation, 211, 320, 325, 382, 446, Eigenschaften algebra, 520
543,624 eigenspinor, 396, 414
inner, 320, 544 Eilenberg-MacLane space, 75
symmetric, 540 Einstein-Hilbert action, 270,492, 507,
determinant, 131, 217, 580 512
diagonal submanifold, 357 Eisenstein series, 335, 549
diagram, 627 elementary C*-algebra, 167, 373
dictionary of spaces and algebras, 15 elementary partides, 519
diffeomorphism group, 12, 240, 631 elliptic curve, 398, 549
differential, 322 elliptic operator, 263, 276, 327, 400
differential form, 325, 371 differential, 392, 425, 488
differential operator, 136, 298 pseudodifferential, 158,293,303
elliptic, 427, 504 endomorphism algebra, 83
first-order, 502 entire cochain, 449, 454
dimension spectrum, 480 entire cydic cohomology, 449
8-invariant functional, 636, 637 enveloping C*-algebra, 524
Dirac K-cyde, 479 Epstein-Glaser renormalization, 592,
Dirac adjoint, 583 630
Dirac equation, 559, 573 equivalence bimodule, 162, 375, 377
Subject Index 677

equivalence of categories, 11 Fourier transform, 6, 271, 309, 331


equivalent projectors, 91 frame, 90
essential extension, 84 frame bundle, 631
essential ideal, 13-15, 84 Frechet algebra, 9, 135, 357,469,483
essentially selfadjoint operator, 393 Frechetspace, 12,467,535
E-theory, 110, 130 Fredholm index, 142, 397
Euler angles, 417 Fredholm module, 326, 400, 444, 446,
Euler characteristic, 361, 428 462,493
Euler vector field, 264, 266, 361 Fredholm operator, 142, 226,446,480,
evaluation map, 5, 10, 43, 45, 63 485
exact couple, 441 free Hamiltonian, 5 76
exact sequence free module, 79
augmentation, 22 full C*-module, 65
of vector bundles, 51 full corner, 165
split, 25, 51, 58 functional calculus, 28, 29, 274, 450
excision, 103, 446 functor, 9, 95
extended orthogonal group, 236 continuous, 98
extension of C*-algebras, 22 contravariant, 9
extension of scalars, 60 exact, 58
exterior algebra, 172, 184 faithful, 58
exterior bundle, 3 70 full, 58
externallines, 627 halfexact, 98, 105
homotopy-invariant, 98, 105
faithful state, 30, 529, 535 K-theory, 98, 128
fermion field, 249 normalized, 98
Feynman graph, 627 stable, 98
Feynman rules, 5 78 fundamental1-form, 254, 337
Feynman slash notation, 559 fundamental dass, 499
filtration, 618
finite-rank operator, 142, 284, 463, gamma matrices, 545, 562
471,496 gauge fixing, 385, 409, 411
finite-rank projector, 143 gauge potential, 386
finitely generated module, 79 Gaussian, 208, 217
finitely generated projective module, Gaussian elimination, 135
59, 74,101,164,482,489, Gelfand spectrum, 5
500,554,556 Gelfand topology, 5, 9
first order condition, 489 Gelfand transformation, 5, 10
first-order differential calculus, 321 Gelfand-Fuchs cohomology, 640
flag manifold, 420 generalized Dirac operator, 420
Fock representation, 187, 215 generalized limit, 288, 290, 464,473,
Fock space, 186, 215, 399 474
bosonic, 201 geometric quantization, 421
polarized, 186 r functor, 57, 59, 370
Fock state, 215 GLS symbol, 117, 275
forest, 628 GNS construction, 30, 167, 214, 541
formal quantum field theory, 582 good locally convex algebra, 135,469
formally selfadjoint operator, 391 graded bialgebra, 38, 620
Fourier kernel, 114, 118 graded differential algebra, 322, 348
Fourier series, 526, 528 universal, 323
678 Subject Index

graded module, 188 Hochschild cohomology, 349, 430, 438


graded tensor product, 210 continuous, 356
gradient, 252, 505 Hochschild cycle, 484,489,493, 514,
grading operator, 188, 212, 330, 372, 515, 547
424,488 Hochschild homology, 345, 447, 513
graph, 627 Hodge star operator, 423
Grassmann variables, 203 Hodge-de Rham Laplacian, 261, 425
Grassmannian, 53, 185 Hodge-Dirac operator, 426
Grossman-Larson Hopf algebra, 614, Hölderinequality, 289,312,316,453,
618,626 458,465,500
Grossmann-Royer reflection operators, holomorphic functional calculus, 28,
115 134
Grothendieck group, 92 stability und er, 134
ground state, 122 homeomorphism, 10, 15
group action, 138, 523, 631 homogeneaus distribution, 271, 306
ergodic, 537 homogeneaus function, 265, 266, 306
group algebra, 40 homology module, 105, 344
group C*-algebra, 525 homology theory, 105
twisted, 526, 533 *-homomorphism, 10, 12, 136
group-like element, 40, 46, 624, 636 homotopic morphisms, 16
homotopic projectors, 91
Haagerup tensor product, 33, 46, 166 homotopy dass, 17, 20, 111
Haar functional, 43, 46 homotopyequivalence, 16, 17, 53,137
Haar measure, 43, 418 homotopy group, 77, 121
halfexact cofunctor, 103, 109 homotopy invariance, 341
halfexact functor, 98, 105 Hopf algebra, 34, 39, 601
Hamiltonian density, 582 Hopf C*-algebra, 46
harmonic form, 425, 428 Hopf module algebra, 4 7
harmonic oscillator, 122 Horn's inequality, 289
heat kernel, 280, 284 hyperbolic automorphism, 533
expansion, 283 hyperplane bundle, 78, 408
Heaviside function, 275, 566 hypertrace, 465, 498
Heisenberg equation, 198
Heisenberg group, 67 idealizer, 14
Hermitian connection, 338, 383 idempotent, 18, 81, 87, 93, 137, 343
Hermitian metric, 65 implementation problem, 224, 573,
Hermitian pairing, 65, 159, 370, 371 594
hexagon, 126, 343 implementor, 191, 562
Hilbert algebra, 179 index, 142,145,149,227,397
Hilbert module, 65 of an odd operator, 397
Hilbert space, 30, 65 index formula, 243
virtual, 422 index map, 130, 144, 154, 234
Hilbert transform, 330 index pairing, 446, 480, 485, 547
Hilbert-Schmidt operator, 217, 225, inductive Iimit, 56, 117, 3 58
240, 248, 311' 328, 495, 5 73 infinitesimal element, 141
Hochschild boundary, 348 infinitesimal operator, 284
Hochschild coboundary, 349, 430, 599 infinitesimal spin representation, 195,
truncated, 430 221,383
Hochschild cocycle, 349, 470, 476 inner automorphism, 215
Subject Index 679

integral, 326, 333 scalar, 398, 511


of a function, 297 spinor, 398, 511
interaction Hamiltonian, 576 leaf of a rooted tree, 601
interaction representation, 5 71 Lebesgue measure dass, 494
internallines, 627 Leibniz rule, 47, 253, 320, 321, 335,
interpolation inequality, 317, 464 382,383,385,502,632
invariance under Hopf action, 636 length of a curve, 388
invertible element, 18, 93, 134, 135, length of a tree vertex, 601
147,482 Levi-Civita connection, 254, 337, 382,
involution, 27, 175, 338, 483 409,427
irrational rotation algebra, 523 Lichnerowicz formula, 395,398,413,
irreducible complexity, 126 419, 511
irreducible representation, 31, 184, Lie algebra, 38, 139, 182
215,418 restricted orthogonal, 221
isotropic subspace, 185 Lie algebra cohomology, 223
Lie derivative, 253
Jacobi identity, 211, 223, 484 Lie group, 121, 138, 420
JLO cocycle, 451 Lie superalgebra, 211
joint spectrum, 494 line bundle, 20, 75
Jordan-Pauli function, 567 canonical, 421
Hopf, 121
K 0 cofunctor, 101 tautological, 78, 121, 408
K 0 functor, 165
Liouville measure, 113
Kähler differentials, 322 Lipschitz function, 389
Kaplansky's formula, 88, 122, 147,482 local frame, 50, 55, 631
K -cycle, 400
local orthonormal basis of 1-forms,
unbounded,401 382,406,423
kernel of a pseudodifferential opera-
local section, 50
tor, 271, 277, 299
local uniform closure, 67
ketbra, 71, 374
locality of Hochschild homology, 3 57
K-homology, 400,446,485
locally compact space, 4
Killing form, 420
locally convex topology, 134, 356,469
K -orientation, 3 79
logarithmic element, 18, 128, 183
KR-homology, 400, 483
long exact sequence, 19, 104,378,436
KRi-cycle, 483, 489
Lorentz product, 5 58
reduced,403,405, 505
Lorentz transformations, 558
unreduced,401
Kronecker flow, 523, 553
Kronecker foliation algebra, 553 Majorana field, 244
K-theory functor, 98, 128 Majorana spinor, 407
K-theory of tori, 133, 556 manifold, 251, 370, 487
mapping cone, 23
L 2 -spinors, 389 unreduced, 26
Laplace expansion, 219 mapping cylinder, 23, 25
Laplace transform, 273 matrix algebra, 32, 93, 503
Laplace-Beltrami operator, 259, 260, maximal C* cross-norm, 33, 113
425 maximal ideal, 7
Laplacian, 259, 272, 291 Mayer-Vietoris sequence, 127
generalized, 263 Ma<;aev ideal, 316, 317
680 Subject Index

measurable operator, 288, 293, 489, noncommutative integral, 285, 297,


493, 545 398,399,493,546,548
Mellin transform, 283 noncommutative pullback bundle, 62
metaplectic representation, 116, 182 noncommutative residue, 267
metaplectic structure, 382 noncommutative space, 8
metrizable space, 15 noncommutative spin geometry, 485,
minimal coupling, 263 492, 545
Minkowski space, 558 noncommutative torus, 102, 529, 535
mod-2 reduction, 378 noncommutative Yang-Mills theories,
modular pair, 46, 636, 638 593
module morphism, 61 nonunital C*-algebra, 14, 97, 128
Möbius transformation, 417 normal coordinates, 257, 361
moment asymptotic expansion, 118, normally ordered product, 230
275 n-simplex, 438
moment of a distribution, 118, 284 n-skeleton, 109
Moritaequivalence, 162,164,166,177, nuclear C*-algebra, 33
375,482,491, 502, 504, 549, number operator, 122, 197, 509
553 quantized,242
Morita invariance, 103, 165, 556
morphism, 10, 15 one-point compactification, 4, 11
bialgebra, 40 operator ideal, 311, 313
bundle, 52 operator module, 166
module, 61 opposite algebra, 61, 175, 353, 481,
of C*-algebras, 10, 29, 112, 136 488, 542
of C*-modules, 70 ordered group, 96
of Hopf algebras, 39 orientable manifold, 371, 499
restriction, 12, 63 orientation dass, 422
unital, 29 orientation cycle, 488, 493
orthogonal complex structure, 184
VVoronowicz, 15,46
orthogonal group, 180
Moyal algebra, 117, 539
orthogonal projector, 54
Moyal asymptotic morphism, 120, 122
orthonormal basis, 54
Moyal product, 116, 539, 593
oriented, 179, 188
Moyal pseudodifferential calculus, 115
out-vacuum vector, 228, 233, 242
Moyal quantization, 67, 113, 120,201
Moyal quantizer, 113, 115 pairing, 65, 159, 423, 448, 471, 500,
Moyal quantum mechanics, 111, 123 551,553,622
Mpc structure, 382 paracompact space, 50
multiplication map, 34, 46, 124, 320 parametrix, 303, 327, 393
multiplication operator, 14, 328, 495 parity, 210
matrix-valued, 501 Parseval-Plancherel formula, 418
multiplicity function, 495 partial trace, 514
multiplier algebra, 14, 15, 84, 117 partition of unity, 20, 50
musical isomorphisms, 252 path component, 128, 137
Paulimatrices, 76,333,411,560,595
natural growth operator, 609 PBVV basis for 'U(g), 37, 621
natural transformation, 10, 105 Penrose tiling, 96
neutrino field, 561 periodic cyclic cohomology, 444, 540
neutrino paradigm, 562 periodicity operator, 353, 440, 541
Subject Index 681

Pfaffian, 204, 217 Dyson, 567


phase of the scattering matrix, 585, Feynman, 566, 579
586 quantum, 586
phase operator, 446, 545 retarded, 566, 5 75
phase space, 113 unitary, 571
Picard group, 74, 3 76 proper ideal, 27
Pimsner-Voiculescu embedding, 531 proper map, 10, 15
Pimsner-Voiculescu hexagon, 556 pseudo-Riemannian manifold, 252
Planck's constant, 111 pseudodifferential operator, 266, 298,
Poincare duality, 361, 426, 485, 490, 327, 331
547 classical, 266
Poincare group, 5 58 elliptic, 506
point at infinity, 4 on a manifold, 304
pointed space, 11, 17, 23, 101 order of, 299
Poisson bracket, 302 properly supported, 300
polar decomposition, 143, 233, 310, pseudoinverse, 146, 149
314 pullback, 21, 64
polarization, 185, 421 pullback bundle, 52
restricted, 226 Puppe sequence, 25
polarized Fock space, 186 of C*-algebras, 104
Pontryagin duality, 46, 599 of spaces, 108
positive element of a C*-algebra, 29 pure state, 8, 30, 391
positive linear functional, 15, 29, 290
positive square root, 29, 73 Q-algebra, 135
pre-C* -algebra, 134, 469, 482, 489, quadratic form, 173, 503
529 positive definite, 179, 510
pre-C*-module, 65, 159 quantization, 113,250,422,538
pre-Fredholm module, 327, 400 quantization map, 173, 396
primitive element, 38, 610, 615, 619 quantum dimension, 594
primitive graph, 628 quantum electrodynamics, 5 73
primitive of a distribution, 275 quantum field theory, 522
principal bundle, 234, 264, 631 quantumHall effect, 521
principal homogeneaus space, 3 77 quantum plane, 36
principal symbol, 301, 393, 422, 502, quite irrational matrix, 537
506 quotient space, 11, 167
principal value distribution, 307
product of spin geometries, 486 range of a projector, 86, 88
projective module, 80, 336, 353 rank of a CW-complex, 109
projective representation, 114, 115, rank of a vector bundle, 50
194,236 rank-one projector, 122
projective resolution, 353, 360 rapidly decreasing sequence, 528, 535
projective tensor product, 33, 357 real structure, 483
projector, 54, 71, 76, 86, 112, 185, realification, 238
529 reduced K0 -ring, 101
minimal, 503 reduced suspension, 109
of rank one, 3 73 reflection, 180
Powers-Rieffel, 529, 547 regular Operator, 146, 151
propagator, 564 regularization of distributions, 306,
advanced, 566 589
682 Subject Index

renormalization, 592, 627 Schwinger term, 222, 223, 298, 351


representation a-compact space, 15
irreducible, 31, 184, 191, 215 second-quantized operators, 196
of SL(2, ([), 560 section, 50
of a C*-algebra, 31, 159, 167 along a map, 62
type of, 189, 407 of a continuous field, 67, 111,
representative function, 34, 42, 623 373
reproducing kerne!, 113 of a vector bundle, 56
reproducing property, 114 smooth, 60, 339
residue calculus, 566 selfadjoint element, 6, 28
resolvent equation, 273 selfadjoint operator, 273
restricted orthogonal group, 225, 236 separating duality, 623
restricted orthogonal Lie algebra, 221 separating vector, 31, 541
restriction of scalars, 60 series reversion, 278, 634
Ricci tensor, 256 sesquilinear form, 65
Riemann curvature tensor, 384 shift operator, 525
Riemann sphere, 75, 292, 408, 515 short exact sequence, 19, 131, 235
Riemann zeta function, 250, 283 of C*-algebras, 21, 98, 105
Riemannian curvature, 255, 256, 339 of groups, 181
Riemannian density, 258, 389, 399, of modules, 80
507 of sheaves, 375
Riemannian manifold, 252 split, 22, 25, 80, 108
Riemannianmetric, 76,252,388,492, shuffle permutation, 190, 205, 424
504 shuffle product, 348, 356
Riesz operator, 332 signature of a quadratic form, 1 74
Riesz theorem, 160, 380 signature operator, 427
rooted tree, 601, 613 similarity of idempotents, 89, 91
rotation, 194, 559 simple cut, 602
rotation group, 180 simplicial category, 438
row vectors, 66, 80 singular value, 217, 248, 284, 311,
Rules, 578 464
Runge-Kutta method, 617 skewadjoint operator, 195, 541
skewgroup, 43
scalar curvature, 256, 270, 283, 395, skewsymmetric matrix, 204
414, 511 skewsymmetrization, 347, 355, 363,
scalar Laplacian, 398, 511 490,499
scalarproduct, 66, 179, 184, 186,423, smash product, 23, 47, 622, 632
500 smooth domain, 276, 468
scattering matrix smooth element, 138
classical, 245, 572 smooth rapidly decreasing functions,
quantum, 245, 247, 576, 582 135
Schatten p-class, 285,311,317, 328, smoothing operator, 266, 299
462,482 Sobolev space, 300, 306, 335, 466
Schur's Iemma, 187, 193 Sobolev's Iemma, 300
Schwartz space, 115, 135, 535, 549 spectral asymmetry, 281
Schwarz inequality, 30, 69, 84, 119, spectral density, 274
389 spectral function, 272
generalized, 537 spectral measure, 495
Schwinger function, 567 spectral projector, 112, 471
Subject Index 683

spectral radius, 6, 7, 28, 111 subgraph, 628


spectral sequence, 441 subpfaffian, 20S, 208
spectral triple, 401, 446, 487 superalgebra, 210
p+-summable, 4SO, 4S7, 46S, 482 supercommutator, 173, 17S, 211, 328
real, 481 superficial degree of divergence, 627
regular, 466, 482, 489 supermechanics, 62
tarne, 46S superrepresentation, 188, 211
e-summable, 450 superspace, 188,210
spectrum, S, 28, 134, 136, 167, 310 supertrace, 202, 212, 330
absolutely continuous, 49S suspension, 24, 108, 109
spherical harmonics, 41 S unreduced, 26
spin connection, 383, 411 symbol, 298
twisted, 42 7 Grossmann-Loupias-Stein, 117,
spin curvature, 384 27S
spin geometry, 48S, 488 of spectral density, 276
0-dimensional, 486 symbol calculus, 269, 301
irreducible, 491 symbol map, 173, 267
spin group, 181 symmetric algebra, 38, 620
spin manifold, 381, 421 symmetric bilinear form, 172, S76
spin representation, 192, 23S, S8S symmetric bimodule, 320
spin structure, 381, 408, 421, 487, symmetric gauge function, 314
492, SOS symmetric norm, 288, 313, 463
spinc group, 181 symmetric operator, 391
spinc manifold, 3 78 symmetrically normed ideal, 288, 313,
spinc structure, 378, S04, S12 496
spinor, 186, 409 symmetry, 86, 314, 326
spinor bundle, 379, SOO via Hopf algebras, 47
spinor harmonics, 408, 41 S symplectic cone, 264
spinor Laplacian, 394, 398, 411, 413, symplectic form, 11 S
419, 511 symplectic group, 116, 182
spinor module, 409, 487 symplectic homogeneaus space, 114
spinor space, 389, 487 symplectic manifold, 264
spinorial clock, 177
stable C*-algebra, 8S tangent algebra, 324
stable homotopy, 121 Tannaka-Krein duality, 46, 617
stable range, 121 tempered distribution, 298, 300
stably equivalent C* -algebras, 8S, 165 tensor algebra, 37
stably equivalent vector bundles, 102 tensor product, 85
stably quasiisomorphic C*-modules, algebraic, 24, 32, 67, 73
149 Haagerup, 33, 46
Standard Model, S19 of C*-algebras, 24
state, 30, 215, 391 of C*-bimodules, 162
Stereographie projection, 7S of C*-modules, 1S9, 166
Stiefel-Whitney dass, 381, 421 of asymptotic morphisms, 113
Stone-Cech compactification, 13, 1 S of connections, 2SS, 336, 38S
cr-trace, 636 of Hilbert spaces, 32, 33
Stratonovich-Weyl quantizer, 114 of modules, S7
subbundle, 51 of vector bundles, 100
subdiagram, 627 projective, 33, 357
684 Subject Index

spatial, 33 Milnor-Moore, 615,621,626


;I2-graded, 189 Mishchenko-Fomenko, 158
theorem Myers-Steenrod, 388
Arens-Royden, 20 Nash embedding, 337
Atiyah-Jänich, 121, 144, 244 Nelson, 216
Atkinson, 141 open mapping, 112, 142
Banach-Alaoglu, 5 Panaite, 626
Bartle-Graves, 112, 120 Peter-Weyl, 45
Bass, 102 Plymen, 375
Borel-Weil, 421 Poincare-Birkhoff-Witt, 37
Bott periodicity, 121, 125 Rellich, 300, 306
Calkin, 313 Riesz, 160, 380
Chern isomorphism, 343, 491 Riesz-Markov, 16
classification, 64 Serre-Swan, 59, 91, 101, 360, 370,
Connes character, 479, 595 500
Connes spin manifold, 492, 513 Shale-Stinespring, 216, 232
Connes trace, 293, 479, 506 Shilov,20
Connes-Langmann, 333,479 spectral, 8, 273
divergence, 259, 392 stable range, 102
Elliott, 96 Stokes,259,326,364,499
Euler, 265 Stone-Weierstrass, 7, 16, 45, 138,
family index, 158 324
Forster, 75 Tannaka-Krein, 624
Fourier integral, 115 Tikhonov, 5, 44
Frobenius-Schur, 191 Toeplitz-Schur, 475
Furry, 581 Tomita, 542
Gelfand-Naimark, 7, 29, 31,288 Voiculescu, 496
Hahn-Banach,8,31 von Neumann, 483, 542
Hardy-Littlewood, 296 Weierstrass, 146
Hochschild-Kostant-Rosenberg, Weyl-von Neumann, 495
356 time-ordered product, 5 72
Hochschild-Kostant-Rosenberg- Tomita conjugation, 542
Connes, 355,363,442,485, topological A-module, 357
499 topological group, 9
Hurewicz, 75 topological K-theory, 93
lkehara-Wiener, 296 topological stable rank, 102
Kasparov absorption, 147 topologically projective module, 3 57
Kastler-Kalau-Walze, 270, 280, torsion group, 339
283, 511 torsion tensor, 254, 337
Kato-Rosenblum, 495 torus, 291, 398, 527
Klee, 12 total complex of a bicomplex, 432
Kuiper, 143, 145 total volume of a geometry, 548
Kuiper-Mingo, 156 trace, 30,179,268,269,285,312,343,
Kuroda,495 349,479,529,538,554,636
Lichnerowicz, 395 traceclass operator, 115,285,311,495
Lidskii, 580 tracial state, 30, 529, 535, 537
Liouville, 28 transgression formula, 455, 459
local index, 480, 635 transition function, 50, 55, 408
Milnor, 137, 156 transpose, 200, 598
Subject Index 685

tree factorial, 605 Voiculescu's modulus, 496


triangle inequality, 69, 390 volume element, 258
trunk of a simple cut, 602 volume form, 342, 423, 489, 500, 515,
Tsygan bicomplex, 432 546
twisting, 3 76 volume of a sphere, 269
of vector bundles, 3 79 von Neumann algebra, 129, 483, 542
type 11 1 factor, 214 von Neumann factor, 214
type of the spin representation, 407
weak* topology, 5, 391
U (1) gauge field, 262 Weyl elements, 534, 535
ultraviolet divergences, 593 Weyl neutrino equations, 561
unbounded K-cycle, 401 Weyl operators, 115, 539
unilateral shift, 158 Weyl spinors, 397
unital ring, 60, 79, 92, 166 Weyl's estimate, 272, 279, 293
unitary element, 6, 28, 180, 536 Whitney product formula, 421
unitary equivalence, 86, 216, 224, 233 Whitney sum, 52
of C*-modules, 85, 87, 147 of spin geometries, 491, 503
of geometries, 486 Wiek rotation, 568
unitary group, 54, 156, 185 Wick-ordered product, 230
unitary operator, 143, 239 Wightman functions, 567
on a C* -module, 85 winding number, 75
restricted, 240 Wodzickiresidue, 251,267,282,479,
unitization, 14, 15 492, 506
universal1-form, 321 density, 267, 280
universal connection, 336
universal enveloping algebra, 37, 42, ~2-grading, 174, 210, 371

139, 618, 623 zeta function, 273, 281


unreduced cone, 26 zeta operator, 274
unreduced suspension, 108, 109 zeta residue, 251, 282
Zimmermann forest formula, 608, 628
vacuum functional, 237
vacuum persistence amplitude, 245,
248, 576, 580
vacuum polarization, 585
vacuum vector, 186, 208, 215, 226,
228,242
vector bundle, 49
dual, 57
equivalence, 50, 55
Euclidean, 370
tautological, 54, 64
trivial, 50, 52, 101
vector field, 136, 325
horizontal, 632
vertical, 632
Virasoro group, 258
virtual bundle, 100, 121
virtual Hilbert space, 142
virtual rank, 101

You might also like